Download as pdf or txt
Download as pdf or txt
You are on page 1of 596

THE ROUTLEDGE INTERNATIONAL

HANDBOOK OF NEUROAESTHETICS

The Routledge International Handbook of Neuroaesthetics is an authoritative reference work that provides the
reader with a wide-ranging introduction to this exciting new scientific discipline. The book brings together
leading international academics to offer a well-balanced overview of this burgeoning field while addressing
two questions central to the field: how the brain computes aesthetic appreciation for sensory objects and how
art is created and experienced.
The editors, Martin Skov and Marcos Nadal, have compiled a neuroscientific, physiological, and
psychological overview of the systems underlying the evaluation of sensory objects and aesthetic appreciation.
Covering a variety of art forms mediated by vision, audition, movement, and language, the handbook puts
forward a critical review of the current research to explain how and why perceptual and emotional processes
are essential for art production. The work also unravels the interaction of art with expectations, experience
and knowledge, and the modulation of artistic appreciation through social and contextual settings, eventually
bringing to light the potential of art to influence mental states, health, and well-being. The concepts are
presented through research on the neural processes enabling artistic creativity, artistic expertise, and the
evolution of symbolic cognition.
This handbook is a compelling read for anyone interested in making a first venture into this exciting
new area of study and is best suited for students and researchers in the fields of neuroaesthetics, perceptual
learning, and cognitive psychology.

Martin Skov is Senior Researcher at Copenhagen Business School and the Danish Research Centre for
Magnetic Resonance. His research focuses on understanding the neurobiological mechanisms of sensory
liking. He has published extensively on neuroaesthetics, including the book Neuroaesthetics (2009) and an
influential series of papers on the conceptual foundations of the field.

Marcos Nadal is Associate Professor at the Department of Psychology of the University of the Balearic
Islands, Spain. His research is devoted to characterizing the psychological, neural, and evolutionary
foundations of aesthetic appreciation. His contributions earned him the Baumgarten Award from the
International Association of Empirical Aesthetics and the Daniel Berlyne Award from the American
Psychological Association Division 10.
THE ROUTLEDGE INTERNATIONAL HANDBOOK SERIES

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF PSYCHOSOCIAL EPIDEMIOLOGY


Edited by Mika Kivimaki, David G. Batty, Ichiro Kawachi, Andrew Steptoe

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF HUMAN AGGRESSION


Current Issues and Perspectives
Edited by Jane L. Ireland, Philip Birch, Carol A. Ireland

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF JUNGIAN FILM STUDIES


Edited by Luke Hockley

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF DISCRIMINATION, PREJUDICE


AND STEREOTYPING
Edited by Cristian Tileagă, Martha Augoustinos and Kevin Durrheim

THE ROUTLEDGE HANDBOOK OF PUBLIC HEALTH AND THE COMMUNITY


Edited by Ben Y. F. Fong and Martin C. S. Wong

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF PERINATAL MENTAL HEALTH


DISORDERS
Edited by Amy Wenzel

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF COMMUNITY PSYCHOLOGY


Facing Global Crises with Hope
Edited by Carolyn Kagan, Jacqui Akhurst, Jaime Alfaro, Rebecca Lawthom, Michael Richards, and Alba Zambrano

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF MORALITY, COGNITION, AND


EMOTION IN CHINA
Edited by Ryan Nichols

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF COMPARATIVE PSYCHOLOGY


Edited by Todd M. Freeberg, Amanda R. Ridley and Patrizia d’Ettorre

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF NEUROAESTHETICS


Edited by Martin Skov and Marcos Nadal

THE ROUTLEDGE INTERNATIONAL HANDBOOK OF PSYCHOANALYSIS AND


PHILOSOPHY
Edited by Aner Govrin and Tair Caspi
THE ROUTLEDGE
INTERNATIONAL HANDBOOK
OF NEUROAESTHETICS

Edited by Martin Skov and Marcos Nadal


Cover image: Based on a drawing by Santiago Ramón y Cajal
First published 2023
by Routledge
4 Park Square, Milton Park, Abingdon, Oxon OX14 4RN
and by Routledge
605 Third Avenue, New York, NY 10158

Routledge is an imprint of the Taylor & Francis Group, an informa business


© 2023 selection and editorial matter, Martin Skov and Marcos Nadal;
individual chapters, the contributors
The right of Martin Skov and Marcos Nadal to be identified as the
authors of the editorial material, and of the authors for their individual
chapters, has been asserted in accordance with sections 77 and 78 of the
Copyright, Designs and Patents Act 1988.
All rights reserved. No part of this book may be reprinted or reproduced or
utilised in any form or by any electronic, mechanical, or other means, now
known or hereafter invented, including photocopying and recording, or in
any information storage or retrieval system, without permission in writing
from the publishers.
Trademark notice: Product or corporate names may be trademarks or
registered trademarks, and are used only for identification and explanation
without intent to infringe.
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library
Library of Congress Cataloging-in-Publication Data
Names: Skov, Martin, 1968– editor. | Nadal, Marcos, editor.
Title: The Routledge international handbook of
neuroaesthetics / edited by
Martin Skov and Marcos Nadal.
Description: Milton Park, Abingdon, Oxon ;
New York, NY : Routledge, 2023. | Series: Routledge international
handbooks | Includes bibliographical references and index.
Identifiers: LCCN 2022015583 (print) |
LCCN 2022015584 (ebook) |
ISBN 9781032348803 (paperback) | ISBN 9780367442743 (hardback) |
ISBN 9781003008675 (ebook)
Subjects: LCSH: Aesthetics—Psychological aspects. |
Neuropsychology. | Visual perception.
Classification: LCC BH301.P45 R68 2023 (print) |
LCC BH301.P45 (ebook) |
DDC 111/.85—dc23/eng/20220630
LC record available at https://lccn.loc.gov/2022015583
LC ebook record available at https://lccn.loc.gov/2022015584
ISBN: 978-0-367-44274-3 (hbk)
ISBN: 978-1-032-34880-3 (pbk)
ISBN: 978-1-003-00867-5 (ebk)
DOI: 10.4324/9781003008675
Typeset in Bembo
by Apex CoVantage, LLC
CONTENTS

List of contributors viii


List of figures xiv
List of tables xvii
Preface xviii

1 Neuroaesthetics as a scientific discipline: an intellectual history 1


Martin Skov

PART I
Aesthetic liking 29

2 Sensory liking: how nervous systems assign hedonic value to sensory objects 31
Martin Skov

3 The neurobiology of liking 63


Eloise Stark, Kent C. Berridge and Morten L. Kringelbach

4 Disliking: from adaptive disgust to ugliness 71


Christoph Klebl, Michael Donner and Indra Bishnoi

5 The influence of interoceptive signals on the processing of external sensory stimuli 89


Alejandro Galvez-Pol, Enric Munar and James M. Kilner

6 Neural correlates of visual aesthetic appeal 103


Edward A. Vessel, Tomohiro Ishizu and Giacomo Bignardi

v
Contents

  7 Auditory pleasure elicited by music 134


Ernest Mas-Herrero

  8 Odour aesthetics: hedonic perception of olfactory stimuli 148


Gulce Nazli Dikecligil and Jay A. Gottfried

  9 Movement appreciation 172


Kohinoor M. Darda, Ionela Bara and Emily S. Cross

10 How architectural design influences emotions, physiology, and behavior 194


Alex Coburn, Adam Weinberger and Anjan Chatterjee

11 Sexual selection, aesthetic appreciation and mate choice 218


Michael J. Ryan

12 Aesthetic sensitivity: origin and development 240


Ana Clemente

13 The evolution of sensory valuation systems 254


Esther Ureña and Marcos Nadal

PART II
Art293

14 Perception and cognition in visual art experience 295


Rebecca Chamberlain

15 The music system 315


Amy M. Belfi and Psyche Loui

16 Watching and engaging in dance 337


Beatriz Calvo-Merino

17 Making sense of space: the neuroaesthetics of architecture 346


Zakaria Djebbara, Lars Brorson Fich and Giovanni Vecchiato

18 Literature and poetry 366


Arthur M. Jacobs

19 Narrative 379
Franziska Hartung

vi
Contents

20 Music-evoked emotions: their contribution to aesthetic experiences, health, and


well-being 397
Liila Taruffi and Stefan Koelsch

21 The health benefits of art experience 410


Claire Howlin

22 Experiencing art in museums 423


Aniko Illes and Pablo P. L. Tinio

23 Context and complexity of aesthetic experiences: a neuroscientific view 438


Julia Crone and Helmut Leder

24 Experiencing art in social settings 448


Haeeun Lee and Guido Orgs

25 Top-down processes in art experience 461


Aenne A. Brielmann

26 Preferences need inferences: learning, valuation, and curiosity in aesthetic


experience 475
Sander Van de Cruys, Jo Bervoets and Agnes Moors

27 Neuroscience of artistic creativity 507


Oshin Vartanian

28 Expertise and the brain of the performing artist 526


Fredrik Ullén

29 The emergence of symbolic cognition 539


Francesco d’Errico and Ivan Colagè

30 Neuropsychology of art and aesthetics 555


Alejandro Dorado and Marcos Nadal

Index 571

vii
CONTRIBUTORS

Ionela Bara
Department of Psychology
Bangor University
Bangor, UK

Amy M. Belfi
Department of Psychological Science
Missouri University of Science and Technology
Missouri, USA

Kent C. Berridge
Department of Psychology
University of Michigan
Ann Arbor, Michigan, USA

Jo Bervoets
Department of Philosophy
University of Antwerp
Antwerpen, Belgium

Giacomo Bignardi
Language and Genetics Department
Max Planck Institute for Psycholinguistics
Nijmegen, The Netherlands

Indra Bishnoi
Department of Neuroscience
University of Western Ontario
London, Canada

viii
Contributors

Aenne A. Brielmann
Department of Computational Neuroscience
Max Planck Institute for Biological Cybernetics
Tübingen, Germany

Beatriz Calvo-Merino
Department of Psychology
City, University of London
London, UK

Rebecca Chamberlain
Department of Psychology
Goldsmiths, University of London
London, UK

Anjan Chatterjee
Penn Center for Neuroaesthetics
University of Pennsylvania
Philadelphia, Pennsylvania, USA

Ana Clemente
Institute of Neurosciences
University of Barcelona
Barcelona, Spain

Alex Coburn
Penn Center for Neuroaesthetics
University of Pennsylvania
Philadelphia, Pennsylvania, USA

Ivan Colagè
Pontifical University of the Holy Cross
DISF Research Centre and Faculty of Philosophy
Rome, Italy

Julia Crone
Vienna Cognitive Sciences Hub
University of Vienna
Vienna, Austria

Emily S. Cross
Institute of Neuroscience & Psychology
University of Glasgow
Glasgow, UK

ix
Contributors

Kohinoor M. Darda
Penn Center for Neuroaesthetics
University of Pennsylvania
Philadelphia, Pennsylvania, USA

Francesco d’Errico
University of Bordeaux, PACEA, UMR 5199
Pessac, France

Gulce Nazli Dikecligil


Department of Neurology and Psychology
University of Pennsylvania
Philadelphia, Pennsylvania, USA

Zakaria Djebbara
Department of Architecture and Media Technology
Aalborg University
Aalborg, Denmark

Michael Donner
School of Psychological Sciences
University of Melbourne
Melbourne, Australia

Alejandro Dorado
Department of Psychology
University of the Balearic Islands
Palma, Spain

Lars Brorson Fich


Department of Architecture and Media Technology
Aalborg University
Aalborg, Denmark

Alejandro Galvez-Pol
Department of Psychology
University of the Balearic Islands
Palma, Spain

Jay A. Gottfried
Department of Neurology and Psychology
University of Pennsylvania
Philadelphia, Pennsylvania, USA

Franziska Hartung
Department of Psychology

x
Contributors

Newcastle University
Newcastle, UK

Claire Howlin
Autism Research Centre,
University of Cambridge
Cambridge, UK

Aniko Illes
Moholy-Nagy University of Art and Design
Budapest, Hungary

Tomohiro Ishizu
Department of Psychology
Kansai University
Osaka, Japan

Arthur M. Jacobs
Center for Cognitive Neuroscience
Freie Universität Berlin
Berlin, Germany

James. M. Kilner
Institute of Neurology
University College London
London, UK

Christoph Klebl
School of Psychology
University of Queensland
Brisbane, Australia

Stefan Koelsch
Department of Biological and Medical Psychology
University of Bergen
Bergen, Norway

Morten L. Kringelbach
Center for Music in the Brain
Aarhus University
Aarhus, Denmark

Helmut Leder
Faculty of Psychology
University of Vienna
Vienna, Austria

xi
Contributors

Haeeun Lee
Department of Psychology
Goldsmiths, University of London
London, UK

Psyche Loui
Department of Music
Northeastern University
Boston, Massachusetts, USA

Ernest Mas-Herrero
Institute of Neurosciences
University of Barcelona
Barcelona, Spain

Agnes Moors
Research Group of Quantitative Psychology and Individual Differences
KU Leuven
Leuven, Belgium

Enric Munar
Department of Psychology
University of the Balearic Islands
Palma, Spain

Marcos Nadal
Department of Psychology
University of the Balearic Islands
Palma, Spain

Guido Orgs
Department of Psychology
Goldsmiths, University of London
London, UK

Michael J. Ryan
Department of Integrative Biology
University of Texas at Austin
Austin, Texas, USA

Martin Skov
Danish Research Centre for Magnetic Resonance
Copenhagen University Hospital Hvidovre
Copenhagen, Denmark

Eloise Stark
Department of Psychiatry

xii
Contributors

Oxford University
Oxford, UK

Liila Taruffi
Department of Music
Durham University
Durham, UK

Pablo P. L. Tinio
Educational Foundations Department
Montclair State University
Montclair, New Jersey, USA

Fredrik Ullén
Department of Cognitive Neuropsychology
Max Plank Institute for Empirical Aesthetics
Frankfurt, Germany

Esther Ureña
Department of Psychology
University of the Balearic Islands
Palma, Spain

Sander Van de Cruys


Laboratory of Experimental Psychology
KU Leuven
Leuven, Belgium

Oshin Vartanian
Department of Psychology
University of Toronto
Toronto, Canada

Giovanni Vecchiato
Institute of Neuroscience
Italian National Research Council
Rome, Italy

Edward A. Vessel
Max Plank Institute for Empirical
Aesthetics
Frankfurt, Germany

Adam Weinberger
Penn Center for Neuroaesthetics
University of Pennsylvania
Philadelphia, Pennsylvania, USA

xiii
FIGURES

1.1A Data from a search of Google Ngrams’s corpus of publications in English. 2


1.1B Data from Pubmed showing the number of publications indexed as “neuroaesthetics” in
the period between 1965 and 2020. 2
1.2 A timeline of the development of aesthetics as a conceptual category. 5
1.3 Data from a bibliometric analysis of the development in the way the word “aesthetic”
was used together with other conceptual terms. 7
1.4 A figure adapted from Descartes’ book L’homme (Paris: Girard, 1677) depicting the
physiological system underpinning vision. 10
1.5 Chatterjee’s (2003) and Leder et al.’s (2004) two models of the neural processes involved
in aesthetic experience combined into one representation. 15
2.1 Sensory liking evaluations always occur in the context of behavioural tasks. 33
2.2 A schematic depiction of the functional components that constitute sensory liking
systems in biological organisms. 35
2.3 Midsagittal representation of the human brain showing the approximate location of the
neuroanatomical structures that make up the human evaluative system. 37
2.4 A schematic overview of the human evaluative system that shows how computational
mechanisms map onto structures in the mesocorticolimbic reward circuitry. 42
2.5 Model of computational mechanisms known to be involved in human sensory liking
evaluations, based on the empirical evidence reviewed in the chapter. 52
3.1 The Pleasure Cycle. 65
4.1 Illustration of the human brain with the bilateral insula highlighted in dark orange. 72
5.1 Interoceptive processes and their influence on stimulus processing. 93
5.2 Active sensing of stimuli in interoception. 97
6.1 Methodological “sources of variation.” 105
6.2 Amount of “shared taste” across observers varies by visual aesthetic domain. 108
6.3 The visual system and aesthetic appreciation. 111
6.4 Prefrontal and subcortical structures implicated in visual aesthetic appeal. 116
6.5 The default-mode network (DMN) is a network of highly interconnected brain regions
that are thought to support self-referential and inwardly directed thought and are
typically suppressed by tasks that require external focus. 120

xiv
Figures

8.1 Illustration of two differing models of odour valence perception. 151


8.2 Neuroanatomy of the human olfactory regions. 153
9.1 The representation of motion in photography, paintings, and sculpture. 173
9.2 Typical motion cues used in visual art. 175
9.3 Brain networks implicated by previous research in movement appreciation include the
perceptual/visual areas, sensorimotor network, reward network, and regions of the
default-mode network. 184
10.1 The Aesthetic Triad. 200
11.1 Examples of sexually selected traits. 219
11.2 The túngara frog has an unusual larynx characterized by a large fibrous mass
(FM) that protrudes from the vocal cords (VC). 225
11.3 A power spectrum of human speech normalized to the frequency with the greatest
amplitude, and consonance rankings of musical dyads as a function of the normalized
spectrum of speech sounds. 231
12.1 Frequency of books mentioning aesthetic sensitivity (blue), esthetic sensitivity (red),
aesthetic sensitiveness (green), and esthetic sensitiveness (orange) from 1800 to 2019. 241
13.1 Simplified representation, in the form of a sequentially branching tree, of the
evolutionary relations among the major groups of organisms mentioned in the text. 259
13.2 The figure shows a representation of how neurons evolved from individual
choanoflagellates as they joined in colonies and became able to exchange the contents
of vesicles in a regular fashion. 261
13.3 The photograph shows an exemplar of the ctenophore species Mertensia ovum swimming
in the dark. 262
13.4 Schematic representation of the central nervous system of the vertebrate ancestor with
differentiated diencephalon, mesencephalon, rhombencephalon, and nerve cord. 263
13.5 Representation of the telencephalon of an amphibian (salamander, top row), reptilian
(turtle, middle row), and mammal (hedgehog, bottom row). 266
13.6 Phylogenetic tree showing the relations among major groups of mammals and
illustrating the extension of primary and secondary visual, auditory, and somatosensory
cortical areas. 269
13.7 Illustration of the evolutionary relations among hominoids species, with macaque
monkeys as outgroup, with a comparison of adult brain sizes. 272
13.8 Distributed association zones are disproportionately expanded in humans. 273
13.9 Differences between humans and macaques in cortico-striatal resting-state functional
connectivity of the left and right dorsal caudate. 276
14.1 Tinio’s (2013) Mirror model of art. 298
14.2 Leder et al.’s (2004) model of aesthetic appreciation and aesthetic judgements. 299
14.3 The Aesthetic Triad (Chatterjee & Vartanian, 2014). 302
17.1 The anatomy of a neuron. 347
17.2 Magnetic resonance imaging. 348
17.3 The proposed network relevant to the identification and interaction with the built
environment is based on the reviewed studies. 358
18.1 Extension of the neurocognitive poetics model sketching the likely main neural
correlates of subprocesses involved in implicit and explicit fiction processing. 373
19.1 Hierarchical processing memory for narratives by Hasson et al. 381
21.1 The cognitive mechanisms identified in the literature and how they fit
together.414

xv
Figures

25.1 A simplified schema illustrating the interplay of top-down and bottom-up processes
involved in perception and evaluation. 462
25.2 Schematic overview of the development of models of aesthetic experiences. 465
26.1 Art, a prime example of appreciated stimuli, often breaks order or simplicity. 477
26.2 Two-tone or so-called Mooney images are created by blurring and thresholding
greyscale photographs. 483
26.3 A photograph of a frog that has been used to create Figure 26.2.484
26.4 A squiggly line that actually makes up the boundary of profiles of two different
opposing faces. 491
29.1 Occurrence of cultural innovations in four regions of the world during the last 850 ky. 543
29.2 Frequency of cultural innovations in the last 850,000 years at the regional scale and
(insert) at the global scale. 545
29.3 Schematic representation of the “top-down-also” view. 546
30.1 Two examples of Loring Hughes’ artistic work. 559
30.2 Four examples of Federico Fellini’s performances in line-bisection tasks, with
characteristic personal cartoons. 560
30.3 William Utermohlen. Series of Self-Portraits.562
30.4 Change in painting style observed after the start of dopaminergic treatment in a patient
with Parkinson’s disease. 564
30.5 Results of Bromberger and colleagues’ (2011) voxel lesion symptom analyses showing
areas where damage was associated with significant deviations of art attribute judgments. 566

xvi
TABLES

17.1 A brief overview of the main disadvantages and advantages of fMRI and EEG that the
reader must consider when further reading through the different studies. 349
17.2 An overview of the positively involved brain regions as reported by the reviewed studies. 356

xvii
PREFACE

When we first dipped our feet into neuroaesthetics, in the early 2000s, it was a new area of inquiry on the
fringes of neuroscience and psychology. The people doing the research were few and scattered about, and
it wasn’t difficult to read all the relevant literature in a single semester. The picture today is very different:
papers, chapters, and talks on neuroaesthetics can be found in major psychology and neuroscience journals,
in handbooks, and in mainstream neuroscience conferences such as Human Brain Mapping or Society for
Neuroscience. We now count our colleagues in the hundreds, and it has become difficult to keep up the pace
with everything that is being published. In only 20 years, neuroaesthetics has become a true scientific field
that is attracting a new generation of scientists.
This publication is the first handbook to provide a broad and comprehensive overview of neuroaesthetics.
As editors, we have been confronted with the daunting task of deciding the best way to present the scientific
accomplishments of a discipline that is still very much in its infancy. We decided that, rather than focusing
on prominent theories or authors, it would be better to organize the book around a selection of the most
important topics addressed by neuroaesthetics. We selected only topics that have generated a substantial body
of experimental evidence. That we managed to assemble a list of 30 topics that warranted inclusion in the
book shows the amount of progress neuroaesthetics has made in just two decades.
For historical reasons explained in Chapter 1, aesthetics refers to two different issues: the experience of
art, or how and why artworks are created and appreciated, and hedonic valuation, or how and why objects
are liked or disliked. Accordingly, neuroaesthetics has sought to identify the neurobiological mechanisms
underpinning both the experience of art and hedonic valuation. But it is important to realize that, although
in some cases art and aesthetic liking overlap, the two problems are not identical. Engaging with music,
dance, or narrative storytelling involves the activation of many neural processes that are not related to aes-
thetic evaluations. In contrast, the human brain assesses the hedonic value of many sensory objects other
than artworks. We have therefore chosen to divide the handbook into two parts that deal with each problem
separately.
In the first part, on aesthetic liking, we have collected chapters that provide a solid introduction to
what is known about the neurobiology of sensory liking. The first chapters in Part 1 present the general
context for the study of neuroaesthetics and the general principles of sensory liking, reviewing findings that
explain how liking and disliking occur as a function of perceptual computations occurring in different sen-
sory modalities, as well as projections from cognitive and interoceptive systems. The next chapters present
a detailed account of how the human brain uses this capacity for hedonic evaluation in different artforms:

xviii
Preface

music, visual art, dance, architecture, and so on. We have also included a chapter that discusses the ongoing
effort to understand why aesthetic evaluations vary not only across the human population but also in the
same individual as a consequence of experience and context. Finally, we have included two chapters in Part 1
that specifically review examples of sensory liking in other animals, both in the context of mate choice and in
contexts that are perhaps less obviously adaptive. While comparative work remains limited, we are convinced
that developing a common theory of hedonic evaluation that can explain how sensory liking and disliking
work in different functional contexts and in different nervous systems will be the next frontier in “aesthetic”
neuroscience.
Part 2, on art, provides a comprehensive overview of the existing research on the mechanisms underlying
the experience of works of art. Each chapter presents what is known about the neural mechanisms involved
in the perceptual, cognitive, and affective aspects of our engagement with music, visual art, dance, built
environments, and literature. This part also includes chapters on the way art is experienced in social situa-
tions, such as museums or concerts, and how art experiences may serve to elicit physiological and emotional
responses that can benefit health and well-being. Furthermore, we have included several chapters that discuss
research showing that art experiences are profoundly shaped by contextual conditions, including knowledge,
experience, and expectations. Finally, the last four chapters of the book survey work on different aspects
of art production, including questions of why humans began creating art, what artistic creativity is, how the
brains of musicians differ from non-musicians, and how art production can break down or change in patients
with brain damage or neurodegenerative diseases.
Our overarching aim in editing the book has been to provide a first entry into the field of neuroaesthet-
ics with introductions that are comprehensive, reliable, and succinctly written. We especially hope that the
handbook will find use as a textbook in some of many academic courses that have started to appear over
the last few years around the world. But we also hope the book will enjoy wider use. We both believe that
neuroaesthetics has reached a point where there is a need for taking stock of what is empirically known about
perennial questions such as whether aesthetic liking is driven by stimulus properties or modulated by con-
textual conditions. Furthermore, the continuing progress of neuroaesthetics requires a collective enterprise
with a common framework. This includes finding common ground with respect to how key concepts are
conceived and employed in experiments. We also hope that the handbook can make a modest contribution
in helping neuroaesthetics take these next steps in its development.
It has not been easy putting together a volume with 30 contributions at a time when the lives of many
people have been severely disrupted. We want to thank the contributors who, living through a global pan-
demic, persevered and wrote excellent chapters. We think the end result is much more than could have
been expected under such circumstances and a true testament to the vitality of neuroaesthetics as a scientific
discipline and to the motivation of the researchers that make it all happen.
Martin Skov
Copenhagen Business School & Copenhagen
University Hospital Hvidovre,
Copenhagen, Denmark

Marcos Nadal
University of the Balearic Islands,
Palma, Spain
January 2022

xix
1
NEUROAESTHETICS AS A
SCIENTIFIC DISCIPLINE
An intellectual history

Martin Skov

A Google Ngram search reveals that use of the word neuroaesthetics was almost non-existent before the year
2000. It first became part of academic discourse between 2000 and 2005 and was only widely adopted after
2010 (Figure 1.1A). A similar trend can be observed if one searches for publications indexed with the key-
word “neuroaesthetics” in databases such as PubMed or Web of Science: Before 2000, few such publications
exist, if any. They then start to appear in increasingly frequent numbers during the 2000s, and multiply year
by year throughout the last decade (Figure 1.1B).
What do these numbers tell us about the history of neuroaesthetics? First, they tell us that there was no
scientific enterprise that referred to itself as neuroaesthetics before 2000. The discipline that exists today
under this name was born between 2000 and 2010. Second, they tell us that the first 20  years of neu-
roaesthetics can best be described as two historically different periods: an initial era of incubation, where
neuroaesthetics was established as a viable idea (2000–2010), followed by a mature period much richer in
scientific output (2010–2021).
The new scientific discipline of neuroaesthetics did not, however, appear out of thin air. As I will show
in this chapter, it is possible to trace an effort to furnish aesthetics with a neuroscientific basis back to the
18th century. We must therefore ask: what changed around 2000? Why did a new discipline dedicated to
the exploration of the neurobiological basis of aesthetic experience emerge, and what was “new” about it?
The answer, I will suggest, is that neuroaesthetics arose as a concerted effort to transform what had previously
been almost exclusively a speculative and theoretical endeavour into an actual experimental science. What set
neuroaesthetics apart from earlier attempts to craft a neuroscientific basis for aesthetics was the development,
in the 1980s and 1990s, of non-invasive neuroimaging methods that allowed for the exploration of neural
activity associated with complex human thought. What was new about the neuroaesthetics that emerged in
the 2000s was an ability to probe the human brain as it engaged in aesthetic experiences.
It is, however, important to note that this new scientific enterprise did not simply break with the histori-
cal tradition that preceded it. Rather, the way the new field of neuroaesthetics conceived of its own mission
and the problems it wanted to pursue was highly influenced by ideas that had developed much earlier, start-
ing with the conceptual invention of the category of aesthetics in the 18th and 19th centuries. In my view,
this prehistory remains both unknown and unexplored. I will therefore try to show how, over the course
of the 18th and 19th centuries, notions of beauty and taste were fused with that of fine art to form the idea

DOI: 10.4324/9781003008675-1 1
Martin Skov

Figure 1.1A Data from a search of Google Ngrams’s corpus of publications in English.

Figure 1.1B Data from Pubmed showing the number of publications indexed as “neuroaesthetics” in the period between
1965 and 2020.

that humans have distinctive experiences called aesthetic experiences and why this idea prompted a quest to
understand the psychological and physiological nature of these experiences that endures to this day. In some
detail I trace how generations of researchers tried to apply emerging insights into the neurobiology of the
human brain to (speculative) explanations of how aesthetic experience arises. My aim is to give the reader an

2
History of neuroaesthetics as a discipline

idea of how these theories served as a direct inspiration for the establishment of neuroaesthetics, influencing
both theoretical models and experimental work in the early years of the new discipline’s existence.
I should perhaps stress that this chapter should not be thought of as a traditional introduction to neuroaes-
thetics. Its purpose is not to review findings or provide a representative overview of the work conducted over
the last 20 years; the rest of the book covers that material. Instead, it is meant to be a historical study of the
intellectual ideas and theories that prefaced neuroaesthetics. I hope it will give the reader an idea of where
the impetus to study the neurobiological basis of aesthetics came from and why the scientific enterprise of
neuroaesthetics took the form it did, especially in its first incarnation.

Inventing the concept of aesthetics


Neuroaesthetics calls for a neuroscientific study of aesthetics, but what does this mean? Most introductions
to neuroaesthetics take “aesthetics” to refer to the study of mental states that are associated with the experi-
ence of art or evaluating sensory objects for their aesthetic value. For example, Chatterjee (2011) defines
aesthetics as a term “used broadly to encompass the perception, production, and response to art, as well
as interactions with objects and scenes that evoke an intense feeling, often of pleasure” (p. 53). Upon this
definition, neuroaesthetics can be viewed as a branch of cognitive neuroscience “focused on understanding
the biological bases of aesthetic experiences” (Chatterjee & Vartanian, 2016, p. 172). Aesthetic experiences,
more specifically, “are an emergent property of the interaction of the sensory-motor, emotion-valuation, and
knowledge-meaning neural systems” (Chatterjee & Vartanian, 2016, p. 178).
It is worth pondering where this idea comes from. Why do we believe that the human brain experiences a
special category of mental states called aesthetic experiences that are related to the engagement with works of
art and evaluative appraisals? There are many forms of behaviour and experience that seem specific to Homo
sapiens that are not singled out for similar psychological and neuroscientific scrutiny. For instance, sport is an
activity that is every bit as particular to human behaviour and experience as art and aesthetic appraisals, yet
there is no existing field of neurosports. Most neuroscientists simply consider the experience of a football game
the emergent property of neural systems that are common to any other experience that involves perception,
emotion, cognition, and so on.
Aesthetics is treated differently. Not only are aesthetic experiences considered mental traits that are critical
to an understanding of the human mind, but they have been the subject of scientific inquiry for more than
150 years (Nadal & Ureña, 2022). While there is no concept in psychology and neuroscience of, say, special-
ized sports emotions, most psychologists and neuroscientists are convinced that distinctive aesthetic emotions
exist (Menninghaus et al., 2019; Skov & Nadal, 2020b). Indeed, the ability to entertain aesthetic experiences
has been routinely used in fields such as palaeontology or ethology to distinguish the psychology of Homo
sapiens from that of other species, suggesting that modern humans have evolved a dedicated mental “faculty”
for generating aesthetic experiences (Ayala, 2017; Dobzhansky, 1962; Klein, 2002; see also Chapter 13).
What motivates this conviction? Since many of the assumptions associated with the idea of specialized
aesthetic experiences are unsupported by empirical knowledge (Skov  & Nadal, 2020a, 2020b), we can
surmise that it does not stem from psychological or neuroscientific findings. For example, while there is
ample evidence that aesthetic evaluations rely causally on executive functions such as working memory (e.g.,
Cattaneo et al., 2014; Che et al., 2021; Sherman et al., 2015; see also Chapter 2), no theory of neuroaesthet-
ics has developed a concept of “aesthetic working memory” akin to that of aesthetic emotions. Instead, the
conception of aesthetics that continues to inform contemporary psychology and neuroscience was almost
entirely invented by philosophers working before modern psychology and neuroscience. It is this complex
of ideas, developed over the course of the 18th and 19th centuries, that not only accounts for the contin-
ued belief in aesthetic experiences but also, by and large, determines how we still conceive of these today
(i.e., that aesthetic experiences entail aesthetic emotional states but not aesthetic working memory states).

3
Martin Skov

It goes beyond this chapter to trace all the historical roots of the ideas that constitute the conception
of aesthetics (see, e.g., Dickie, 1996; Kivy, 2003; Shiner, 2001). However, several events are critical to the
understanding of why Western philosophy became convinced that engagements with art and evaluative
appraisals elicit specialized aesthetic experiences and why explaining these as a function of neurobiological
mechanisms became an enduring preoccupation.
First, in the 16th century, Italian philosophers began using the word taste to describe the mental power
governing judgments of liking and disliking (Tonelli, 2003). Before, classical and medieval philosophers had
considered judgments of beauty as responses to specific objective properties (Dieckmann, 1974; Tatarkiewicz,
1972). This view implied that objects could either be beautiful or not and that judgments of beauty therefore
had to be universal. Greek philosophers did not develop anything like a psychological explanation of how
humans detect the beauty of objects, but as a general principle, they believed that mental judgments involved
the application, by a faculty of reason, of concepts to sense impressions. Under this assumption, beauty
judgments could be thought of as mental acts involving the recognition that the perceived properties of an
object adhered to the principle of beauty (specifically that they were organized in a particularly harmonious
and orderly fashion; Tatarkiewicz, 1972). Greek philosophers agreed that successful beauty judgments were
accompanied by feelings of pleasure (Dieckmann, 1974).
An enduring problem for the classical conception of beauty was the fact that people often did not agree
on whether an object was beautiful. This observation cast doubt on the idea that beauty is an inherent quality
of sensory objects which the human mind recognizes and responds to. To many Renaissance philosophers,
it seemed more reasonable to assume that beauty is an evaluative reaction to sensory impressions that varies
from person to person. They introduced the concept of taste to capture this change in understanding of how
beauty judgments worked: from an objectivist relation between object and evaluative power to a subjectivist
conception. Rather than universal judgments, the philosophers of taste proposed that the evaluative power of
the human mind produced different tastes that varied from one mind to another (Tonelli, 2003).
The concept of taste had two significant consequences for the development of a notion of aesthetic
experiences. Both of these consequences sprung from the subjectivist implication of the concept of taste:
that liking and disliking are based on individual evaluations, not universal judgments. Accepting this theory
invariably raised the problem of a “standard” of taste. If beauty is a creation of subjective taste, how can we
know which tastes are “good” or “bad”? During the 17th and 18th centuries, European intellectuals became
fascinated by this question. Every aspect of social life was subjected to debates over what constituted good
and bad taste, with “good” manners, “eloquent” speech, “haute” cuisine, and so on, being codified in nor-
mative treatises that the nobility and emerging bourgeoisie devoured to fit in with “good” society (Shiner,
2001; Smith, 1997). Deciding what counted as good taste was seen as crucial to Bildung—the edification
both of the person and the nation (Shiner, 2001; Smith, 1997): only objects and behaviours determined to
embody good taste were considered “good” for you.
The first consequence of this development was the creation of a category of “Fine Arts” (Kristeller, 1951,
1952). As the philosopher Paul Oskar Kristeller (1951) has shown, before the 18th century, there was no
Western concept of art that comprised only “the five major arts of painting, sculpture, architecture, music
and poetry” (p. 497), “all by themselves, clearly separated by common characteristics from the crafts, the
sciences and other human activities” (p. 498). When the Greeks had spoken of techné and Latin philosophers
of ars, they meant “all kinds of human activities which we would call crafts or sciences . . . something that
can be taught and learned” (Kristeller, 1951, p. 498). Neither was art specifically associated with beauty or
other forms of evaluative judgments. Only in the late 17th century did art emerge as a “modern system”
(Kristeller, 1951, 1952) that invested certain objects with a value and prestige they had hitherto not been
assumed to possess and set them apart from nonart objects. This development was very much interwoven
with the contemporaneous societal quest for standards of “good” taste: The new concept of fine art posited

4
History of neuroaesthetics as a discipline

that, in contrast to “mechanical arts” (Kristeller, 1952, p.  21) and other human endeavours, objects that
qualified as art were thought to conjure the especially refined and elevated kind of pleasure that character-
ized “good” taste.
The other consequence was a gradual reckoning with the psychological causes of taste. If beauty and
other judgments of taste were subjective in nature, then the evaluative power they were the result of must
operate by some discernible principle. The human mind had to contain some form of mechanism that
determined the taste response to a given sensory input. In parallel to the invention, around 1700, of fine art
as a sui generis category, philosophical attempts to provide a foundation for taste resulted in a psychological
invention as well: the idea that taste is the result of its own “faculty” or “sense.” The Scottish philosopher
Francis Hutcheson (1725/1973) was the first to argue the human mind comprises an “inner sense” spe-
cialized for beauty judgments. It is this sense of beauty, he argued, that gives rise to our individual taste
experiences by acting upon the impressions we receive from the outer senses. Furthermore, Hutcheson
distinguished the sense of beauty from a moral sense, introducing the notion—later to be much further
developed—that beauty judgments only apply to certain sense impressions (Costelloe, 2018; Dickie, 1996;
Kivy, 2003). His compatriot David Hume soon adopted Hutcheson’s idea that taste responses are grounded
in a specialized faculty and extended Hutcheson’s account to the question of a standard of taste (Hume,
1757/1985). While taste is a subjective response to sense impressions, some tastes are better than others.
How can this be possible? Because, Hume argued, the faculty of taste can be “improved by practice, per-
fected by comparison, and cleared of all prejudice” (1757/1985, p. 241). In other words, people can acquire
“good” and “bad” taste by means of experience and learning, presumably through some sort of reshaping
of the beauty sense.

Figure 1.2 A timeline of the development of aesthetics as a conceptual category. Between 1500 and 1700, European
philosophers invented two new concepts: taste and fine art. When Baumgarten introduced the word aesthetics
into the literature in 1735, it was adopted as the general name for the study of judgments of taste, especially
through the influence of the work of Kant. In the 19th century, spearheaded by Hegel, aesthetic judgments,
especially beauty, came to be seen as primarily connected to an engagement with fine art objects. This novel
interpretation created the idea that the human mind enjoys specialized aesthetic experiences that are elicited by
fine art objects.

Now, none of these intellectual events happened under the banner of “aesthetics.” That word did not exist
until 1735, when it was invented by a German philosopher, Alexander Gottlieb Baumgarten (1735/1951;
see Figure  1.2). Indeed, as originally conceived by Baumgarten, aesthetics did not even refer specifically
to existing questions of taste, beauty, or art. Rather, Baumgarten (1735/1951) had argued that philosophy
needed a general study of those sensory judgments that are a not accounted for by logic; that is to say, all
forms of judgments that involved “lower-level epistemology” (Allesch, 2018). Aesthetics was the name of
this proposed study. It only became applied to the debates regarding the nature of taste discussed previously

5
Martin Skov

several decades later (Reiss, 1994) and primarily through the influence of Kant’s third critique, Critique of
Judgment (Kant, 1790/2001). Kant embraced Baumgarten’s nomenclature and used the term aesthetic judgment
to distinguish judgments of taste from judgments of knowledge and moral judgments (the topics of his two
other critiques). Furthermore, Kant (1790/2001) introduced a set of criteria for defining different kinds of
judgments of taste, distinguishing between, amongst others, judgments of beauty, judgments of the sublime,
and judgments of the agreeable. Most famously, he argued that pure judgments of beauty would be disinterested,
universal, exhibit “purposiveness without an end,” and be necessary (Kant, 1790/2001).
The influence of Kant was so pronounced that, by the beginning of the 19th century, aesthetics had taken
over as the de facto name of philosophical inquiries into taste. At first this seemed to be nothing more than
a rebranding of the ideas developed in the preceding decades. Instead of speaking of a sense of beauty, the
norm became to speak of an aesthetic sense as the mechanism of evaluative taste. However, it soon was clear
that under the name of aesthetic evaluations, the phenomenon of taste became invested with a more specific
and restricted meaning. Evaluations rooted in the aesthetic sense were considered to involve specific mental
states, first and foremost feelings of pleasure or pain. But aesthetic evaluations also seem to involve distinct
perceptual and cognitive states (concepts that philosophers at the time would not have used). For example,
Kant insisted that aesthetic judgments were characterized not only by evoking states of subjective feelings but
also by aspiring to be correct. They were subjective but also “universally valid” (Stolnitz, 1961). Similarly,
Joshua Reynolds (1798) claimed that aesthetic evaluations were shaped by learning and experience. They
depend on “our skill in selecting, and our care in digesting, methodizing, and comparing our observa-
tions” (pp. 212–13). The idea took hold that aesthetic evaluations apply a special form of contemplation
to the object being evaluated, a unique type of “gaze” characterized by a particular kind of attention and
“sentiment.”
At first, this emerging concept of aesthetic appreciation as a distinct kind of taste evaluation, produced
by a special aesthetic sense, did not make any claims about being applicable only to certain aesthetic
objects. However, this changed as the 19th century unfolded. In 1820, the German philosopher Hegel
gave a series of lectures later collected in a book called Philosophy of Fine Art (1920). In the course of these
lectures, he redefined the concept of aesthetics to mean the study of “schöne Künste”: Art, he claimed, is
the principal means for “revealing” beauty (I, p. 125) because of its unique ability to embody “the Sensous
Semblance [das sinnliche Scheinen] of the Idea” (I, p. 154). Hegel’s argument merged the idea of “aesthetic
taste” with the new system of “Fine Arts.” In consequence, aesthetic experiences—a novel concept—came
to be seen as a special class of experiences that human can have that involved the application of aesthetic
appreciation to a limited set of Fine Art objects (Carroll, 2008; Shiner, 2001). By the end of the 19th cen-
tury, aesthetics had ceased to be understood as a (broad) study of taste and was now exclusively known as
the study of art and the aesthetic experiences art objects were believed to afford (Kranjec & Skov, 2021;
Figure 1.3).
Collectively, the events recounted here forged the concept of aesthetics we still employ. The assumptions
and theories with which we continue to invest aesthetic concepts were inherited from the philosophic ideas
described previously. This is especially true of the idea that the human mind is equipped with a distinct
aesthetic sense, a mechanism for producing specific aesthetic evaluations of sensory objects. Even among
contemporary neuroaesthetics researchers, it is common to see the notion of aesthetic evaluation character-
ized as a specific form of appreciative contemplation that is principally directed at fine art objects. Yet this
idea was invented and formulated by philosophers who lived before the study of the human mind evolved
into an empirical and experimental science (Reed, 1997). The assumptions underlying this idea continue to
inform work in psychology and neuroscience on aesthetics. Indeed, this idea has guided empirical research
unceasingly since experimental psychology and neuroscience first became possible: We focus our investiga-
tions on certain properties (e.g., aesthetic emotions) and disregard others (e.g., aesthetic working memory),

6
History of neuroaesthetics as a discipline

Epoch 1800–1850 1851–1900 1901–1950 1951–2000

RANK
1 culture sense experience experience
2 feeling judgment value(s) value(s)
3 education value sense appreciation
4 character pleasure theory theory
5 criticism point pleasure pleasure
6 principles culture appreciation quality(ities)
7 value taste enjoyment sense
8 pleasure feeling(s) judgment object
9 judgment nature emotion
10 part

Figure 1.3 Data from a bibliometric analysis of the development in the way the word “aesthetic” was used together
with other conceptual terms. The topmost panel shows the correlation of the frequency of “aesthetic” with
questions of art and beauty over the period between 1800 and 2000, demonstrating that the notion of aesthet-
ics went from being associated with notion of beauty to becoming increasingly associated with the notion
of art. The lower panel shows the top-ranked nouns modified by “aesthetic” in the same period. This data
demonstrates how aesthetics went from being used to describe concepts related to taste (“aesthetic feeling,”
“aesthetic character”) to primarily describe a general experience (“aesthetic experience”). Figure adapted
from Kranjec and Skov (2021).

because we still—explicitly or tacitly—accept the basic ideas of aesthetics set out in the course of the 18th
and 19th centuries.1

A neuroscience of aesthetics before neuroaesthetics


While the concept of aesthetics was motivated by pre-scientific ideas, it evidently made claims about human
psychology and physiology. For example, every philosopher involved in the invention of aesthetics as a
concept agreed that judgments of beauty involved the generation of pleasure ( Judgments of ugliness were
thought to involve the generation of pain.) As we have seen, the central new intellectual invention inspir-
ing theories of aesthetics was the claim that the human mind is equipped with a mechanism for producing
aesthetic evaluations—an aesthetic sense. This claim implied that such an evaluative mechanism operated

7
Martin Skov

according to specific principles—that the aesthetic sense was characterized by what we today would call a
computational function with a physiological implementation. Even though 18th- and 19th-century philoso-
phers did not speak in such modern terms, they understood that there had to be “laws” of some sort that
determined if a sensory input was experienced as liked or disliked and that the aesthetic sense, if it existed,
had to occupy a specific place within the architecture of the human mind (linking, for example, sensation to
emotions). In consequence, they tried to explain how the aesthetic sense worked.
Naturally, these explanations were wholly uninformed by modern ideas about human brain structure and
function. They were uninformed by modern knowledge of anatomy and molecular biology. They obvi-
ously were all speculative, based on analytic arguments rather than empirical findings. When, for instance,
Hutcheson (1725/1973) argued that the “power of receiving” (p. 34) the idea of beauty consists of the mind
forming “sensible ideas” (p. 36) that are characterized by “uniformity amidst Variety” (p. 40), he clearly did
not advance a psychological theory that conforms to any contemporary models of human perception or
cognition, nor did he bother to test if people do in fact experience beauty as a cause of the “uniformity
amidst variety” principle his theory advanced. Still, it should be acknowledged, I  think, that Hutcheson
(1725/1973) was trying to explain how the evaluative mechanism—the beauty sense—he claims to exist
functions. He understood that there is a psychological mystery at the heart of the new mental faculty he had
proposed, and he realized that if we want to understand taste as a subjective response to sensory stimuli, we
need to flesh out the computational rules according to which these evaluations operate. The same can be said
of Hume, Baumgarten, Kant, or any of the other 18th- and 19th-century philosophers who helped found
the concept of aesthetics.
In this sense, we can talk of a psychology of aesthetics avant la lettre. Indeed, from the very start, theories
of aesthetic evaluation took their departure from existing models of the human mind. Thus, Hutcheson’s
theory, to stick to this example, owed its notion of an “inner sense,” dedicated to beauty evaluations, almost
completely to Locke’s theory of perception (Dickie, 1996; Kivy, 2003), which had radically upended the
classical view of how sensations interface with reason (Reed, 1997). It has been argued that before Fechner,
direct calls for a “psychological aesthetics” were limited to a few, largely ignored publications such as Johann
Heinrich Zschokke’s 1793 book Ideen zur psychologischen Aesthetik (e.g., Allesch, 2018). This argument, how-
ever, misrepresents the degree to which all theories of aesthetic evaluation were conditioned by the existing
(natural philosophical) understanding of human psychology.2 Of course, only with the advent of a scientific
psychology in the 19th century (Reed, 1997) did theories of aesthetics begin to incorporate knowledge of
psychological functions and processes that had their origin in empirical studies of the human mind (Allesch,
1987, 2018; Nadal & Ureña, 2022).
Similarly, the development of aesthetic experience as a category gave rise to a neuroscience of aesthetics
avant la lettre, first in the form of speculation about the physiological properties of the aesthetic sense and
later in the form of what Chatterjee and Vartanian (2014) have dubbed a descriptive neuroaesthetics. The term
descriptive neuroaesthetics, in Chatterjee and Vartanian’s definition, encompasses all attempts to account for
aesthetic phenomena that make use of facts about the brain derived from mainstream neuroscience research.
In contrast, Chatterjee and Vartanian (2014) label experiments that specifically examine neurobiological
processes thought to play a causal role in the generation of aesthetic experiences experimental neuroaesthet-
ics. If we accept this terminology, I think it is appropriate to call the numerous papers and books published
before 2000 that tried to incorporate findings produced by the nascent field of experimental neuroscience
into theories of aesthetics examples of descriptive neuroaesthetics. These writings only rarely relied on actual
examples of experimental neuroaesthetics, but they did attempt to connect the hypothesized functions of
aesthetic experience to revealed functions of the brain.
To keep my discussion of this literature manageable, I will here only touch upon a few cases that illustrate
this point. Common to all of them is the fact that they demonstrate how efforts to develop a neuroscience of
aesthetics begin with a received conception of what aesthetics is. These publications all assumed the existence

8
History of neuroaesthetics as a discipline

of an aesthetic sense that produces aesthetic experiences. What they wanted to understand were the neural
mechanisms of this putative mental faculty, and they tried to do so by applying relevant findings from the
emerging field of neuroscience to the problems they were interested in.
As described previously, the concept of aesthetics presented theorists with several psychological problems
that arose from the contention that humans experience distinct aesthetic experiences. These problems were
copious. For example, the notion of aesthetic experience raised the question of how aesthetic experiences
are to be distinguished from non-aesthetic experiences. It also posed the developmental question of why
people have different aesthetic experiences for the same sensory objects. Finally, it presented theorists with
the thorny question of how it is possible to define fine art as a psychologically or neuroscientifically relevant
category. At one point or another, in the course of the 19th and 20th centuries, theories of aesthetics sought
inspiration from discoveries in neuroscience and biology to provide possible explanations for these problems.
Here, I will only deal with theories that tried to address the most central problem created by the concept of
aesthetics: explaining the psychological mechanism that must underlie the process of aesthetic evaluation.
As noted, aesthetic theory posited that this mechanism took the form of an “inner” sense—a “power” or
“faculty,” and so on—that acted upon information it received from the body’s outer senses. This inner sense
was thought to embody an evaluative function that reacted to individual sense impressions with either pain
or pleasure. As early as the second half of the 18th century, philosophers began to seek inspiration from
the emerging field of physiology to craft material accounts that would explain how this putative function
worked. In doing so, they established a literature of physiological aesthetics that used models of nerve action
to speculate on the way transfer of aesthetic information would happen and why different neural processes
resulted in feelings of pleasures and pain.
Physiology had become a distinct science in the 17th century with the work of René Descartes (Ochs,
2004). Descartes’ theory of human physiology built on his mechanical physics and described sense impres-
sions as the movement of animal spirits through hollow nerve tubules (Ochs, 2004). Observing the general
laws of mechanics, this action was thought to happen when external objects acted on the body’s sense
organs and forced animal spirits to flow through the nerve fibres (in a way akin to a hydraulic system; see
Figure 1.4). Famously, Descartes retained a dualistic view of the human mind, so he did not believe think-
ing could be explained by the actions of physiological processes. Instead, he proposed that sense impressions
terminated in the pineal gland, where they were somehow transferred to the soul that did the actual thinking.
This account began to break down in the 18th century, though. This change can be seen, for instance, in the
publication, in 1767, of Albrech von Haller’s Elementa physiologia corpris humania. In this eight-volume tome,
von Haller claimed—partly informed by his own observations—that the human body contains two different
kinds of nerve actions: On the one hand, some parts of the human body, in line with Descartes, contracted
automatically when stimulated. Von Haller called these parts irritable. But other, sensible, body parts appeared
to respond according to their own power of movement. In contrast to the irritable parts of the body, the sen-
sible parts were not simply caused to act by the external stimulation but seemed to respond to external sense
impressions by obeying some intrinsic power. The reason for the difference between irritable and sensible
reactions, von Haller conjectured, was to be found in their physiological make-up, especially the type and
number of fibres that constituted the two different kinds of sensation (Ochs, 2004).
The new ideas of von Haller and other 18th-century physiologists broke with Descartes by claiming that
not all nerve actions occur as mechanistic bottom-up processes. Some have an inherent sensibility that makes
them react to sense impressions according to their own intrinsic principles (Fastrup, 2007; Ochs, 2004). This
“Vitalistic” account (Rey, 2002) greatly inspired physiological theories of aesthetic evaluation by suggest-
ing that different parts of the sensory system can exhibit different sensibilities. To philosophers interested in
aesthetics, the concept of sensibility suggested that “aesthetic” reactions to sensory objects might consist of
specific forms of nerve movement. For example, in Philosophical Enquiry into the Origin of our Ideas of the Sublime
and the Beautiful, Edmund Burke (1757) proposed that pain and pleasure arose from either a “stretching” or a

9
Martin Skov

Figure 1.4 A  figure adapted from Descartes’ book L’homme (Paris: Girard, 1677) depicting the physiological system
underpinning vision. Descartes envisioned the outer senses as a system of hollow nerve tubes through which
animal spirits travelled according to the rules of mechanical physics when stimuli engaged the sensory sys-
tems. In the later tradition of physiological aesthetics, scholars attempted to work out the way that aesthetic
experiences could arise from this material basis. The majority of this work concentrated on identifying on
one or more of three possible physiological causes of aesthetic experience: a specific sensitivity or movement
characterizing the sensory organs by which aesthetic objects are perceived (“Outer sense”); a specific mecha-
nism for apprehension of sense impressions (“Beauty sense”); or a specialized way in which we experience
feelings of pleasure and pain as a result of the aesthetic experience.

“relaxation” of nerve fibres. These nerve actions, in turn, had their origin in particular responses of sensible
nerves to external objects. For instance, Burke hypothesized that because large objects strike the retina all at
once, the stretching of nerves leads to the elicitation of pain responses that further self-preservation. How-
ever, in cases where such objects are perceived from a safe distance, the nerves can become relaxed, prompt-
ing feelings of awe and astonishment, a form of pleasurable pain (Burke, 1757). The emotional state we feel
forms the basis of our aesthetic judgments: feelings of pleasure yield the idea of beauty, while feelings of pain
yield the idea of ugliness. As a kind of intermediate response, feelings of “pleasurable pain” were thought to
yield the idea of the sublime.
The latter part of the 18th century saw the publication of a slew of books arguing for different models
of physiological aesthetics. In addition to Burke’s, important treatises were published by Théophile Bordeu,
Denis Diderot, Henri Fouquet, Daniel Webb, and Uvedale Price, among others (Fastrup, 2007; Rey, 2002).
All of these models used ideas of nerve sensibilities to root the psychology of aesthetic evaluations in a
material physiology. Proponents of physiological aesthetics were convinced that mental states (“Ideas”) were
generated by certain nerve acts, especially individual kinds of fibre movements. This, of course, amounted

10
History of neuroaesthetics as a discipline

to a “psycho-physiology.” However, an empirical exploration of this relationship between physiology and


psychology seemed doomed, since it was difficult to see how it was possible to differentiate between indi-
vidual thoughts. Kant (1786/1970) had made this very argument in the Metaphysical Foundations of Natural
Science, declaring any empirical science of psychology impossible. Then, in 1834, the German physiologist
Ernst Weber discovered that for the sensations of two different weights to be perceptually noticeable, the
difference in intensity of the larger weight to the smaller weight had to obey a specific ratio, ΔI/I = k (Ochs,
2004). Weber’s fortuitous observation demonstrated that it was in fact possible to relate different psychologi-
cal states to physiological measurement and that this relationship could even be expressed in mathematical
terms.
Weber’s work inspired Gustav Fechner and Wilhelm Wundt to create psychology as an experimental sci-
ence (Reed, 1997). By measuring responses to controlled sensory stimuli, Fechner and Wundt hoped to dis-
cover “psychophysical” laws akin to the one unearthed by Weber (Ochs, 2004; Reed, 1997). Such laws laid
the foundations of psychology in the physiological substrate of the brain, explaining how “complex ideas”
arise from the processes of perception. Both Fechner and Wundt considered aesthetic experiences amena-
ble to this approach. For example, in his textbook Grundzüge der physiologischen Psychologie, Wundt (1873)
explained how aesthetic feelings arise as a consequence of the psychological constitution of the human
brain. In Wundt’s theory, sensations (Empfindungen) constituted the interface between the physical and the
physic by forming the elemental basis of mental representations (Vorstellungen). Mental representations, in
turn, comprised the content of consciousness (Kim, 2016; Ochs, 2004; Wundt, 1873). Wundt argued that,
as physiological processes, sensations had three properties: quality, intensity, and a “feeling-tone” (Gefüh-
lston), all determined by the “peculiar laws of excitation of the neural matter” (Wundt, 1873, I, p. 390).
However, for individual sensations to come together as psychological ideas, they had to be subjected to a
process of “psychological synthesis of sensations” (Wundt, 1873, II, p. 256). This process, termed appercep-
tion, used acts of separation and combination to compound individual associations into specific representations
(Kim, 2016; Wundt, 1873). Some such representations produced aesthetic feelings, as explained by Wundt
(1873) in chapter 17 of his book: These were feelings of liking and disliking (“Gefallens und Missfallens”;
p. 691) that accompanied particular mental representations. Like representations, aesthetic feelings were not
identical to sensations but psychological constructs built through the apperception of certain sense quali-
ties. For example, Wundt (1873) claimed that rhythm is experienced as pleasurable because recurring sense
impressions of equal intensity or quality are experienced as regularly occurring auditory representations
(Gehörsvorstellungen).
Fechner took a different approach to psychology to Wundt’s (Höge, 1995; Kim, 2016; Nadal & Ureña,
2022). For him, sensory activity and psychological states are interrelated, two manifestations of the same phe-
nomenon (Nadal & Ureña, 2022). In other words, Fechner did not view psychological acts as independent
of physiological activity. In his work on aesthetics, he argued that for a stimulus to be liked, it must elicit a
certain level of pleasure (Fechner, 1876). Based on his empirical findings, Fechner argued that this “pleasing-
ness” (Wohlgefälligkeit) could be caused by either direct factors, namely the physical properties of the object
being sensed, or by associative factors, the knowledge and experience marshalled by the individual’s memory
(Fechner, 1876). Thus, even though Fechner did not conceive of psychological acts as different from physi-
ological processes in the way Wundt did, he did admit that aesthetic evaluations must involve some form of
“aesthetic apprehension” that interacted with the “direct impressions of the matter” (Fechner, 1876, p. 94).
How the mechanics of this interaction worked, Fechner had to admit he did not know. Describing its physi-
ological nature was a job for future generations.
Neither Wundt nor Fechner considered aesthetic evaluations a unique psychological mechanism. The
mechanisms of perception, psychological apprehension, and pleasure associated with aesthetic experiences
were all be understood as functions of general principles of sensation, cognition, and emotion. Other writers

11
Martin Skov

did, however, try to use physiology to explain how aesthetic liking and disliking might emerge as sui generis
psychological states. For example, the Canadian Darwinist Grant Allen (1877) published a book, Physiological
Aesthetics, in which he presented a theory of “Aesthetic Feelings” as “special cases” (p. 1) of emotions. Allen’s
work, which was strongly informed by the Herbert Spencer’s “Æstho-Physiology,” contended that aesthetic
feelings are those “pleasures or pain . . . which result from the contemplation of the beautiful or the ugly,
in art or nature” (1877, p. 3). Since “the physical states that accompany each such Aesthetic Feeling are the
same in kind as those which characterize all other pleasures or pains” (p. 2), Allen needed to explain how
“the special class of pleasures and pains known as aesthetic differ from the remainder of their genus” (p. 2).
The answer he provided was that aesthetic pleasures and pain are differentiated from ordinary pleasures and
pain by “their remoteness from life-serving function and their having pleasure alone as their immediate end”
(p. 33). This difference he attributed to a difference in physiological activity: Ordinary pleasure is “the con-
comitant of the healthy action of any or all of the organs or members supplied with afferent cerebro-spinal
nerves, to an extent not exceeding the ordinary powers of reparation possessed by the system” (p. 21). In
contrast, aesthetic pleasure was “the subjective concomitant of the normal amount of activity, not directly
connected with life-serving function, in the peripheral end-organs of the cerebro-spinal nervous system”
(Allen, 1877, p. 34). Humans were believed to “exercise” the latter kind of process by contemplating either
“Art” or “nature generally” (p. 36).
As a whole, 18th- and 19th-century works on physiological aesthetics represent an enduring dream
of understanding the “aesthetic sense” in materialist terms (Morgan, 2017). As should be clear from the
examples I  have discussed, the questions they sought to answer arose not from an investigation into the
physiological constitution of the nervous system, but from pre-physiological assumptions about the human
mind. Theories of physiological aesthetics consistently treated this mechanism as distinct, in the sense that it
was thought to be characterized by a specialized set of properties: sensory, cognitive, or emotional acts that
distinguished aesthetic experiences as psychological acts.
What is also clear is that the success of these theories was rather limited. While the description of the
nervous system improved considerably during the 19th century, a fundamental gap remained between what
was understood at the level of nerve activity and what was understood at the level of psychological function
(Ochs, 2004). Theories of aesthetics inspired by physiology were often greeted by initial enthusiasm but
were then eventually found lacking. Perhaps no example illustrates this trajectory better than the notion of
Einfühlung (empathy), launched by Robert Vischer in 1873 as a key psychological function meant to explain
the existence of aesthetic experiences (Allesch, 2017). Vischer argued that aesthetic experiences were not
the result of “mere” perception; rather “the contemplative eye . . . reconstructs the forces expressed by the
motion of objects through ‘reproductive empathy’ [reproduktive Einfühlung]” (Allesch, 2017, p. 229). This re-
constructive process, Vischer claimed, involved an object-matching activation of perceptual, emotional, and
motor systems. The concept of Einfühlung was soon taken up by Theodor Lipps and proved very popular for
several years, indeed well into the 20th century (Allesch, 2017). However, as it became clear that the emo-
tional content of works of art rarely produced an identical correlate in the observer, it was abandoned again,
only to seem, by the 1930s, never to have existed at all (Allesch, 2017).
At the same time, the central ambition of physiological aesthetics, to explain aesthetic experiences as
a function of physiological mechanism, came under attack by philosophers who questioned the ability
of physiological aesthetics to reduce mental states to a physiological psychology. For example, reacting to
Fechner’s aesthetics “from below,” the Austrian philosopher Franz Brentano argued that it was impossible to
explain the mental acts involved in aesthetic experiences by recourse to physiology; Fechner’s approach was
consequently doomed to failure (Höge, 1995). This was in fact a common criticism levelled at any attempt
to reduce the human mind to psychological explanation that eventually, around 1900, erupted into the so-
called Psychologismusstreit (Kush, 1995). Philosophers such as Gotlob Frege and Edmund Husserl claimed that
higher-order thought, as a matter of principle, could not be reduced to psychology; any adequate description

12
History of neuroaesthetics as a discipline

of, say, logic would require its own special science. This anti-psychologist argument was soon adopted by
other scholars in the humanities, leading to a rapidly expanding rejection of psychology and neuroscience as
means to explain human thought and behaviour (Smith, 1997).
Together, the general failure of physiological theories of aesthetics and the mounting anti-psychologism
conspired to make any neuroscience of aesthetics seem like a pipe dream. The complex constitution of
the aesthetic experience appeared far removed from the explorations of cell structure, anatomy, and neural
activity that were taking place in early 20th-century neuroscience. As a result, by the 1930s, attempts to do
descriptive neuroaesthetics had dwindled to almost nothing and seemed to no longer play any role in the
further advancement of aesthetics—even amongst psychologists engaged in empirical aesthetics research.3
This situation only started to change in the 1960s when the English psychologist Daniel Berlyne began to
publish on what he would himself come to call a new aesthetics (1966, 1974). Berlyne proposed that aesthetic
evaluation could be understood as a specific form of exploratory behaviour, with hedonic liking determined
by degree of autonomic arousal (1971). Depending on how surprising, complex, or novel a sensory object
was perceived to be, the brain would respond with different levels of arousal. Those objects able to elicit high
levels—albeit not abnormally high levels—of arousal were experienced as pleasing (Berlyne, 1971).
Berlyne’s theory of aesthetics was directly inspired by new neurophysiological work that had begun
to connect a range of autonomous processes—skin conductance, pupil dilation, respiration, and muscular
tension—to neural activity in localized parts of the brain, especially the reticular formation of the brain stem,
the hypothalamus, and the thalamus (Fuller et al., 2011; Swanson, 2000). Building on this work, Berlyne
was able to formulate a theory that actually attributed psychological function to a specific neurobiological
mechanism. No more vague invocations of “nerve movements,” “normal amount of activity,” or “aesthetic
apprehension.” Berlyne’s theory postulated a physiological cause of aesthetic liking that could be proven
either right or wrong. However, while Berlyne’s work certainly had a galvanizing effect on experimental aes-
thetics research in the 1960s and 1970s, it failed to produce a neuroscientific effort to test its core hypothesis
(Martindale, 2007; Machotka, 1980; Silvia, 2005). Examination of Berlyne’s theory remained purely behav-
ioural (e.g., Berlyne, 1974; Martindale et al., 1990; Nadal et al., 2010), at least until experimental scrutiny of
arousal and reward became possible in the context of experimental neuroaesthetics (see Chapter 2).
A second finding from the neurosciences that came to greatly influence theories of aesthetics was the
realization that the visual system is functionally specialized (Felleman  & van Essen, 1991; Livingstone  &
Hubel, 1988; Schiller, 1997; Zeki, 2005). By mapping neuronal responses to a variety of stimulus features,
vision neuroscientists had discovered that nuclei located in different parts of the visual system coded for
different stimulus features. Moreover, the first (posterior) cortical areas to receive projections from the eyes
appeared to break down and analyse the visual scene in terms of individual features such as edges, brightness,
movement, or colour. Only as information was projected forward to later (anterior) parts of the occipital and
temporal lobes were these elements synthesized into meaningful object recognition (Livingstone & Hubel,
1988; Schiller, 1997; Zeki, 2005). Collectively, these observations fostered two important principles that
would inform aesthetics researchers’ understanding of both visual art in particular and aesthetic experiences
more generally: First, the principle that individual functional properties of perception could be attributed to
segregated anatomical regions and, second, the principle that perceptual-cognitive experiences seemed to be
constructed by a sequence of processing steps, where information was projected from module to module.
The first principle, finally, opened the door to an understanding of how the brain represents the content
of aesthetic objects, especially works of art (Changeux, 1994; Gregory et al., 1995; Latto, 1995; Livingstone,
1988; Solso, 1994). With a detailed description of the way sensory stimuli were computed by the brain, it
became possible to ask how particular objects such as representative or abstract paintings engaged these neural
nodes. Such a “framework,” in the words of Chatterjee (2003, p. 55), offered the “promise of sorting out the
perceptual and cognitive aspects of aesthetic viewing, as well as the emotional response to beauty” (p. 59).
The second principle provided a key to solving the conundrum of how aesthetic experiences could arise

13
Martin Skov

from a combination of both “bottom-up” and “top-down” processes, thus possibly revealing which parts of
brain underpinned Fechner’s direct and associate factors, as well as explaining how they came to interact.
The first question was pursued by two comparable but slightly different approaches. Both utilized the
advances in vision neuroscience to provide a description of the way the human brain constructed aesthetic
experiences from the physical properties inherent to artworks but had different conceptions of the nature of
these experiences. The first approach, spearheaded by Margaret Livingstone, Patrick Cavanagh, and others,
considered art a visual object on par with other visual objects, subject to the same computational rules as
other stimuli. What set art objects apart from ordinary visual objects was the way artists could create novel,
sometimes unique, strategies of presenting visual information that were able to exploit these general rules.
As such, the composition of art objects often contravened the physics of everyday perception and gave rise
to forms of visual “illusions” (Cavanagh, 2005; Conway & Livingstone, 2007; Grossberg, 2005; Livingstone,
1988, 2002). By mining insights from vision neuroscience, the ambition was to explain how such “illu-
sions” worked. The other approach, principally associated with Semir Zeki (1999a, 1999b), suggested that
art objects were better understood as visual stimuli that served a specific perceptual and cognitive function.
Rather than perpetually inventing novel ways to exploit the brain’s computational mechanisms, Zeki (1999a)
hypothesized that artists created objects that were specifically designed to provoke the acquisition of knowl-
edge about the world. The function of art, as he wrote in the book Inner Vision, was

to represent the constant, lasting essential and enduring features of objects, surfaces, faces, situations, and so
on, and thus allow us to acquire knowledge not only about the particular object, or face, or condition
represented on the canvas but to generalize about a wide category of objects or faces.
(Zeki, 1999b, pp. 9–10; italics in the original)

In this way, the ambition of Zeki was not only to explain how individual artworks elicited particular
experiences by engaging specific neural mechanisms but also to provide an overarching theory of what art
is—an ambition not altogether dissimilar from the way theorists of evolutionary aesthetics have tried to iden-
tify the adaptive purpose of art (Davies, 2012).
What united both approaches, though, was a preoccupation with individual stimulus features. Attempts
to explain art experiences as emerging from an architecture of functionally specialized modules favoured
bottom-up descriptions of how individual stimulus features come to engage different pathways of the visual
system. At the same time, it was clear that top-down processes associated with knowledge and experience
had to play an important role in mediating responses to perceived stimulus features. The first theorist to sug-
gest a theory that attempted to account for the role played by top-down processes in aesthetic experiences
was Robert Solso (1994), who attributed the “meaning, or semantic value, derived from . . . basic forms” to
the “(internal) context in which art is viewed” (p. 102). These contextual components, he claimed, aroused
expectations that influenced the way the perceptual properties of an artwork is computed. Solso (1994)
applied the information-processing notion of schemata to argue that later (more anterior) parts of the visual
system activate “art schemata” that impose subjective knowledge on the interpretation of what is seen by
modulating the processing of bottom-up information (Solso, 2003).
The work on crafting models to account for aesthetic experience based on a visual information-
processing architecture came to its conclusion with the publication by two papers by Anjan Chatterjee
(2003) and Helmut Leder and colleagues (2004). Both models were ambitious in their goals by explicitly
aiming to suggest a framework with which it was possible to understand aesthetic experiences in their total-
ity. They accomplished this goal by enumerating computational modules hypothesized to be specific to
aesthetic responses (see Figure 1.5). Both models included modules involved in perceptual analysis, modules
involved in cognitive “mastering” of the artwork, and modules subserving two forms of “output”: On the
one hand aesthetic “decisions” (Chatterjee, 2003) or “judgments” (Leder et  al., 2004) and on the other

14
History of neuroaesthetics as a discipline

Figure 1.5 Chatterjee’s (2003) and Leder et al.’s (2004) two models of the neural processes involved in aesthetic experience
combined into one representation. Both models take their departure from the finding that the visual system is
functionally specialized. Despite a slight disagreement—Chatterjee’s modules are depicted in white, Leder and
colleagues’ in grey—with respect to the precise nature of the functional modules hypothesized to underpin
aesthetic experience, the two models concur that aesthetic responses involve three hierarchically organized
processing stages: an initial stage of perceptual processing, a later stage of cognitive “mastering,” and a decision
stage leading to outputs of aesthetic judgments and emotions. Figure adapted from Vartanian and Nadal (2007).

“emotional responses” (Chatterjee, 2003) or “aesthetic emotions” (Leder et al., 2004). Together, the two
models articulated earlier intuitions by codifying three “stages” of any aesthetic experience—perceptual
analysis, meaning construction, and aesthetic judgment—and naming candidate processes that might make
up the individual stages. In combination they set up a potential new research program outlining “a general
framework for aesthetic experience at the psychological level” that could also “be tested experimentally using
biological methods at a more micro level” (Vartanian & Nadal, 2007, p. 442).

Neuroaesthetics as an experimental science


Even a brief survey of historical efforts to apply neuroscience to theories of aesthetic experience makes two
things clear: (1) The quest to explain the mechanisms and functions of aesthetic experience in neurobio-
logical terms originated with the philosophical invention of aesthetics as a concept. Indeed, all the main
explananda pursued by any neuroscience of aesthetics derive from the philosophical formulation of aesthetics
as a psychological category. (2) It is far from a novel enterprise. To claim, as is sometimes done in articles
on neuroaesthetics, that inquiries into the neural bases of aesthetic experiences only began with advent
of modern neuroaesthetics is historically misleading and dilutes the degree to which much contemporary
research consists in a re-examination of problems already honed by previous generations. To name just three
examples, all of the following problems first posed by physiological aesthetics are still considered central to
contemporary neuroaesthetics research:

1 Are aesthetic responses determined by the physical properties of the aesthetic object or by the internal conditions
of the neural processes that react to the information acquired from the stimulus? As we have seen, this
question has recurred in different forms over the last two and a half centuries, first in the form of an
interaction between sense impressions and nerve sensibilities, then as an interaction between Fechner’s
direct and associative factors, and more recently as a transfer of information between “bottom-up” and

15
Martin Skov

“top-down” processing modules. It has gone on to form a cornerstone of contemporary neuroaesthet-


ics, with the vast majority of studies of aesthetic evaluation dedicated to understanding the role played
by neural processes associated with perceptual analysis, memory, conceptual knowledge, or framing, in
determining aesthetic liking (see reviews of this body of work in the other chapters collected in Part 1
of the book). To what degree liking and disliking are primarily encoded by stimulus features or by the
“eye of the beholder” remains an unresolved and contentious topic.
2 Is it possible to distinguish aesthetic experiences from other psychological states? The idea that the human brain
might contain neural mechanisms that code for distinct aesthetic responses to sensory objects emerges
from the philosophical proposition that aesthetic evaluations spring from the application of a distinct
evaluative mechanism to a specific set of sensory qualities. It immediately gave rise to an enduring quest
to identify the neural processes that are specific to the “aesthetic” processing of sensory stimuli. As
described previously, this endeavour has historically focused on modelling those perceptual-cognitive
and emotional processes that are unique to aesthetic experiences, especially processes that are exclusive
to the representation of art objects and aesthetic emotions. If anything, the field of neuroaesthetics is
founded on the belief inherited from these predecessors that these questions merit an answer. Thus, the
experimental work published under the moniker of neuroaesthetics can be viewed as being concerned
with two overarching questions: First, how different categories of art—first music and visual art, later
also dance, architecture, and literature—are represented by the perceptual, cognitive, and emotional sys-
tems of the brain (see the chapters collected in Part 2 of the book) and, second, how the brain is engaged
when it evaluates these and other objects considered aesthetic for their perceived hedonic value (Part 1).
3 Are aesthetic experiences intrinsically connected to works of art? The first theories of sensory evaluation to be
called theories of aesthetics, after Baumgarten’s invention of the word, did not view aesthetics as specifi-
cally concerned with art objects. The establishment of the art object as the main elicitor of aesthetic
experiences only happened as a consequence of the 19th-century events I  reviewed in the previous
section. For this reason, there has since been a lasting tension in the psychological and neuroscientific
research that seeks to explain the material bases of aesthetic experiences. Since the time of Fechner and
Grant Allen, people have failed to come to an agreement on whether aesthetic evaluations only arise in
response to art objects or if they apply to all categories of stimuli. If we accept that aesthetic experiences
can be elicited by some non-art objects, which are these? Abstract visual forms, faces, everyday artefacts?
There is no accepted answer to this question, even today. To experimental neuroscientists, especially
those preoccupied with measuring neural responses to specific stimuli and tasks, this question continues
to matter a lot. Is neuroaesthetics research defined by the use of artworks as stimuli in experiments? Are
the “outcomes” of aesthetic responses to sensory objects—whether in the form of “judgments” or “aes-
thetic emotions”—unique or general? At the most fundamental level, individual researchers conceive of
neuroaesthetics by harboring different answers to these questions. I will return to this issue below.

In sum, the scientific enterprise that starts to take form in the early 2000s under the name of “neuroaesthet-
ics” is best understood as an extension of an existing aspiration to account for the phenomenon of aesthetic
experience in neurobiological terms. It effectively takes on problems that have crystallized over the preced-
ing two centuries. What is novel about neuroaesthetics is that it succeeds in transforming what had been,
by necessity, a purely descriptive neuroscience of aesthetics into an experimental science. The key to this
change was a revolution in experimental methods that took place over the course of the 1980s and 1990s.
These inventions made it possible to measure neural activity in living humans as they engaged in tasks con-
sidered aesthetic in nature. More than anything, neuroaesthetics in this modern sense owes its existence to
this historical development.
Of greatest importance to the emergence of neuroaesthetics was the invention of a range of non-invasive
neuroimaging methods, especially positron emission tomography (PET), functional magnetic resonance

16
History of neuroaesthetics as a discipline

imaging (fMRI), and magnetoencephalography (MEG), as well as electroencephalography (EEG). But it


is important to stress that neuroimaging methods were not the only advancements in methodology that
helped make problems in aesthetics amenable to experimental scrutiny. In addition to methods that allowed
for access to neural processes, methods that enabled the control and presentation of art stimuli also played a
significant role. These included the availability of technical equipment such as colour scanners that facilitated
the reproduction and handling of visual artworks (Helmut Leder, personal communication) and the devel-
opment of computer programs such as MATLAB and Eprime that enabled a controlled presentation of art
stimuli (Gerry Cupchik and Chris McManus, personal communication). Even as late as the 1980s, such tasks
had been more than a little cumbersome.
Another important development was the appearance of computational methods allowing for a charac-
terization of the physical and perceptual constitution of works of art—infamous for their often inscrutable
complexity. For example, using statistical methods such as Fourier analysis, backed up by ever-increasing
computer power, it became possible to reveal regularities in paintings and to ask how these deviated from
non-art visual scenes (e.g., Graham & Redies, 2010; Redies, Hasenstein et al., 2007; Redies, Hänisch et al.,
2007). Similarly, music researchers demonstrated that computational models can be employed to describe the
statistical relationship of tones in melodies and how these regularities become encoded as predictive models
when people listen to them (e.g., Pearce, 2018; Pearce & Wiggins, 2005). Importantly, understanding the
informational content being perceived in the context of an aesthetic experience can greatly aid the neu-
roscientist in exploring the specific neural correlates of individual art stimuli (Brachmann & Redies, 2017;
Cheung et al., 2019; Gold et al., 2019; Pearce et al., 2009).
Finally, developments in eye-tracking equipment facilitated important explorations into physiological
processes associated with aesthetic experience before the advent of neuroimaging. Since eye movements
reflect the aspects of the visual scene we fixate on, the ability to measure fixations and gaze patterns helps
inform our understanding of what perceptual information the visual system is attending to and processing
(Martinez-Conde et al., 2004). Attempts to build measurement systems able to measure eye movements go
well back to the 19th century. One such early device was used by Stratton (1902) to test the hypothesis that
people like curved contours because their smooth, uninterrupted lines afforded easier gaze patterns than
sharp or angular lines. In violation of his own assumptions Stratton (1902) found both kinds of contour to
elicit jerky and discontinuous microsaccades. Later, in the 1960s Yarbus (1967) invented an eye-tracking
“cap” that could be attached to the eyeball by suction. Using this system, Yarbus (1967) described several per-
ceptual experiments, including one demonstrating that different task instructions modulated the way the vis-
ual system scanned a painting for information. In the 1970s and 1980s, Yarbus’ research inspired several vision
scientists, including Calvin Nodine and Paul Locher, to start using visual art objects to explore how visual
attention and scan patterns are informed by symmetry (Locher & Nodine, 1977), artistic style (Nodine &
McGinnis, 1983), art expertise (Nodine et al., 1993), or degree of abstractness (Zangemeister et al., 1995).
With modern, laser-based eye-trackers, it has since become possible to also measure how dilated the pupil
is at a microsecond interval, allowing for a coupling of visual fixation and autonomic arousal. Several recent
experiments have integrated this advance into the field of neuroaesthetics. For example, Ramsøy and col-
leagues (2012) used the analysis of pupil dilation change to show that paintings experienced in unpredictable
contexts are less liked than when experienced in predictable contexts, a variance in preference explained by
a difference in the ability of predictable and unpredictable contexts to elicit physiological arousal.
Still, these important predecessors notwithstanding, it was the introduction of neuroimaging to the wider
study of human cognitive neuroscience that more than anything helped make neuroaesthetics possible as a
viable experimental study. Cognitive neuroscience itself was a neologism first coined in the 1970s by George
Miller and later popularized by Michael Gazzaniga (1988). It was initially thought of as a program for bridg-
ing the gap between neuroscience and the emerging field of cognitive science (Churchland & Sejnowski,
1988). However, the neural correlates of complex cognitive thought such as language or reasoning were

17
Martin Skov

difficult to study in humans, expect for the rare situations where neuropsychological cases permitted testing
of cognitive states. Such studies were marred, though, by the fact that they rested on the problematic business
of inferring function from brain damage that was not only arbitrary but also could be difficult to describe
accurately.4 Without the development of methods for recording neural activity in a non-invasive manner,
either in the form of electrical and magnetic field recordings (Hillyard, 1993) or deduced from changes in
local blood flow and metabolism (Raichle, 1998), cognitive neuroscience might well have remained a pipe
dream. With such methods starting to emerge in the 1980s, however, the possibility of measuring the loca-
tion, functional character, and temporal unfolding of neural signals while human subjects engaged in cogni-
tive tasks finally became possible. As they became more widely adopted in the 1990s, a virtual revolution in
human neuroscience took place, ushering in a multitude of individual cognitive neuroscience fields dedi-
cated to any conceivable aspect of higher-order human psychology: economics, morality, social behaviour,
politics, and so on. Neuroaesthetics was borne out of this general transformation of the field of neuroscience.
The first aesthetic topic to be systematically probed by use of neuroimaging techniques was music percep-
tion (Peretz & Zatorre, 2005; Tervaniemi, 2009). As a temporal art form, the perception of which relies on
the processing information organized sequentially in time, music lends itself well to studies conducted with
EEG. Since EEG can record electrical activity produced by cortical cells on a millisecond scale (Hillyard,
1993), it is well suited as a tool for tracing neural activity unfolding in responses to music stimuli. In the
1980s, music researchers increasingly began to use EEG—especially by measuring event-related potentials
(ERPs)—to characterize the neural correlates of music perception (e.g., Besson & Macar, 1987; Klein et al.,
1984; Paller et  al., 1988; Pantev et  al., 1989; Wieser  & Mazzola, 1986). These experiments sought to
describe the neural mechanisms underpinning the key perceptual constituents of music such as pitch or
the grouping of tones into chords and melody. The ultimate hope was to ascertain if the neural processes
involved in music perception were distinct or shared with other forms of auditory perception, including
language. Unfortunately, the ability of EEG to localize the anatomical source of neural is limited, so when
access to PET and fMRI improved in the 1990s, these techniques, with their superior ability to localize neu-
ral activity, were soon taken up by music scientists as well (e.g., Griffiths et al., 1999; Hughdal et al., 1999;
Sakai et al., 1999; Zatorre et al., 1992).
Compared to music, neuroimaging studies of visual art perception were much slower to appear. A few
EEG experiments reporting neural responses to complexity in visual images had been published in the 1960s
and 1970s (e.g., Baker & Franken, 1967; Berlyne & McDonnell, 1965; Gale et al., 1971, 1975). But while
these experiments were motivated by Berlyne’s theory of aesthetics, they did not employ traditional examples
of visual art as stimuli, and it was also unclear how to infer function from their reported observations. One
EEG experiment conducted by Nicki and Gale (1977) reported alpha abundance to increase with complex-
ity in abstract paintings. This finding they interpreted, not quite convincingly, as evidence for a depression
of “cortical arousal.” Perhaps disappointed by the obscurity of such results, interest in using neuroimaging
to investigate visual art perception seems to have waned in the 1980s, only to be resurrected by Zeki and
Marini’s (1998) fMRI study on how the visual system represents irregular colour-object combinations in
Mondrian paintings. Even in the 2000s, though, neuroimaging of visual art perception remained spotty, pos-
sibly because it was not considered sufficiently dissimilar to ordinary vision to warrant its own study (though
see, e.g., Fairhall & Ishai, 2007; Kirk, 2008).
Only at the end of the 1990s did the focus on perception widen to encompass other topics of interest to
aesthetics, especially emotional responses elicited by art and the neural correlates of aesthetic liking. Since
most neuroimaging studies centred on these two issues concentrated on understanding what happens when
music or visual art generate pleasure or displeasure, they often pursued both questions without quite distin-
guishing between what it means to experience pleasure as a consequence of different forms of (evaluative
or not) engagements with art. Collectively, PET and fMRI experiments of revealed that aesthetic pleasure
and displeasure correlate with activity in the mesolimbic reward circuit combined with modality-specific

18
History of neuroaesthetics as a discipline

engagements of auditory (Blood et al., 1999; Blood & Zatorre, 2001; Brown et al., 2004; Koelsch et al., 2005;
Menon & Levitin, 2005) or visual processes (Cela-Conde et al., 2004; Kawabata & Zeki, 2004; Vartanian &
Goel, 2004a), depending on the stimulus employed. Most of the music experiments, though, eschewed sub-
jective reports of hedonic experiences, relying instead on a pre-determined categorization of stimuli as either
pleasurable or displeasurable, while the visual art experiments all based their analyses of neural correlates on
locating correlates to liking ratings. Together they raised the question of how much pleasure and displeasure—
as well as, possibly, other emotional states—are modulated by conditions under which artwork are appraised.
A third, related topic was introduced by a series of experiments starting to appear in the early 2000s
(Brattico et al., 2003; Jacobsen & Höfel, 2001, 2003). Using EEG, these studies compared neural activation
elicited by aesthetic and descriptive judgments of graphic patterns and musical cadences. Thus, rather than
asking how the brain computes liking and disliking in the form of hedonic responses, they sought to identify
a general neural system for actively enacting aesthetic evaluations. Findings, supported by results reported
later by fMRI experiments (e.g., Chatterjee et al., 2009; Ishizu & Zeki, 2013; Jacobsen et al., 2006), sug-
gested that sensory and emotional systems are modulated differently when sensory objects are attended to
with the explicit purpose of forming an aesthetic judgment than when they are not.
It is not surprising that the gradual embrace of neuroimaging methods to pursue topics of interest to
aesthetic theory followed a trajectory similar to other ventures into cognitive neuroscience. By concentrating
its initial efforts on perception and memory, the budding cognitive neuroscience program could apply the
risky, new techniques to aspects of human cognition that already were being studied in behavioural labs and
of course were thought to share neural architecture with other animals already under neuroscientific scrutiny.
Because it was possible to scan participants using tasks that were well established by psychological experi-
ments, as well as stimuli that were believed to be well controlled, studies of perception in different functional
contexts seemed the safest bet to get human cognitive neuroscience off the ground. It took most of the 1990s
to edge neuroimaging research closer to the exploration of cognitive states that were considered subjective
and less “rational.” An important force in driving this change was the growing realization, especially fol-
lowing the work on patients with ventromedial prefrontal lesions pioneered by Antonio Damasio, Antoine
Bechara and Daniel Tranel, that emotional states contributed significant inputs to both cognitive reasoning
and behavioural motivation (Dukes et al., 2021). The ability to employ neuroimaging techniques to explore
the neural underpinnings of psychological states considered important to aesthetics benefitted from this gen-
eral evolution, opening research slowly up to include complex subjective responses to works of art.
Still, topics relevant to aesthetics were embraced only incrementally. Very few neuroimaging studies
published in the 1990s were dedicated to the investigation of either art perception or emotional process-
ing related to aesthetic liking. Moreover, it is doubtful if any of the researchers involved in neuroimaging
experiments on such topics pre-2000 thought of their work as part of a common “neuroaesthetics” endeav-
our. Indeed, there are good reasons to believe that none of the music researchers who, in the 1980s and
1990s, took up EEG, PET, or fMRI to explore the neural mechanisms of music experience considered their
research part of a broader inquiry into human aesthetics (Robert Zatorre, personal communication). As
I began this chapter by showing, word usage data suggests that the word neuroaesthetics itself only became
widely adopted after 2000 (Figure 1.1A). This begs the question of how a scattered and uncoordinated series
of ventures into “neuroaesthetics” was transformed into a bona fide scientific discipline and why this process
began to take place in the early 2000s.
While more historical research is needed to provide a clear answer to this question, I will suggest that
three events played an important role. First, the popularisation and general acceptance of the word neuroaes-
thetics was itself important to the fostering of a collective identity. Its invention has often been attributed
to Semir Zeki, but it is doubtful he was the first to use it. The first reference to “neuroaesthetics” I have
been able to find appears in the introduction to the 1988 book Art and the Brain. Here the word is used to
identify an “emerging field” dedicated to research on the “biological foundations of aesthetics” (Rentschler

19
Martin Skov

et al., 1988, p. 10). Art and the Brain grew out of a series of meetings convened by the German Studiengruppe
Biologische Grundlagen der Aesthetik in the early years of the 1980s. It is possible that someone involved in
these meetings was the first to coin the word, although none of the participants I have spoken to remember
who the originator was (Chris McManus and Ernst Pöppel, personal communication). It is true, though,
that Zeki—who first seems to have referred to the neuroscientific study of art as “neuro-esthetics” (with a
hyphen) in Inner Vision (Zeki, 1999b, p. 2)—would go on to play an outsized role in popularizing the use of
the word in years between 2000 and 2005. In any case, regardless of who deserves the accolades for having
invented it, I think it is evident that having one catchy name to rally around greatly helped foster the idea
that the experimental investigation of the neurobiological basis of art and aesthetics should be considered a
single scientific enterprise.
The second factor I would point to is an emerging desire to establish neuroaesthetics as a scientific enter-
prise that happened to gain momentum between 2000 and 2005. It is possible that a growing number of
researchers began to see an experimental neuroaesthetics as a possibility because the advances of the 1990s
had demonstrated the feasibility, at least in principle, of studying the neurobiological basis of art and aesthet-
ics. But it also true, I think, that the advocacy that this desire gave rise to (e.g., Chatterjee, 2003; Hagendoorn,
2003; Ramachandran & Hirstein, 1999; Skov, 2005; Zeki, 1999a, 1999b; Vartanian & Goel, 2004b) in many
ways ended up functioning as a self-fulfilling prophecy. For example, two articles by Ramachandran and
Hirstein (1999) and Zeki (1999a), published together in Journal of Consciousness Studies, had a huge impact
on a broad audience of readers. Both articles argued that any theory of art and aesthetic liking that did not
consider the neural mechanisms underlying aesthetic experience was doomed to fail. As emphatic calls to
arms, they both provoked a lot of excitement and a large number of detractors. In retrospect, neither paper
contained any theoretical idea—such as Zeki’s (1999a) claim that art serves the purpose of acquiring lasting
knowledge or Ramachandran and Hirstein’s (1999) famous hypothesis that the peak shift principle is one
of eight universal rules that account for aesthetic experience—that would go on to influence the course of
experimental neuroaesthetics. As rallying cries, however, they undoubtedly played a crucial role in putting
the idea of neuroaesthetics on the map.
Finally, I will point to the process of working out a theoretical foundation for neuroaesthetics as a third
factor that helped establishing and solidifying neuroaesthetics as an experimental discipline. To clamour for
an experimental neuroscience of aesthetics was to invite several tough questions about the objectives and
aims of such a science. What problems, exactly, should the intended field of neuroaesthetics pursue? On an
abstract level, most evangelists of neuroaesthetics could probably agree that its main topics were to be the
phenomena of art experience and aesthetic liking. But defining the precise scope of these concepts was a
less self-evident endeavour. What should count as “art” for example? As noted, it was not clear to music
researchers that music fell within the scope of neuroaesthetics. Did instances of graphic design and product
packaging qualify? Similarly, it was unclear what was to be meant by “aesthetic liking.” Should the study of
neuroaesthetics include experiments where hedonic pleasure was elicited by faces, artefacts, or even food and
drink? Or was it to be restricted only to the study of pleasure responses elicited by fine art stimuli? Further-
more, the new field had to content with criticisms from philosophers who argued that to even consider using
neuroscience to explain art and aesthetic liking was a category mistake (e.g., Hyman, 2010; Tallis, 2008).
Looking back on the period between 2000 and 2010, it is striking how much energy was channelled into
providing convincing answers to these troublesome questions. A flurry of theoretical papers that sought to
not only justify the existence of neuroaesthetics as an experimental discipline but also to clarify what type of
problems would fall within its purview saw the light of day (e.g., Brown & Dissanayake, 2009; Chatterjee,
2003; Jacobsen, 2006, 2009; Skov, 2009; Skov & Vartanian, 2009; Vartanian & Goel, 2004b; Vartanian &
Nadal, 2007). As findings from neuroimaging experiments began to slowly appear, this debate was further
augmented by discussions about how to make sense of experimental results and to what degree these sup-
ported this or that theoretical model of art experience and aesthetic liking (e.g., Di Dio & Gallese, 2009;

20
History of neuroaesthetics as a discipline

Nadal et  al., 2008; Skov, 2007, 2010; Vartanian, 2009). In my view, this theoretical work played a very
important role in transforming the budding idea of having a neuroscientific study of aesthetics into an actual
scientific discipline concerned with a specific set of problems and a methodological framework for experi-
mentally examining these problems.

Conclusion
Neuroaesthetics is a very young discipline. The historical record informs us that it only began taking shape as
a concerted, experimental science in the 2000s (Figure 1.1). Indeed, most of that decade served as a Gründer-
zeit, establishing the parameters of neuroaesthetics as an experimental science. Only in the last 10 years or
so has neuroaesthetics expanded into a viable scientific discipline as a growing number of researchers and
labs have become attracted to problems associated with art experience and aesthetic liking. Neuroaesthet-
ics was not created ex nihilo, though. Before neuroaesthetics, there were 250 years of sustained attempts to
explain the physiological bases of aesthetic experience by employing results from mainstream neuroscience.
I have tried to show how these efforts were motivated by the belief that the human mind is endowed with
a dedicated psychological mechanism for generating such experiences. This belief did not originate in any
empirical observation of human psychology or neuroscience but was based on philosophical arguments. Of
course, such a historical trajectory is not unique to neuroaesthetics. As pointed out by György Buzsáki (2020)
and others, it is the rule more than the exception that the objective of neuroscientific research has been
conceived of as providing neurobiological mechanisms for psychological and behavioural concepts inherited
from philosophy and folk psychology.
Casting light on the historical roots of neuroaesthetics is not only important because it reveals an intellec-
tual history that is fascinating in itself but also because understanding these roots will help explain why neu-
roaesthetics emerged as an experimental science concerned with a specific set of hypotheses and problems.
For example, the ambition to identify a neural system underpinning aesthetic evaluations, or the conviction
that art objects such as visual art or music elicit unique perceptual, cognitive, and emotional responses, both
had their origin in this prehistory. Indeed, any of the controversies witnessed by neuroaesthetics over the
past two decades can conceivably be understood as attempts to reconcile emerging experimental findings
with the conceptual roots of the field. This includes debates over whether neuroaesthetics is a science of art
or a science of aesthetic liking (or perhaps a science of the aesthetic liking of art; Brattico & Pearce, 2013;
Brown & Dissanayake, 2009; Pearce et al., 2016; Skov & Nadal, 2020a) or whether emotional responses such
as pleasure which are encountered as part of aesthetic experiences can be distinguished from other emotional
states (Chatterjee & Vartanian, 2014, 2016; Skov & Nadal, 2020a, 2020b).
As the other 29 chapters in this book demonstrate, neuroaesthetics has evolved tremendously since its
early years. Where the period between 2000 and 2010 in many ways was a struggle to get a new discipline
off the ground, the last decade has produced a historically unprecedented amount of novel insights into the
neural mechanisms associated with art experience and aesthetic liking. Many of the ancient problems that
have preoccupied scholars for centuries are now finding empirical solutions or are being reformulated based
on the rapid improvement in our understanding of human neurobiology. Some of the assumptions that moti-
vated the intellectual history I have sketched in this chapter are becoming obsolete. But they are only ceasing
to be relevant to theories of art experience and aesthetic liking because neuroaesthetics instigated a program
for experimentally investigating their provenance. In this sense, neuroaesthetics has been a true success story
(Skov, 2019).
Perhaps the best indication of its success as a research program is the way neuroaesthetics has expanded in
both the number of researchers and scientific output over the last 10 years. Especially in the last five years, the
field has seen an unprecedented influx of new researchers and labs, many based outside Europe and North
America. For instance, neuroaesthetics is currently expanding aggressively in Asia, especially in China, where

21
Martin Skov

many universities have begun allocating funding to the establishment of dedicated research centres. These
groups, such as the Neuroaesthetics Center at Guangzhou University, often host a large number of both sen-
ior researchers and students who pursue many different topics (Xianyou He, personal communication). It is
quite possible that, as a result of this development, neuroaesthetics will start to move in quite new directions,
taking up questions that were unknown to its historical ancestors. Parts of this future we are already getting
a glimpse of: Neuroaesthetics researchers are increasingly shifting their focus to the investigation of the pos-
sible health benefits of aesthetic experiences or the existence of aesthetic brain systems in other animals (see
Chapters 2, 11, 13, 20, and 21 for descriptions of this new research effort). A new generation for whom
neuroaesthetics is a given, not a mere possibility, is taking over the field and will get to define its purpose in
new ways. The next 20 years will no doubt be even more interesting than the first 20.

Acknowledgements
I want to thank the following people for generously agreeing to either talk to me, respond to my inquiries,
or comment on an earlier draft of this chapter: Bevil Conway, Anjan Chatterjee, Gerry Cupchik, Xianyou
He, Thomas Jacobsen, Stefan Koelsch, Helmut Leder, Paul Locher, Chris McManus, Gisèle Marty, Marcos
Nadal, Ernst Pöppel, Chris Redies, Oshin Vartanian, and Robert Zatorre. Naturally, none of the historical
interpretations presented in this chapter should be attributed to anyone but me.

Notes
1 It is worth noting that neuroaesthetics is far from the only branch of neuroscience that has inherited most of its central
explananda from philosophy and folk psychology. See Buzsáki (2020) and Pessoa et al. (2021) for important discussions
of how this state of affairs can be argued to have led neuroscience astray in important respects.
2 Unfortunately, to the best of my knowledge, no historical study of the intersection between philosophy of mind and
the development of aesthetic evaluation as a category exists.
3 For example, reading Charles W. Valentine’s (1962) book The Experimental Psychology of Beauty, you will be hard
pressed to find any mention of possible neuroscientific explanations for why we find sensory objects beautiful. Tell-
ingly, this book was Valentine’s attempt to summarize work in empirical aesthetics that primarily spanned the period
between 1920 and 1960.
4 Not surprisingly, this also applies to the use of neuropsychology in aesthetic theory. Over the years, many case studies
of artists and other patients have been reported in the literature (reviewed by Zaidel, 2005). However, the impact of
these studies on aesthetic theory has been minimal. Chatterjee (2009) attributes this lack of influence to the anecdotal
nature of many case reports, their absence of measurement, and the endemic inaccessibility of most publications that
have reported relevant neuropsychological case studies.

References
Allen, G. (1877). Physiological aesthetics. D. Appleton and Co.
Allesch, C. G. (1987). Geschichte der psychologischen Ästhetik. Hogrefe.
Allesch, C. G. (2017). Einfühlung—A key concept of psychological aesthetics. In V. Lux & S. Weigel (Eds.), Empathy,
Palgrave studies in the theory and history of psychology (pp. 223–243). Palgrave Macmillan.
Allesch, C. G. (2018). Psychology in emerging aesthetics. In L. Tateo (Ed.), An old melody in a new song. Theory and history
in the human and social sciences (pp. 33–51). Springer.
Ayala, F. J. (2017). Adaptive significance of ethics and aesthetics. In M. Tibayrenc & F. J. Ayala (Eds.), On human nature.
Biology, psychology, ethics, politics, and religion (pp. 601–623). Academic Press.
Baker, G., & Franken, R. (1967). Effects of stimulus size, brightness and complexity upon EEG desynchronization. Psy-
chonomic Science, 7(9), 289–290. https://doi.org/10.3758/BF03328565
Baumgarten, A. G. (1735/1954). Reflections on poetry: Alexander Gottlieb Baumgarten’s Meditationes philosophicae de nonnullis
ad poema pertinentibus. University of California Press.

22
History of neuroaesthetics as a discipline

Berlyne, D. E. (1966). Curiosity and exploration. Science, 153(3731), 25–33. https://doi.org/10.1126/science.153.


3731.25
Berlyne, D. E. (1971). Aesthetics and psychobiology. Appleton-Century-Crofts.
Berlyne, D. E. (1974). Studies in the new experimental aesthetics: Steps toward an objective psychology of aesthetic appreciation.
Hemisphere Publishing Corporation.
Berlyne, D. E., & McDonnell, P. (1965). Effects of stimulus complexity and incongruity on duration of EEG desynchro-
nization. Electroencephalography and Clinical Neurophysiology, 8, 156–161.
Besson, M., & Macar, F. (1987). An event-related potential analysis of incongruity in music and other non-linguistic
contexts. Psychophysiology, 24(1), 14–25. https://doi.org/10.1111/j.1469-8986.1987.tb01853.x
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Blood, A. J., Zatorre, R. J., Bermudez, P., & Evans, A. C. (1999). Emotional responses to pleasant and unpleasant music
correlate with activity in paralimbic brain regions. Nature Neuroscience, 2(4), 382–387. https://doi.org/10.1038/7299
Brachmann, A., & Redies, C. (2017). Computational and experimental approaches to visual aesthetics. Frontiers in Com-
putational Neuroscience, 11, 102. https://doi.org/10.3389/fncom.2017.00102
Brattico, E., Jacobsen, T., De Baene, W., Nakai, N., & Tervaniemi, M. (2003). Electrical brain responses to descrip-
tive versus evaluative judgments of music. Annals of the New York Academy of Sciences, 999, 155–157. https://doi.
org/10.1196/annals.1284.018
Brattico, E., & Pearce, M. (2013). The neuroaesthetics of music. Psychology of Aesthetics, Creativity, and the Arts, 7(1),
48–61. https://doi.org/10.1037/a0031624
Brown, S., & Dissanayake, E. (2009). The arts are more than aesthetics: Neuroaesthetics as narrow aesthetics. In M. Skov &
O. Vartanian (Eds.), Neuroaesthetics (pp. 43–57). Baywood Publishing.
Brown, S., Martinez, M. J., & Parsons, L. M. (2004). Passive music listening spontaneously engages limbic and paralimbic
systems. NeuroReport, 15(13), 2033, 2037. https://doi.org/10.1097/00001756-200409150-00008
Burke, E. (1757). A philosophical enquiry into the origin of our ideas of the sublime and beautiful. R. and J. Dodsley.
Buzsáki, G. (2020). The brain-cognitive behavior problem: A retrospective. eNeuro, 7(4), 1–8.
Carroll, N. (2008). Aesthetic experience, art and artists. In R. Shusterman  & A. Tomlin (Eds.), Aesthetic experience
(pp. 145–165). Routledge.
Cattaneo, Z., Lega, C., Flexas, A., Nadal, M., Munar, E.,  & Cela-Conde, C. J. (2014). The world can look better:
Enhancing beauty experience with brain stimulation. Social Cognitive and Affective Neuroscience, 9(11), 1713–1721.
https://doi.org/10.1093/scan/nst165
Cavanagh, P. (2005). The artist as neuroscientist. Nature, 434(7031), 301–307. https://doi.org/10.1038/434301a
Cela-Conde, C. J., Marty, G., Maestú, F., Ortiz, T., Munar, E., Fernández, A., Roca, M., Rosselló, J., & Quesney, F.
(2004). Activation of the prefrontal cortex in the human visual aesthetic perception. Proceedings of the National Academy
of Sciences of the United States of America, 101(16), 6321–6325. https://doi.org/10.1073/pnas.0401427101
Changeux, J.-P. (1994). Art and neuroscience. Leonardo, 27(3), 189–201. https://doi.org/10.2307/1576051
Chatterjee, A. (2003). Prospects for a cognitive neuroscience of visual aesthetics. Bulletin of Psychology and the Arts, 4(2),
55–60.
Chatterjee, A. (2009). Prospects for a neuropsychology of visual art. In M. Skov & O. Vartanian (Eds.), Neuroaesthetics
(pp. 131–143). Baywood Publishing.
Chatterjee, A. (2011). Neuroaesthetics: A coming of age story. Journal of Cognitive Neuroscience, 23(1), 53–62. https://doi.
org/10.1162/jocn.2010.21457
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Chatterjee, A., & Vartanian, O. (2016). Neuroscience of aesthetics. Annals of the New York Academy of Sciences, 1369(1),
172–194. https://doi.org/10.1111/nyas.13035
Chatterjee, A., Thomas, A., Smith, S. E., & Aguirre, G. K. (2009). The neural response to facial attractiveness. Neuropsy-
chology, 23(2), 135–143. https://doi.org/10.1037/a0014430
Che, J., Sun, X., Skov, M., Vartanian, O., Rosselló, J., & Nadal, M. (2021). The role of working memory capacity in
evaluative judgments of liking and beauty. Cognition and Emotion, 35(7), 1407–1415. https://doi.org/10.1080/0269
9931.2021.1947781
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology,
29(23), 1–9. https://doi.org/10.1016/j.cub.2019.09.067

23
Martin Skov

Churchland, P. S.,  & Sejnowski, T. J. (1988). Perspectives on cognitive neuroscience. Science, 242(4879), 741–745.
https://doi.org/10.1126/science.3055294
Conway, B. R., & Livingstone, M. S. (2007). Perspectives on science and art. Current Opinion in Neurobiology, 17(4),
476–482. https://doi.org/10.1016/j.conb.2007.07.010
Costelloe, T. M. (2018). The faculty of taste. In J. A. Harris (Ed.), The Oxford handbook of British philosophy in the eighteenth
century. https://doi.org/10.1093/oxfordhb/9780199549023.013.019. Oxford University Press.
Davies, S. (2012). The artful species. Oxford University Press.
Dickie, G. (1996). A century of taste. Oxford University Press.
Di Dio, C., & Gallese, V. (2009). Neuroaesthetics: A review. Current Opinion in Neurobiology, 19, 1–6.
Dieckmann, H. (1974). Theories of beauty to the mid-nineteenth century. In P. P. Wiener (Ed.), Dictionary of the history
of ideas, I (pp. 195–206). Charles Scribner’s Sons.
Dobzhansky, T. (1962). Mankind evolving. The evolution of the human species. Yale University Press.
Dukes, D., Abrams, K., Adolphs, R., Ahmed, M. E., Beatty, A., Berridge, K. C., Broomhall, S., Brosch, T., Campos, J. J., Clay,
Z., Clément, F., Cunningham, W. A., Damasio, A., Damasio, H., D’Arms, J., Davidson, J. W., de Gelder, B., Deonna,
J., de Sousa, R., . . Sander, D. (2021). The rise of affectivism. Nature Human Behaviour, 5(7), 816–820. https://doi.org/
10.1038/s41562-021-01130-8
Fairhall, S. L., & Ishai, A. (2008). Neural correlates of object indeterminacy in art compositions. Consciousness and Cogni-
tion, 17(3), 923–932. https://doi.org/10.1016/j.concog.2007.07.005
Fastrup, A. (2007). Sensibilitetens bevægelse: Denis Diderots fysiologiske æstetik [The Movement of Sensibility: The Physi-
ological Aesthetics of Denis Diderot]. København: Museum Tusculanum Press.
Fechner, G. T. (1876). Vorschule der Ästhetik. Breitkopf und Hörtel.
Felleman, D. J., & van Essen, D. C. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cerebral
Cortex, 1(1), 1–47. https://doi.org/10.1093/cercor/1.1.1-a
Fuller, P. M., Sherman, D., Pedersen, N. P., Saper, C. B., & Lu, J. (2011). Reassessment of the structural basis of the
ascending arousal system. Journal of Comparative Neurology, 519(5), 933–956. https://doi.org/10.1002/cne.22559
Gale, A., Christie, B., & Penfold, V. (1971). Stimulus complexity and the occipital EEG. British Journal of Psychology,
62(4), 527–531. https://doi.org/10.1111/j.2044-8295.1971.tb02068.x
Gale, A., Spratt, G., Christie, B.,  & Smallbone, A. (1975). Stimulus complexity, EEG abundance gradients, and
detection efficiency in a visual recognition task. British Journal of Psychology, 66(3), 289–298. https://doi.
org/10.1111/j.2044-8295.1975.tb01464.x
Gazzaniga, M. S. (1988). Life with George: The birth of the Cognitive Neuroscience Institute. In W. Hirst et al. (Eds.),
The making of cognitive science: Essays in honor of George A. Miller (pp. 230–241). Cambridge University Press.
Gold, B. P., Pearce, M. T., Mas-Herrero, E., Dagher, A.,  & Zatorre, R. J. (2019). Predictability and uncertainty in
the pleasure of music: A reward for learning? Journal of Neuroscience, 39(47), 9397–9409. https://doi.org/10.1523/
JNEUROSCI.0428-19.2019
Graham, D. J., & Redies, C. (2010). Statistical regularities in art: Relations with visual coding and perception. Vision
Research, 50(16), 1503–1509. https://doi.org/10.1016/j.visres.2010.05.002
Gregory, R., Harris, J., Heard, P., & Rose, D. (1995). The artful eye. Oxford University Press.
Griffiths, T. D., Johnsrude, I., Dean, J. L.,  & Green, G. G. R. (1999). A  common neural substrate for the analy-
sis of pitch and duration pattern in segmented sound? NeuroReport, 10(18), 3825–3830. https://doi.org/10.1097/
00001756-199912160-00019
Grossberg, S. (2005). The art of seeing and painting. Spatial Vision, 21(3–5), 463–486. https://doi.org/10.1163/15685
6808784532608
Hagendoorn, I. (2003). Cognitive dance improvisation: How study of the motor system can inspire dance (and vice
versa). Leonardo, 36(3), 221–228. https://doi.org/10.1162/002409403321921442
Hegel, G. W. F. (1920). The philosophy of fine art (F. Osmaston, trans.). G. Bell & Sons.
Hillyard, S. A. (1993). Electrical and magnetic brain recordings: Contributions to cognitive neuroscience. Current Opin-
ion in Neurobiology, 3(2), 217–224. https://doi.org/10.1016/0959-4388(93)90213-i
Höge, H. (1995). Fechner’s experimental aesthetics and the golden section hypothesis today. Empirical Studies of the Arts,
13(2), 131–148. https://doi.org/10.2190/UHTQ-CFVD-CAU2-WY1C
Hugdahl, K., Brönnick, K., Kyllingsbaek, S., Law, I., Gade, A., & Paulson, O. B. (1999). Brain activation during dichotic
presentations of consonant-vowel and musical instrument stimuli: A 15O-PET study. Neuropsychologia, 37(4), 431–
440. https://doi.org/10.1016/s0028-3932(98)00101-8
Hume, D. (1757/1985). Essays: Moral, political, and literary (rev. ed.) by E. F. Miller. Liberty Classics.
Hutcheson, F. (1725/1973). An inquiry concerning beauty, order, harmony, design. Edited, with an introduction and notes by
Peter Kivy. Martinus-Nijhoff.

24
History of neuroaesthetics as a discipline

Hyman, J. (2010). Art and neuroscience. In R. Frigg & M. C. Hunter (Eds.), Beyond mimesis and convention. Representation
in art and science (pp. 245–261). Springer.
Ishizu, T., & Zeki, S. (2013). The brain’s specialized systems for aesthetic and perceptual judgment. European Journal of
Neuroscience, 37(9), 1413–1420. https://doi.org/10.1111/ejn.12135
Jacobsen, T. (2006). Bridging the arts and sciences: A  framework for the psychology of aesthetics. Leonardo, 39(2),
155–162. https://doi.org/10.1162/leon.2006.39.2.155
Jacobsen, T. (2009). Neuroaesthetics and the psychology of aesthetics. In M. Skov & O. Vartanian (Eds.), Neuroaesthetics
(pp. 27–42). Baywood Publishing.
Jacobsen, T., & Höfel, L. (2001). Aesthetics electrified: An analysis of descriptive symmetry and evaluative aesthetics
judgment using event-related potentials. Empirical Studies of the Arts, 19(2), 177–190.
Jacobsen, T.,  & Höfel, L. (2003). Descriptive and evaluative judgment processes: Behavioral and electrophysiological
indices of processing symmetry and aesthetics. Cognitive, Affective, & Behavioral Neuroscience, 3(4), 289–299.
Jacobsen, T., Schubotz, R. I., Höfel, L., & von Cramon, D. Y. (2006). Brain correlates of aesthetic judgment of beauty.
NeuroImage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010
Kant, I. (1786). Metaphysical foundations of natural science ( J. Ellington, Trans.). Bobbs-Merrill.
Kant, I. (1790/2001). Critique of the power of judgment (P. Guyer & E. Matthews, Trans., Rev. ed.). Cambridge University
Press.
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Kim, A. (2016). Wilhelm Maximilian Wundt. The Stanford Encyclopedia of Philosophy (Fall 2016 ed.), Edward N.
Zalta (Ed.). https://plato.stanford.edu/archives/fall2016/entries/wilhelm-wundt/.
Kirk, U. (2008). The neural basis of object-context relationships on aesthetic judgment. PLOS ONE, 3(11), e3754.
https://doi.org/10.1371/journal.pone.0003754
Kivy, P. (2003). The seventh sense: Francis Hutcheson and eighteenth-century British aesthetics. Oxford University Press.
Klein, M., Coles, M. G. H., & Donchin, E. (1984). People with absolute pitch process tone without producing a P300.
Science, 223(4642), 1306–1309. https://doi.org/10.1126/science.223.4642.1306
Klein, R. G. (2002). The dawn of human culture. John Wiley & Sons.
Koelsch, S., Fritz, T., V Cramon, D. Y., Müller, K., & Friederici, A. D. (2005). Investigating emotion with music: An
fMRI study. Human Brain Mapping, 27(3), 239–250. https://doi.org/10.1002/hbm.20180
Kranjec, A., & Skov, M. (2021). Visualizing aesthetics across two centuries. Empirical Studies of the Arts, 39(1), 78–100.
https://doi.org/10.1177/0276237420905308
Kristeller, P. O. (1951). The modern system of the arts: A study in the history of aesthetics part I. Journal of the History of
Ideas, 12(4), 496–527. https://doi.org/10.2307/2707484
Kristeller, P. O. (1952). The modern system of the arts: A study in the history of aesthetics (II). Journal of the History of
Ideas, 13(1), 17–46. https://doi.org/10.2307/2707724
Kusch, M. (1995). Psychologism: A case study in the sociology of philosophical knowledge. Routledge.
Latto, R. (1995). The brain of the beholder. In G. Gregory et al. (Eds.), The artful eye (pp. 66–94). Oxford University Press.
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811
Livingstone, M. S. (1988). Art, illusion and the visual system. Scientific American, 258(1), 78–85. https://doi.org/10.1038/
scientificamerican0188-78
Livingstone, M. S. (2002). Vision and art. The biology of seeing. Harry N. Abrams.
Livingstone, M. S., & Hubel, D. (1988). Segregation of form, color, movement, and depth: Anatomy, physiology, and
perception. Science, 240(4853), 740–749. https://doi.org/10.1126/science.3283936
Locher, P. J., & Nodine, C. F. (1977). Symmetry catches the eye. In J. K. O’Reagan & A. Levy-Schoen (Eds.), From
physiology to cognition. North-Holland Press.
Machotka, P. (1980). Daniel Berlyne’s contributions to empirical aesthetics. Motivation and Emotion, 4(2), 113–121.
https://doi.org/10.1007/BF00995192
Martindale, C. (2007). The foundation and future of the society for the psychology of aesthetics, creativity, and the arts.
Psychology of Aesthetics, Creativity, and the Arts, 1(3), 121–132. https://doi.org/10.1037/1931-3896.1.3.121
Martindale, C., Moore, K., & Borkum, J. (1990). Aesthetic preference: Anomalous findings for Berlyne’s psychobiologi-
cal theory. American Journal of Psychology, 103(1), 53–80. https://doi.org/10.2307/1423259
Martinez-Conde, S., Macknik, S. L., & Hubel, D. H. (2004). The role of fixational eye movements in visual perception.
Nature Reviews Neuroscience, 5(3), 229–240. https://doi.org/10.1038/nrn1348
Menninghaus, W., Wagner, V., Wassiliwizky, E., Schindler, I., Hanich, J., Jacobsen, T., & Koelsch, S. (2019). What are
aesthetic emotions? Psychological Review, 126(2), 171–195. https://doi.org/10.1037/rev0000135

25
Martin Skov

Menon, V., & Levitin, D. J. (2005). The rewards of music listening: Response and physiological connectivity of the mes-
olimbic system. Neuroimage, 28(1), 175–184. https://doi.org/10.1016/j.neuroimage.2005.05.053
Morgan, B. (2017). The outward mind: Materialist aesthetics in victorian science and literature. University of Chicago Press.
Nadal, M., Munar, E., Capó, M. A., Rosselló, J., & Cela-Conde, C. J. (2008). Towards a framework for the study of the neural
correlates of aesthetic preference. Spatial Vision, 21(3–5), 379–396. https://doi.org/10.1163/156856808784532653
Nadal, M., Munar, E., Marty, G., & Cela-Conde, C. J. (2010). Visual complexity and beauty appreciation: Explaining the
divergence of results. Empirical Studies of the Arts, 28(2), 173–191. https://doi.org/10.2190/EM.28.2.d
Nadal, M., & Ureña, E. (2022). One hundred years of empirical aesthetics: Fechner to Berlyne (1876–1976). In M.
Nadal & O. Vartanian (Eds.), The Oxford handbook of empirical aesthetics. Oxford University Press.
Nicki, R. M., & Gale, A. (1977). EEG, measures of complexity, and preference for nonrepresentational works of art.
Perception, 6(3), 281–286. https://doi.org/10.1068/p060281
Nodine, C. F., Locher, P. J., & Krupinski, E. A. (1993). The role of formal art training on perception and aesthetic judg-
ment of art compositions. Leonardo, 26(3), 219–227. https://doi.org/10.2307/1575815
Nodine, C. F., & McGinnis, J. H. (1983). Artistic style, compositional design, and visual scanning. Visual Arts Research,
12, 1–9.
Ochs, S. (2004). A history of nerve functions. Cambridge University Press.
Paller, K. A., McCarthy, G., & Wood, C. C. (1988). ERPs predictive of subsequent recall and recognition performance.
Biological Psychology, 26(1–3), 269–276. https://doi.org/10.1016/0301-0511(88)90023-3
Pantev, C., Hoke, M., Lehnertz, K., & Lütkenhöner, B. (1989). Neuromagnetic evidence of an amplitopic organiza-
tion of the human auditory cortex. Electroencephalography and Clinical Neurophysiology, 72(3), 225–231. https://doi.
org/10.1016/0013-4694(89)90247-2
Pearce, M. T. (2018). Statistical learning and probabilistic prediction in music cognition: Mechanisms of stylistic encul-
turation. Annals of the New York Academy of Sciences, 1423(1), 378. https://doi.org/10.1111/nyas.13654
Pearce, M. T., Ruiz, M. H., Kapasi, S., Wiggins, G. A.,  & Bhattacharya, J. (2010). Unsupervised statistical learning
underpins computational, behavioural, and neural manifestations of musical expectation. NeuroImage, 50(1), 302–313.
https://doi.org/10.1016/j.neuroimage.2009.12.019
Pearce, M. T., & Wiggins, G. A. (2005). Expectation in melody: The influence of context and learning. Music Perception,
23(5), 377–405. https://doi.org/10.1525/mp.2006.23.5.377
Pearce, M. T., Zaidel, D. W., Vartanian, O., Skov, M., Leder, H., Chatterjee, A., & Nadal, M. (2016). Neuroaesthetics:
The cognitive neuroscience of aesthetic experience. Perspectives on Psychological Science, 11(2), 265–279. https://doi.
org/10.1177/1745691615621274
Peretz, I., & Zatorre, R. J. (2005). Brain organization for music processing. Annual Review of Psychology, 56, 89–114.
https://doi.org/10.1146/annurev.psych.56.091103.070225
Pessoa, L., Medina, L., & Desfilis, E. (2021). Refocusing neuroscience: Moving away from mental categories and towards
complex behaviours. Philosophical Transactions of the Royal Society of London B, 377, 20200534.
Raichle, M. E. (1998). Behind the scenes of functional brain imaging: A historical and physiological perspective. Pro-
ceedings of the National Academy of Sciences, 95(3), 765–772. https://doi.org/10.1073/pnas.95.3.765
Ramachandran, V. S., & Hirstein, W. (1999). The science of art: A neurological theory of aesthetic experience. Journal
of Consciousness Studies, 6(6–7), 15–51.
Ramsøy, T. Z., Friis-Olivarius, M., Jacobsen, C., Jensen, S. B., & Skov, M. (2012). Effects of perceptual uncertainty on
arousal and preference across different visual domains. Journal of Neuroscience, Psychology, and Economics, 5(4), 212–226.
https://doi.org/10.1037/a0030198
Redies, C., Hänisch, J., Blickhan, M., & Denzler, J. (2007). Artists portray human faces with the Fourier statistics of
complex natural scenes. Network, 18(3), 235–248. https://doi.org/10.1080/09548980701574496
Redies, C., Hasenstein, J., & Denzler, J. (2007). Fractal-like image statistics in visual art: Similarity to natural scenes.
Spatial Vision, 21(1–2), 137–148. https://doi.org/10.1163/156856807782753921
Reed, E. S. (1997). From soul to mind. The emergence of psychology from Erasmus Darwin to William James. Yale University
Press.
Reiss, H. (1994). The “naturalization” of the term “Ästhetik” in eighteenth-century German: Alexander Gottlieb Baum-
garten and his impact. Modern Language Review, 89(3), 645–658. https://doi.org/10.2307/3735122
Rentschler, I., Herzberger, B., & Epstein, D. (1988). Biology and art. Biological aspects of aesthetics. Springer.
Rey, R. (2002). Psyche, soma, and the Vitalist philosophy of medicine. In John P. Wright & Paul Potter (Eds.), Psyche and
soma: Physicians and metaphysicians on the mind-body problem from antiquity to enlightenment (pp. 255–265). Clarendon Press.
Reynolds, Sir J. (1798). The works of Sir Joshua Reynolds, knight; . . . containing his discourses. In Idlers, Etc. (with his
last corrections and additions) (2nd ed., 3 vols.). T. Cadell, Jun. and W. Davies Publishing.

26
History of neuroaesthetics as a discipline

Sakai, K., Hikosaka, O., Miyauchi, S., Takino, R., Tamada, T., Iwata, N. K., & Nielsen, M. (1999). Neural representa-
tion of a rhythm depends on its interval ratio. Journal of Neuroscience, 19(22), 10074–10081. https://doi.org/10.1523/
JNEUROSCI.19-22-10074.1999
Schiller, P. H. (1997). Past and present ideas about how the visual scene is analyzed by the brain. Cerebral Cortex, 12,
59–90. https://doi.org/10.1007/978-1-4757-9625-4_2
Sherman, A., Grabowecky, M.,  & Suzuki, S. (2015). In the working memory of the beholder: Art appreciation is
enhanced when visual complexity is compatible with working memory. Journal of Experimental Psychology. Human
Perception and Performance, 41(4), 898–903. https://doi.org/10.1037/a0039314
Shiner, L. (2001). The invention of Art: A cultural history. University of Chicago Press.
Silvia, P. J. (2005). Emotional responses to art: From collation and arousal to cognition and emotion. Review of General
Psychology, 9(4), 342–357. https://doi.org/10.1037/1089-2680.9.4.342
Skov, M. (2005). Hvad er neuroæstetik? [What is neuroaesthetics?]. Kritik, 174, 1–10.
Skov, M. (2007). Følelser og æstetik [Emotions and aesthetics]. In M. Skov & T. W. Jensen (Eds.), Følelser og Kognition
(pp. 167–196). Museum Tusculanum.
Skov, M. (2009). Neuroaesthetic problems. A framework for neuroaesthetic research. In M. Skov & O. Vartanian (Eds.),
Neuroaesthetics (pp. 9–26). Baywood Publishing.
Skov, M. (2010). The pleasure of art. In M. Kringelbach & K. Berridge (Eds.), Pleasures of the brain (pp. 270–283). Oxford
University Press.
Skov, M. (2019). Aesthetic appreciation: The view from neuroimaging. Empirical Studies of the Arts, 37(2), 220–248.
https://doi.org/10.1177/0276237419839257
Skov, M., & Nadal, M. (2020a). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Skov, M., & Nadal, M. (2020b). There are no aesthetic emotions: Comment on Menninghaus et al. (2019). Psychological
Review, 127(4), 640–649. https://doi.org/10.1037/rev0000187
Skov, M., & Vartanian, O. (2009). Neuroaesthetics. Baywood Publishing.
Smith, R. (1997). The human sciences. W.W. Norton & Company.
Solso, R. L. (1994). Cognition and the visual arts. MIT Press.
Solso, R. L. (2003). The psychology of art and the evolution of the conscious brain. MIT Press.
Stolnitz, J. (1961). On the origins of “aesthetic disinterestedness”. The Journal of Aesthetics and Art Criticism, 20(2),
131–144. https://doi.org/10.1111/1540_6245.jaac20.2.0131
Stratton, G. M. (1902). Eye-movements and the aesthetics of visual form. Philosophical Studies, 20, 336–359.
Swanson, L. W. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Research, 886(1–2), 113–164.
https://doi.org/10.1016/s0006-8993(00)02905-x
Tallis, R. (2008). The limitations of a neurological approach to art. Lancet, 372(9632), 19–20. https://doi.org/10.1016/
S0140-6736(08)60975-7
Tatarkiewicz, W. (1972). The great theory of beauty and its decline. Journal of Aesthetics and Art Criticism, 31(2), 165–180.
https://doi.org/10.1111/1540_6245.jaac31.2.0165
Tervaniemi, M. (2009). Musical sounds in the human brain. In M. Skov & O. Vartanian (Eds.), Neuroaesthetics (pp. 221–
231). Baywood Publishing.
Tonelli, G. (2003). Taste in the history of aesthetics from the Renaissance to 1770. In P. P. Wiener (Ed.), Dictionary of the
history of ideas, 4. Charles Scribner’s Sons.
Valentine, C. W. (1962). The experimental psychology of beauty. Methuen, and Co.
Vartanian, O. (2009). Conscious experience of pleasure in art. In M. Skov & O. Vartanian (Eds.), Neuroaesthetics (pp. 261–
273). Baywood Publishing.
Vartanian, O., & Goel, V. (2004a). Neuroanatomical correlates of aesthetic preference for paintings. NeuroReport, 15(5),
893–897. https://doi.org/10.1097/00001756-200404090-00032
Vartanian, O., & Goel, V. (2004b). Emotion pathways in the brain mediate aesthetic preference. Bulletin of Psychology and
the Arts, 5(1), 37–42.
Vartanian, O., & Nadal, M. (2007). A biological approach to a model of aesthetic experience. In L. Dorfman, C. Mar-
tindale & V. Petrov (Eds.), Aesthetics and innovation (pp. 429–443). Cambridge Scholars Press.
Wieser, H. G., & Mazzola, G. (1986). Musical consonances and dissonances: Are they distinguished independently by
the right and left hippocampi? Neuropsychologia, 24(6), 805–812. https://doi.org/10.1016/0028-3932(86)90079-5
Wundt, W. (1873). Grundzüge der physiologischen Psychologie, 2 vols. Engelmann.
Yarbus, A. L. (1967). Eye movements and vision. Plenum Press.
Zaidel, D. W. (2005). Neuropsychology of art. Psychology Press.

27
Martin Skov

Zangemeister, W. H., Sherman, K.,  & Stark, L. (1995). Evidence for a global scanpath strategy in viewing abstract
compared with realistic images. Neuropsychologia, 33(8), 1009–1025. https://doi.org/10.1016/0028-3932(95)00014-t
Zatorre, R. J., Evans, A. C., Meyer, E., & Gjedde, A. (1992). Lateralization of phonetic and pitch discrimination in
speech processing. Science, 256(5058), 846–849. https://doi.org/10.1126/science.1589767
Zeki, S. (1999a). Art and the brain. Journal of Consciousness Studies, 6(6–7), 76–96.
Zeki, S. (1999b). Inner vision. An exploration of art and the brain. Oxford University Press.
Zeki, S. (2005). The Ferrier Lecture 1995 behind the seen: The functional specialization of the brain in space and time.
Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 360(1458), 1145–1183. https://doi.
org/10.1098/rstb.2005.1666
Zeki, S., & Marini, L. (1998). Three cortical stages of colour processing in the human brain. Brain, 121(9), 1669–1685.
https://doi.org/10.1093/brain/121.9.1669

28
PART I

Aesthetic liking
2
SENSORY LIKING
How nervous systems assign hedonic
value to sensory objects

Martin Skov

All biological organisms show preferences for sensory objects they encounter in their physical environment.
Such preferences are manifested in the organism’s behavioural response to the object: some objects the
organism will devote resources to approach, while others it will avoid at all costs. These behavioural prefer-
ences are rooted in some form of evaluative mechanism that tags sensory stimuli as either liked or disliked.
The scientific study of this evaluative function is called the study of sensory liking. This chapter is intended
as an introduction to what is known about the neurobiological processes that the human brain uses when it
engages in such sensory liking evaluations.
Historically, the study of sensory liking has pursued three main questions. The first concerns the relation
of the stimulus to sensory liking: which stimulus properties are liked or disliked, and what is the cause of such
relationships? Are liking and disliking responses to a given sensory percept governed by universal laws, or are
they subject to subjective and contextual variation? These questions have intrigued Western philosophy for a
very long time, at least since the Pythagoreans, 3000 years ago. This esoteric group of Greek mathematicians
and philosophers originated the view that beauty responses are governed by objective relations and believed
that all objects organized according to principles of proportion and harmony are experienced as pleasurable
(Dieckmann, 1974). Later, Renaissance philosophers began to oppose this view, arguing that people like dif-
ferent things and that liking responses to stimulus properties therefore vary according to individual “tastes”
(Tonelli, 2003; see also Chapter 1). When experimental psychology emerged in the second half of the 19th
century, it also felt compelled to examine the relation between stimulus and liking response (Nadal & Ureña,
2022). To this day, most experimental paradigms used to investigate sensory liking consist of measuring
reported liking ratings for stimuli whose features are systematically manipulated.
The second question concerns the genesis of stimulus-liking relationships: why are certain stimulus prop-
erties liked while others are disliked? If humans or some other species like the colour red, where does this
preference come from? The first to systematically collect observations to support a scientific answer to this
question was Charles Darwin, who, in the book The Descent of Man, and Selection in Relation to Sex (1871),
argued that biological organisms have evolved a “power of discrimination and taste” (p. 246) that serves adap-
tive needs. Specifically, Darwin (1871) hypothesized that, in sexual species, females evolve preferences for
male sensory traits that signal fecundity, that is, a greater likelihood of reproductive success if the female chose
to mate with males that embody the preferred trait. According to this evolutionary argument, a preference
for stimulus features such as the colour red can arise because females who chose, say, males with red plum-
age as sexual partners have more viable progeny than females who do not. Since Darwin, this evolutionary

DOI: 10.4324/9781003008675-3 31
Martin Skov

thesis has been adopted as a possible general explanation for the existence of stimulus-liking relationships:
biological organisms evolve preferences for stimulus features that are statistically associated with behavioural
choices that have benefitted survival (Adolphs & Andler, 2018; Damasio & Carvalho, 2013; Panksepp, 1998).
This theory also explains why different species have different stimulus preferences: since individual species
have different survival needs and inhabit different physical habitats, sensory liking will evolve to take different
forms across biological taxa (Rosenthal, 2017).
Finally, the third question that has dominated the study of sensory liking concerns the nature of the
mechanism that implements stimulus-liking relationships—Darwin’s putative “power of taste.” The idea that
the human mind must contain a dedicated psychological mechanism for assigning liking and disliking to
sensory inputs was first introduced in the 18th century (see Chapter 1). Using an array of different meth-
ods, psychologists and neuroscientists have tried to identify the psychological and physiological processes
involved in generating liking and disliking for stimulus properties. This has revealed a set of neurobiological
processes that take place throughout a network of mesocorticolimbic structures that encode states of pleasure
and displeasure in response to perceptual representations of sensory stimuli (Becker et al., 2019; Berridge &
Kringelbach, 2015; Martínez-García & Lanuza, 2018; O’Connell & Hofmann, 2011; Panksepp, 1998; Skov,
2020). The character of these affective responses appears to determine if a stimulus is liked or disliked.
A popular way to unify the work prompted by these questions is to think of sensory liking as a psycho-
logical and neurobiological reaction to sensory input. If a given stimulus property elicits states of pleasure, it is
experienced as liked. If it elicits states of displeasure, it is experienced as disliked. This coupling of stimulus
and hedonic response is thought to be determined by evolution: stimuli that prove beneficial to survival are
linked to pleasure through innate mechanisms; stimuli that prove harmful to survival are linked to displeas-
ure. However, even though this account of sensory liking contains a kernel of truth, it fails to explain many
of the empirical findings produced by recent psychology and neuroscience. Specifically, it does not accord
well with the ubiquitous observation that fixed liking and disliking responses to individual stimulus features
are not the rule but rather the exception, especially in organisms with complex nervous systems (Coppin &
Sander, 2012; Corradi et al., 2020; Hayes, 2020; Rosenthal, 2017; Skov, 2020). Instead, studies of sensory
liking consistently find that liking and disliking vary with endogenous and contextual factors.
A possible explanation for variation in stimulus-liking responses is learning. To proponents of the classical
theory of sensory liking, learning has served as a possible modulating factor that explains why individuals
differ in their reactions to a given stimulus feature: whereas pleasure and displeasure reactions to stimulus
features are generally innately determined, they can be modulated through learning mechanisms. Stud-
ies of reinforcement, personality and knowledge differences, and cross-cultural variation certainly suggest
that learning plays a central role in shaping liking and disliking (Che et al., 2018; Pessiglione & Lebreton,
2015; Rosenthal, 2017; Schultz, 2015). Yet admitting learning as a modulating factor does not explain why
stimulus-liking reactions need to be modulated to fit individual and contextual conditions. In this chapter,
I  present an alternative account of sensory liking that takes the fact that, for many biological organisms,
sensory liking and disliking are flexible responses to sensory objects as its starting explanandum. Based on
multiple findings, I show that sensory liking should be considered not a stereotypical reaction to stimuli but
an evaluative event whereby biological organisms seek to establish the potential benefit or harm of a sensory
object relative to their current physiological, behavioural, and environmental conditions.
This alternative account suggests that, rather than being solely determined by the sensory nature of the
stimulus, liking and disliking emerge from a conjunction of evaluative processes that are not only driven
by stimulus information but also by a range of other factors. The functional purpose of this computational
system is to generate liking and disliking constructs that service the organism’s current physiological and
behavioural needs. In order to do so, sensory liking evaluations assess the hedonic value of a stimulus as it
pertains to the organism’s immediate concerns: Can the sensory object being evaluated help restore the met-
abolic balance of the organism’s body? Is it possibly toxic? Has the organism encountered the object before?

32
Sensory liking

Figure 2.1 Sensor y liking evaluations always occur in the context of behavioural tasks. For this reason, how much an
organism likes or dislikes a given stimulus can yield different survival benefits under different physiological
and environmental circumstances. For example, when foraging for something to eat in an unknown terrain,
decisions about whether to venture into an environment with low visibility hinge on several considerations:
Does the organism believe, based on previous experiences, that the forest contains delicious fruit? Are there
grounds to believe that the forest hides predators? What is the risk of searching for fruits in the forest (i.e.,
how certain is the organism that it will be rewarded by deciding to go into the unknown?). Is the organism
very hungry? Only by factoring in these factors can the organism compute how valuable a stimulus is vis-à-
vis its actual physiological needs and current behavioural undertakings.

Is there a risk associated with working to obtain and ingest the object? These and other considerations are
factored into the evaluation of the object’s hedonic value (Figure 2.1). It is this computational constitution
that explains the flexible nature of sensory liking evaluations: under some circumstances, a stimulus can be
highly liked; under other circumstances, it can be less liked; and when certain conditions come together, it
can even be disliked.
In what follows, I will review experimental findings that cast light on how the human brain implements
such evaluative events. As we shall see, central to the human evaluative system’s ability to factor endogenous
and contextual conditions into its computation of sensory liking and disliking is the existence of a number
of different functional processes that are able to represent and integrate information about the organism’s
interoceptive states, cognitive knowledge, and learned expectations. Mounting empirical evidence suggests
that liking and disliking outcomes are determined by the specific coordination and integration of these dif-
ferent mechanisms that characterize individual evaluation events. A possible implication of this observation is
that human evaluative events may have systematic characteristics that reflect the different purposes they serve.
For example, some evaluative events might rely more on learned expectations, while others involve greater
explicit attention to the stimulus being assessed. At the end of the chapter, I will briefly discuss how this idea
lends itself to new neuroscientific investigations. But I first expand on how to understand sensory liking as a
general biological phenomenon.

Sensory liking
When we speak of “liking something” in everyday talk, we usually think of the conscious states of pleasure
that eating a chocolate bar or watching a great Netflix show can produce. This folk psychological usage
suggests that the word “liking” primarily refers to a specific experiential state that is unrelated to behaviour.

33
Martin Skov

In the context of biology, however, the term sensory liking describes a broader physiological event, the pur-
pose of which is to assess the hedonic value of sensory objects and motivate behaviour that promotes survival
(Adolphs & Andler, 2018; Berridge & Kringelbach, 2015; Damasio & Carvalho, 2013; Rosenthal, 2017).
The function of these hedonic states is not to give rise to introspective experiences of feelings but to elicit
action programs that allow biological organisms to detect and respond to adaptive opportunities and threats
(Damasio  & Carvalho, 2013; LeDoux, 2012; Panksepp, 1998; Pessiglione  & Lebreton, 2015; Rosenthal,
2017). While evidence suggests that sensory objects become liked through the generation of positive affective
emotions and disliked through the generation of negative affective emotions, these emotional processes can
occur without conscious awareness (Berridge & Winkielman, 2003). Indeed, some researchers have claimed
that only humans represent sensory liking and disliking in the form of conscious feelings (e.g., Adolphs &
Andler, 2018; LeDoux, 2012, 2021).
We can illustrate the relationship between sensory liking evaluations and behaviour with the example of
sweet and bitter tastes. In many animals, the sensation of sweet and bitter compounds leads to liking and
disliking responses that manifest themselves in a specific repertoire of behavioural acts (Hayes, 2020; Peciña,
2008, Steiner et al., 2001). In the case of liking responses, these typically take the form of upwards tongue
protrusions, smiles, and the initiation of motor outputs associated with ingestion of foods or fluids (Hayes,
2020; Peciña, 2008; Steiner et al., 2001). Disliking responses, in contrast, elicit downwards tongue protru-
sions, mouth gaping, grimacing, and a rejection of food and fluid objects (Hayes, 2020; Steiner et al., 2001;
Dinehart et al., 2006). Because the former set of responses leads to engagement with sweet compounds, we
say that these objects are liked. Similarly, because the latter set of responses prompt the organism to avoid
bitter compounds, we say that those are disliked.
Importantly, sweet and bitter tastes derive their hedonic values from the way they signal different sur-
vival consequences of engagement with sweet and bitter objects. For example, organisms like sweet-tasting
compounds such as sugar or sucrose because these contain molecules that are important to the restoration of
metabolic homeostasis (Hayes, 2020; Zimmerman et al., 2017). Conversely, bitter compounds are disliked
because they are usually encountered in objects that also contain toxic chemicals that are a danger to the
organism (Hayes, 2020; Katz & Sadacca, 2011; Zimmerman et al., 2017). Liking and disliking evaluations tag
sweet-tasting compounds as conducive to survival and bitter-tasting compounds as threatening and help elicit
appropriate behavioural actions. Experiments with mice show that stimulating the different cortical fields
in the insula that represent sweet and bitter tastes—even absent any sensory input—is sufficient to initiate
appetitive and aversive behaviours (Peng et al., 2015).
Because liking and disliking function as assessments of an object’s survival value, the hedonic value of
a stimulus depends on the adaptive relevance it holds for a given organism. Organisms differ in terms of
their physiological constitution and survival needs and therefore will exhibit different liking and disliking
responses. For example, not all species like sweet-tasting compounds, and even those that do can develop
disliking responses to stimuli such as glucose if environmental conditions change. Thus, Wada-Katsumata and
colleagues (2013) found a population of German cockroaches that have been exposed to poisoned glucose-
containing baits to have evolved disliking responses for glucose through a reorganization of their gustatory
system: where glucose previously stimulated sweet-taste receptors, it now stimulates bitter-taste receptors.
Even in species with an innate preference for sweet compounds, liking is frequently modulated by the
homeostatic state of the organism, yielding different degrees of liking depending on satiety signals that are
transmitted from blood and gut receptors to the brain (Han et al., 2018; Führer et al., 2008; Thomas et al.,
2015). In short, hedonic values communicate the survival value of the object to the individual organism and
as such vary greatly across biological taxa and as a function of physiological and behavioural conditions.
In terms of mechanisms, all sensory liking systems consist of three basic functional components (Fig-
ure 2.2): a sensory system that allows the organism to detect and represent physical objects that are of relevance
to the organism, an evaluative system that computes a hedonic value based on the perceptual output of the

34
Sensory liking

Figure 2.2 A  schematic depiction of the functional components that constitute sensory liking systems in biological
organisms. Regardless of their physiological organization, all organisms assign hedonic values to parts of their
physical surroundings (A) when stimulation of sensory receptors (B) elicits projection of information to an
evaluative mechanism (C). This evaluative system computes a hedonic value for the stimulus that forms the
basis of decisions the organism makes with respect to behaviours that involve the sensory object (D). As a
general rule, the way biological organisms engage with sensory objects reflects their preference for these
objects. The three core functional components (B, C, D) are instantiated by different physiological processes
across species, but in many species, including all vertebrates, the mechanisms of sensory representation, valu-
ation, and choice are implemented by different neuroanatomical systems.

sensory system, and a decision-making system that initiates behaviour on basis of the hedonic value assigned to
the stimulus (Damasio & Carvalho, 2013; LeDoux, 2012; Panksepp, 1998; Rosenthal, 2017).
The way these systems are constituted varies across taxa as a function of the individual species’ evolutionary
history. The computational mechanisms employed by a species can change as a consequence of evolutionary
innovations affecting any of the three systems. For example, the make-up of a species’ sensory systems deter-
mines the physical information it is capable of detecting and responding to and thus conditions the range of
sensory objects it is able to compute liking or disliking evaluations for. Mate choice research has shown that
changes to a species’ sexual preferences can follow directly from evolutionary alterations to its sensory sys-
tems that are driven by changes in its ecological environment (Cummings & Endler, 2018; Rosenthal, 2017;
Ryan & Cummings, 2013; see also Chapter 11). Similarly, comparative work suggests that the evaluative sys-
tem is subject to evolutionary modification. As described by Esther Ureña and Marcos Nadal in Chapter 13,
a series of evolutionary innovations have expanded the ability of vertebrates to compute hedonic evaluations
that integrate information representing the organism’s physiological state, prior experiences, and ongoing
task requirements (see also Cisek, 2021; Martínez-García & Lanuza, 2018; O’Connell & Hofmann, 2011;
Swanson, 2000). The evolutionary benefit of these changes appears to have been an increased control over
behavioural responses to a broader range of sensory objects, improving the odds of survival.
An important consequence of this variation in physiological constitution is how fixed a species’ liking and
disliking responses are. As a general rule, the simpler a sensory liking system, the more automatic and stereo-
typical the liking and disliking outcomes it produces. For example, most single-cell bacteria produce appeti-
tive and aversive behaviours that are primarily determined by the way receptors on the surface of the cell
are stimulated (van Duijn et al., 2006). Thus, in E. coli cells, a chemotaxis connects sensory receptors on the
cell surface with propeller-like motor structures (Falke et al., 1997). Depending on the way these receptors
are stimulated, chains of protein complexes are elicited, terminating in one of two movements of the cell’s
flagella: “running” behaviour that moves the cell towards attractive stimuli, and “tumbling” behaviour that
moves it away from repellent stimuli (Falke et al., 1997; van Duijn et al., 2006). Similarly, in male silkmoths,
stimulation of olfactory receptors by the female pheromone bombykol is sufficient to elicit fixed appetitive
behaviours related to courtship (Rosenthal, 2017; Sakurai et al., 2011).
In contrast, more complex sensory liking systems allow for flexible hedonic evaluations that are less deter-
mined by sensory stimulation. Such systems often consist of a multitude of cell assemblies that are able to

35
Martin Skov

represent endogenous and contextual factors that are of relevance to the organism’s survival (Rosenthal, 2017;
Skov, 2020). These processes contribute inputs to the neural mechanisms that compute hedonic evaluations.
As a consequence, liking and disliking outcomes are modulated by factors other than the sensory stimulus.
For instance, studies of vertebrate mate choice have revealed that irrespective of their “innate” attractiveness,
courters can remain unattractive to choosers unless they are able to engage the chooser’s sensory system in a
way that makes them stand out from conspecifics and other environmental stimuli (Rosenthal, 2017). Thus,
examination of courtship behaviour in túngara frogs by Michael Ryan (1980) has shown that male túngara
frogs combine two sounds—a “chuck” and a “whine”—to make themselves better heard by—and more
attractive to—female frogs in their vicinity. Male túngara frogs that add chucks to whines are more effortlessly
detected by female frogs and as a consequence become more liked (Griddi-Papp et al., 2006). Investigating
the computational reasons for this outcome, Hoke and colleagues (2004) found that, compared to other
sound combinations, listening to whine-chuck combinations elicited increased neural activation in the audi-
tory nucleus of female túngara frogs and enhanced functional connectivity between auditory brain regions
and the reward circuit. Findings such as these suggest that how liked a male túngara frog is experienced to be
by a female chooser is partly dependent on the way her auditory system is able to attend to and represent calls
amidst a cacophony of other sounds and that projections from her sensory system to her evaluative system
are mediated by this auditory process.
Similarly, processes involved in the computation of decision outputs can have profound effects on the
outcome of sensory liking. For example, whereas as a general rule, decision-making involves the identifica-
tion and choosing of sensory objects with the highest hedonic value (Pessiglione & Lebreton, 2015; Rangel
et al., 2008), exceptions to this rule exist. Both when foraging for food and when choosing between dif-
ferent mate options, many species will occasionally, under certain contextual circumstances, choose a less-
desired stimulus (Kolling et al., 2016; Ryan et al., 2018). One example of this phenomenon is the so-called
decoy effect where organisms exhibit different preferences for two options, A and B, depending on whether a
third option, C, is present (Bateson et al., 2002; Lea & Ryan, 2015; Shafir et al., 2002). Another contextual
circumstance that affects choice is time discounting. Many organisms will accept a less desired option over a
more attractive option if the latter is expected only to occur at a later point or perhaps not at all (Calhoun &
Hayden, 2015). For example, Lynch and colleagues (2005, 2006) found that female túngara frogs will ignore
an unattractive synthetic mating call early in the evening but then begin to approach it as the night comes
to an end, presumably because they become more permissive of less valued options when the opportunity
to mate starts to dwindle. In such cases, liking outcomes are influenced by neural activity in the decision-
making system that weighs not only the hedonic output of the evaluative system but also other factors when
selecting which behavioural act to implement.
In sum, sensory liking evaluations vary significantly with the physiological constitution of the sensory
liking system. The computational complexity of the sensory liking system affects how flexible the rela-
tion of liking and disliking outcomes to a stimulus can be. Organisms with simple sensory liking systems
produce more fixed liking and disliking outcomes characterized by stereotypical behavioural responses to
a stimulus. In contrast, organisms with more complex sensory liking systems produce more flexible behav-
ioural responses where liking and disliking outcomes are influenced by endogenous and contextual factors
that go beyond the stimulus. It is worth stressing that this difference does not suggest that the latter form
of sensory liking represents a qualitative improvement over the former form of sensory liking. All sensory
liking systems are adapted to the species’ habitat and have their pros and cons. Thus, while the neuroana-
tomical innovations to vertebrate evaluative systems have expanded the factors sensory liking evaluations can
take into account—and hence potentially make liking and disliking evaluations better fitted to fluctuating
or uncertain circumstances—they have also increased the risk of producing non-adaptive or “irrational”
outcomes (Camerer, 1995; Santos & Chen, 2009). In humans, for example, it is well known that many lik-
ing and disliking outcomes can be detrimental to survival. For instance, humans will decline free money if

36
Sensory liking

they deem such offers “unfair” (Henrich et al., 2001) or will willingly ingest disgusting (and thus potentially
dangerous) food items (Herz, 2012).
Human sensory liking evaluations are especially flexible. Over the last 30 years, experimental research
has revealed that this flexibility can be explained by the ability of neural mechanisms to represent and
integrate interoceptive information reflecting the physiological state of the body and the individual’s prior
experience with the stimulus being evaluated, as well as cognitive knowledge that pertain to the evaluative
context. These neural processes are distributed widely across the human brain in structures that are involved
in perception, cognition, emotion, and motor control. During evaluative events, they are integrated in idi-
osyncratic ways that reflect the contextual conditions of the specific evaluative event: what the person knows
about the object being evaluated, what her physiological needs are, what the demands are of the task she is
presently engaged in, and so on. The result is liking and disliking outcomes that are tailored to these condi-
tions. Although we do not yet know the computational details of how this integration of evaluative processes
happens, we do know quite a bit about the neural mechanisms the brain uses to represent information that
is of relevance to contextual sensory liking evaluations.

Hedonic evaluation
In humans, for a stimulus to acquire a positive or a negative hedonic value, it must engage neural nuclei in
a network of mesocorticolimbic structures (Figure 2.3). Specifically, whether the stimulus is tagged as liked
or disliked depends on the nature of emotional responses generated by this network of neural processes: lik-
ing outcomes require the generation of states of pleasure, whereas disliking outcomes require the generation

Figure 2.3 Midsagittal representation of the human brain showing the approximate location of the neuroanatomical
structures that make up the human evaluative system: anterior cingulate cortex (ACC), amygdala, insula,
striatum (including nucleus accumbens and pallidum), and the ventromedial prefrontal cortex (vmPFC),
including orbitofrontal cortex (OFC).

37
Martin Skov

of states of negative emotions such as disgust, fear or pain. Experimental evidence suggests that specific
patches embedded in the striatum, amygdala, insula, and ventromedial prefrontal cortex (vmPFC) encode
these hedonic states but that they probably do so through coordinated activity that involves the activation of
processes in all structures. In addition to pleasure and displeasure, the human evaluative system also encodes
a number of other functional mechanisms. These include processes that predict how rewarding a stimulus is
expected to be, processes that calculate risk contingencies, and processes that compare hedonic values associ-
ated with competing behavioural options. Furthermore, all the mesocorticolimbic structures that make up
the evaluative system receive projections from other neural systems, and these projections will often modulate
processes involved in the computation of hedonic value.

Evidence that reward structures encode hedonic values


The bulk of our understanding of the human evaluative system comes from neuroimaging studies where
participants are presented with some stimulus and are asked to rate how much they like or dislike it (Becker
et al., 2019; Kringelbach, 2005; Skov, 2020). Several meta-analyses have pinpointed regions of the human
brain that are commonly activated across such studies (Bartra et al., 2013; Brown et al., 2011; Kühn & Gal-
linat, 2012; Sescousse et al., 2013). Collectively, they have found that subjective liking co-varies with neural
activity that occurs in the striatum, the amygdala, the orbitofrontal cortex (OFC), the anterior cingulate cor-
tex (ACC), and the insula (Figure 2.3). Thus, findings from human neuroimaging suggest that nuclei located
in these structures are involved in the encoding of hedonic valuation. However, neuroimaging results cannot
prove that a relation between observed behaviour and neural activity is causal. We therefore need additional
information to determine if activation of nuclei in the mesocorticolimbic reward system is necessary and
sufficient to paint a stimulus as pleasant or unpleasant.
One line of research that helps cast light on the causal role of the evaluative system consists of exami-
nations of neuropsychological patients with acquired anhedonia, the inability to experience pleasure from
stimuli that used to elicit pleasure. Several psychiatric disorders, including depression, Parkinson’s disease, or
schizophrenia, can cause patients to lose the ability to form pleasure responses (e.g., Gold et al., 2008; Hol-
comb & Rowland, 2007). Studies show that this damage to pleasure responses is associated with structural
and functional alterations to precisely the mesocorticolimbic reward circuitry (e.g., Holcomb & Rowland,
2007; Keller et al., 2013; Schlaepfer et al., 2008), lending support to the notion that engagement of this
system is in fact necessary for the successful generation of sensory liking.
Especially compelling evidence comes from recent work examining the phenomenon of specific musical
anhedonia (SMA). SMA is characterized by a loss of pleasure experienced for musical stimuli even as the
patient’s capacity for generating positive hedonic responses to other stimuli remains intact (Belfi & Loui,
2020). For example, Mas-Herrero and colleagues (2014) demonstrated that a cohort of SMA participants
evinced normal pleasure responses to monetary rewards but reduced pleasure ratings and a lack of autonomic
responses to music. In a follow-up study, Martínez-Molina et al. (2016) found that this dissociation could be
explained by a difference in the way music and money engaged nucleus accumbens (NAcc) in SMA subjects.
Whereas monetary gains elicited normal NAcc responses, the ability of music to activate this structure was
almost non-existent (Martínez-Molina et  al., 2016). Furthermore, modelling of functional connectivity
between auditory cortex and ventral striatum, the anatomical location of NAcc, showed decreased functional
connectivity, a finding subsequent diffusion tensor imaging (DTI) experiments have attributed to a signifi-
cantly lower volume of axonal tracts connecting auditory cortex and NAcc, as well as OFC, in SMA patients
compared to healthy controls (Loui et al., 2017; Sachs et al., 2016).
Even further corroboration of the notion that hedonic evaluations are encoded by neural activity centred
on mesocorticolimbic structures comes from the observation that people with SMA appear fully able to per-
ceive and represent musical stimuli despite their inability to experience them as pleasurable. Thus, comparing

38
Sensory liking

SMA participants to controls Martínez-Molina and colleagues (2016) found no difference in activation of
the superior temporal gyrus that could explain the former’s inability to generate normal hedonic responses
for music. Similarly, neuropsychological testing has found music perception to be intact in SMA patients
(Satoh et  al., 2011, 2016) and the ability to experience pleasure from music preserved in many patients
with auditory agnosia (Matthews et al., 2009), suggesting that the ability to perceive music and the ability
to evaluate its hedonic value are dissociated. Together, these findings strongly support the hypothesis that
hedonic evaluations of sensory inputs require a transfer of activity from sensory systems to nuclei located in
the mesocorticolimbic reward circuitry.
In addition to neuropsychological evidence, direct manipulations of neurons in different parts of the
mesocorticolimbic reward circuitry also support the notion that this network is responsible for generating
hedonic values. Though such experiments are rare in human subjects, extensive work conducted in nonhu-
man animals has demonstrated that targeted excitation or inhibition of neurons in striatum, insula, amygdala,
or OFC are sufficient to elicit liking and disliking responses. For example, in the study by Peng and colleagues
(2015) mentioned previously, photostimulation of sweet and bitter cortical fields in the insula of transduced
mice was sufficient to evoke appetitive and aversive behaviours, even in the absence of any sensation of actual
bitter and sweet tastants. Additionally, Berridge and his group have published a multitude of experiments on
rats that have found neurochemical stimulation of nuclei in the ventral striatum, ventral pallidum, the insula,
and the OFC to enhance or diminish liking and disliking (e.g., Castro & Berridge, 2017; Smith & Berridge,
2005, 2007; Smith et al., 2010). Although it is important not to simply equate animal neurobiology with
human neurobiology, evidence from studies where human subjects are administered naltrexone, an µ-opioid
antagonist, suggests a similar pattern (Nummenmaa  & Tuominen, 2017). Ingestion of naltrexone blocks
opioid receptors in the mesolimbic reward system, causing an attenuation of hedonic liking experienced for
stimuli such as faces (Chelnokova et al., 2014), sweet tastants (Eikemo et al., 2016), money (Petrovic et al.,
2008), music (Mallik et al., 2017), or erotic images (Buchel et al., 2018), as well as diminished neural activity
in regions such as ventral striatum, amygdala, OFC, and ACC as measured by functional magnetic resonance
imaging (fMRI) (Buchel et al., 2018; Petrovic et al., 2008).

The emotional basis of liking


Based on the research reviewed previously, a general consensus holds that liking outcomes occur when
the brain’s evaluative system applies a pleasure gloss onto a sensation (e.g., Becker et al., 2019; Berridge &
Kringelbach, 2015; Kringelbach, 2005; Panksepp, 1998; Smith et  al., 2010). What “pleasure” is, from a
neuroscientific point of view, remains unclear, though. Psychologists and neuroscientists speak of pleasure as
an emotion or an affective process. However, there is disagreement over how to best conceive of emotions
(e.g., Adolphs & Andler, 2018; Barrett, 2017; Berridge, 2018; Damasio & Carvalho, 2013; LeDoux, 2012;
Panksepp, 1998). One way to characterize pleasure is to say that it is an emotional process that entails the
generation of states of positive valence, but in some respects, this feels like exchanging one poorly defined
concept for another. Berridge has called pleasure a “niceness” gloss (Kringelbach & Berridge, 2010, p. 10), a
characterization that seems to tap into our folk psychological understanding of emotions as conscious phe-
nomenological states. As noted previously, it is still unclear whether conscious representations are necessary
for, or even a prevalent component of, emotional processes (see Adolphs & Andler, 2018; Barrett, 2017;
LeDoux, 2012; Panksepp, 1998 for opposing positions on this issue). Certainly, pleasure processes appear to
occur unconsciously, even in humans (Berridge & Winkielman, 2003). For example, Flexas and colleagues
(2013) displayed faces exhibiting happiness—a stimulus known to elicit pleasure—for 20 ms before the pres-
entation of an abstract art stimulus they had asked participants to rate. Even though the participants had no
conscious knowledge of having seen the face, it still influenced their appreciation of the painting, with happy
faces significantly enhancing subjective liking.

39
Martin Skov

It is quite likely that the folk psychological words we use to name emotional processes are too invested
with pre-scientific associations to be of much help in describing neurobiological processes (Buzsáki, 2019;
Pessoa et al., 2021). Indeed, rather than thinking of “pleasure” as one mental function that maps onto one
specific neurobiological process or mechanism, it seems better to think of the word as shorthand for a range
of different neurobiological events. First of all, states of pleasure appear to prompt liking outcomes. As men-
tioned, stimulation of nuclei believed to encode pleasure elicits motor outputs associated with appetitive
behaviour (Castro & Berridge, 2017; Peng et al., 2015; Smith & Berridge, 2005, 2007). Thus, one function
of pleasure computations seems to be to make sensory objects more liked: the more pleasure is generated fol-
lowing exposure to a stimulus, the more liked this stimulus becomes. In humans, this principle also appears
to hold for reported liking ratings. For example, neuroimaging experiments that have investigated the rela-
tion between satiety and food liking have found neural activity in OFC to scale parametrically with reported
pleasure and liking ratings (Kringelbach et al., 2003; Small et al., 2001; Thomas et al., 2015). Indeed, gen-
eration of a pleasure response seems to be a prerequisite for humans to experience sensory objects as likeable
at all, as illustrated by the literature on anhedonia patients reviewed above. Together, these results suggest
that humans probably use the magnitude of pleasure responses to a stimulus to decide how it rates on some
liking scale.
Another function of pleasure computations is the modulation of autonomic nervous system activity.
Regulation of sympathetic arousal, heart rate, respiration, or body temperature is believed to be a central
component in the generation of appetitive motivation (Bradley et  al., 2001; Lang  & Bradley, 2010). In
short, these motive processes help mobilize the organism and play a crucial role in initiating approach and
avoidance behaviour. In humans, reported pleasure for a stimulus is associated with physiological arousal (as
measured by changes in skin conductance) and increased heart and respiration rates but decreased tempera-
ture and pulse amplitude (Bradley et al., 2001; Salimpoor et al., 2009). Interestingly, the different degrees
of physiological activation also scale with how pleasurable and liked a stimulus is reported to be (Salimpoor
et al., 2009), suggesting that the way we consciously experience pleasure depends, in part, on changes in
body states. For example, musical “chills” that are experienced as intensely pleasurable exhibit a physiological
profile that can be distinguished from that of musical pieces that are experienced as merely “highly” pleasur-
able (Salimpoor et al., 2009). Possibly, our phenomenological experience of pleasure computations reflects
physiological changes to our body’s state rather than generation of positive hedonic value per se.
Viewed together, current evidence suggests that we use pleasure to refer to several different neurobiologi-
cal processes, the purpose of which is to enact decision-making and motor outputs associated with appetitive
behaviour as well as to motivate the organism to engage in such acts by modulating autonomic physiological
activity. For animals such as humans, who can consciously represent aspects of these processes, pleasure can
also be accessed through introspection—which allows us to report how much we like an object—and phe-
nomenological feelings. In both cases, the computational rule seems to be that enhanced pleasure predicts
sensory liking: the more pleasure is generated for a sensory object, the more motivated the organism is to
approach and engage appetitively with this object.

The emotional basis of disliking


While the relation between pleasure and sensory liking is somewhat well described—albeit not fully
understood—much less is known about the relation of sensory disliking to negative emotions: “while pleas-
ure and positive valence of hedonics came into the focus of affective neuroscience in recent years, there is
little work on the core processes of displeasure” (Becker and colleagues, 2019, p. 224). Of course, there is
an abundant literature on negative emotions such as pain, fear, and disgust, but curiously little experimental
effort has been expended to determine which of these emotions, if any, underwrite sensory disliking. In the
same way that pleasure states appear to determine liking outcomes, it is to be expected that generation of

40
Sensory liking

negative emotion is necessary for disliking outcomes to occur. Yet the question of how this happens remains
unanswered.
Fear, pain, and disgust are all aversive emotions whose functional purpose is to aid organisms in detecting
and avoiding harmful sensory objects (e.g., Cain, 2018; Corder et al., 2019; Fox & Shackman, 2019; Hart &
Hart, 2018; Sharvit et al., 2015). In this sense, they can all be said to cause sensory disliking as a general prin-
ciple. However, at least in humans there are many sensory objects for which it is unclear if states of fear, pain,
or disgust code for disliking. For example, my dislike for Kanye West’s music does not seem to follow from
any generation of acute fear or pain, at least as far as I am able to perceive consciously. Furthermore, humans
are able to express liking for sensory objects that do elicit fear, pain, or disgust responses, as witnessed by our
willingness to engage in recreational fear (Andersen et al., 2020), sexual sadomasochism (Georgiadis et al.,
2012), or the eating of fermented food items (Herz, 2012). Thus, the computational role negative emotions
play in encoding (human) sensory disliking is not straightforward.
One possibility is that different disliking evaluations engage different negative emotions. Emerging evi-
dence suggests that disliking for different categories of stimuli is associated with the activation of different
negative emotions. Thus, stimuli that signal monetary losses appear to tap into the brain’s pain network
(Delgado et al., 2006; Knutson et al., 2007; Seymour et al., 2007; Seymour, Singer et al., 2007). For exam-
ple, the cost of buying retail goods co-varies with activity in a part of the insula (Knutson et al., 2007) that
is also known to be involved in the encoding of nociceptive pain (Baliki & Apkarian, 2015). Similarly, neu-
roimaging studies have found common activity in the striatum to correlate with acute pain and financial loss
(Delgado, 2007; Seymour, Singer et al., 2007; Tom et al., 2007). Though such results do not prove conclu-
sively that disliking elicited by monetary losses is caused by the activation of the human brain’s pain system,
they suggest aversive outcomes elicited by monetary losses share a neurocomputational architecture with the
encoding of nociceptive pain (Delgado et al., 2006).
In contrast, disliking for other categories of stimuli seems to occur as a consequence of the brain’s disgust
network being engaged. Thus, in the experiment where Flexas and colleagues (2013) found 20-ms presen-
tations of facial happiness to enhance liking ratings for abstract art, they also found similar brief displays of
facial disgust to reduce participants’ reported liking. Other experiments have shown that experienced disgust
for facial disfigurements predicts judgments of ugliness (Bull & Rumsey, 1988; Klebl et al., 2020). In a recent
experiment, Klebl and colleagues (2020) extended this finding to also include the evaluation of buildings,
demonstrating that buildings that were rated as ugly elicited greater disgust responses than buildings rated as
beautiful. Building on these findings, Dorado and colleagues (2022) tested if liking evaluations of different
types of sensory stimuli do in fact engage different negative emotions. First they assessed the participants’
individual sensitivity to the emotions sadness, fear, anger, and disgust. Then they asked the participants to
evaluate stimuli displaying degrees of moral transgressions, more or less symmetrical geometrical patterns,
and rooms varying in tidiness (Dorado et al., 2022). Results showed that individuals who were more sensitive
to anger and fear disliked moral transgressions more than individuals with less sensitivity to anger and fear,
while individuals who were more sensitive to disgust disliked asymmetrical patterns and untidy rooms more
than individuals with less sensitivity to disgust (Dorado et al., 2022). Thus, some categories of sensory objects
appear to become disliked through the engagement of the brain’s pain system, while others become disliked
through the engagement of fear and anger network, and yet others become disliked through the engagement
of the brain’s disgust system.
Unfortunately, neuroscientific exploration of the emotional causes of sensory disliking remains limited.
There is, for instance, an almost complete dearth of experiments that manipulate states of fear, pain, or
disgust in order to test how excitation or inhibition of these systems affects disliking evaluations. Whereas,
as reviewed, we have some evidence that magnitude of pleasure scales with liking outcomes, there is to the
best of my knowledge no study that has investigated if degree of induced pain or disgust scales with disliking
outcomes. Indeed, when it comes to evaluations of disliking for non-adaptive stimuli, it is far from clear that

41
Martin Skov

we have definitive evidence that these are caused by activation of core defensive emotions, even though this
remains the best and the most parsimonious explanation.

Neural correlates of liking and disliking


A central objective of the neuroscientific study of sensory liking is to describe the computational mechanisms
used by the evaluative system. What neuroscientists especially hope to understand are the neurobiological
causes of sensory liking. This aim entails identifying the neurobiological processes that determine whether a
sensory stimulus is experienced as liked or disliked. Evidence from both human and nonhuman animal stud-
ies suggests that such neural processes are in fact widely distributed across more or less the whole evaluative
system, with regions coding for liking and disliking found in most of the anatomical structures that consti-
tute the mesocorticolimbic reward circuitry (Figure 2.4). Moreover, experiments have revealed that sensory
liking evaluations also elicit neural processes that encode other functions than just the positive and negative
affective states underwriting liking and disliking.

Figure 2.4 A schematic overview of the human evaluative system that shows how computational mechanisms map onto
structures in the mesocorticolimbic reward circuitry. Coloured boxes indicate that experiments have found
empirical evidence for the encoding of pleasure (red boxes), displeasure (blue boxes), motivational processes
(grey boxes), or the integration of different source of information pertaining to value outcomes (green
boxes). The overview makes clear that nuclei coding for pleasure and displeasure are distributed across all
structures in the mesocorticolimbic reward circuit. In contrast, nuclei encoding motivational processes such
as wanting or incentive salience appear to be primarily located in the striatum, amygdala, and insula, whereas
processes involved in the integration and computation of value outcomes seem centred on structures in the
ventromedial prefrontal cortex (vmPFC), including the orbitofrontal cortex and anterior cingulate cortex.
Triangles indicate that processes in brainstem and hypothalamic structures project interoceptive information
reflecting the body’s physiological state to mechanisms in the evaluative system.

42
Sensory liking

Much of our general understanding of the neurobiology involved in sensory evaluations comes from
studies of nonhuman animals. Such studies allow for the systematic mapping of different functions by the
stimulation of either individual neurons or small cell assemblies. Manipulating different parts of the evalua-
tive system allows for observing the effect that activation or silencing of individual nuclei has on liking and
disliking responses. As already mentioned, an influential example of this approach is the work conducted by
Berridge and his group. A series of experiments from the Berridge lab has demonstrated that stimulation
of small (1–6 mm3) regions of rat NAcc, ventral pallidum, insula, and OFC is sufficient to enhance liking
for a stimulus, as indexed by orofacial reactions (for reviews, see Smith et al., 2010; Berridge, 2018; see also
Chapter  3). Berridge himself has dubbed these regions hedonic hotspots. Stimulation of other surround-
ing regions (coldspots) appears to suppress liking responses to stimuli, while stimulation of yet other regions
enhances disliking reactions to stimuli such as bitter tastants (Berridge, 2018; Berridge & Kringelbach, 2015;
Castro et al., 2016; Smith et al., 2010). Hotspots that enhance disliking seem especially prominent in the
amygdala and insula but are also found in parts of the striatum and OFC (Cain, 2018; Corder et al., 2019;
Fox & Shackman, 2019; Hart & Hart, 2018; Sharvit et al., 2015).
It remains unclear precisely how these putative hotspots and coldspots for liking and disliking work
together in the course of an evaluative event. Findings from many studies suggest that the different functional
patches interact and modulate the activity of each other (Berridge, 2018; Haber  & Knutson, 2010). For
example, Wang and colleagues (2018) found projections from the gustatory insula cortex—thought to be
involved in the encoding of sweet taste—to the basolateral nucleus of the amygdala to mediate taste liking
but projections from the posterior insula—thought to be involved in the encoding of bitter taste—to the
central nucleus of the amygdala to mediate taste disliking (Wang et al., 2018). Furthermore, it is still unclear
whether evaluative events rely on the activation of single functional regions or the coordination of multiple
hotspots and coldspots across the whole evaluative system, although the latter hypothesis seems more likely.
Berridge (2018) has suggested that stimulating one hotspot recruits other hotspots as well. He has recently
proposed that the “unanimous activation of multiple hotspots together appears required in order to amplify
sensory pleasures” (Berridge, 2018, p. 10), pointing to the observation that simultaneously suppressing one
hotspot while stimulating another fail to elicit liking (Berridge, 2018).
While human neuroscience does not as readily allow for studies that manipulate neural activity, neuro-
imaging experiments have broadly confirmed that subjective ratings of liking and disliking correlate with a
distributed network of neural activity occurring in the striatum, amygdala, insula, and OFC (Bartra et al.,
2013; Brown et al., 2011; Delgado et al., 2006; Knutson et al., 2007; Kühn & Gallinat, 2012; Sescousse
et al., 2013; Seymour et al., 2007; Seymour, Singer et al., 2007). It has been suggested that liking and dis-
liking might map onto dissociable regions in the human brain as well, but this has proven quite difficult to
show with the imaging methods we currently have at our disposal. For example, in an early meta-analysis
conducted by Kringelbach and Rolls (2004), activity related to rewarding outcomes was primarily located in
the medial part of the OFC, whereas activity related to punishing outcomes was mostly located in the lateral
part of the OFC. However, not all neuroimaging findings have supported such a clean-cut dissociation. For
instance, fMRI studies that have modelled neural activity as a continuous function of reported subjective
liking have found blood oxygen level–dependent (BOLD) activity in the medial OFC to scale parametrically
with both positive and negative ratings (e.g., Small et al., 2001; Ishizu & Zeki, 2011; Kringelbach et al.,
2003; Lebreton et al., 2009). It is quite likely that the inability of neuroimaging to probe activity below
large volumes impedes the adequate distinction between different computational functions associated with
liking and disliking evaluations. Moreover, evidence suggests that neural activity occurring in the evaluative
system is mediated by the specific task conditions of the individual evaluation event examined. For exam-
ple, Mas-Herrero and colleagues (2021) conducted a meta-analysis of imaging studies that reported neural
correlates for subjective liking of food and music. Comparisons of evaluations involving these two different
categories of stimuli found some regions of the evaluative system to be commonly activated by both types of

43
Martin Skov

evaluations (NAcc, OFC, insula) but also found certain regions to be more activated for music evaluations
than for food evaluations (ventral striatum) and some regions to be more activated by food evaluations than
by music evaluations (insula, putamen, amygdala). I will return to this issue later.
Another problem associated with identifying neural mechanisms involved in hedonic evaluation is that it
is difficult to distinguish between computational functions that encode the affective gloss (pleasantness, dis-
pleasure) assigned to a stimulus and the motivational impact of such hedonic values. Evidence from animal
studies suggests that dissociable neural processes likely implement different functions. Thus, in the work by
Berridge and his group, stimulation of a different region of NAcc than the hotspot encoding pleasure triggered
increased ingestion of food rather than modulation of orofacial reactions (Smith et al., 2010; Berridge, 2018).
These two patches in the NAcc therefore appear to code for different mechanisms associated with sensory
liking evaluations: the liking hotspot computes the positive hedonic impact of the stimulus, and the wanting
hotspot computes the motivational impact of the stimulus. Activity in the former patch is responsible for
assigning pleasantness to the stimulus, while activity in the latter is responsible for eliciting appetitive behaviour
(Berridge et al., 2009). Similarly, in the amygdala, there seem to be distinct neural ensembles in the basolateral
region that encode the unpleasantness of pain (e.g., Corder et al., 2019) or fear (LeDoux, 2012, 2021), while
other nuclei in the extended amygdala complex code for different types of aversion behaviours (Cain, 2018).
The motivational processes elicited by sensory liking evaluations condition how the organism reacts
behaviourally to stimuli it encounters in the environment. Mammalian brains exhibit a diversity of moti-
vational responses, running the gamut from withdrawal reflexes, freezing, flight, or fighting behaviours to
foraging, ingestion, and courtship behaviours that can be initiated depending on character of the hedonic
value assigned to a stimulus. How an organism reacts to a sensory object in a given situation seems to depend
on the activation of individual motivational mechanisms. In all vertebrates, furthermore, the neural processes
that instantiate these motivational responses are susceptible to learning. For example, single cell recordings
in the ventral striatum of monkeys have revealed that certain cells respond to the unexpected occurrence of
rewards (Schultz et al., 1992). This prediction error signal appears to serve as a neural learning mechanism
that continuously helps update the contingency between a given sensory object and its experienced reward
outcome (Schultz et al., 1997). Similar learning mechanisms exist for teaching the organism if behavioural
engagement with an object will prove punishing (Cain, 2018; LeDoux, 2012). This observation suggests that
motivational responses function as predictors of how rewarding or punishing a sensory object is expected to
be based on previous experiences (Pessiglione & Lebreton, 2015).
In humans, results from neuroimaging studies suggest that exposure to stimuli that are expected to elicit
pleasure activates ventral striatum (e.g., Knutson et  al., 2001, 2005; O’Doherty et  al., 2002; Knutson  &
Greer, 2008; Salimpoor et  al., 2011). Nuclei in this part of the evaluation system may therefore encode
reward predictions that serve to enhance incentive salience and appetitive motivation for the stimulus being
evaluated. Supporting this idea, research by Brian Knutson and others has found that the magnitude of
observed neural activity in this region predicts behaviours such as decisions to buy retail goods or invest in
monetary gambles (Knutson & Genevsky, 2018; Salimpoor et al., 2013). Human imaging experiments have
also found evidence for the encoding of reward prediction errors in the NAcc (e.g., Gold et al., 2019), impli-
cating striatal activity in stimulus-reward learning. Interestingly, these stimulus-reward contingencies also
appear to influence liking and disliking evaluations. Thus, studies of music listening have revealed that reward
expectations modulate the degree of pleasantness of musical stimuli, even though motivational and hedonic
outcomes of the psychological states might be mediated by different processes in ventral striatum and other
structures (Cheung et al., 2019; Gold et al., 2019; Salimpoor et al., 2011; Shany et al., 2019).
In addition to encoding motivational outcomes of sensory liking evaluations, the mesocorticolimbic
evaluative system also harbours processes that track hedonic values of different sensory objects in order for
the organism to be able to choose a preferred option. Electrophysiological recordings in nonhuman animals
and human neuroimaging have consistently implicated the ventromedial prefrontal cortex, and especially the

44
Sensory liking

OFC, in such comparison computations (e.g., Grabenhorst et al., 2010; Lebreton et al., 2009; Lopez-Persem
et al., 2017, 2020; Padoa-Schioppa & Assad, 2006). Existing evidence suggests that OFC neurons contrast
the different hedonic values of available options by using a common linear scale, with decisions generally
driven by the magnitude of the value assigned to each option (Lebreton et al., 2009; Lopez-Persem et al.,
2020; Padoa-Schioppa & Assad, 2006).
As noted in the previous section, however, organisms do not always prefer the sensory object they assign
the greatest hedonic value to. The reason for this is that behavioural decision-making not only considers
hedonic values but also other information of relevance to a given choice dilemma. For example, in addition
to promising an expected reward, a given choice option may also come with a potential loss or punishment.
In a famous fMRI experiment, Knutson and colleagues (2007) showed participants images of chocolates
with and without price information and asked them to decide if they wanted to purchase the chocolate item.
Results found NAcc activity to reflect the hedonic value the participants attributed to individual chocolates,
while insula activity correlated with the expected loss associated with the price tag. Behavioural choices
revealed that purchase decisions weighed both evaluative signals, with participants declining to buy choco-
lates they desired if these were deemed too expensive. Activity in the OFC tracked this cost-benefit analysis,
suggesting that neurons located in this part of the evaluation system compute the integration of different
pieces of contextual information that help inform how beneficial or harmful a sensory object is relative to
the organism’s current situation.
In sum, currently available evidence suggests that several different mesocorticolimbic processes are acti-
vated in the course of sensory liking evaluations. These processes appear to code for different functions,
including the generation of positive and negative affective states, the elicitation of motivational programs
that modulate autonomic physiology, decision-making, motor outputs, and the computation of discrepan-
cies between expected and experienced rewards and punishments. Intriguingly, nuclei distributed across all
of the evaluative system’s core anatomical structures, such as striatum, amygdala, insula, and OFC, seem to
contribute to the encoding of these mechanisms, although processes located in ventral striatum, amygdala,
and insula may be more important to the computation of predictive motivational states compared to pro-
cesses located in the ventromedial prefrontal cortex that, in turn, play a greater role in integrating contex-
tual information and implementing decisions about how to respond to a stimulus based on sensory liking
evaluations (Figure 2.4). However, while it is possible to experimentally distinguish between these different
processes, it is unlikely that they occur in isolation from each other during actual evaluative events. Rather,
our current understanding of sensory liking computations suggests that individual evaluation events marshal
all the mechanisms reviewed here and that the way the different neural processes unfold reflects conditions
of relevance to the evaluative event.

The modulation of evaluation events by interoceptive processes


A central input to the evaluative system comes from afferent projections from the interior of the body that
are routed through brainstem, thalamic, and hypothalamic nuclei (Craig, 2002, 2010; see also Chapter 5).
Together, these interoceptive signals constitute a form of visceral sensation that represents body states, includ-
ing metabolism, muscular sensations, temperature, or touch. Since survival depends on the successful regu-
lation of these physiological states, a core function of sensory liking evaluations is to identify and promote
engagement with parts of the environment that can contribute to energy and thermic regulation, pain anal-
gesia, and so on (Berridge, 2004). Obviously, the ability of the evaluative system to take information about
internal states into account when assessing the hedonic value of a stimulus improves the organism’s chance
of selecting behavioural acts that serve to uphold homeostasis. In humans, evidence suggests that interocep-
tive signals can significantly influence liking and disliking outcomes by modulating neural activity related to
hedonic evaluation in the mesocorticolimbic system.

45
Martin Skov

As described in the previous section, one prominent example of the way sensory liking evaluations incor-
porate interoceptive information is found in the effect of satiation on food and drink liking (Hayes, 2020).
Typically, experienced pleasure for a food item is enhanced when energy deficits are high and diminished
when energy levels rise (Small et al., 2001; Kringelbach et al., 2003). Energetic states related to fluid balance,
glucose metabolism, and so on are signalled in several different ways to the evaluative system. For example,
changes in plasma osmolarity and angiotensin II, a hormone that signals water and salt deficiencies, are
projected to neurons in the lamina terminalis of the forebrain (Zimmerman et al., 2017). From here they
project to the brainstem and hypothalamus, two anatomical structures that are crucial to homeostatic control
(Swanson, 2000). Outputs from these energy balance controllers modulate neural activity in the evaluative
system, in part through dopaminergic projections from the ventral tegmental area (VTA) to the striatum and
amygdala (Zheng & Berthoud, 2007). Glycemic control, similarly, involves afferent projections to neural cir-
cuits in the hypothalamus that provides input to VTA via the parabrachial nucleus (Sternson & Eiselt, 2017).
Using fMRI, Morville and colleagues (2021) have shown that neural activity in human VTA, substantia
nigra, and the parabrachial nucleus varies systematically with fluctuations in serum glucose.
Representations of energy homeostasis in the evaluative system modulate both computations of pleasure
and displeasure for food and drink tastes (Kringelbach et al., 2003; Small et al., 2001; Thomas et al., 2015) as
well as liking responses to food- and drink-related visual cues (Führer et al., 2008; Siep et al., 2009; Thomas
et al., 2015). Thus, diminished pleasantness ratings for stimuli such as chocolate or strawberry liquids obtained
during satiated states correlate with reduced BOLD signals in NAcc, insula, and OFC (Kringelbach et al.,
2003; Small et al., 2001; Thomas et al., 2015). In contrast, enhanced pleasure experienced for similar stimuli
when energy levels are depleted correlates with elevated neural activation in the same regions (Kringelbach
et al., 2003; Small et al., 2001; Thomas et al., 2015).
By enabling variable hedonic evaluations of a food or drink stimulus, the evaluative system can make
better decisions about how to expend resources to pursue and consume specific food items that will meet
the organism’s physiological needs (Hayes, 2020; Zheng & Berthoud, 2007). Not only can flexible liking
and disliking outcomes regulate intake (Hayes, 2020), variations in liking and disliking for individual chemi-
cal compounds can also tailor behavioural responses to specific metabolic needs—for example by favouring
foraging for salt over foraging for glucose when sodium levels are low but serum glucose levels are high
(Zheng & Berthoud, 2007).
Another example of endogenous physiological states modulating sensory liking evaluations is found in
the context of mate choices (Ryan & Jordan, 2017). In most sexual species, reproduction is highly regulated,
with fecundity being determined, at least partially, by fluctuations in the production of gonadal sex hor-
mones such as androgen, testosterone, oestrogen, or progesterone (Roney, 2018). For instance, in females, a
reproductive cycle triggers the production of oestrogen and progesterone at specific moments that not only
influence courtship and copulative behaviours but also are critical to parental behaviours such as egg laying,
lactation, or parental care (Cheng, 2008; Ryan & Jordan, 2017).
Animal experiments also find that oestrogen and progesterone influence how attractive females find male
traits (Rosenthal, 2017; Ryan & Jordan, 2017). This effect seems to be a result of oestrogen and progesterone
modulating neural activity in the mesocorticolimbic evaluative system. Thus, oestradiol levels have been
found in multiple experiments to modulate dopamine production, especially in the striatum (Yoest et al.,
2018; Diekhof, 2018). Studies suggest that this modulation of neural activity is associated with enhanced
sensitivity, and reaction, to rewards (Sakaki & Mather, 2012; Yoest et al., 2018; Diekhof, 2018). In humans, a
few neuroimaging experiments have found activity in the ventral striatum and OFC to be higher for reward-
ing stimuli during the follicular phase than the luteal phase (Dreher et al., 2007; Frank et al., 2010; Rupp
et al., 2009a, 2009b). But a full account of the computational role played by endocrinological inputs to the
evaluative system remains a work in progress.

46
Sensory liking

There is some evidence that changes in hormone production affect the way women’s liking and disliking
evaluations unfold. Hence, some studies have found women to attend more and exhibit heightened reac-
tions to liked images during the follicular phase, suggesting that increases in oestrogen enhance motivational
responses to sensory objects (Pilarczyk et al., 2019; Roney, 2018). In contrast, women appear to experience
enhanced disliking responses to negative stimuli, especially objects and events that signal threat or disgust,
during the luteal phase (Fleischman & Fessler, 2011; Masataka & Shibasaki, 2012). However, attempts to
show that oestrogen and progesterone directly influence how liked or disliked specific stimulus properties are
experienced to be by women have so far proven controversial. For instance, findings suggesting that mas-
culine facial features are more liked by women when oestrogen levels are high (e.g., Johnston et al., 2001;
Penton-Voak et al., 1999; Penton-Voak & Perrett, 2000) have failed to replicate (Marcinkowska et al., 2016;
Jones et al., 2018; Dixson et al., 2018).
On balance, the studies I have reviewed here support the idea that interoceptive information is projected
to the evaluative system, where it modulates neural activity related to the encoding of pleasure, displeasure,
and motivational outputs. By incorporating information about physiological states such as energy homeo-
stasis or fecundity, the evaluative system can tailor sensory liking evaluations to individual adaptive scenarios.

The modulation of evaluation events by perceptual and cognitive processes


In addition to interoceptive signalling, there is mounting evidence that evaluative conditions are reflected by
perceptual and cognitive processes involved in the representation of the stimulus being evaluated. These pro-
cesses allow organisms to represent and integrate contextual information that is of importance to the assess-
ment of a given object’s hedonic value. For example, if a stimulus is unknown, it might be advantageous to
assume that it is potentially dangerous and thus code it as unpleasant. Similarly, having specialized knowledge
about an object, including knowledge about its provenance, precise identity, or how other conspecifics rate
it, can help increase the accuracy of the organism’s evaluative predictions. Research suggests that the human
brain contains numerous mechanisms that allow for the representation of such conditions and that liking and
disliking evaluations are affected by their input to the evaluative system.

Expectations
Predictive coding plays a fundamental role in the representation of contextual conditions (Friston, 2010).
Based on previous experiences, perceptual systems predict what information the organism expects to
encounter. These predictions constitute a gating mechanism that enables the organism to attend to sensory
information that is of relevance to current physiological and behavioural concerns (Alink et al., 2010; Engel
et al., 2013; Egner et al., 2010; Murray et al., 2004; Oliva & Torralba, 2007). Predictive coding, in the con-
text of sensory liking, serves to allocate attention to those parts of the sensory Umwelt that are potentially
rewarding or punishing and to predict the relevance of a stimulus to ongoing survival needs. Thus, sensory
objects predicted to be rewarding attract visual attention (di Pellegrino et al., 2011; Valuch et al., 2015) and
are allocated greater computational resources than other sensory objects (Chen et al., 2012; Li et al., 2016;
Liu & Chen, 2012). In the auditory domain, work on music perception has revealed that the auditory sys-
tem represents musical stimuli in part based on predictions of how tones and chords are temporally related
(Pearce, 2018). These predictions are rooted in the learning of statistical regularities from repeated expo-
sure to music. Over time, learning of such regularities accumulates into a perceptual model of probabilistic
chord, pitch, or tone progressions (Pearce, 2018). Computational modelling has shown that the brain uses
these models to predict the onset on musical events when perceiving new musical stimuli, yielding variable
states of perceptual uncertainty and surprise depending on whether predictions match sensory events or not

47
Martin Skov

(Hansen & Pearce, 2014; Pearce et al., 2010). Deviations from model expectations elicit neural activation
in a connected network of auditory, hippocampal, and evaluative structures, including the amygdala and
NAcc (Cheung et al., 2019; Koelsch et al., 2008). Analysing the relation between perceptual uncertainty,
surprise, and liking, Cheung and colleagues (2019) found the interaction between uncertainty and surprise
about chord progression to elicit pleasure when a surprising chord occurred with low uncertainty or when
an unsurprising chord occurred with high uncertainty.
Examination of the visual system in animal models has revealed that neural activity in the visual cortex
encodes the reward history of a stimulus (Hikosaka et al., 2006; Serences, 2008; Summerfield & de Lange,
2014). This causes neurons in V1 and V4 to change firing rates for visual stimuli depending on the organ-
ism’s reward and punishment history with a given sensory object (Baruni et al., 2015; Gavornik et al., 2009;
Goltstein et al., 2018; Shuler & Bear, 2006). Human eye-tracking work has found visual attention to track
aspects of the visual scene that the individual finds pleasing (Goller et al., 2019; Leder et al., 2016). There is
also emerging evidence that merely attending longer to a stimulus, as indexed by gaze duration, can enhance
liking for a visual stimulus (Shimojo et al., 2003).
Famously, “mere exposure” to a visual stimulus is enough to modulate sensory liking evaluations, with
repeated exposure enhancing how liked a sensory object is experienced to be (Bornstein, & D’Agostino,
1992; Zajonc, 1968). It has been suggested that this effect is a function of biological organisms’ aversion
to uncertain or unpredictable sensory events (Zajonc, 1980). In a combined electrophysiology and fMRI
study, Herry and colleagues (2007) demonstrated that mice and humans responded aversively to different
task assignments, as quantified by enhanced place avoidance in mice and elevated attention to angry faces in
humans, when exposed to sequences of neutral sound pulses varying in temporal predictability. This increase
in aversive behaviour during unpredictable sound events correlated with higher engagement of the amygdala
(Herry et al., 2007). In a follow-up experiment, Ramsøy and colleagues (2012) showed that exposing human
participants to the same pulse sequences while they rated how much they liked brand logos and artworks
yielded lower liking outcomes during unpredictable perceptual events than during predictable perceptual
events. Together, these results suggest that humans and other animals interpret perceptual uncertainty as
indicative of potential danger that must be avoided and that, consequently, they evaluate sensory objects
affected by this uncertainty as less liked.
In sum, research investigating the influence of perceptual predictions on sensory liking evaluations shows
that perceptual systems represent how familiar a stimulus is even at the perceptual level. By matching model
predictions to actual sensory information, perceptual processes encode if a stimulus is novel, surprising, or
known to elicit pleasure or displeasure and project this information to the evaluative system, where it is inte-
grated into evaluations of how pleasing or displeasing a stimulus is.

Knowledge
In addition to predictive coding, representations of stimuli are heavily modulated by conceptual knowledge.
Semantic and episodic memory of an object influence how it is perceived by activating associative networks
of conceptual nodes based on the nature of the sensory input (Buckley & Gaffan, 2006; Miyashita, 2004).
These associative networks are encoded by a distributed network of nuclei in different parts of the tempo-
ral lobe, including nuclei located in the hippocampal formation, parahippocampal and perirhinal cortices,
and fusiform cortex (Devlin & Price, 2007; Henke, 2010). Through activations of this network, sensory-
cognitive associations can individualize the object in ways that allow for more precise sensory liking evalua-
tions that fit the individual’s needs and behavioural agenda. For example, whether it is prudent to approach
or avoid an object can depend on the object’s provenance, identity, or social status.
There is abundant evidence that the conceptual knowledge an individual brings to bear on the perceptual
and cognitive representation of a stimulus can influence how liked or disliked it is. For example, acquired

48
Sensory liking

knowledge about a stimulus affects the degree to which it is evaluated to be liked or disliked, as evidenced by
comparisons of experimental subjects who vary in expertise (Kozbelt, 2020). For instance, while most peo-
ple like sensory objects with curved contours more than objects with angular contours (Chuquichambi et al.,
2022), architects and designers report diminished liking for curved buildings combined with an enhanced
liking for angular architectural structures (Palumbo et al., 2020; Vartanian et al., 2019). Kirk and colleagues
(2009a) used fMRI to compare neural activity occurring during sensory liking evaluations of buildings in
architects and non-architects and found greater activation of hippocampus and precuneus in the architects,
a result that might suggest recruitment of stored associations acquired through extensive engagement with
architectural objects. This idea is partially corroborated by findings from other neuroimaging experiments
that have demonstrated that the density and remoteness of associations generated by object perception can be
predicted from the degree to which the entorhinal, parahippocampal, perirhinal, and fusiform cortices are
engaged (Friis-Olivarius et al., 2017; Hulme et al., 2014). It should be emphasized, though, that it remains
largely unknown how processes encoding stimulus expertise modulate liking and disliking mechanisms in
the evaluative system.
Experiments have also shown that it is possible to influence liking and disliking for sensory stimuli by
manipulating a person’s immediate state of knowledge with respect to a stimulus. This can be demonstrated
by furnishing people engaged in sensory liking evaluations with different items of information about the
stimulus being evaluated (Okamoto & Dan, 2013). For instance, being told that a stimulus is exclusive rather
than generally available, expensive rather than cheap, or fabricated by a luxury manufacturer rather than by a
less prestigious producer are all pieces of information that have been experimentally shown to enhance liking
despite the fact that the stimulus remains the same under both conditions (Krishna, 2012; Okamoto & Dan,
2013; Fernqvist & Ekelund, 2014; Piqueras-Fiszman & Spence, 2018). Several neuroimaging studies have
investigated what happens during such evaluative events and found that object-external information modu-
lates neural activity in parts of the mesocorticolimbic reward circuit, including the OFC (e.g., McClure
et al., 2004; Plassmann et al., 2008; Kirk et al., 2009b). In their study, Kirk and colleagues (2009b) also found
evidence that the different pieces of semantic information used to describe the stimulus elicited variable
activity in the entorhinal cortex and the temporal pole, key anatomical structures involved in the encoding
of memory and conceptual knowledge.

Evaluative task conditions


Finally, sensory liking evaluations are also modulated by neural mechanisms that represent the task conditions
of a particular evaluation event. I have already noted some examples of this phenomenon. For instance, how
sexually attractive a courter is deemed to be can vary depending on whether the chooser makes the evalua-
tion early in the evening or just before closing hours (Rosenthal, 2017; Ryan & Jordan, 2017). Similarly, lik-
ing outcomes are also modulated by the availability of options: Is the apple being evaluated the only possible
source of energy, or are there other, perhaps better, options around? As a general principle, sensory liking
evaluations are always embedded within specific task conditions that the organism needs to take into account
to make the best behavioural decision. How the human brain accomplishes this computational task remains
somewhat unclear, but it is likely that prefrontal structures construct a model of relevant task conditions—for
example, what behavioural act the sensory liking evaluation is in service of, what the evaluative anchor being
used is, how risky and potentially rewarding or punishing the available options are, and so on—that is used to
modulate sensory and evaluative processes (Dayan & Berridge, 2014; O’Doherty et al., 2017).
Neuroimaging experiments that have compared explicit evaluation tasks to other non-evaluation tasks
have found that attending to a stimulus with the instruction to rate how liked it is according to a specific
evaluative anchor enhances neural activity in a broad network of anatomical structures that encode both per-
ceptual information and hedonic values (e.g., Chatterjee et al., 2009; Grabenhorst & Rolls, 2008; Ishizu &

49
Martin Skov

Zeki, 2013; Jacobsen et al., 2006; Kim et al., 2007; Lebreton et al., 2009). This finding suggests that task
conditions modulate both the way the stimulus is represented and computations of liking and disliking.
Although it is still unclear what functional mechanisms these activation patterns reflect, emerging evidence
from experiments examining the impact of different task conditions on liking outcomes suggests that one
mechanism involved in the computation of explicit evaluations is a matching of perceptual and emotional
states with a model representing task conditions.
For example, Sherman and colleagues (2015) measured individual working memory capacity (WMC)
for a cohort of participants and asked how different levels of WMC affected explicit evaluations of liking for
paintings. Results revealed that participants with greater visual WMC reported elevated ratings for paintings
with high degrees of visual complexity. The implication of this finding is that humans hold visual informa-
tion relevant to the evaluative task present in working memory while making explicit judgment decisions.
Che and colleagues (2021) tested this hypothesis directly by asking participants to rate visual art according to
two different explicit judgments, beauty and liking. This evaluation task was embedded in a working memory
task where participants saw a matrix with 1, 3, or 5 dots before the painting. They were instructed to remem-
ber the first matrix while viewing and rating the paintings because they would then have to reproduce it
after. This manipulation loaded the participants’ visual working memory during sensory liking evaluations
and allowed for a direct test of the idea that people hold visual information relevant to an evaluation task
during the decision phase. Results of Che et al.’s (2021) study confirmed that loading of working memory
did indeed influence processes associated with evaluation: The participants took significantly longer time to
make beauty judgments than to make liking judgments. This finding suggests that beauty evaluations make
greater demand on visual working memory than liking evaluations, a conclusion that is further supported
by another finding from the experiment, that participants with greater WMC were faster to complete the
working memory task after liking judgments than after beauty judgments (Che et al., 2021).
One possible reason explicit beauty judgments engage working memory processes more than liking
judgments is that people consider beauty a more complex, and possibly more restrictive, evaluative model
than liking (Skov & Nadal, 2021). We know from Brielmann and Pelli’s (2019) work that to be judged as
beautiful, a stimulus must elicit greater pleasure than to be judged as likeable. This finding indicates that peo-
ple conceive of beauty as a more restrictive category than liking and that explicit evaluations of beauty use
introspection to decide if the pleasure felt for a stimulus matches this high end of the spectrum. It is similarly
possible that people conceive of beautiful objects as having specific object properties and that explicit evalu-
ations of beauty involve a process of matching the actual perceptual information to this model. This process,
as the previous evidence suggests, likely is mediated by working memory mechanisms.
A recent experiment by Che and colleagues (2022) lends tentative support to the idea that people con-
ceive of beauty and liking as different evaluative anchors and that cognitive models of what counts as beauti-
ful or likeable inform explicit sensory liking evaluations that make use of these evaluative anchors. Che et al.
(2022) asked participants to rate faces and paintings using both beauty and liking as reported judgments. In
one part of the experiment, participants were left to conceive of beauty and liking using their own intui-
tions. In another part, Che and colleagues (2022) gave the participants specific instructions with respect to
how they should think of beauty and liking as evaluative anchors. Liking judgments, the participants were
told, are subjective and based on inner feelings of how pleasant an object is. Beauty judgments, in contrast,
they were instructed to think of as objective, based on the object’s order or proportion. Comparing ratings
from the two test conditions, Che et al. (2022) found beauty judgments to take a markedly longer time and
liking judgments to take a noticeably shorter time during the instruction condition than during the non-
instruction condition. Beauty ratings were also higher with instructions than without, while liking ratings
were lower with instructions than without instructions. It is difficult to see how these results would emerge
unless participants model beauty as a more specialized evaluative anchor than liking, with only certain per-
ceptual features and greater pleasure allowing a stimulus to be judged as beautiful. Determining if a stimulus

50
Sensory liking

matches model requirements therefore takes a longer time when people are asked to assess if it is beautiful
than when they are asked to decide if it is simply pleasant (Skov & Nadal, 2021).
While it remains to be worked out precisely how cognitive mechanisms represent task conditions and
how such representations come to modulate evaluative computations in the mesocorticolimbic system, it
seems clear that the conditions under which a stimulus is being evaluated are factored into assessments of how
liked or disliked it is. Together with perceptual predictions and conceptual associations, such representations
of tasks requirements and goals help further contextualize sensory liking evaluations by integrating informa-
tion about the organism’s previous experiences with the object, its acquired knowledge, and the parameters
of the current behavioural task that the sensory liking evaluation is meant to serve.

Conclusion
I started this chapter by considering the historical assumption that the key to understanding why humans
and other biological organisms like some sensory objects and dislike others lies in unearthing either innate
or learned relationships between stimulus features and hedonic reactions. As my overview of experimental
work has shown, this assumption is not tenable. Rather, existing evidence strongly compels us to view liking
and disliking outcomes as the product of an integration of multiple factors that include, but are not limited
to, stimulus information: the physiological states of the organism, its motivational needs, prior experiences
with the stimulus, cognitive knowledge, and the weighing of goals and requirements associated with the
behavioural act the organism is presently engaged in. The adaptive value of such flexible liking and disliking
outcomes is obvious: Instead of being forced to respond in a stereotypical manner to sensory stimulation,
flexible liking and disliking responses allow for behavioural responses that are tailored to the organism’s fluc-
tuating survival needs and changes in the environment it inhabits.
Centrally, liking and disliking outcomes are flexible because sensory liking evaluations consist of large-
scale integration of computational mechanisms that represent these contextual conditions (Figure 2.5). As my
review of experimental findings demonstrates, how much a stimulus is liked or disliked in a given situation
depends upon the specific, combined contribution from all, or most, of these mechanisms. If a stimulus is
unknown, it can be less liked than if it is well known. If the organism has previously liked a stimulus, that
experience boosts expectations that the stimulus will also be found pleasant in the current situation, and so
on. Individual computational mechanisms make contributions that are modulated by the particular state of
the evaluative event. It is therefore misleading to conceive of sensory liking evaluations as reflexive reactions
to stimulus information. Sensory liking evaluations are temporal events where information from multiple
computational nodes is projected back and forth under influence of both the stimulus and the evaluative
context.
This revised explanatory framework, where sensory liking evaluations are understood as computational
events with variable inputs from different mechanisms, raises the question whether it is possible to categorize
evaluation events according to their function. In other words, instead of characterizing the function of dif-
ferent forms of sensory liking evaluation by basing definitions on the eliciting stimulus—as we normally do
by speaking of food hedonics, sexual preferences, economic utility, or art appreciation—it may make more
analytical sense to see different sensory liking evaluations as determined by properties inherent to individual
events. For instance, some evaluation events might be more reliant on projections from mechanisms rep-
resenting energy homeostasis (e.g., when considering what to make for lunch), while others may engage
mechanisms representing endocrinological levels to a higher degree (e.g., when looking though a dating app
for potential dates). To date, little work, however, has been conducted to answer this interesting question.
A shift in focus from the stimulus to evaluative events could be a way for the field of empirical aesthet-
ics to make progress on the longstanding question of whether “aesthetic evaluations” constitute a sui generis
form of sensory liking (see Chapter 1). Attempts to define aesthetic liking as distinct from other kinds of

51
Martin Skov

Figure 2.5 Model of computational mechanisms known to be involved in human sensory liking evaluations, based on
the empirical evidence reviewed in the chapter. Activation of each mechanism holds the power to modulate
liking and disliking for a given stimulus. Current evidence suggests that the specific way the whole system
of computational nodes is activated during an evaluative event depends on physiological, environmental,
and behavioural circumstances. For example, engagement of processes coding for perceptual and reward
predictions is a function of the organism’s previous experience with the stimulus being evaluated. Similarly,
whether the organism attends to the stimulus with the express purpose of assessing if it fits task requirements
of a prospective behavioural act or if it “passively” computes the object’s hedonic value while focused on
another task is determined by the contextual conditions of the evaluation event. Projections from mecha-
nisms representing endogenous and exogenous states of relevance to the individual evaluative event project
to neural mechanisms in the evaluative system that code for pleasure, displeasure, and motivational outputs.
Liking and disliking outcomes emerge as a result of the connected pattern of neural activity.

sensory liking have traditionally motivated their arguments by appealing to the idea that aesthetic “quali-
ties” constitute a special class of stimulus properties. For example, aesthetic evaluations have been considered
distinct because they are directed at certain stimulus features (e.g., Menninghaus et al., 2019). Other theories
have proposed that aesthetic stimuli “afford” specific evaluative responses, including especially intense states
of pleasure (e.g., Makin, 2017) or more contemplative hedonic states of “being moved” (e.g., Vessel, 2020).
However, examining the scientific evidence accumulated over the last 20 years, reveals no empirical support
for such claims (e.g., Skov & Nadal, 2020, 2022).
An alternative hypothesis could be that certain evaluative events qualify as aesthetic. There is for instance
a substantial amount of evidence that emotional responses are attenuated during certain types of sensory
liking evaluations. Thus, the encoding of fear appears to be diminished in evaluative situations where peo-
ple expect the stimulus to be fictive (e.g., Mocaiber et al., 2010; Van Dongen et al., 2016). Similarly, it has
been hypothesized that pleasure responses are not accompanied by normal motivational outputs when they
occur in the context of liking evaluations people believe are directed at art stimuli (e.g., Sarasso et al., 2020;

52
Sensory liking

but see Skov & Nadal, 2022). If this idea turns out to be true—and the jury is still out on whether it is—it
might make theoretical sense to designate evaluation events where motive responses are regulated as a conse-
quence of the belief that potential liking and disliking outcomes do not concern survival needs as “aesthetic”
evaluations.
Of course, neither emotional regulation nor stimulus attention is a computational mechanism that is
unique to the representation of specific behavioural contexts in the same way energy homeostasis might be
unique to the representation of food consumption or gonadal hormone levels might be unique to the rep-
resentation of sexual courtship and reproduction. Regulation of adaptive emotions seems equally important
to sensory liking evaluations that take place in service of the appreciation of horror movies and evaluations
that occur in the context of enjoying fermented cheeses. Whether we feel the need to think of sensory liking
evaluations that rely especially on the attenuation of adaptive emotions as “aesthetic” rests entirely on how
we define the concept of aesthetics.
In any case, understanding why a sensory stimulus is liked or disliked involves factoring in the computa-
tional context of the individual evaluation event. Sensory liking always serves physiological and behavioural
needs. The human brain has evolved a large number of neural mechanisms that respond to the varying
conditions of different evaluation events. Only by understanding the way these mechanisms communicate
with each other and integrate information relative to the circumstances of individual events can we hope to
develop a computational theory of why a stimulus become liked or disliked.

References
Adolphs, R.,  & Andler, D. (2018). Investigating emotions as functional states distinct from feelings. Emotion Review,
10(3), 191–201. https://doi.org/10.1177/1754073918765662
Alink, A., Schwiedrzik, C. M., Kohler, A., Singer, W., & Muckli, L. (2010). Stimulus predictability reduces responses
in primary visual cortex. Journal of Neuroscience, 30(8), 2960–2966. https://doi.org/10.1523/JNEUROSCI.3730-
10.2010
Andersen, M. M., Schjoedt, U., Price, H., Rosas, F. E., Scrivner, C., & Clasen, M. (2020). Playing with fear: A field
study in recreational horror. Psychological Science, 31(12), 1497–1510. https://doi.org/10.1177/0956797620972116
Baliki, M. N., & Apkarian, A. V. (2015). Nociception, pain, negative moods, and behavior selection. Neuron, 87(3),
474–491. https://doi.org/10.1016/j.neuron.2015.06.005
Barrett, L. F. (2017). How emotions are made: The secret life of the brain. Houghton Mifflin Harcourt.
Bartra, O., McGuire, J. T., & Kable, J. W. (2013). The valuation system: A coordinate-based meta-analysis of BOLD fMRI
experiments examining neural correlates of subjective value. Neuroimage, 76, 412–427. https://doi.org/10.1016/j.
neuroimage.2013.02.063
Baruni, J. K., Lau, B., & Salzman, C. D. (2015). Reward expectation differentially modulates attentional behavior and
activity in visual area V4. Nature Neuroscience, 18(11), 1656–1663. https://doi.org/10.1038/nn.4141
Bateson, M., Healy, S. D., & Hurly, T. A. (2002). Irrational choices in hummingbird foraging behaviour. Animal Behav-
iour, 63(3), 587–596. https://doi.org/10.1006/anbe.2001.1925
Becker, S., Bräscher, A. K., Bannister, S., Bensafi, M., Calma-Birling, D., Chan, R. C. K., Eerola, T., Ellingsen, D. M.,
Ferdenzi, C., Hanson, J. L., Joffily, M., Lidhar, N. K., Lowe, L. J., Martin, L. J., Musser, E. D., Noll-Hussong, M.,
Olino, T. M., Pintos Lobo, R., & Wang, Y. (2019). The role of hedonics in the human affectome. Neuroscience and
Biobehavioral Reviews, 102, 221–241. https://doi.org/10.1016/j.neubiorev.2019.05.003
Belfi, A. M., & Loui, P. (2020). Musical anhedonia and rewards of music listening: Current advances and a proposed
model. Annals of the New York Academy of Sciences, 1464(1), 99–114. https://doi.org/10.1111/nyas.14241
Berridge, K. C. (2004). Motivation concepts in behavioral neuroscience. Physiology and Behavior, 81(2), 179–209. https://
doi.org/10.1016/j.physbeh.2004.02.004
Berridge, K. C. (2018). Evolving concepts of emotion and motivation. Frontiers in Psychology, 9, 1647. https://doi.
org/10.3389/fpsyg.2018.01647
Berridge, K. C.,  & Kringelbach, M. L. (2015). Pleasure systems in the brain. Neuron, 86(3), 646–664. https://doi.
org/10.1016/j.neuron.2015.02.018
Berridge, K. C., Robinson, T. E., & Aldridge, J. W. (2009). Dissecting components of reward: “liking”, “wanting”, and
learning. Current Opinion in Pharmacology, 9(1), 65–73. https://doi.org/10.1016/j.coph.2008.12.014

53
Martin Skov

Berridge, K. C., & Winkielman, P. (2003). What is an unconscious emotion?(The case for unconscious “liking”). Cogni-
tion and Emotion, 17(2), 181–211. https://doi.org/10.1080/02699930302289
Bornstein, R. F., & D’Agostino, P. R. (1992). Stimulus recognition and the mere exposure effect. Journal of Personality and
Social Psychology, 63(4), 545–552. https://doi.org/10.1037//0022-3514.63.4.545
Bradley, M. M., Codispoti, M., Cuthbert, B. N., & Lang, P. J. (2001). Emotion and motivation I: Defensive and appeti-
tive reactions in picture processing. Emotion, 1(3), 276–298. https://doi.org/10.1037/1528-3542.1.3.276
Brielmann, A. A., & Pelli, D. G. (2019). Intense beauty requires intense pleasure. Frontiers in Psychology, 10, 2420. https://
doi.org/10.3389/fpsyg.2019.02420
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. Neuroimage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Buchel, C., Miedl, S., & Sprenger, C. (2018). Hedonic processing in humans is mediated by an opioidergic mechanism
in a mesocorticolimbic system. eLife, 7, e39648. https://doi.org/10.7554/eLife.39648
Buckley, M. J., & Gaffan, D. (2006). Perirhinal cortical contributions to object perception. Trends in Cognitive Sciences,
10(3), 100–107. https://doi.org/10.1016/j.tics.2006.01.008
Bull, R., & Rumsey, N. (1988). The social psychology of facial disfigurement. In The social psychology of facial appearance
(pp. 179–215). Springer.
Buzsáki, G. (2019). The brain from inside out. Oxford University Press.
Cain, C. K. (2018). Avoidance problems reconsidered. Current Opinion in Behavioral Sciences, 26, 9–17. https://doi.
org/10.1016/j.cobeha.2018.09.002
Calhoun, A. J., & Hayden, B. Y. (2015). The foraging brain. Current Opinion in Behavioral Sciences, 5, 24–31. https://doi.
org/10.1016/j.cobeha.2015.07.003
Camerer, C. (1995). Individual decision making. In J. H. Kagel & A. E. Roth (Eds.), Handbook of experimental economics
(pp. 587–703). Princeton University Press.
Castro, D. C., & Berridge, K. C. (2017). Opioid and orexin hedonic hotspots in rat orbitofrontal cortex and insula.
Proceedings of the National Academy of Sciences of the United States of America, 114(43), E9125–E9134. https://doi.
org/10.1073/pnas.1705753114
Castro, D. C., Terry, R. A., & Berridge, K. C. (2016). Orexin in rostral hotspot of nucleus accumbens enhances sucrose
‘liking’ and intake but scopolamine in caudal shell shifts ‘liking’ toward ‘disgust’ and ‘fear’. Neuropsychopharmacology,
41(8), 2101–2111. https://doi.org/10.1038/npp.2016.10
Chatterjee, A., Thomas, A., Smith, S. E., & Aguirre, G. K. (2009). The neural response to facial attractiveness. Neuropsy-
chology, 23(2), 135–143. https://doi.org/10.1037/a0014430
Che, J., Sun, X., Gallardo, V., & Nadal, M. (2018). Cross-cultural empirical aesthetics. In Progress in Brain Research (Vol.
237). Elsevier. https://doi.org/10.1016/bs.pbr.2018.03.002
Che, J., Sun, X., Skov, M., & Nadal, M. (2022, In press). Judging beauty and liking: The effects of personal intuitions
and task instructions. Psychology of Aesthetics, Creativity and the Arts.
Che, J., Sun, X., Skov, M., Vartanian, O., Rosselló, J., & Nadal, M. (2021). The role of working memory capacity in
evaluative judgments of liking and beauty. Cognition and Emotion, 35(7), 1407–1415. https://doi.org/10.1080/0269
9931.2021.1947781
Chelnokova, O., Laeng, B., Eikemo, M., Riegels, J., Løseth, G., Maurud, H., Willoch, F., & Leknes, S. (2014). Rewards
of beauty: The opioid system mediates social motivation in humans. Molecular Psychiatry, 19(7), 746–747. https://doi.
org/10.1038/mp.2014.1
Chen, W., Liu, C. H., & Nakabayashi, K. (2012). Beauty hinders attention switch in change detection: The role of facial
attractiveness and distinctiveness. PLOS ONE, 7(2), e32897. https://doi.org/10.1371/journal.pone.0032897
Cheng, M. F. (2008). The role of vocal self-stimulation in female responses to males: Implications for state-reading. Hor-
mones and Behavior, 53(1), 1–10. https://doi.org/10.1016/j.yhbeh.2007.08.007
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology,
29(23), 4084–4092.e4. https://doi.org/10.1016/j.cub.2019.09.067
Chuquichambi, E. G., Vartanian, O., Corradi, G., Silvia, P., Nadal, M., Skov, M., & Munar, E. (2022). How universal is
preference for visual curvature? A systematic review and meta-analysis. Under review.
Cisek, P. (2021). Evolution of behavioural control from chordates to primates. Philosophical Transactions of the Royal Society
of London Series B, 377(1844).
Coppin, G., & Sander, D. (2012). The flexibility of chemosensory preferences. In R. J. Dolan & T. Sharot (Eds.), Neu-
roscience of preferences and choice (pp. 257–275). Academic Press.
Corder, G., Ahanonu, B., Grewe, B. F., Wang, D., Schnitzer, M. J., & Scherrer, G. (2019). An amygdalar neural ensem-
ble that encodes the unpleasantness of pain. Science, 363(6424), 276–281. https://doi.org/10.1126/science.aap8586

54
Sensory liking

Corradi, G., Chuquichambi, E. G., Barrada, J. R., Clemente, A., & Nadal, M. (2020). A new conception of visual aes-
thetic sensitivity. British Journal of Psychology, 111(4), 630–658. https://doi.org/10.1111/bjop.12427
Craig, A. D. (2002). How do you feel? Interoception: The sense of the physiological condition of the body. Nature
Reviews. Neuroscience, 3(8), 655–666. https://doi.org/10.1038/nrn894
Craig, A. D. (2010). The sentient self. Brain Structure and Function, 214(5–6), 563–577. https://doi.org/10.1007/
s00429-010-0248-y
Cummings, M. E., & Endler, J. A (2018). 25 years of sensory drive: The evidence and its watery bias. Current Zoology,
64(4), 471–484.
Damasio, A., & Carvalho, G. B. (2013). The nature of feelings: Evolutionary and neurobiological origins. Nature Reviews.
Neuroscience, 14(2), 143–152. https://doi.org/10.1038/nrn3403
Darwin, C. (1871). The descent of man and selection in relation to sex. Murray.
Dayan, P., & Berridge, K. C. (2014). Model-based and model-free Pavlovian reward learning: Revaluation, revision, and
revelation. Cognitive, Affective and Behavioral Neuroscience, 14(2), 473–492. https://doi.org/10.3758/s13415-014-0277-8
Delgado, M. R. (2007). Reward-related responses in the human striatum. Annals of the New York Academy of Sciences,
1104, 70–88. https://doi.org/10.1196/annals.1390.002
Delgado, M. R., Labouliere, C. D., & Phelps, E. A. (2006). Fear of losing money? Aversive conditioning with secondary
reinforcers. Social Cognitive and Affective Neuroscience, 1(3), 250–259. https://doi.org/10.1093/scan/nsl025
Devlin, J. T., & Price, C. J. (2007). Perirhinal contributions to human visual perception. Current Biology, 17(17), 1484–
1488. https://doi.org/10.1016/j.cub.2007.07.066
Dieckmann, H. (1974). Theories of beauty to the mid-nineteenth century. In P. P. Wiener (Ed.), Dictionary of the history
of ideas, I (pp. 195–206). Charles Scribner’s Sons.
Diekhof, E. K. (2018). Estradiol and the reward system in humans. Current Opinion in Behavioral Sciences, 23, 58–64.
https://doi.org/10.1016/j.cobeha.2018.03.010
Dinehart, M. E., Hayes, J. E., Bartoshuk, L. M., Lanier, S. L., & Duffy, V. B. (2006). Bitter taste markers explain variabil-
ity in vegetable sweetness, bitterness, and intake. Physiology and Behavior, 87(2), 304–313. https://doi.org/10.1016/j.
physbeh.2005.10.018
di Pellegrino, G., Magarelli, S., & Mengarelli, F. (2011). Food pleasantness affects visual selective attention. Quarterly
Journal of Experimental Psychology, 64(3), 560–571. https://doi.org/10.1080/17470218.2010.504031
Dixson, B. J. W., Blake, K. R., Denson, T. F., Gooda-Vossos, A., O’Dean, S. M., Sulikowski, D., Rantala, M. J., &
Brooks, R. C. (2018). The role of mating context and fecundability in women’s preferences for men’s facial masculin-
ity and beardedness. Psychoneuroendocrinology, 93, 90–102. https://doi.org/10.1016/j.psyneuen.2018.04.007
Dorado, A., Skov, M., Rosselló, J., & Nadal, M. (2022). Defensive emotions and evaluative judgments: Sensitivity to
anger and fear predicts moral judgments, whereas sensitivity to disgust predicts aesthetic judgments. British Journal of
Psychology.
Dreher, J. C., Schmidt, P. J., Kohn, P., Furman, D., Rubinow, D., & Berman, K. F. (2007). Menstrual cycle phase modu-
lates reward-related neural function in women. Proceedings of the National Academy of Sciences of the United States of
America, 104(7), 2465–2470. https://doi.org/10.1073/pnas.0605569104
Egner, T., Monti, J. M.,  & Summerfield, C. (2010). Expectation and surprise determine neural population
responses in the ventral visual stream. Journal of Neuroscience, 30(49), 16601–16608. https://doi.org/10.1523/
JNEUROSCI.2770-10.2010
Eikemo, M., Løseth, G. E., Johnstone, T., Gjerstad, J., Willoch, F.,  & Leknes, S. (2016). Sweet taste pleasantness is
modulated by morphine and naltrexone. Psychopharmacology, 233(21–22), 3711–3723. https://doi.org/10.1007/
s00213-016-4403-x
Engel, A. K., Maye, A., Kurthen, M., & König, P. (2013). Where’s the action? The pragmatic turn in cognitive science.
Trends in Cognitive Sciences, 17(5), 202–209. https://doi.org/10.1016/j.tics.2013.03.006
Falke, J. J., Bass, R. B., Butler, S. L., Chervitz, S. A., & Danielson, M. A. (1997). The two-component signaling path-
way of bacterial chemotaxis: A molecular view of signal transduction by receptors, kinases, and adaptation enzymes.
Annual Review of Cell and Developmental Biology, 13, 457–512. https://doi.org/10.1146/annurev.cellbio.13.1.457
Fernqvist, F., & Ekelund, L. (2014). Credence and the effect on consumer liking of food—A review. Food Quality and
Preference, 32, 340–353. https://doi.org/10.1016/j.foodqual.2013.10.005
Fleischman, D. S.,  & Fessler, D. M. T. (2011). Progesterone’s effects on the psychology of disease avoidance: Sup-
port for the compensatory behavioral prophylaxis hypothesis. Hormones and Behavior, 59(2), 271–275. https://doi.
org/10.1016/j.yhbeh.2010.11.014
Flexas, A., Rosselló, J., Christensen, J. F., Nadal, M., Olivera La Rosa, A., & Munar, E. (2013). Affective priming using
facial expressions modulates liking for abstract art. PLOS ONE, 8(11), e80154. https://doi.org/10.1371/journal.
pone.0080154

55
Martin Skov

Fox, A. S., & Shackman, A. J. (2019). The central extended amygdala in fear and anxiety: Closing the gap between
mechanistic and neuroimaging research. Neuroscience Letters, 693, 58–67.
Frank, T. C., Kim, G. L., Krzemien, A., & Van Vugt, D. A. (2010). Effect of menstrual cycle phase on corticolimbic brain
activation by visual food cues. Brain Research, 1363, 81–92. https://doi.org/10.1016/j.brainres.2010.09.071
Friis-Olivarius, M., Hulme, O. J., Skov, M., Ramsøy, T. Z., & Siebner, H. R. (2017). Imaging the creative unconscious:
Reflexive neural responses to objects in the visual and parahippocampal region predicts state and trait creativity. Sci-
entific Reports, 7(1), 14420. https://doi.org/10.1038/s41598-017-14729-7
Friston, K. (2010). The free-energy principle: A  unified brain theory? Nature Reviews. Neuroscience, 11(2), 127–138.
https://doi.org/10.1038/nrn2787
Führer, D., Zysset, S., & Stumvoll, M. (2008). Brain activity in hunger and satiety: An exploratory visually stimulated
fMRI study. Obesity, 16(5), 945–950. https://doi.org/10.1038/oby.2008.33
Gavornik, J. P., Shuler, M. G., Loewenstein, Y., Bear, M. F., & Shouval, H. Z. (2009). Learning reward timing in cortex
through reward dependent expression of synaptic plasticity. Proceedings of the National Academy of Sciences of the United
States of America, 106(16), 6826–6831. https://doi.org/10.1073/pnas.0901835106
Georgiadis, J. R., Kringelbach, M. L., & Pfaus, J. G. (2012). Sex for fun: A synthesis of human and animal neurobiology.
Nature Reviews. Urology, 9(9), 486–498. https://doi.org/10.1038/nrurol.2012.151
Gold, B. P., Mas-Herrero, E., Zeighami, Y., Benovoy, M., Dagher, Al., & Zatorre, R. J. (2019). Musical reward predic-
tion errors engage the nucleus accumbens and motivate learning. Proceedings of the National Academy of Sciences of the
United States of America, 116(8), 3310–3315. https://doi.org/10.1073/pnas.1809855116
Gold, J. M., Waltz, J. A., Prentice, K. J., Morris, S. E.,  & Heerey, E. A. (2008). Reward processing in schizophre-
nia: A deficit in the representation of value. Schizophrenia Bulletin, 34(5), 835–847. https://doi.org/10.1093/schbul/
sbn068
Goller, J., Mitrovic, A., & Leder, H. (2019). Effects of liking on visual attention in faces and paintings. Acta Psychologica,
197, 115–123. https://doi.org/10.1016/j.actpsy.2019.05.008
Goltstein, P. M., Meijer, G. T., & Pennartz, C. M. (2018). Conditioning sharpens the spatial representation of rewarded
stimuli in mouse primary visual cortex. eLife, 7, e37683. https://doi.org/10.7554/eLife.37683
Grabenhorst, F., & Rolls, E. T. (2008). Selective attention to affective value alters how the brain processes taste stimuli.
European Journal of Neuroscience, 27(3), 723–729. https://doi.org/10.1111/j.1460-9568.2008.06033.x
Grabenhorst, F., & Rolls, E. T. (2010). Attentional modulation of affective versus sensory processing: Functional con-
nectivity and a top-down biased activation theory of selective attention. Journal of Neurophysiology, 104(3), 1649–1660.
https://doi.org/10.1152/jn.00352.2010
Gridi-Papp, M., Rand, A. S., & Ryan, M. J. (2006). Animal communication: Complex call production in the túngara
frog. Nature, 441(7089), 38. https://doi.org/10.1038/441038a
Haber, S. N., & Knutson, B. (2010). The reward circuit: Linking primate anatomy and human imaging. Neuropsychophar-
macology, 35(1), 4–26. https://doi.org/10.1038/npp.2009.129
Han, W., Tellez, L. A., Perkins, M. H., Shammah-Lagnado, S. J., de Lartigue, G., & de Araujo, I. E. (2018). A neural
circuit for gut-induced reward. Cell, 175, 665–678. https://doi.org/10.1016/j.cell.2018.08.049
Hansen, N. C., & Pearce, M. T. (2014). Predictive uncertainty in auditory sequence processing. Frontiers in Psychology, 5,
1052. https://doi.org/10.3389/fpsyg.2014.01052
Hart, B. L., & Hart, L. A. (2018). How mammals stay healthy in nature: The evolution of behaviours to avoid parasites
and pathogens. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 373(1751), 20170205.
https://doi.org/10.1098/rstb.2017.0205
Hayes, J. E. (2020). Influence of sensation and liking on eating and drinking. In H. L. Meiselman (Ed.), Handbook of eating
and drinking (pp. 1–25). Springer.
Henke, K. (2010). A model for memory systems based on processing modes rather than consciousness. Nature Reviews.
Neuroscience, 11(7), 523–532. https://doi.org/10.1038/nrn2850
Henrich, J., Boyd, R., Bowles, S., Camerer, C., Fehr, E., Gintis, H., & McElreath, R. (2001). In search of Homo eco-
nomicus: Behavioral experiments in 15 small-scale societies. American Economic Review, 91(2), 73–78. https://doi.
org/10.1257/aer.91.2.73
Herry, C., Bach, D. R., Esposito, F., Di Salle, F., Perrig, W. J., Scheffler, K., Lüthi, A., & Seifritz, E. (2007). Processing
of temporal unpredictability in human and animal amygdala. Journal of Neuroscience, 27(22), 5958–5966. https://doi.
org/10.1523/JNEUROSCI.5218-06.2007
Herz, R. (2012). That’s disgusting. Unravelling the mysteries of repulsion. W.W. Norton & Company.
Hikosaka, O., Nakamura, K.,  & Nakahara, H. (2006). Basal ganglia orient eyes to reward. Journal of Neurophysiology,
95(2), 567–584. https://doi.org/10.1152/jn.00458.2005

56
Sensory liking

Hoke, K. L., Burmeister, S. S., Fernald, R. D., Rand, A. S., Ryan, M. J., & Wilczynski, W. (2004). Functional map-
ping of the auditory midbrain during mate call reception. Journal of Neuroscience, 24(50), 11264–11272. https://doi.
org/10.1523/JNEUROSCI.2079-04.2004
Holcomb, H. H., & Rowland, L. M. (2007). How schizophrenia and depression disrupt reward circuitry. Current Treat-
ment Options in Neurology, 9(5), 357–362
Hulme, O. J., Skov, M., Chadwick, M. J., Siebner, H. R., & Ramsøy, T. Z. (2014). Sparse encoding of automatic visual
association in hippocampal networks. Neuroimage, 102(2), 458–464. https://doi.org/10.1016/j.neuroimage.2014.
07.020
Ishizu, T., & Zeki, S. (2011). Toward a brain-based theory of beauty. PLOS ONE, 6(7), e21852. https://doi.org/10.1371/
journal.pone.0021852
Ishizu, T., & Zeki, S. (2013). The brain’s specialized systems for aesthetic and perceptual judgment. European Journal of
Neuroscience, 37(9), 1413–1420. https://doi.org/10.1111/ejn.12135
Jacobsen, T., Schubotz, R. I., Höfel, L., & von Cramon, D. Y. (2006). Brain correlates of aesthetic judgment of beauty.
NeuroImage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010
Johnston, V. S., Hagel, R., Franklin, M., Fink, B.,  & Grammer, K. (2001). Male facial attractiveness: Evidence for
hormone-mediated adaptive design. Evolution and Human Behavior, 22(4), 251–267. https://doi.org/10.1016/
S1090-5138(01)00066-6
Jones, B. C., Hahn, A. C., Fisher, C. I., Wang, H., Kandrik, M., Han, C., Fasolt, V., Morrison, D., Lee, A. J., Holzleit-
ner, I. J., O’Shea, K. J., Roberts, S. C., Little, A. C., & DeBruine, L. M., . . . , DeBruine. (2018). No compelling
evidence that preferences for facial masculinity track changes in women’s hormonal status. Psychological Science, 29(6),
996–1005. https://doi.org/10.1177/0956797618760197
Katz, D. B., & Sadacca, B. F. (2011). Taste. In J. A. Gottfried (Ed.), Neurobiology of sensation and reward. CRC Press.
Keller, J., & Young, C. B. (2013). Trait anhedonia is associated with reduced reactivity and connectivity of mesolim-
bic and paralimbic reward pathways. Journal of Psychiatric Research, 47(10), 1319–1328. https://doi.org/10.1016/j.
jpsychires.2013.05.015
Kim, H., Adolphs, R., O’Doherty, J. P., & Shimojo, S. (2007). Temporal isolation of neural processes underlying face
preference decisions. Proceedings of the National Academy of Sciences of the United States of America, 104(46), 18253–
18258. https://doi.org/10.1073/pnas.0703101104
Kirk, U., Skov, M., Christensen, M. S., & Nygaard, M. S. (2009a). Brain correlates of aesthetic experience: A parametric
fMRI study. Brain and Cognition, 69, 306–315
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009b). Modulation of aesthetic value by semantic con-
text: An fMRI study. Neuroimage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009
Klebl, C., Greenaway, K. H., Rhee, J. J. S., & Bastian, B. (2020). Ugliness judgments alert us to cues of pathogen pres-
ence. Social Psychological and Personality Science, 12(5), 617–628. https://doi.org/10.1177/1948550620931655
Knutson, B., Adams, C. M., Fong, G. W., & Hommer, D. (2001). Anticipation of increasing monetary reward selec-
tively recruits nucleus accumbens. Journal of Neuroscience, 21(16), RC159–RC159. https://doi.org/10.1523/
JNEUROSCI.21-16-j0002.2001
Knutson, B., & Genevsky, A. (2018). Neuroforecasting aggregate choice. Current Directions in Psychological Science, 27(2),
110–115. https://doi.org/10.1177/0963721417737877
Knutson, B., & Greer, S. M. (2008). Anticipatory affect: Neural correlates and consequences for choice. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 363(1511), 3771–3786. https://doi.org/10.1098/
rstb.2008.0155
Knutson, B., Rick, S., Wimmer, G. E., Prelec, D., & Loewenstein, G. (2007). Neural predictors of purchases. Neuron,
53(1), 147–156. https://doi.org/10.1016/j.neuron.2006.11.010
Knutson, B., Taylor, J., Kaufman, M., Peterson, R., & Glover, G. (2005). Distributed neural representation of expected
value. Journal of Neuroscience, 25(19), 4806–4812. https://doi.org/10.1523/JNEUROSCI.0642-05.2005
Koelsch, S., Fritz, T., & Schlaug, G. (2008). Amygdala activity can be modulated by unexpected chord functions during
music listening. NeuroReport, 19(18), 1815–1819. https://doi.org/10.1097/WNR.0b013e32831a8722
Kolling, N., Behrens, T. E. J., Wittmann, M. K., & Rushworth, M. F. S. (2016). Multiple signals in anterior cingulate
cortex. Current Opinion in Neurobiology, 37, 36–43. https://doi.org/10.1016/j.conb.2015.12.007
Kozbelt, A. (2020). The influence of expertise on aesthetics. In M. Nadal & O. Vartanian (Eds.), The Oxford handbook of
empirical aesthetics. Oxford University Press.
Kringelbach, M. L. (2005). The human orbitofrontal cortex: Linking reward to hedonic experience. Nature Reviews.
Neuroscience, 6(9), 691–702. https://doi.org/10.1038/nrn1747
Kringelbach, M. L., & Berridge, K. C. (2010). Pleasures of the brain. Oxford University Press.

57
Martin Skov

Kringelbach, M. L., O’Doherty, J., Rolls, E. T., & Andrews, C. (2003). Activation of the human orbitofrontal cortex to
a liquid food stimulus is correlated with its subjective pleasantness. Cerebral Cortex, 13(10), 1064–1071. https://doi.
org/10.1093/cercor/13.10.1064
Kringelbach, M. L.,  & Rolls, E. T. (2004). The functional neuroanatomy of the human orbitofrontal cortex: Evi-
dence from neuroimaging and neuropsychology. Progress in Neurobiology, 72(5), 341–372. https://doi.org/10.1016/j.
pneurobio.2004.03.006
Krishna, A. (2012). An integrative review of sensory marketing: Engaging the senses to affect perception, judgment and
behavior. Journal of Consumer Psychology, 22(3), 332–351. https://doi.org/10.1016/j.jcps.2011.08.003
Kühn, S., & Gallinat, J. (2012). The neural correlates of subjective pleasantness. Neuroimage, 61(1), 289–294. https://doi.
org/10.1016/j.neuroimage.2012.02.065
Lang, P. J., & Bradley, M. M. (2010). Emotion and the motivational brain. Biological Psychology, 84(3), 437–450. https://
doi.org/10.1016/j.biopsycho.2009.10.007
Lea, A. M., & Ryan, M. J. (2015). Irrationality in mate choice revealed by túngara frogs. Science, 349(6251), 964–966.
https://doi.org/10.1126/science.aab2012
Lebreton, M., Jorge, S., Michel, V., Thirion, B.,  & Pessiglione, M. (2009). An automatic valuation system in the
human brain: Evidence from functional neuroimaging. Neuron, 64(3), 431–439. https://doi.org/10.1016/j.
neuron.2009.09.040
Leder, H., Mitrovic, A., & Goller, J. (2016). How beauty determines gaze! Facial attractiveness and gaze duration in
images of real world scenes. i-Perception, 7(4), 2041669516664355. https://doi.org/10.1177/2041669516664355
LeDoux, J. E. (2012). Rethinking the emotional brain. Neuron, 73(4), 653–676. https://doi.org/10.1016/j.
neuron.2012.02.004
LeDoux, J. E. (2021). As soon as there was life, there was danger: The deep history of survival behaviours and the shal-
lower history of consciousness. Philosophical Transactions of the Royal Society of London Series B, 377, 20210292
Li, J., Oksama, L., & Hyonä, J. (2016). How facial attractiveness affects sustained attention. Scandinavian Journal of Psychol-
ogy, 57(5), 383–392. https://doi.org/10.1111/sjop.12304
Liu, C. H., & Chen, W. (2012). Beauty is better pursued: Effects of attractiveness in multiple-face tracking. The Quarterly
Journal of Experimental Psychology, 65(3), 553–564. https://doi.org/10.1080/17470218.2011.624186
Lopez-Persem, A., Bastin, J., Petton, M., Abitbol, R., Lehongre, K., Adam, C., Navarro, V., Rheims, S., Kahane, P.,
Domenech, P., & Pessiglione, M. (2020). Four core properties of the human brain valuation system demonstrated in
intracranial signals. Nature Neuroscience, 23(5), 664–675. https://doi.org/10.1038/s41593-020-0615-9
Lopez-Persem, A., Rigoux, L., Bourgeois-Gironde, S., Daunizeau, J., & Pessiglione, M. (2017). Choose, rate or squeeze:
Comparison of economic value functions elicited by different behavioral tasks. PLOS Computational Biology, 13(11),
e1005848. https://doi.org/10.1371/journal.pcbi.1005848
Loui, P., Patterson, S., Sachs, M. E., Leung, Y., Zeng, T., & Przysinda, E. (2017). White matter correlates of musical
anhedonia: Implications for evolution of music. Frontiers in Psychology, 8, 1664. https://doi.org/10.3389/fpsyg.2017.
01664
Lynch, K. S., Crews, D., Ryan, M. J.,  & Wilczynski, W. (2006). Hormonal state influences aspects of female mate
choice in the túngara frog (Physalaemus pustulosus). Hormones and Behavior, 49(4), 450–457. https://doi.org/10.1016/j.
yhbeh.2005.10.001
Lynch, K. S., Stanely Rand, A. S., Ryan, M. J., & Wilczynski, W. (2005). Plasticity in female mate choice associated with
changing reproductive states. Animal Behaviour, 69(3), 689–699. https://doi.org/10.1016/j.anbehav.2004.05.016
Makin, A. (2017). The gap between aesthetic science and aesthetic experience. Journal of Consciousness Studies, 24,
184–213.
Mallik, A., Chanda, M. L., & Levitin, D. J. (2017). Anhedonia to music and mu-opioids: Evidence from the administra-
tion of naltrexone. Scientific Reports, 7, 41952. https://doi.org/10.1038/srep41952
Marcinkowska, U. M., Ellison, P. T., Galbarczyk, A., Milkowska, K., Pawlowski, B., Thune, I., & Jasienska, G. (2016).
Lack of support for relation between woman’s masculinity preference, estradiol level and mating context. Hormones
and Behavior, 78, 1–7. https://doi.org/10.1016/j.yhbeh.2015.10.012
Martínez-García, F., & Lanuza, E. (2018). Evolution of vertebrate survival circuits. Current Opinion in Behavioral Sciences,
24, 113–123. https://doi.org/10.1016/j.cobeha.2018.06.012
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J.,  & Marco-Pallarés, J. (2016). Neural
correlates of specific musical anhedonia. Proceedings of the National Academy of Sciences of the United States of America,
113(46), E7337–E7345. https://doi.org/10.1073/pnas.1611211113
Masataka, N., & Shibasaki, M. (2012). Premenstrual enhancement of snake detection in visual search in healthy women.
Scientific Reports, 2, 307. https://doi.org/10.1038/srep00307

58
Sensory liking

Mas-Herrero, E., Maini, L., Sescousse, G., & Zatorre, R. J. (2021). Common and distinct neural correlates of music
and food-induced pleasure: A coordinate-based meta-analysis of neuroimaging studies. Neuroscience and Biobehavioral
Reviews, 123, 61–71. https://doi.org/10.1016/j.neubiorev.2020.12.008
Mas-Herrero, E., Zatorre, R. J., Rodriguez-Fornells, A., & Marco-Pallarés, J. (2014). Dissociation between musical and
monetary reward responses in specific musical anhedonia. Current Biology, 24(6), 699–704. https://doi.org/10.1016/j.
cub.2014.01.068
Matthews, B. R., Chang, C. C., De May, M., Engstrom, J.,  & Miller, B. L. (2009). Pleasurable emotional response
to music: A  case of neurodegenerative generalized auditory agnosia. Neurocase, 15(3), 248–259. https://doi.
org/10.1080/13554790802632934
McClure, S. M., Li, J., Tomlin, D., Cypert, K. S., Montague, L. M., & Montague, P. R. (2004). Neural correlates of behav-
ioral preference for culturally familiar drinks. Neuron, 44(2), 379–387. https://doi.org/10.1016/j.neuron.2004.09.019
Menninghaus, W., Wagner, V., Wassiliwizky, E., Schindler, I., Hanich, J., Jacobsen, T., & Koelsch, S. (2019). What are
aesthetic emotions? Psychological Review, 126(2), 171–195. https://doi.org/10.1037/rev0000135
Miyashita, Y. (2004). Cognitive memory: Cellular and network machineries and their top-down control. Science,
306(5695), 435–440. https://doi.org/10.1126/science.1101864
Mocaiber, I., Pereira, M. G., Erthal, F. S., Machado-Pinheiro, W., David, I. A., Cagy, M., Volchan, E., & de Oliveira, L.
(2010). Fact or fiction? An event-related potential study of implicit emotion regulation. Neuroscience Letters, 476(2),
84–88. https://doi.org/10.1016/j.neulet.2010.04.008
Morville, T., Madsen, K. H., Siebner, H. R., & Hulme, O. J. (2021). Reward signalling in brainstem nuclei under fluc-
tuating blood glucose. PLOS ONE, 16(4), e0243899. https://doi.org/10.1371/journal.pone.0243899
Murray, S. O., Schrater, P., & Kersten, D. (2004). Perceptual grouping and the interactions between visual cortical areas.
Neural Networks, 17(5–6), 695–705. https://doi.org/10.1016/j.neunet.2004.03.010
Nadal, M., & Ureña, E. (2022). One hundred years of empirical aesthetics: Fechner to Berlyne (1876–1976). In M.
Nadal & O. Vartanian (Eds.), The Oxford handbook of empirical aesthetics. Oxford University Press.
Nummenmaa, L., & Tuominen, L. (2017). Opioid system and human emotions. British Journal of Pharmacology, 175(14),
2737–2749. https://doi.org/10.1111/bph.13812
O’Connell, L. A., & Hofmann, H. A. (2011). The vertebrate mesolimbic reward system and social behavior network:
A  comparative synthesis. The Journal of Comparative Neurology, 519(18), 3599–3639. https://doi.org/10.1002/
cne.22735
O’Doherty, J. P., Cockburn, J., & Pauli, W. M. (2017). Learning, reward, and decision making. Annual Review of Psychol-
ogy, 68, 73–100. https://doi.org/10.1146/annurev-psych-010416-044216
O’Doherty, J. P., Deichmann, R., Critchley, H. D., & Dolan, R. J. (2002). Neural responses during anticipation of a
primary taste reward. Neuron, 33(5), 815–826. https://doi.org/10.1016/s0896-6273(02)00603-7
Okamoto, M.,  & Dan, I. (2013). Extrinsic information influences taste and flavor perception: A  review from psy-
chological and neuroimaging perspectives. Seminars in Cell and Developmental Biology, 24(3), 247–255. https://doi.
org/10.1016/j.semcdb.2012.11.001
Oliva, A., & Torralba, A. (2007). The role of context in object recognition. Trends in Cognitive Sciences, 11(12), 520–527.
https://doi.org/10.1016/j.tics.2007.09.009
Padoa-Schioppa, C.,  & Assad, J. A. (2006). Neurons in orbitofrontal cortex encode economic value. Nature, 441,
223–226.
Palumbo, L., Rampone, G., Bertamini, M., Sinico, M., Clarke, E., & Vartanian, O. (2020). Visual preference for abstract
curvature and for interior spaces: Beyond undergraduate student samples. Psychology of Aesthetics, Creativity, and the
Arts. https://doi.org/10.1037/aca0000359
Panksepp, J. (1998). Affective neuroscience. Oxford University Press.
Pearce, M. T. (2018). Statistical learning and probabilistic prediction in music cognition: Mechanisms of stylistic encul-
turation. Annals of the New York Academy of Sciences, 1423, 378–395. https://doi.org/10.1111/nyas.13654
Pearce, M. T., Ruiz, M. H., Kapasi, S., Wiggins, G. A.,  & Bhattacharya, J. (2010). Unsupervised statistical learning
underpins computational, behavioural, and neural manifestations of musical expectation. Neuroimage, 50(1), 302–313.
https://doi.org/10.1016/j.neuroimage.2009.12.019
Peciña, S. (2008). Opioid reward “liking” and “wanting” in the nucleus accumbens. Physiology and Behavior, 94(5),
675–680. https://doi.org/10.1016/j.physbeh.2008.04.006
Peng, Y., Gillis-Smith, S., Jin, H., Tränkner, D., Ryba, N. J., & Zuker, C. S. (2015). Sweet and bitter taste in the brain
of awake behaving animals. Nature, 527(7579), 512–515. https://doi.org/10.1038/nature15763
Penton-Voak, I. S., & Perrett, D. I. (2000). Female preference for male faces changes cyclically. Evolution and Human
Behavior, 21(1), 39–48. https://doi.org/10.1016/S1090-5138(99)00033-1

59
Martin Skov

Penton-Voak, I. S., Perrett, D. I., Castles, D. L., Kobayashi, T., Burt, D. M., Murray, L. K., & Minamisawa, R. (1999).
Menstrual cycle alters face preference. Nature, 399(6738), 741–742. https://doi.org/10.1038/21557
Pessiglione, M., & Lebreton, M. (2015). From the reward circuit to the valuation system: How the brain motivates behav-
ior. In G. H. E. Gendolla et al. (Eds.), Handbook of biobehavioral approaches to self-regulation (pp. 157–173). Springer.
Pessoa, L., Medina, L., & Desfilis, E. (2021). Refocusing neuroscience: Moving away from mental categories and towards
complex behaviours. Philosophical Transactions of the Royal Society of London Series B, 377, 20200534.
Petrovic, P., Pleger, B., Seymour, B., Klöppel, S., de Martino, B., Critchley, H., & Dolan, R. J. (2008). Blocking central
opiate function modulates hedonic impact and anterior cingulate response to rewards and losses. Journal of Neuroscience,
28(42), 10509–10516. https://doi.org/10.1523/JNEUROSCI.2807-08.2008
Pilarczyk, J., Schwertnet, E., Woloszyn, K., & Kuniecki, M. (2019). Phase of menstrual cycle affects engagement of
attention with emotional images. Psychoendocrinology, 104, 25–32.
Piqueras-Fiszman, B., & Spence, C. (2018). Sensory expectations based on product-extrinsic food cues: An interdisci-
plinary review of the empirical evidence and theoretical accounts. Food Quality and Preference, 40, 165–179. https://
doi.org/10.1016/j.foodqual.2014.09.013
Plassmann, H., O’Doherty, J., Shiv, B., & Rangel, A. (2008). Marketing actions can modulate neural representations of
experienced pleasantness. Proceedings of the National Academy of Sciences of the United States of America, 105(3), 1050–
1054. https://doi.org/10.1073/pnas.0706929105
Ramsøy, T. Z., Friis-Olivarius, M., Jacobsen, C., Jensen, S. B., & Skov, M. (2012). Effects of perceptual uncertainty on
arousal and preference across different visual domains. Journal of Neuroscience, Psychology, and Economics, 5(4), 212–226.
https://doi.org/10.1037/a0030198
Rangel, A., Camerer, C., & Montague, P. R. (2008). A framework for studying the Neurobiology of value-based deci-
sion making. Nature Reviews. Neuroscience, 9(7), 545–556. https://doi.org/10.1038/nrn2357
Roney, J. R. (2018). Functional roles of gonadal hormones in human pair bonding and sexuality. In O. C. Schultheiss &
P. H. Mehta (Eds.), Routledge international handbook of social neuroendocrinology (pp. 239–255). Routledge.
Rosenthal, G. G. (2017). Mate choice. The evolution of sexual decision making from microbes to humans. Princeton University
Press.
Rupp, H. A., James, T. W., Ketterson, E. D., Sengelaub, D. R., Janssen, E., & Heiman, J. R. (2009a). Neural activation
in women in response to masculinized male faces: Mediation by hormones and psychosexual factors. Evolution and
Human Behavior, 30(1), 1–10. https://doi.org/10.1016/j.evolhumbehav.2008.08.006
Rupp, H. A., James, T. W., Ketterson, E. D., Sengelaub, D. R., Janssen, E., & Heiman, J. R. (2009b). Neural activation
in the orbitofrontal cortex in response to male faces increases during the follicular phase. Hormones and Behavior, 56(1),
66–72. https://doi.org/10.1016/j.yhbeh.2009.03.005
Ryan, M. J. (1980). Female mate choice in a neotropical frog. Science, 209(4455), 523–525. https://doi.org/10.1126/
science.209.4455.523
Ryan, M. J., & Cummings, M. E. (2013). Perceptual biases and mate choice. Annual Review of Ecology, Evolution, and
Systematics, 44(1), 437–459. https://doi.org/10.1146/annurev-ecolsys-110512-135901
Ryan, M. J., & Jordan, L. A. (2017). Courtship and mate choice. In J. Call et al. (Eds.), APA handbook of comparative
psychology (Vol. I, pp. 765–786). American Psychological Association.
Ryan, M. J., Page, R. A., Hunter, K. L., & Taylor, R. C. (2019). ‘Crazy love’: Nonlinearity and irrationality in mate
choice. Animal Behaviour, 147, 189–198. https://doi.org/10.1016/j.anbehav.2018.04.004
Sachs, M. E., Ellis, R. J., Schlaug, G., & Loui, P. (2016). Brain connectivity reflects human aesthetic responses to music.
Social Cognitive and Affective Neuroscience, 11(6), 884–891. https://doi.org/10.1093/scan/nsw009
Sakaki, M., & Mather, M. (2012). How reward and emotional stimuli induce different reactions across the menstrual
cycle. Social and Personality Psychology Compass, 6(1), 1–17. https://doi.org/10.1111/j.1751-9004.2011.00415.x
Sakurai, T., Mitsuno, H., Haupt, S. S., Uchino, K., Yokohari, F., Nishioka, T., Kobayashi, I., Sezutsu, H., Tamura, T., &
Kanzaki, R. (2011). A single sex pheromone receptor determines chemical response specificity of sexual behavior in
the silkmoth Bombyx mori. PLOS Genetics, 7(6), 1–10.
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Salimpoor, V. N., Benovoy, M., Longo, G., Cooperstock, J. R., & Zatorre, R. J. (2009). The rewarding aspects of music
listening are related to degree of emotional arousal. PLOS ONE, 4(10), e7487. https://doi.org/10.1371/journal.
pone.0007487
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059

60
Sensory liking

Santos, L. R.,  & Chen, M. K. (2009). The evolution of rational and irrational economic behaviour: Evidence and
insight from a nonhuman primate species. In P. W. Glimcher et al. (Eds.), Neuroeconomics. Decision making and the brain
(pp. 81–93). Academic Press.
Sarasso, P., Neppi-Modona, M., Sacco, K.,  & Ronga, I. (2020). ‘Stopping for knowledge’: The sense of beauty in
the perception-action cycle. Neuroscience and Biobehavioral Reviews, 118, 723–738. https://doi.org/10.1016/j.
neubiorev.2020.09.004
Satoh, M., Kato, N., Tabei, K. I., Nakano, C., Abe, M., Fujita, R., Kida, H., Tomimoto, H., & Kondo, K. (2016). A case
of musical anhedonia due to right putaminal hemorrhage: A disconnection syndrome between the auditory cortex
and insula. Neurocase, 22(6), 518–525. https://doi.org/10.1080/13554794.2016.1264609
Satoh, M., Nakase, T., Nagata, K., & Tomimoto, H. (2011). Musical anhedonia: Selective loss of emotional experience
in listening to music. Neurocase, 17(5), 410–417. https://doi.org/10.1080/13554794.2010.532139
Schlaepfer, T. E., Cohen, M. X., Frick, C., Kosel, M., Brodesser, D., Axmacher, N., Joe, A. Y., Kreft, M., Lenartz, D., &
Sturm, V. (2008). Deep brain stimulation to reward circuitry alleviates anhedonia in refractory major depression.
Neuropsychopharmacology, 33(2), 368–377. https://doi.org/10.1038/sj.npp.1301408
Schultz, W. (2015). Neuronal reward and decision signals: From theories to data. Physiological Reviews, 95(3), 853–951.
https://doi.org/10.1152/physrev.00023.2014
Schultz, W., Apicella, P., Scarnati, E., & Ljungberg, T. (1992). Neuronal activity in monkey ventral striatum related to
the expectation of reward. Journal of Neuroscience, 12(12), 4595–4610. https://doi.org/10.1523/JNEUROSCI.12-
12-04595.1992
Schultz, W., Dayan, P.,  & Montague, P. R. (1997). A  neural substrate of prediction and reward. Science, 275(5306),
1593–1599. https://doi.org/10.1126/science.275.5306.1593
Serences, J. T. (2008). Value-based modulations in human visual cortex. Neuron, 60(6), 1169–1181. https://doi.
org/10.1016/j.neuron.2008.10.051
Sescousse, G., Caldú, X., Segura, B., & Dreher, J. C. (2013). Processing of primary and secondary rewards: A quantita-
tive meta-analysis and review of human functional neuroimaging studies. Neuroscience and Biobehavioral Reviews, 37(4),
681–696. https://doi.org/10.1016/j.neubiorev.2013.02.002
Seymour, B., Daw, N., Dayan, P., Singer, T., & Dolan, R. (2007). Differential encoding of losses and gains in the human
striatum. Journal of Neuroscience, 27(18), 4826–4831. https://doi.org/10.1523/JNEUROSCI.0400-07.2007
Seymour, B., Singer, T.,  & Dolan, R. (2007). The neurobiology of punishment. Nature Reviews Neuroscience, 8(4),
300–311. https://doi.org/10.1038/nrn2119
Shafir, S., Waite, T. A., & Smith, B. H. (2002). Context-dependent violations of rational choice in honeybees (Apis mel-
lifera) and gray jays (Perisoreus canadensis). Behavioral Ecology and Sociobiology, 51(2), 180–187. https://doi.org/10.1007/
s00265-001-0420-8
Shany, O., Singer, N., Gold, B. P., Jacoby, N., Tarrasch, R., Hendler, T., & Granot, R. (2019). Surprise-related activation
in the nucleus accumbens interacts with music-induced pleasantness. Social Cognitive and Affective Neuroscience, 14(4),
459–470. https://doi.org/10.1093/scan/nsz019
Sharvit, G., Vuilleumier, P., Delplanque, S.,  & Corradi-Dell’Acqua, C. (2015). Cross-modal and modality-specific
expectancy effects between pain and disgust. Scientific Reports, 5, 17487. https://doi.org/10.1038/srep17487
Sherman, A., Grabowecky, M.,  & Suzuki, S. (2015). In the working memory of the beholder: Art appreciation is
enhanced when visual complexity is compatible with working memory. Journal of Experimental Psychology. Human
Perception and Performance, 41(4), 898–903. https://doi.org/10.1037/a0039314
Shimojo, S., Simion, C., Shimojo, E., & Scheier, C. (2003). Gaze bias both reflects and influences preference. Nature
Neuroscience, 6, 1317–1322.
Shuler, M. G., & Bear, M. F. (2006). Reward timing in the primary visual cortex. Science, 311(5767), 1606–1609. https://
doi.org/10.1126/science.1123513
Siep, N., Roefs, A., Roebroeck, A., Havermans, R., Bonte, M. L., & Jansen, A. (2009). Hunger is the best spice: An
fMRI study of the effects of attention, hunger and calorie content on food reward processing in the amygdala and
orbitofrontal cortex. Behavioural Brain Research, 198(1), 149–158. https://doi.org/10.1016/j.bbr.2008.10.035
Skov, M. (2020). The neurobiology of sensory valuation. In M. Nadal & O. Vartanian (Eds.), The Oxford handbook
of empirical aesthetics (pp.  1–40). https://doi.org/10.1093/oxfordhb/9780198824350.013.7. Oxford University
Press.
Skov, M., & Nadal, M. (2020). There are no aesthetic emotions: Comment on Menninghaus et al. (2019). Psychological
Review, 127(4), 640–649. https://doi.org/10.1037/rev0000187
Skov, M., & Nadal, M. (2021). The nature of beauty: Behavior, cognition, and neurobiology. Annals of the New York
Academy of Sciences, 1488(1), 44–55. https://doi.org/10.1111/nyas.14524
Skov, M., & Nadal, M. (2022). Is aesthetic pleasure adaptive? A review of the literature. Under review.

61
Martin Skov

Small, D. M., Zatorre, R. J., Dagher, A., Evans, A. C., & Jones-Gotman, M. (2001). Changes in brain activity related to
eating chocolate: From pleasure to aversion. Brain, 124(9), 1720–1733. https://doi.org/10.1093/brain/124.9.1720
Smith, K. S., & Berridge, K. C. (2005). The ventral pallidum and hedonic reward: Neurochemical maps of sucrose “liking”
and food intake. Journal of Neuroscience, 25(38), 8637–8649. https://doi.org/10.1523/JNEUROSCI.1902-05.2005
Smith, K. S.,  & Berridge, K. C. (2007). Opioid limbic circuit for reward: Interaction between hedonic hotspots of
nucleus accumbens and ventral pallidum. Journal of Neuroscience, 27(7), 1594–1605. https://doi.org/10.1523/
JNEUROSCI.4205-06.2007
Smith, K. S., Mahler, S. V., Pecina, S., & Berridge, K. C. (2010). Hedonic hotspots: Generating sensory pleasure in the
brain. In M. L. Kringelbach & K. C. Berridge (Eds.), Pleasures of the brain (pp. 27–49). Oxford University Press.
Steiner, J. E., Glaser, D., Hawilo, M. E., & Berridge, K. C. (2001). Comparative expression of hedonic impact: Affective
reactions to taste by human infants and other primates. Neuroscience and Biobehavioral Reviews, 25(1), 53–74. https://
doi.org/10.1016/s0149-7634(00)00051-8
Sternson, S. M., & Eiselt, A. K. (2017). Three pillars for the neural control of appetite. Annual Review of Physiology, 79,
401–423. https://doi.org/10.1146/annurev-physiol-021115-104948
Summerfield, C.,  & de Lange, F. P. (2014). Expectation in perceptual decision making: Neural and computational
mechanisms. Nature Reviews. Neuroscience, 15(11), 745–756. https://doi.org/10.1038/nrn3838
Swanson, L. W. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Research, 886(1–2), 113–164.
https://doi.org/10.1016/s0006-8993(00)02905-x
Thomas, J. M., Higgs, S., Dourish, C. T., Hansen, P. C., Harmer, C. J., & McCabe, C. (2015). Satiation attenuates BOLD
activity in brain regions involved in reward and increases activity in dorsolateral prefrontal cortex: An fMRI study in
healthy volunteers. American Journal of Clinical Nutrition, 101(4), 697–704. https://doi.org/10.3945/ajcn.114.097543
Tom, S. M., Fox, C. R., Trepel, C., & Poldrack, R. A. (2007). The neural basis of loss aversion in decision-making under
risk. Science, 315(5811), 515–518. https://doi.org/10.1126/science.1134239
Tonelli, G. (2003). Taste in the history of aesthetics from the Renaissance to 1770. In Dictionary of the History of Ideas (Vol.
4). Charles Scribner’s Sons.
Valuch, C., Pflüger, L. S., Wallner, B., Laeng, B., & Ansorge, U. (2015). Using eye tracking to test for individual differ-
ences in attention to attractive faces. Frontiers in Psychology, 6, 42. https://doi.org/10.3389/fpsyg.2015.00042
Van Dongen, N. N. N., Van Strien, J. W., & Dijkstra, K. (2016). Implicit emotion regulation in the context of viewing
artworks: ERP evidence in response to pleasant and unpleasant pictures. Brain and Cognition, 107, 48–54. https://doi.
org/10.1016/j.bandc.2016.06.003
van Duijn, M., Keijzer, F., & Franken, D. (2006). Principles of minimal cognition: Casting cognition as sensorimotor
coordination. Adaptive Behavior, 14(2), 157–170. https://doi.org/10.1177/105971230601400207
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Rostrup, N., Skov, M., Corradi,
G., & Nadal, M. (2019). Preference for curvilinear contour in interior architectural spaces: Evidence from experts
and nonexperts. Psychology of Aesthetics, Creativity, and the Arts, 13(1), 110–116. https://doi.org/10.1037/aca0000150
Vessel, E. A. (2020). Neuroaesthetics. In Reference module in neuroscience and biobehavioral psychology. Elsevier. https://doi.
org/10.1016/B978-0-12-809324-5.24104-7
Wada-Katsumata, A., Silverman, J., & Schal, C. (2013). Changes in taste neurons support the emergence of an adaptive
behavior in cockroaches. Science, 340(6135), 972–975. https://doi.org/10.1126/science.1234854
Wang, L., Gillis-Smith, S., Peng, Y., Zhang, J., Chen, X., Salzman, C. D., Ryba, N. J. P., & Zuker, C. S. (2018). The
coding of valence and identity in the mammalian taste system. Nature, 558(7708), 127–131. https://doi.org/10.1038/
s41586-018-0165-4
Yoest, K. E., Quigley, J. A., & Becker, J. B. (2018). Rapid effects of ovarian hormones in dorsal striatum and nucleus
accumbens. Hormones and Behavior, 104, 119–129. https://doi.org/10.1016/j.yhbeh.2018.04.002
Zajonc, R. B. (1968). Attitudinal effects of mere exposure. Journal of Personality and Social Psychology, Monongraph sup-
plement No. 2, 9(2, Pt. 2), 1–27. https://doi.org/10.1037/h0025848
Zajonc, R. B. (1980). Feeling and thinking: Preferences need no inferences. American Psychologist, 35(2), 151–175.
https://doi.org/10.1037/0003-066X.35.2.151
Zheng, H., & Berthoud, H. R. (2007). Eating for pleasure or calories. Current Opinion in Pharmacology, 7(6), 607–612.
https://doi.org/10.1016/j.coph.2007.10.011
Zimmerman, C. A., Leib, D. E., & Knight, Z. A. (2017). Neural circuits underlying thirst and fluid homeostasis. Nature
Reviews. Neuroscience, 18(8), 459–469. https://doi.org/10.1038/nrn.2017.71

62
3
THE NEUROBIOLOGY OF LIKING
Eloise Stark, Kent C. Berridge and Morten L. Kringelbach

In order to sustain human life, it is an evolutionary imperative that stimuli which promote individual and
species survival (such as infant care, food, and sex) are prioritized above less relevant or less important stimuli
(Berridge & Kringelbach, 2008, 2015). It is therefore logical to assume that there must be an affective core
within the brain, which has the role of paying attention to and evaluating stimuli, assessing their valence as
positive or negative, and making them available for conscious appraisal and decision-making (Barrett et al.,
2007; Frijda, 1986; Kringelbach & Phillips, 2014; Russell, 2003).
The affective valence of a stimulus or event is the quality it has of being perceived as “good” versus “bad,”
or hedonically “liked” or “disliked.” The key here is the process of transforming perception into affect, which
is an inferential process that actively generates the valence based upon past experience and present context
(Friston et al., 2006; see also Chapter 2). This percept is actively generated by the brain and is subsequently
translated into hedonic and motivational aspects of valenced reaction (Berridge & Kringelbach, 2015). In this
chapter, we concern ourselves with the hedonic impact of a stimulus, namely “liking.” “Liking” is a crucial
component of both reward and emotion and is fundamental to human survival by encouraging approach
behaviour and consummation of basic rewards, such as food and sex, and higher-order rewards, such as music
or visual arts (Kringelbach & Berridge, 2010).
Affective valence can be experienced consciously as a subjective experience, which is hard to capture
except via subjective self-report but also influences both behaviour and physiology, which can be measured
objectively. For example, we can tell whether newborn infants “like” the taste of different foods by giving
them a taste and watching their behavioural response. Infants demonstrate facial expressions of “liking” in
response to pleasurable tastes, for example, indicated by a relaxed facial expression and rhythmic tongue and
mouth movements (Steiner, 1973). We can contrast this to bitter tastes, which elicit “disgust” reactions indi-
cated by a gaping mouth and turning one’s body away. These affective facial expressions to different tastes are
conserved across several species, including human infants, apes and monkeys, and even rats (Grill & Norgren,
1978; Steiner et al., 2001).
The objective behavioural or physiological reaction may occur with or without conscious subjective
feelings (Anderson & Adolphs, 2014; Berridge, 2018; Damasio & Carvalho, 2013; Frijda & Parrott, 2011;
Winkielman et al., 2005; Winkielman & Gogolushko, 2018). It is therefore important to distinguish between
the two elements of positive affective valence by denoting objective affective reactions, or the core hedonic
impact, with quotation marks (“liking”) and subjective hedonic feelings as liking.

DOI: 10.4324/9781003008675-4 63
Stark, Berridge and Kringelbach

In this chapter, we will review the evidence for the “liking” phase of the Pleasure Cycle and explore how
positive affective valence is constructed by the brain. The research clearly shows that hedonic hotspots and
coldspots mediate “liking” within the brain and dispel the popular myth that this is a process controlled by
dopamine, which is more related to “wanting.” Finally, we build on Aristotle’s distinction between “hedonia”
(pleasure) and “eudaimonia” to suggest that eudaimonia and human flourishing might usefully be construed
as a brain state which could perhaps best be described as “meaningful pleasure.”

The Pleasure Cycle: “wanting,” “liking,” and “satiety”


The Pleasure Cycle can be used to conceptualize the “wanting,” “liking,” and “satiety” from pleasurable
stimuli (see Figure 3.1A). Each element has a dissociable neurobiological basis which contributes in a time-
varying and interactive way. These phases are not necessarily mutually exclusive, as “liking” and “wanting”
can simultaneously overlap. However, each phase may dominate in turn at different temporal moments. The
focus of this chapter is on the “liking” phase, which follows on from “wanting” and involves consummation
of a stimulus and online evaluation of the stimuli’s ongoing reward value. This is the part of reward process-
ing during which pleasure is at its highest, and may even involve a peak, such as the experience of orgasms
(Georgiadis & Kringelbach, 2012; Georgiadis et al., 2012) and “chills” when listening to music (Blood &
Zatorre, 2001; Laeng et al., 2016; Panksepp, 1995).
Before we go into the neurobiological basis of the “liking” phase, it is important to dispel the common
myth that this is coordinated by the brain’s mesolimbic dopamine system. The evidence clearly shows that
dopamine is neither necessary nor sufficient for the generation of “liking” within the brain. In particular, the
facial expressions that denote a “liking” response are not altered by suppression or activation of mesolimbic
dopamine systems. In rodent studies, while intra-accumbens dopamine agonists (e.g., amphetamine micro-
injections) lead to increases in “wanting,” “liking” remains unaffected by such changes (Wyvell & Berridge,
2000). Other studies have used dopamine antagonists, such as neuroleptic drugs, finding that while “wanting”
is decreased, there are no changes in “liking” (Peciña et al., 1997). Large 6-hydroxydopamine (6-OHDA)
lesions, which cause vast destruction of ascending dopamine neurons, lead to profound aphagia (refusal to
swallow), suggesting diminished “wanting,” but have consistently failed to suppress hedonic reaction pat-
terns to sweet tastes indicative of “liking” (Berridge et al., 1989). Likewise in humans, dopamine-receptor
antagonists frequently fail to suppress subjective ratings of pleasure for cigarettes (Brauer et al., 2001), cocaine
(Leyton et al., 2005), or amphetamines (Wachtel et al., 2002).
Together, these findings strongly contradict the popular view that dopamine is the “neurotransmitter of
pleasure,” in the sense that it does not appear to mediate what most people associate most with a pleasurable
response: the subjective hedonic value, or “liking.” The true function of dopamine is beyond the scope of the
current chapter, but indications are that it has an important role in the “wanting” phase of the pleasure cycle
to provide incentive salience to important stimuli (Berridge & Kringelbach, 2008).

Hedonic hotspots and coldspots


The brain system subserving “liking” reactions and generating intense pleasure is a far smaller and more func-
tionally fragile system than the sizeable and robust “wanting” system in the brain. The generation of “liking”
is more restricted than wanting, both neurochemically and anatomically. Neurochemically, opioid stimula-
tion (but not dopamine stimulation) in specific subregions of brain structures can enhance “liking,” whereas
“wanting” is enhanced by both. Anatomically, “liking” is enhanced by opioid-stimulating microinjections
in subregional “hedonic hotspots” within an anatomical structure but not in remaining subregions of the
same anatomic structure—even if stimulation anywhere in the entire structure can enhance “wanting.” The
generation of a “liking” response is also more restricted as a brain circuit, requiring unanimous activation of

64
The neurobiology of liking

Figure 3.1 The Pleasure Cycle. (A) Fundamental (i.e., rewards associated with behaviour necessary for survival of an
individual or the species) and higher-order pleasures are associated with a cyclical time course. Typically,
rewarding moments go through a phase of expectation or wanting for a reward, which sometimes leads to a
phase of consummation or liking of the reward which can have a peak level of pleasure (e.g., encountering a
loved one, a tasty meal, sexual orgasm, drug rush, winning a gambling bet). This can be followed by a satiety
or learning phase, where one learns and updates predictions for the reward, but note that learning obviously
can take place throughout the cycle. These various phases have been identified at many levels of investiga-
tion; for example, recent research on the computational mechanisms underlying prediction, evaluation, and
prediction error are particularly interesting. Note, however, that a very few rewards might possibly lack a
satiety phase (suggested candidates for brief or missing satiety phase have included money, some abstract
rewards, and some drug and brain stimulation rewards that activate dopamine systems rather directly). (B)
The hedonic hotspots (red) and coldspots (blue) in the rat brain, shown on the coronal, sagittal, and hori-
zontal planes and in 3D fronto-lateral perspective view (clockwise from top left). (C) Similarly, the rendering
shows putative human hotspots, extrapolated from neuroimaging literature and the rat causal hotspots. In the
perspective views, the tentative interconnected networks between the different hotspots and coldspots have
been added to give an impression of the topology of the pleasure network.

65
Stark, Berridge and Kringelbach

multiple hotspots simultaneously, whereas “wanting” can be enhanced by a single hotspot. In short, enhance-
ment of pleasure “liking” is restricted and fragile, and brain pleasure systems are relatively recalcitrant to
activation compared to “wanting” systems. Consequently, our brain mechanisms may consign us more often
to states of desire than of pleasure.
The “hedonic hotspots” that mediate “liking” anatomically have been compared to “islands” of brain
tissue contained within the larger “sea” of a full limbic structure, such as the nucleus accumbens or ventral
pallidum (Smith et al., 2010). The size of each hotspot discovered so far is only approximately one cubic
millimetre in volume of the brain of a rat. In the human brain, a hotspot is expected to be about a cubic
centimetre in volume, extrapolated by considering the difference in ratio between the whole-brain size of
rats and humans.
One key methodological advance in exploring these hedonic hotspots was the finding that they can
generate increases in pleasure “liking” reactions to sweetness when stimulated with apposite neurochemical
microinjections (Berridge, 2019). For instance, pleasure-enhancing neurochemicals stimulate opioid recep-
tors, which detect heroin-like neurochemicals in the hotspots. Or they may stimulate endocannabinoid
receptors that detect marijuana-like neurochemicals. Importantly, no “liking” enhancement occurs if the
same drug microinjections are moved outside the precise boundaries of the hedonic hotspots, even if they are
administered within the same brain structure. If this does happen, and the regions surrounding the hotspots
are stimulated instead, the microinjections uniformly stimulate intense “wanting” but without evidence of
enhanced “liking.”
Several of these hedonic hotspots that mediate the “liking” response have been found, dispersed through
the rodent brain from the cortex to the brainstem (see Figure 3.1B). They appear to be functionally inter-
connected like an archipelago of interacting islands. Hotspots are found in the limbic prefrontal cortex,
nucleus accumbens, ventral pallidum (the chief target of nucleus accumbens), and the brainstem pons.
The entire network may need to activate together as a single integrated circuit in order to magnify sen-
sory pleasures, which thus involves collaboration between the various hotspots. For example, activation
of one hotspot by an opioid microinjection automatically recruits activation in other hotspots in different
brain structures (Smith & Berridge, 2007). Pleasure magnification requires coordination among all opioid
hotspots in the nucleus accumbens and ventral pallidum (Smith & Berridge, 2007). If this coordination is
prevented by the suppression of another hotspot with an opioid-opposing drug, a simple hotspot opioid
activation will not boost pleasure. Although opioid stimulation of either hotspot would normally be suf-
ficient to increase “liking,” this will not be the case if the larger circuit is not activated. Even if “liking”
enhancement is disabled, stimulation of “wanting” persists after either hotspot is triggered. Thus, while
partial activation of the limbic circuit is sufficient to induce strong desire, complete activation is required to
generate intense pleasure.
A growing corpus of research in humans has identified that many rewards, as different and diverse as
species-specific pleasures such music and art, compared with fundamental pleasures such as sex and food,
all activate and share a reward network comprised of overlapping brain regions (Cacioppo et  al., 2012;
Georgiadis & Kringelbach, 2012; Kringelbach et al., 2012; Salimpoor et al., 2011; Vartanian & Skov, 2014;
Veldhuizen et al., 2010; Vuust et al., 2021; Vuust & Kringelbach, 2010; this literature is further discussed in
Chapters 7–12) (Figure 3.1C). This has sometimes been called the “common currency” reward network, of
which one important implication is that experiments exploring one type of pleasure, such as food and sex,
should apply to other kinds of pleasure, such as species-specific pleasures like music. This strongly argues
against the view of lower and higher pleasures (Crisp & Kringelbach, 2017). Regions within this “common
currency” network include anatomical regions of the prefrontal cortex, such as the anterior cingulate, insula,
and orbitofrontal cortices, as well as subcortical structures including the ventral pallidum, amygdala, and
nucleus accumbens.

66
The neurobiology of liking

Anhedonia
Anhedonia refers to a absence or severe reduction in the experience of pleasure and is a significant feature
of neuropsychiatric disorders such as mood disorders, eating disorders, and addiction (Rømer Thomsen
et al., 2015). It constitutes a significant burden to such disorders; for example, anhedonia in the context of
depression is a predictor of poor treatment response (Spijker et al., 2001), and in addiction it is a predictor of
relapse (Koob and Le Moal, 2001; Volkow et al., 2002). It is prescient to study anhedonia as a transdiagnostic
symptom, given that it is arguably more likely to be linked to a specific neurobiological underpinning than
heterogeneous psychiatric diagnostic categories, such as schizophrenia (Hyman & Fenton, 2003; Insel et al.,
2010).
The different components of the pleasure cycle require efficient state transitions to move between the dif-
ferent phases, such as from “wanting” to consummation and “liking.” It has been proposed that anhedonia in
affective disorders results from perturbations to the orchestration of such state transitions (Rømer Thomsen
et al., 2015). To illustrate this proposal, one example of a perturbation to the pleasure system comes from
rodent studies where hedonic hotspots are selectively damaged. If you ablate the hedonic hotspot in rodent
posterior ventral pallidum, these animals will no longer display positive hedonic reactions when given sweet-
tasting foods, instead displaying a “disliking” reaction such as mouth gapes, which are usually reserved for
bitter or noxious tastes (Aldridge & Berridge, 2010; Cromwell & Berridge, 1993; Khan et al., 2020). One
case study of a human who underwent a hypoxic episode and subsequent bilateral lesions of the globus pal-
lidus also illustrates this perturbation to the pleasure system, as the patient subsequently reported anhedonia,
including a diminished pleasurable response to alcohol (Miller et al., 2006).

From pleasure to meaningful pleasure and human flourishing


The transition from the experience of pleasure to well-being is not straightforward (Kringelbach & Berridge,
2009). A lot of pleasure rarely, if ever, translates into states of well-being. Rather, excessive pleasure seeking
can often get stuck in maladaptive cyclical addictive behaviours which are seldom pleasurable over the long
term.
Aristotle distinguished between hedonia and eudaimonia, where the latter is perhaps best thought of as
virtue or trait: a way to live well, thrive, and, in the end, have a good life. Equally, however, it also possible to
think of eudaimonia and human flourishing as meaningful experiences which tend to be fleeting, transitory
states which are hard to reliably invoke. Take for example patients with a life-threatening cancer diagnosis
who can be helped by a small dose of psychedelics (Grob et al., 2011). Such carefully controlled psychedelic
experiences are often rated by people as among their five most meaningful experiences (Griffiths et al., 2008;
Griffiths et al., 2011). As an example, one participant reported “to ‘let go’ and become enveloped in the
beauty of—in this case music—was enormously spiritual.” Another reported “I realised I was glad to be alive.
I’ve always thought I wouldn’t be able to feel that.”
This state of eudaimonia is clearly important to ease these patients’ severe depression and anxiety but,
given their transitory nature, difficult to study with scientific means. Over the last few years, we have started
to investigate robust ways of getting human participants engaged in meaningful eudaimonic experiences,
which has involved special practices and stimuli, including psychedelics but also social interactions, music,
and meditation. This experimental approach is then combined with sophisticated whole-brain modelling of
neuroimaging data, which allows us to draw causal mechanistic inferences on the underlying brain mecha-
nisms and networks (Deco et al., 2015; Kringelbach & Deco, 2020).
Over the coming years we hope to make significant progress in understanding how and why humans are
optimized to search for meaning in everything we encounter. Beyond fundamental pleasures keeping us alive,

67
Stark, Berridge and Kringelbach

we have a deep need to discover meaning in our interactions with the world through our senses, interpret-
ing neural signals from the enchanted loom of the billions of neurons constituting our brain. As such we go
through the cycles of life (Ahrends et al., 2021), constantly predicting what might happen and extracting
meaning from the events we encounter and appraising them with further meaning.
The arts exemplify this process: we inject meaning and complexity into the music, dance, visual art, and
poetry we encounter through anticipatory frameworks and neuroaesthetic reactions. A “meaningful” life is
one of the distinguishing characteristics of eudaimonia. Importantly, many people claim that they cannot live
without music or the arts, that they value it and that it provides them with more than just pleasure. Perhaps
art is only incidental, but it may also endure because it exploits the fundamental nature of what makes us
human.

References
Ahrends, C., Vuust, P., & Kringelbach, M. L. (2021). Predictive intelligence for learning and optimization. In A. K.
Barbey, S. Karama & R. J. Haier (Eds.), The Cambridge handbook of intelligence and cognitive neuroscience (pp. 162–188).
Cambridge University Press.
Aldridge, J. W., & Berridge, K. C. (2010). Neural coding of pleasure: “Rose-tinted glasses” of the ventral pallidum. In
M. L. Kringelbach & K. C. Berridge (Eds.), Pleasures of the brain (pp. 62–73). Oxford University Press.
Anderson, D. J.,  & Adolphs, R. (2014). A  framework for studying emotions across species. Cell, 157(1), 187–200.
https://doi.org/10.1016/j.cell.2014.03.003
Barrett, L. F., Mesquita, B., Ochsner, K. N., & Gross, J. J. (2007). The experience of emotion. Annual Review of Psychol-
ogy, 58, 373–403. https://doi.org/10.1146/annurev.psych.58.110405.085709
Berridge, K. C. (2018). Evolving concepts of emotion and motivation. Frontiers in Psychology, 9, 1647. https://doi.
org/10.3389/fpsyg.2018.01647
Berridge, K. C. (2019). A liking versus wanting perspective on emotion and the brain. In The Oxford handbook of positive
emotion and psychopathology (p. 184). Oxford University Press.
Berridge, K. C., & Kringelbach, M. L. (2008). Affective neuroscience of pleasure: Reward in humans and animals. Psy-
chopharmacology, 199(3), 457–480. https://doi.org/10.1007/s00213-008-1099-6
Berridge, K. C.,  & Kringelbach, M. L. (2015). Pleasure systems in the brain. Neuron, 86(3), 646–664. https://doi.
org/10.1016/j.neuron.2015.02.018
Berridge, K. C., Venier, I. L., & Robinson, T. E. (1989). Taste reactivity analysis of 6-hydroxydopamine-induced apha-
gia: Implications for arousal and anhedonia hypotheses of dopamine function. Behavioral Neuroscience, 103(1), 36–45.
https://doi.org/10.1037//0735-7044.103.1.36
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Brauer, L. H., Cramblett, M. J., Paxton, D. A.,  & Rose, J. E. (2001). Haloperidol reduces smoking of both
nicotine-containing and denicotinized cigarettes. Psychopharmacology, 159(1), 31–37. https://doi.org/10.1007/s00213
0100894
Cacioppo, S., Bianchi-Demicheli, F., Frum, C., Pfaus, J. G., & Lewis, J. W. (2012). The common neural bases between
sexual desire and love: A multilevel kernel density fMRI analysis. Journal of Sexual Medicine, 9(4), 1048–1054. https://
doi.org/10.1111/j.1743-6109.2012.02651.x
Crisp, R., & Kringelbach, M. L. (2017, In press). Higher and lower pleasures revisited: Evidence from neuroscience.
Neuroethics. https://doi.org/10.1007/s12152-017-9339-2
Cromwell, H. C.,  & Berridge, K. C. (1993). Where does damage lead to enhanced food aversion: The ven-
tral pallidum/substantia innominata or lateral hypothalamus? Brain Research, 624(1–2), 1–10. https://doi.
org/10.1016/0006-8993(93)90053-p
Damasio, A., & Carvalho, G. B. (2013). The nature of feelings: Evolutionary and neurobiological origins. Nature Reviews.
Neuroscience, 14(2), 143–152. https://doi.org/10.1038/nrn3403
Deco, G., Tononi, G., Boly, M., & Kringelbach, M. L. (2015). Rethinking segregation and integration: Contributions of
whole-brain modelling. Nature Reviews. Neuroscience, 16(7), 430–439. https://doi.org/10.1038/nrn3963
Frijda, N. H. (1986). The emotions. Cambridge University Press.
Frijda, N. H., & Parrott, W. G. (2011). Basic emotions or ur-emotions? Emotion Review, 3(4), 406–415. https://doi.
org/10.1177/1754073911410742

68
The neurobiology of liking

Friston, K., Kilner, J., & Harrison, L. (2006). A free energy principle for the brain. Journal of Physiology—Paris, 100(1–3),
70–87. https://doi.org/10.1016/j.jphysparis.2006.10.001
Georgiadis, J. R., & Kringelbach, M. L. (2012). The human sexual response cycle: Brain imaging evidence linking sex to
other pleasures. Progress in Neurobiology, 98(1), 49–81. https://doi.org/10.1016/j.pneurobio.2012.05.004
Georgiadis, J. R., Kringelbach, M. L., & Pfaus, J. G. (2012). Sex for fun: A synthesis of human and animal neurobiology.
Nature Reviews. Urology, 9(9), 486–498. https://doi.org/10.1038/nrurol.2012.151
Griffiths, R. R., Johnson, M. W., Richards, W. A., Richards, B. D., McCann, U., & Jesse, R. (2011). Psilocybin occa-
sioned mystical-type experiences: Immediate and persisting dose-related effects. Psychopharmacology, 218(4), 649–665.
https://doi.org/10.1007/s00213-011-2358-5
Griffiths, R., Richards, W., Johnson, M., McCann, U., & Jesse, R. (2008). Mystical-type experiences occasioned by
psilocybin mediate the attribution of personal meaning and spiritual significance 14 months later. Journal of Psychop-
harmacology, 22(6), 621–632. https://doi.org/10.1177/0269881108094300
Grill, H. J., & Norgren, R. (1978). The taste reactivity test. II. Mimetic responses to gustatory stimuli in chronic thalamic
and chronic decerebrate rats. Brain Research, 143(2), 281–297. https://doi.org/10.1016/0006-8993(78)90569-3
Grob, C. S., Danforth, A. L., Chopra, G. S., Hagerty, M., McKay, C. R., Halberstadt, A. L., & Greer, G. R. (2011). Pilot
study of psilocybin treatment for anxiety in patients with advanced-stage cancer. Archives of General Psychiatry, 68(1),
71–78. https://doi.org/10.1001/archgenpsychiatry.2010.116
Hyman, S. E., & Fenton, W. S. (2003). Medicine. What are the right targets for psychopharmacology? Science, 299(5605),
350–351. https://doi.org/10.1126/science.1077141
Insel, T., Cuthbert, B., Garvey, M., Heinssen, R., Pine, D. S., Quinn, K., Sanislow, C., & Wang, P. (2010). Research
domain criteria (RDoC): Toward a new classification framework for research on mental disorders. American Journal
of Psychiatry, 167(7), 748–751.
Khan, H. A., Urstadt, K. R., Mostovoi, N. A., & Berridge, K. C. (2020). Mapping excessive “disgust” in the brain:
Ventral pallidum inactivation recruits distributed circuitry to make sweetness “disgusting”. Cognitive, Affective and
Behavioral Neuroscience, 20(1), 141–159. https://doi.org/10.3758/s13415-019-00758-4
Koob, G. F.,  & Le Moal, M. (2001). Drug addiction, dysregulation of reward, and allostasis. Neuropsychopharmacol-
ogy: Official Publication of the American College of Neuropsychopharmacology, 24(2), 97–129. https://doi.org/10.1016/
S0893-133X(00)00195-0
Kringelbach, M. L., & Berridge, K. C. (2009). Towards a functional neuroanatomy of pleasure and happiness. Trends in
Cognitive Sciences, 13(11), 479–487. https://doi.org/10.1016/j.tics.2009.08.006
Kringelbach, M. L., & Berridge, K. C. (2010). Pleasures of the brain. Oxford University Press.
Kringelbach, M. L.,  & Deco, G. (2020). Brain states and transitions: Insights from computational neuroscience. Cell
Reports, 32(10), 108128. https://doi.org/10.1016/j.celrep.2020.108128
Kringelbach, M. L., & Phillips, H. (2014). Emotion: Pleasure and pain in the brain. Oxford University Press.
Kringelbach, M. L., Stein, A.,  & van Hartevelt, T. J. (2012). The functional human neuroanatomy of food pleasure
cycles. Physiology and Behavior, 106(3), 307–316. https://doi.org/10.1016/j.physbeh.2012.03.023
Laeng, B., Eidet, L. M., Sulutvedt, U., & Panksepp, J. (2016). Music chills: The eye pupil as a mirror to music’s soul.
Consciousness and Cognition, 44, 161–178. https://doi.org/10.1016/j.concog.2016.07.009
Leyton, M., Casey, K. F., Delaney, J. S., Kolivakis, T.,  & Benkelfat, C. (2005). Cocaine craving, euphoria, and self-
administration: A preliminary study of the effect of catecholamine precursor depletion. Behavioral Neuroscience, 119(6),
1619–1627. https://doi.org/10.1037/0735-7044.119.6.1619
Miller, J. M., Vorel, S. R., Tranguch, A. J., Kenny, E. T., Mazzoni, P., van Gorp, W. G., & Kleber, H. D. (2006). Anhe-
donia after a selective bilateral lesion of the globus pallidus. American Journal of Psychiatry, 163(5), 786–788. https://
doi.org/10.1176/ajp.2006.163.5.786
Panksepp, J. (1995). The emotional sources of “Chills” Induced by Music. Music Perception, 13(2), 171–207. https://doi.
org/10.2307/40285693
Peciña, S., Berridge, K. C., & Parker, L. A. (1997). Pimozide does not shift palatability: Separation of anhedonia from
sensorimotor suppression by taste reactivity. Pharmacology, Biochemistry, and Behavior, 58(3), 801–811. https://doi.
org/10.1016/s0091-3057(97)00044-0
Rømer Thomsen, K., Whybrow, P. C., & Kringelbach, M. L. (2015). Reconceptualizing anhedonia: Novel perspec-
tives on balancing the pleasure networks in the human brain. Frontiers in Behavioral Neuroscience, 9, 49. https://doi.
org/10.3389/fnbeh.2015.00049
Russell, J. A. (2003). Core affect and the psychological construction of emotion. Psychological Review, 110(1), 145–172.
https://doi.org/10.1037/0033-295x.110.1.145
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726

69
Stark, Berridge and Kringelbach

Smith, K. S.,  & Berridge, K. C. (2007). Opioid limbic circuit for reward: Interaction between hedonic hotspots of
nucleus accumbens and ventral pallidum. Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 27(7),
1594–1605. https://doi.org/10.1523/JNEUROSCI.4205-06.2007
Smith, K. S., Mahler, S. V., Pecina, S., & Berridge, K. C. (2010). Hedonic hotspots: Generating sensory pleasure in the
brain. In M. L. Kringelbach & K. C. Berridge (Eds.), Pleasures of the brain (pp. 27–49). Oxford University Press.
Spijker, J., Bijl, R. V., de Graaf, R., & Nolen, W. A. (2001). Determinants of poor 1-year outcome of DSM-III-R major
depression in the general population: Results of the Netherlands Mental Health Survey and Incidence Study (NEM-
ESIS). Acta Psychiatrica Scandinavica, 103(2), 122–130. https://doi.org/10.1034/j.1600-0447.2001.103002122.x
Steiner, J. E. (1973). The gustofacial response: Observation on normal and anencephalic newborn infants. Symposium on
Oral Sensation and Perception, 4(4), 254–278.
Steiner, J. E., Glaser, D., Hawilo, M. E., & Berridge, K. C. (2001). Comparative expression of hedonic impact: Affective
reactions to taste by human infants and other primates. Neuroscience and Biobehavioral Reviews, 25(1), 53–74. https://
doi.org/10.1016/s0149-7634(00)00051-8
Vartanian, O., & Skov, M. (2014). Neural correlates of viewing paintings: Evidence from a quantitative meta-analysis
of functional magnetic resonance imaging data. Brain and Cognition, 87, 52–56. https://doi.org/10.1016/j.bandc.
2014.03.004
Veldhuizen, M. G., Rudenga, K. J., & Small, D. M. (2010). The pleasure of taste, flavor and food. In M. L. Kringel-
bach & K. C. Berridge (Eds.), Pleasures of the brain (pp. 146–168). Oxford University Press.
Volkow, N. D., Fowler, J. S., Wang, G. J., & Goldstein, R. Z. (2002). Role of dopamine, the frontal cortex and memory
circuits in drug addiction: Insight from imaging studies. Neurobiology of Learning and Memory, 78(3), 610–624. https://
doi.org/10.1006/nlme.2002.4099
Vuust, P., Heggli, O. A., Friston, K. J., & Kringelbach, M. L. (2021). Music in the brain. Nature Reviews. Neuroscience,
23(5), 287–305. https://doi.org/10.1038/s41583-022-00578-5
Vuust, P., & Kringelbach, M. L. (2010). The pleasure of music. In M. L. Kringelbach & K. C. Berridge (Eds.), Pleasures
of the brain (pp. 255–269). Oxford University Press.
Wachtel, S. R., Ortengren, A.,  & De Wit, H. (2002). The effects of acute haloperidol or risperidone on subjective
responses to methamphetamine in healthy volunteers. Drug and Alcohol Dependence, 68(1), 23–33. https://doi.org/
10.1016/s0376-8716(02)00104-7
Winkielman, P., Berridge, K. C., & Wilbarger, J. L. (2005). Unconscious affective reactions to masked happy versus
angry faces influence consumption behavior and judgments of value. Personality and Social Psychology Bulletin, 31(1),
121–135. https://doi.org/10.1177/0146167204271309
Winkielman, P., & Gogolushko, Y. (2018). Influence of suboptimally and optimally presented affective pictures and words
on consumption-related behavior. Frontiers in Psychology, 8, 2261. https://doi.org/10.3389/fpsyg.2017.02261
Wyvell, C. L., & Berridge, K. C. (2000). Intra-accumbens amphetamine increases the conditioned incentive salience
of sucrose reward: Enhancement of reward “wanting” without enhanced “liking” or response reinforcement. Jour-
nal of Neuroscience: the Official Journal of the Society for Neuroscience, 20(21), 8122–8130. https://doi.org/10.1523/
JNEUROSCI.20-21-08122.2000

70
4
DISLIKING
From adaptive disgust to ugliness

Christoph Klebl, Michael Donner and Indra Bishnoi

Humans and other animals have evolved various mechanisms to deal with threats. For example, the emo-
tion fear has evolved to help us defend against imminent physical threat posed by predators (e.g., lions or
snakes), aggressive conspecifics or heights by motivating a fight-or-flight response (LaBar, 2016). Another
major selection pressure that has been posed on the evolution of all species, including humans, is threat
from pathogens—that is, organisms such as viruses and bacteria that can cause disease ( Janeway et al., 2001;
Wolfe et al., 2007). Pathogen threat differs from physical threat posed by predators or aggressive conspecifics
because pathogens cannot travel large distances easily. In order to infect a new organism, pathogens often
require physical contact between the past and future host through their mouths, skins, anuses or genitals
(Tybur et al., 2013).
In response to these recurring threats, natural selection has fashioned humans (as well as other animals)
with a suite of flexible systems that function to reduce the probability of pathogen infection. Most impor-
tantly, animals have evolved immune systems—complex network of cells, proteins and tissues that remove
pathogens from the body ( Janeway et  al., 2001)—in order to defend their bodies against diseases. The
human immune system is very flexible and can acquire immune responses to defend against novel pathogens
( Janeway et al., 2001). It is, however, energetically costly and might not be able to successfully fight novel
pathogens at first exposure. Therefore, humans have evolved other mechanisms—particularly the emotion
disgust—which serve as an early line of defence against pathogens (Tybur et al., 2013).

Pathogen disgust
With broad agreement that the emotion disgust functions to facilitate disease avoidance (Curtis et al., 2011;
Oaten et al., 2009; Rozin et al., 2008; Tybur et al., 2013), it is unsurprising that research into the physiologi-
cal, psychological and behavioural mechanisms underlying pathogen avoidance has largely been guided by
research on disgust. Much of the contemporary theorizing about the function of disgust has been informed
by careful consideration of the range of cues that elicit disgust (for a review see, Stevenson et al., 2019).
Though there is some disagreement about the number of different elicitor domains of disgust (and thus disa-
greement about the different functions of disgust), theorists generally agree that disgust is elicited by disease-
related cues (e.g., rotting flesh, blood, faeces) and that detection of these cues initiates withdrawal responses
including oral expulsion (e.g., spitting and vomiting) as well as facial movements that guard the face, mouth
and eyes (e.g., wrinkled nose, lip retraction and squinted eyes) (for a review, see Oaten et al., 2009). Disgust

DOI: 10.4324/9781003008675-5 71
Klebl, Donner and Bishnoi

is also characterized by feelings of nausea, prompted as a result of ingesting (or coming close to ingesting)
pathogenic substances, as well as decreased heart rate and increased skin conductance response (Stark, Walter
et al., 2005). Perhaps most importantly, disgust facilitates the avoidance of physical contact with the disgust
elicitor (i.e., removing or distancing oneself from the presence of pathogens), thus neutralizing the possibil-
ity of pathogen transmission. These functional, behavioural, physiological and phenomenological properties
capture what is commonly referred to as pathogen disgust (the term that we also adopt in this chapter; Tybur
et al., 2013) and what others have termed core disgust (Rozin et al., 2008).
It is important to note, however, that the mere presence of pathogens does not always lead to a pathogen
disgust response in all contexts. Following Tybur et al.’s (2013) evolutionary-computational approach to the
emotion of disgust, estimates of the cost of contact with a pathogen elicitor are integrated with estimates
of the costs of avoiding contact with the pathogen elicitor (i.e., the benefits associated with contact). For
example, the caloric benefit of eating decayed food might outweigh the potential costs of pathogen contami-
nants in the food for someone in a state of acute hunger. Similarly, mothers are less disgusted by their own
baby’s faeces-soiled diapers compared to those of some else’s baby (Case et al., 2006), and people are more
comfortable with contacting infectious individuals whom they value interpersonally (Tybur et al., 2020).
Thus, whilst the overall function of the pathogen disgust system is to coordinate the avoidance of infectious
disease vectors, the system also evaluates a range of other inputs that might confer benefits to an individual,
thus allowing flexible responding.

Neurobiology of pathogen disgust


The anterior insular cortex is the primary gustatory cortex (see Figure 4.1), involved in taste and flavour
(Rolls & Baylis, 1994). As pathogens often require physical contact through orifices like the mouth for infec-
tion (Tybur et al., 2013), this cortex has been associated with pathogen-related disgust in mammals (Tuerke
et al., 2012; Wicker et al., 2003). The anterior insular cortex takes sensory input and converts it to physiologi-
cal sensations of disgust, such as nausea and vomiting (Wicker et al., 2003). In mice, the optogenetic activation
of the anterior insular cortex elicits disgust-related responses, such as facial expressions (Dolensek et al., 2020).
Additionally, lesioning the insular cortex prevents conditioned gaping, a disgust behaviour in rats (Kiefer &
Orr, 1992). In humans, electrical stimulation of the anterior insular cortex triggers nausea and sensations of
sickness (Krolak-Salmon et al., 2003; Penfield & Faulk, 1955). In functional magnetic resonance imaging
(fMRI) studies, visually eliciting pathogen disgust leads to an increase in activation of the insular cortex and

Figure 4.1 Illustration of the human brain with the bilateral insula highlighted in dark orange. The insula is located deep
within the lateral sulcus separating the temporal lobe from the parietal and frontal lobes.

72
Disliking

the amygdala (Regenbogen et  al., 2017; Wright et  al., 2004), the brain region responsible for emotional
responses and emotion-based learning and memory (Gallagher & Chiba, 1996; Maren, 1999). Projections
from the anterior insular cortex to the medial and basolateral amygdala have been found to be involved in
disgust-related behaviours, such as aversive taste learning (Kayyal et al., 2019). This relationship may be medi-
ated by the elevation of serotonin in the insular cortex after toxin exposure (Limebeer et al., 2018; Tuerke
et  al., 2012). Furthermore, the area postrema, located in the brainstem, is an important brain region for
toxin disgust specifically because of its structure, which has a thin blood brain barrier, making it particularly
easy for blood borne toxins to infect this region (Dempsey, 1973). Gastrointestinal defence mechanisms like
nausea and vomiting are triggered by the activation of the area postrema in humans and rodents (Borison,
1989; Miller & Leslie, 1994), such that lesions to the area postrema inhibit toxin-induced conditioned disgust
behaviours in rodents (Eckel & Ossenkopp, 1993; Ossenkopp & Eckel, 1995; Ossenkopp et al., 1994).
The induction and expression of disgust involves a number of brain regions within the social decision-
making network (SDMN). The SDMN merges the mesolimbic reward network with the social brain net-
work (Newman, 1999; O’Connell & Hofmann, 2011), which includes the cingulate and prefrontal cortices,
nucleus accumbens, thalamus, dorsal hippocampus, paraventricular nucleus of the hypothalamus, ventral
tegmental area, insular cortex, amygdala and sensory regions such as the piriform cortex and occipital lobe
(Becker et al., 2016; Borg et al., 2013; Goodson, 2013; Johnson et al., 2017; Marlin & Froemke, 2017).
These regions are responsible for encoding sensory and social cues that are useful for the expression of disgust
behaviours and their respective reward values (Kavaliers et al., 2021). It is likely that pathogen disgust, as well
as sexual and moral disgust, is elicited by coordinating across both overlapping and distinct regions within
this network.

Sexual disgust
Whilst there is a strong correspondence between cues that indicate disease and cues that elicit disgust, many
theorists have suggested that cues that do not contain obvious traces of disease can also activate a disgust
response (Ackerman et al., 2007; Lieberman et al., 2007; Rozin et al., 1999, 2008; Tybur et al., 2009, 2013).
Indeed, it has been argued that the role of disgust has been expanded, through the evolutionary process of
exaptation, to certain components of sexuality (e.g., incest, unwelcomed sexual attention). Building on the
pre-existing pathogen disgust system, a sexual disgust system evolved to perform the function of avoiding
contact with sexual partners who might jeopardize one’s reproductive fitness (Tybur et al., 2013). Similar to
the computational structure of pathogen disgust, the sexual disgust system is designed to estimate the fitness
value of an individual as a sexual partner by weighing up the potential costs (e.g., exposure to pathogens
through intercourse) and benefits (e.g., reproduction) of sexual contact. From this perspective, every sexual
encounter carries with it the risk of novel pathogen exposure, and these costs can potentially outweigh any
reproductive benefits associated with having an unrestricted mating strategy (i.e., multiple sexual partners;
Tybur et  al., 2013, 2015). Following this logic, sexual contact that risks producing unhealthy offspring
(i.e., sex with kin, sex with the elderly) is particularly likely to elicit sexual disgust (Tybur et al., 2013). Thus,
to avoid individuals posing a reproductive threat, the sexual disgust system should be particularly sensitive
to input cues associated with sex value such as health, age and attractiveness (among others) (Grammer
et al., 2003; Thornhill & Gangestad, 2006). Given that the nature of the inputs for the sexual disgust system
are qualitatively different from the inputs for the pathogen disgust system (due to their somewhat distinct
adaptive problems), pathogen and sexual disgust likely have somewhat different functional, behavioural,
physiological and phenomenological properties (Tybur et al., 2013). Nevertheless, while sexual and patho-
gen disgust are somewhat distinct from each other, they likely overlap in everyday disgust experiences. For
example, people report feeling disgusted towards potential sexual partners who display symptoms of illness
(Buss & Schmitt, 1993; Symons, 1979). Moreover, across mammalian species, attraction towards a potential

73
Klebl, Donner and Bishnoi

partner is hindered when paired with a potential pathogen threat ( Jones et al., 2013; Kavaliers et al., 2021;
Kavaliers et al., 2014).
Activation of select regions within the SDMN has been observed in functional magnetic resonance
studies where participants were exposed to images of non-preferred sexual stimuli, that is, erotic images of
partners of the opposite sexual orientation, hardcore pornographic stimuli and sadomasochism. Specifically,
the prefrontal cortex, amygdala, nucleus accumbens, hippocampus, thalamus and olfactory regions are acti-
vated during feelings of sexual disgust (Borg et al., 2014; Stark, Schienle et al., 2005; Zhang et al., 2011).
Hormonal modulation has widespread effects between these regions (Kavaliers et al., 2021). Male mice with
higher testosterone levels, due to sexual arousal and/or higher social hierarchy, show a decreased avoidance of
pathogen threat and higher parasite levels (Barnard et al., 1998; Kavaliers et al., 2001; Kavaliers & Choleris,
2018). Additionally, females, who are more susceptible to direct sexually acquired infection, display stronger
physiological disgust and greater avoidance behaviours towards infectious conspecifics (Kavaliers et al., 2019;
Poirotte & Kappeler, 2019).

Moral disgust
A final category of disgust elicitors that is of interest involves moral norm violations. Given what we know
about the evolved functions of both pathogen and sexual disgust, it should be unsurprising that certain
pathogen-related moral transgressions (e.g., eating worms) and sex-related moral transgressions (e.g., mas-
turbation, homosexuality) elicit disgust. What is striking, however, is that people report feeling disgusted, as
well as displaying the canonical disgust facial expression, to moralized actions that are free of pathogen and
sex stimuli such as fraud, theft, physical assault and disloyalty (Chapman et al., 2009; Hutcherson & Gross,
2011; Rozin et al., 2008; Tybur et al., 2009, 2013). A key question here is whether these moral violations
activate a disgust response (with its signature functional, behavioural, physiological and phenomenological
properties) or whether these responses resemble a metaphorical usage of disgust, a contentious issue that has
been debated for many years (e.g., Bloom, 2004; Nabi, 2002; Royzman & Kurzban, 2011). Setting this issue
aside, it has been proposed by Tybur and colleagues (2009, 2013) that moral disgust functions to coordinate
condemnation of people who violate moral rules. In this view, expressing disgust vocally or facially com-
municates one’s disapproval of a given behaviour (Giner-Sorolla et al., 2018; Kupfer & Giner-Sorolla, 2016;
Royzman & Kurzban, 2011; Tybur et al., 2013). An ability to detect others’ expressions of disgust as com-
municating condemnation also plays a key role in reputation management. Given that maintaining a repu-
tation as a valuable community member is essential for the evolution of cooperation (Nowak & Sigmund,
2005; Trivers, 1972), understanding others’ signals of condemnation (e.g., expressions of moral disgust) ena-
bles one to avoid the individual fitness costs associated with punishment and social exclusion (and reap the
individual benefits of cooperation). With enough support, these expressions of moral disgust can be effective
at coordinating punishment and condemnation of socially undesirable behaviours.
The medial prefrontal cortex has been found to be uniquely related to moral disgust. However, despite
this and its unique function, moral disgust has been found to share certain brain regions related to the pro-
cessing of pathogen and sexual disgust, such as the insula and the cingulate cortex (Borg et al., 2008). These
brain regions are activated when participants view photos of moral transgressions, including incest, public
indecency and murder (Borg et al., 2008; Vicario et al., 2017; Ying et al., 2018). Interestingly, similar facial
expressions are evoked by a disgusting taste, observation of a contaminant and unfair treatment in an eco-
nomic game (Chapman et al., 2009). Additionally, thinking about moral transgressions leads individuals to
perceive a neutral-tasting beverage as disgusting (Eskine et al., 2012). As such, while both neurobiologically
unique and overlapping regions bring about moral disgust, the resulting behavioural and physiological output
is similar to that of other types of disgust. The roles of the insula and medial prefrontal cortex in moral dis-
gust may be to bring about these physiological disgust behaviours and moral cognition and decision-making,

74
Disliking

respectively, while the cingulate cortex has been theorized to play a role in moral reasoning (Fumagalli &
Priori, 2012; Koenigs et al., 2007; Sevinc & Spreng, 2014; Vicario et al., 2017).
Before moving on, it is worthwhile acknowledging that whilst we have attempted to cover some of the
key functions and elicitor domains of the emotion disgust that are common to most theoretical models of
disgust, there have been other domains proposed that arguably do not cleanly map onto the three disgust
categories discussed previously (pathogen, sexual and moral). Of primary relevance here is the specific cat-
egory of contamination disgust found in Rozin and colleagues’ (1987, 1994, 2008) account of disgust (Haidt,
2001), as well as Olatunji and colleagues’ (2007) revised model of Rozin’s account. Contamination disgust
represents the experience of disgust toward a neutral stimulus (e.g., toothbrush) that makes physical (and even
imagined) contact with a disgusting stimulus (e.g., faeces). Given this close relationship between contamina-
tion and disgust, others have argued that rather than representing a domain of disgust altogether, contamina-
tion is best thought of as an accompanying feature of the pathogen disgust system (Tybur et al., 2013).

Individual differences in disgust sensitivity


Validation of the various categories of disgust has been examined using measures of individual differences
in disgust sensitivity. Tybur and colleagues (2009) developed the Three Domain Disgust Scale (TDDS) to
assess individual differences in sensitivity to pathogen, sexual and moral disgust. In line with their general
physiological overlap, individuals with higher sensitivity to pathogen, sexual and moral disgust, on average,
experience more intense disgust in each domain. However, factor analyses of self-report data from multiple
studies confirm the three-factor disgust model (DeBruine et al., 2010; Olatunji et al., 2012; Tybur et al.,
2009, 2011), suggesting that pathogen, sexual and moral disgust may be functionally distinct categories. In
addition, varying sex differences on these individual difference measures of disgust also indicate that there
might be qualitative differences between the domains of disgust. Specifically, although women, on average,
report greater disgust responses than men across all disgust domains, the sex difference is markedly larger
in the sexual disgust domain compared to the pathogen and moral domains (Al-Shawaf et al., 2018; Fleis-
chman, 2014; Tybur et al., 2009)—a finding that is consistent with theorizing about the disparate (sexual)
selection pressures faced by males and females (for reviews, see Buss & Schmitt, 1993; Trivers, 1972). Finally,
evidence that pathogen, sexual and moral disgust are somewhat distinct can also be found by considering the
divergent correlations they have with other self-report measures, such as the Big Five personality traits (for
reviews, see Tybur et al., 2009, 2013; Tybur & Karinen, 2018), further supporting the notion that there may
be functionally distinct domains of disgust (Tybur et al., 2013).

The social consequences of the disease-avoidance system


In the previous sections, we have argued that the disgust systems serve to motivate avoidance of pathogens
and costly sexual encounters and to potentially coordinate reactions to moral infractions. Due to the high
costs of failing to detect a threat when one is present, the signal detection process in each disgust system—
pathogen, sexual and moral—is hypersensitive to treating a range of cues as sources of threat. More precisely,
the disease-avoidance system evolved to minimize errors that would lead to the greatest costs to one’s fitness
(Haselton & Buss, 2000). For example, as it is potentially fatal to judge an infected person to be healthy
and comparably less costly to judge a healthy person to be infectious, the disease-avoidance system is biased
toward avoiding the former error (i.e., false positives; Schaller & Park, 2011).
The disease-avoidance system has been argued to be implicated in the stigmatization of individuals that
are perceived to pose the risk of transmitting infectious diseases. Particularly individuals that have visible cues
of infectious diseases, such as people who suffer from leprosy, are at risk of being subjected to stigmatization
(Kurzban & Leary, 2001). However, due to its over-inclusiveness, the disease-avoidance system can also be

75
Klebl, Donner and Bishnoi

activated by people who do not have a contagious disease (Kurzban & Leary, 2001). For example, being
perceived to be a source of novel pathogens or being less familiar with cultural norms (Schaller  & Park,
2011) can be perceived to pose a greater disease risk, potentially leading to the stigmatization of outgroup
members. Both dispositional and situational feelings of vulnerability to disease positively predict xenophobic
attitudes (Faulkner et al., 2004), and women report greater ethnocentric attitudes in the first trimester of
their pregnancies—where the maternal and fetal vulnerability to disease is the greatest—compared to the
second and third trimester of their pregnancies (Navarrete et  al., 2007). Furthermore, people who have
a greater dispositional proneness to experience disgust tend to intuitively disapprove of gay people more
strongly (Inbar et al., 2009).
It has been argued that the disease-avoidance system also underlies the stigmatization of people that are
perceived as unattractive, particularly those with facial and bodily disfigurement (Park et al., 2003). People
with facial and bodily differences are avoided in everyday social interactions, leading to them feeling mar-
ginalized (Goffman, 1963). For example, people approach individuals with facial disfigurement less often
(Kapp-Simon  & McGuire, 1997), they keep greater physical distance from them (Rumsey et  al., 1982)
and they exhibit a greater cardiovascular reactivity when interacting with them (Blascovich et al., 2001),
compared to individuals without facial disfigurement. Studies have found that people feel disgusted toward
and physically avoid facially disfigured individuals to the same degree as people who have influenza (Ryan
et al., 2012). Additionally, people avoid physical contact with individuals who have physical disabilities, such
as facial birthmarks or obesity (Park et al., 2013). Consistent with this, ugly faces elicit greater disgust than
average-looking and attractive faces (Klebl et al., 2021). This bias has social consequences for individuals.
People not only attribute less socially desirable traits to unattractive compared to attractive individuals (Dion
et al., 1972), but unattractive people also face social disadvantages such as lower success in finding employ-
ment (Hosoda et al., 2003) and more severe punishments in jury decisions (Efran, 1974). People’s bias against
individuals with disfigurements is also reflected in (and perhaps exacerbated by) mainstream media. For
example, the majority of all-time top American film villains, but none of the all-time top American film
heroes, have dermatologic conditions (Croley et al., 2017).
The hypervigilance of the disease-avoidance system (especially to pathogen and sex-related cues) seemingly
also extends to people’s abstract social and political attitudes. Indeed, research over the years has demonstrated
that the disgust systems play an important role in political orientation (Inbar et al., 2009; Navarrete & Fessler,
2006; Smith et  al., 2011). One particularly influential line of work considers the political sentiments of
individuals who vary in their sensitivity to different domains of disgust. This work has reliably demonstrated
that individual differences in disgust sensitivity is associated with greater political conservatism (e.g., Inbar
et al., 2009, 2012; Terrizzi et al., 2013). Having established this general relationship, researchers have more
recently begun to unpack which domains of disgust relate to which dimensions of political conservatism in
order to explain the effect. Of relevance here is the finding that pathogen disgust sensitivity is more closely
tied to the aspects of conservatism that are concerned with greater adherence to traditional social norms of
the ingroup (i.e., the traditionalism facet of right-wing authoritarianism) in contrast to aspects of conserva-
tism that are concerned with negativity towards ethnic outgroups (i.e., social dominance orientation) (Tybur
et al., 2016). This relationship was further clarified by a study that found disgust sensitivity (to pathogens)
was related to anti-immigration sentiments when immigrants were described as not assimilating to cultural
traditions (testing the ingroup norm account), but not when immigrants were described as assimilating to
cultural traditions (testing the outgroup avoidance account) (Karinen et al., 2019). Thus, individuals who
have a higher propensity to be disgusted by pathogen cues are more likely to endorse ideologies that preserve
cultural norms—a key feature of political conservatism. Although researchers have paid close attention to the
dimensionality of conservatism when investigating associations between measures of disgust sensitivity and
measures of ideology (for a review, see Terrizzi et al., 2013), less attention has been paid to the dimensionality
of disgust sensitivity in this relationship. There are a growing number of studies showing that the association

76
Disliking

between pathogen disgust sensitivity and ideological preferences may be mediated by sexual disgust sensitiv-
ity (i.e., individual sexual reproduction strategies) (Tybur et al., 2015; Billingsley et al., 2018; but see Aarøe
et al., 2020). These findings suggest that individuals who are more prone to experiencing pathogen disgust
are also more likely to pursue restricted reproductive strategies, perhaps as a means to minimize the threat
of pathogen transmission from sexual interactions. The downstream consequences of increased motivations
toward sexual restrictiveness are that individuals choose to endorse or oppose particular policies that favour
their reproductive interests. Because political conservatism aligns with rules that set out to constrain others’
sexual behaviours (e.g., contraception or abortion), it has been argued that this might explain, at least in part,
why dispositional sensitivity to disgust is associated with more politically conservative ideologies (Kurzban
et al., 2010; Tybur et al., 2015; Weeden et al., 2008). As such, it is important to consider the dimensionality
of disgust in order to gain a better understanding of the motivational underpinnings of ideological commit-
ments and political sentiments.
Paralleling the work linking disgust sensitivity with an individual’s overall political orientation is a related
field of research demonstrating that disgust is implicated in moral judgments. Research in this tradition was
motivated by sentimentalist theories of moral psychology that claim emotions have an important role to play
in people’s evaluations of third-party immoral behaviours (Haidt, 2001; Pizarro et al., 2011). Seminal studies
in this area found that exposing participants to noxious odours, dirty desk and bitter tastes increases the per-
ceived wrongness of actions that violate a range of moral rules (i.e., care/harm, fairness/cheating, sanctity/
degradation), and these studies have been taken as evidence that disgust plays a causal role in moral judge-
ment (Eskine et al., 2011; Schnall et al., 2008; Ugazio et al., 2012). It is worth noting that the robustness
of these effects has been challenged by a recent meta-analysis which found that inductions of state disgust
(i.e., temporarily induced disgust) have a very small effect on moral judgements (Landy & Goodwin, 2015).
Importantly, however, this body of work is limited by the fact that studies have only used pathogen disgust
elicitors to induce state disgust, thereby overlooking the possibility that other types of disgust—namely sexual
disgust and moral disgust—might be implicated in moral judgment.
Likewise, early work linking trait disgust sensitivity (i.e., individual differences in disgust sensitivity)
to harsher ratings of moral judgments were only concerned with measures of pathogen disgust sensitivity
(e.g., Chapman & Anderson, 2014; Horberg et al., 2009; Inbar et al., 2009; Jones & Fitness, 2008). Con-
sequently, moral judgments in these studies were interpreted as being (at least partially) based on disease-
avoidance motivations. Driven by theoretical developments in disgust research, more recent work investigat-
ing the motivational and emotional roots of moral judgement have paid attention to the dimensionality of
disgust sensitivity. Indeed, in a recent meta-analysis (Donner et al., 2020) it was found that sexual disgust sen-
sitivity is more strongly correlated with condemnation towards a variety of moral transgressions than patho-
gen disgust sensitivity, suggesting that motivations to avoid unwanted sexual contact also have the potential
to structure moral attitudes, beliefs and behaviours. Interestingly, the most robust associations were observed
between sexual disgust and the moral value domains of sanctity (i.e., upholding spiritual and physical purity)
and authority (i.e., respecting hierarchy and tradition), which are among the most robust correlates of politi-
cal conservatism (Graham et al., 2011, 2013; Kivikangas et al., 2020), thus hinting at possible pathways by
which disgust sensitivity may shape broad left–right political orientations (as discussed previously).
Neuroimaging studies reveal that social disgust shares SDMN brain regions related to the processing of
pathogen, sexual and moral disgust. Viewing faces from other races reveals the importance of the insula,
amygdala and anterior cingulate cortex in racially biased disgust perception (Liu et  al., 2015). Viewing
images of stigmatized individuals, including overweight, transsexual and pierced faces, also activates these
brain regions, in addition to the prefrontal cortex (Krendl et al., 2006). In primates and rodents, the pre-
frontal–amygdala pathways play a central role in social decision-making and are influenced by oxytocin
(OT; Gangopadhyay et al., 2021). OT is a neuropeptide that is produced in the hypothalamus, such as the
paraventricular nucleus (Mitre et al., 2016). Acting in SDMN brain regions, especially the medial amygdala,

77
Klebl, Donner and Bishnoi

OT mediates approach and avoidance behavioural responses, as well as parasite recognition (Kavaliers et al.,
2006). Within the nucleus accumbens and ventral tegmental area, OT enables social approach behaviours
(Steinman et al., 2019). Mice with a deletion of the OT gene, or those treated with an OT antagonist, display
an impairment in social recognition and avoidance of parasitic conspecifics (Kavaliers et al., 2003; Kavaliers
et al., 2006; Kavaliers, Colwell & Choleris, 2005; Kavaliers, Choleris, Ågmo et al., 2005).
While the SDMN is crucial for disgust, regions such as the ventromedial prefrontal cortex (vmPFC) are
likely specific to social disgust (Ciaramelli et al., 2013). The vmPFC has been implicated in social cognition,
specifically evaluating social information and social preference (Amodio & Frith, 2006; Mitchell et al., 2006).
Lesions to the medial prefrontal cortex in macaque monkeys decrease interest in social, but not object, stimuli
(Deaner et al., 2005; Rudebeck et al., 2006). In humans, damage to the vmPFC does not impact pathogen
or moral disgust but does increase approach behaviour towards stimuli that elicit interpersonal disgust, while
healthy individuals display avoidance behaviours towards these stimuli (Ciaramelli et al., 2013). The vmPFC
does not activate when non-social stimuli rated as disgusting, like vomit or an overflowing toilet, are viewed
(Harris & Fiske, 2006). Instead, viewing socially desirable people, such as college students and rich people,
increases the activation of the vmPFC. However, the activation is low when viewing stigmatized individuals
rated as disgusting, such as drug addicts and homeless people, indicating a lower social preference (Harris &
Fiske, 2006, 2007). Specific regions like the vmPFC, paired with regions in the SDMN, such as the insula
and amygdala, likely play a role in the social consequences of disgust (Harris & Fiske, 2006).

Ugliness judgments and the disease-avoidance system


In the previous sections, we have outlined the important role that disgust plays in pathogen avoidance, sexual
behaviour, social stigmatization and even morality and politics. We now turn to the domain of aesthetic
judgments. Understanding the nature of the disease-avoidance system is crucial for a better understanding of
the function of aesthetic judgments—especially ugliness judgments.
Based on theorizing on the function of beauty which proposes that beauty judgments alert us to objects
that increased the chances of survival and reproduction of our ancestors (i.e., have adaptive value; Dutton,
2009), one might assume that ugliness judgments function to alert humans to objects that are detrimental to
survival and reproduction. Anecdotally, however, not all dangerous objects are ugly. For example, while most
people would find a rotten corpse (i.e., a disease threat) ugly, many people would judge a tiger (i.e., a physi-
cal harm threat) as neutral in appearance or even beautiful. Indeed, there is emerging evidence that ugliness
judgments are not generally elicited by dangerous objects but instead have a more specific function—that is,
alerting us to objects that may pose a disease threat (Klebl et al., 2021; Park et al., 2013; Ryan et al., 2012).
Indirect support for this comes from studies discussed previously which showed that facial and bodily dis-
figurement provide a heuristic cue for disease threat and elicit the same behavioural and emotional response
as individuals with a contagious disease (Park et al., 2003; Park et al., 2013; Ryan et al., 2012). More direct
evidence linking ugliness judgments and the disease-avoidance system was provided by a recent series of
studies which found that people judge objects that possess cues of pathogen presence as uglier than similar
objects without pathogen cues and that this was not due to a generalized negative affect (Klebl et al., 2021).
For example, a plate containing a yellowish liquid that looked like bodily fluids was judged as uglier than
a plate containing a liquid with a chemical blue dye. Furthermore, ugly human faces were found to elicit
greater disgust, but not greater fear or sadness, than beautiful or average-looking faces when controlling
for the shared negative valence of disgust, sadness and fear (Klebl et al., 2021). This suggests that there is a
unique association between ugliness judgments and the emotion disgust. This association has been found
to extend to animals and, to a lesser degree, to buildings whose aesthetic features parallel those of organic
entities (Coburn et al., 2019). Specifically, ugly animals elicit greater disgust, but not greater fear or sadness,
than beautiful animals, and ugly buildings elicit greater disgust, but not greater fear or sadness, than beautiful

78
Disliking

buildings, when controlling for the shared negative valence between ugliness judgments and the negative
emotions (Klebl et al., 2021). The association between ugliness and disgust extending to entities that may
not pose a risk for infectious disease (i.e., ugly human faces and ugly animals) and non-organic entities that
cannot pose a disease threat (i.e., ugly buildings) is consistent with how the over-inclusive disease-avoidance
system operates (Haselton & Buss, 2000; Schaller & Park, 2011).
To date, the literature provides only initial evidence into the function of ugliness and its role in the
disease-avoidance system. However, it hints at the possibility that ugliness judgments constitute the aesthetic
dimension of the disease-avoidance system. As ugliness judgments involve attention allocation to the objects
perceived as ugly (Faust et al., 2019), they may facilitate the disease-avoidance system response by allocating
our attention to potentially infectious objects.

The neuroscience of ugliness


As beauty and ugliness are brought about by activation in similar brain regions, an aesthetic continuum has
been theorized (Kandel, 2012). Studies have found regions in the SDMN, specifically the occipital lobe,
nucleus accumbens and cingulate and prefrontal cortices, to be active when people view both attractive and
unattractive faces, music or art (see Chapters 6 and 7). However, the occipital lobe is inversely related, with
greater activation in this region when viewing increasingly unattractive faces, while the nucleus accumbens
and cingulate and prefrontal cortices seem to be linearly related to faces, music and art (Ishizu & Zeki, 2011;
Martín-Loeches et al., 2014). As noted, beauty judgments may alert us to objects that have adaptive value
(Dutton, 2009). Like art, attractive faces activate the medial orbitofrontal region of the prefrontal cortex, a
region activated by reward value, to a greater degree than unattractive faces (Critchley & Rolls, 1996; Elliott
et al., 2000; Francis et al., 1999; Ishizu & Zeki, 2011; O’Doherty et al., 2001; O’Doherty et al., 2003). This
effect is strengthened when attractive faces are presented with a smiling facial expression (O’Doherty et al.,
2003). In contrast, faces rated as lowly attractive activate the lateral orbitofrontal region of the prefrontal cor-
tex, and the insula and cingulate cortex, more so than attractive faces (Krendl et al., 2006; O’Doherty et al.,
2003). However, attractiveness is hindered when it is paired with a pathogen threat in mammals, including
humans ( Jones et al., 2013). In mice, females display a reduced preference towards male mice that have been
paired with pathogen infection (Kavaliers et al., 2021; Kavaliers et al., 2014). In line with this work, OT
antagonists have been found to decrease toxin-induced disgust behaviours that were previously associated
with a social partner in male rats (Boulet et al., 2016). Given the neural overlap between the types of disgust
(e.g., pathogen disgust) and the aesthetic continuum, recent findings add support to the theory that ugliness
alerts us to objects that may pose a disease threat (Klebl et al., 2021; Park et al., 2013; Ryan et al., 2012).

Major challenges, goals and suggestions


While the adaptive function of disgust has been extensively investigated in the past 40 years, research on the
nature of ugliness and its relationship to the disease-avoidance system is still in its infancy. There is initial
evidence suggesting that ugliness judgments and the disease-avoidance system are closely linked with each
other (Klebl et al., 2021; Park et al., 2003; Park et al., 2013; Ryan et al., 2012). However, as of yet, there are
more questions than answers about the nature of this relationship. Important issues that should be addressed
by future research pertain to (1) whether ugliness judgments and the disease-avoidance system are causally
related to each other, (2) whether ugliness judgments have evolved as part of the disease-avoidance system,
(3) the nature of the relationship beyond the visual domain, (4) the instances in which ugly entities do not
elicit a disease-avoidance response and (5) individual and cultural differences in ugliness judgments.
First, the exact nature of the link between ugliness judgments and the disease-avoidance system is not
yet understood. One possibility is that ugliness judgments and the disease-avoidance system are independent

79
Klebl, Donner and Bishnoi

systems, but both can be elicited by particular lower-level features of objects. For example, facial disfig-
urement may elicit both disgust and ugliness judgments because it possesses a feature (e.g., the degree of
non-prototypicality of an entity; Halberstadt & Rhodes, 2000) that is both correlated with the presence
of disease (Kurzban & Leary, 2001) and ugliness judgments. Another possibility is that ugliness judgments
and the disease-avoidance system are interacting systems and, as such, are causally related with each other.
As outlined previously, ugliness judgments may have the function of allocating our attention to objects that
potentially pose a pathogen threat (cf. Klebl et al., 2021). Perceiving an object as ugly may thus facilitate
sustained attention allocation to that object which in turn may amplify one’s disgust response and motivated
avoidance of a potentially infectious object. Alternatively, disgust experiences may cause people to judge
objects as ugly.
Second, if ugliness judgments facilitate the disease-avoidance system response, the question of whether
ugliness judgments have evolved as a part of the disease-avoidance system or have been co-opted by the
disease-avoidance system to facilitate the allocation of attention to potentially infectious objects remains
open. Although evolutionary processes cannot be directly investigated, testable predictions can be generated
that could provide evidence for either possibility (cf. Confer et al., 2010). For example, if ugliness judgments
have evolved to alert us to cues of pathogen threat, we would expect them to be more universal for entities
that can be infectious (i.e., organisms such as humans or animals) compared to objects that cannot carry
pathogens and may only resemble infectious organisms (i.e., non-organic entities such as buildings or other
artefacts; cf. Schaller & Park, 2011) due to a greater adaptive pressure to avoid ugly organic entities. This is
consistent with research that showed that there is greater inter-individual variability in people’s preferences
for abstract images than for real-world images (Vessel & Rubin, 2010). If ugliness judgments, however, have
been co-opted by the disease-avoidance system, one would expect that ugliness judgments have a function
across organic and non-organic entities that is independent from pathogen avoidance. For example, ugli-
ness judgments may have the function to identify objects that deviate strongly from prototypical exemplars
(cf. Halberstadt & Rhodes, 2000).
Third, in order to further illuminate the function of ugliness and its relationship with the disease-avoidance
system, future research should investigate ugliness judgments in perceptual domains beyond visual ugliness.
For example, while smells and tastes can elicit basic disgust responses (Rozin et al., 2008), it is unknown
whether they can be perceived as ugly. Furthermore, while sounds can be perceived as ugly (Ishizu & Zeki,
2011), as well as disgusting—for example, sounds of contact with microbial environments such as squelching
(Oaten et al., 2009)—it is unclear whether auditory ugliness is related to auditory disgust.
Fourth, future research should investigate instances in which ugliness judgments do not trigger disease
avoidance. There is anecdotal evidence that people can have positive responses to ugly objects, such as
toward ugliness depicted in art. For example, people may enjoy Marcel Duchamp’s (1887–1986) Fountain
or paintings by Egon Schiele (1890–1918) that depict twisted bodies. While more frequent in modern
art, the aesthetic appreciation of ugliness is not just a modern phenomenon. For example, the Japanese
aesthetic tradition wabi-sabi (侘寂) values imperfections such as repaired broken ceramics with their cracks
highlighted ( Juniper, 2011). One reason ugly art may elicit positive emotions is that art appreciation is not
limited to the aesthetic dimension of beauty and ugliness (Fingerhut & Prinz, 2018). Ugly art may elicit
other experiences such as interest or wonder that may counteract the negative responses linked with ugli-
ness perceptions.
Fifth, individual and cultural differences in ugliness judgments should be further investigated. For exam-
ple, some indigenous cultures in Africa, Melanesia and Australia use scarification (Garve et al., 2017), and
women of the Ethiopian Mursi use lip-plates (Turton, 2004) to attract sexual partners, while the same
practices might be considered unattractive in other cultures. Furthermore, personality traits such as openness
to experience or one’s artistic expertise may influence whether objects are perceived to be ugly (Cotter et al.,
2017; Vartanian et al., 2019).

80
Disliking

Conclusion
Disgust is an emotion that has evolved to protect us from threats posed by pathogens. As such, pathogen
disgust involves withdrawal and oral expulsion of toxic substances or objects. Building on this system, dis-
gust has been co-opted into the sexual domain, performing the function of avoiding contact with fitness-
jeopardizing sexual partners, as well as into the moral domain, functioning to coordinate the condemnation
of people who violate moral rules (Tybur et al., 2013). The evolved functions of the disgust systems have
implications for everyday social lives such as people’s political orientation, moral judgments, xenophobic
attitudes, and the stigmatization of people with facial differences. Now, disgust has also been found to be
closely linked to ugliness judgments (Klebl et al., 2021), opening up new avenues of research into the role
of disgust in aesthetic judgments.

References
Aarøe, L., Petersen, M. B., & Arceneaux, K. (2020). The behavioural immune system shapes partisan preferences in
modern democracies: Disgust sensitivity predicts voting for socially conservative parties. Political Psychology, 41(6),
1073–1091. https://doi.org/10.1111/pops.12665
Ackerman, J. M., Kenrick, D. T., & Schaller, M. (2007). Is friendship akin to kinship? Evolution and Human Behaviour,
28(5), 365–374. https://doi.org/10.1016/j.evolhumbehav.2007.04.004
Al-Shawaf, L., Lewis, D. M. G., & Buss, D. M. (2018). Sex differences in disgust: Why are women more easily disgusted
than men? Emotion Review, 10(2), 149–160. https://doi.org/10.1177/1754073917709940
Amodio, D. M., & Frith, C. D. (2006). Meeting of minds: the medial frontal cortex and social cognition. Nature Reviews
Neuroscience, 7(4), 268–277. https://doi.org/10.1038/nrn1884
Barnard, C. J., Behnke, J. M., Gage, A. R., Brown, H., & Smithurst, P. R. (1998). Maternal effects on the development
of social rank and immunity trade-offs in male laboratory mice (Mus musculus). Proceedings of the Royal Society of Lon-
don. Series B: Biological Sciences, 265(1410), 2087–2093. https://doi.org/10.1098/rspb.1998.0544
Becker, C. A., Flaisch, T., Renner, B., & Schupp, H. T. (2016). Neural correlates of the perception of spoiled food
stimuli. Frontiers in Human Neuroscience, 10, 302. https://doi.org/10.3389/fnhum.2016.00302
Billingsley, J., Lieberman, D.,  & Tybur, J. M. (2018). Sexual Disgust Trumps Pathogen Disgust in Predicting Voter
Behavior During the 2016 U.S. Presidential Election. Evolutionary Psychology, 16(2). https://doi.org/10.1177/
1474704918764170
Blascovich, J., Mendes, W. B., Hunter, S. B., Lickel, B.,  & Kowai-Bell, N. (2001). Perceiver threat in social inter-
actions with stigmatized others. Journal of Personality and Social Psychology, 80(2), 253–267. https://doi.
org/10.1037/0022-3514.80.2.253
Bloom, P. (2004). Descartes’ baby: How the science of child development explains what makes us human. Basic Books.
Borg, C., de Jong, P. J., & Georgiadis, J. R. (2014). Subcortical BOLD responses during visual sexual stimulation vary as
a function of implicit porn associations in women. Social Cognitive and Affective Neuroscience, 9(2), 158–166. https://
doi.org/10.1093/scan/nss117
Borg, C., de Jong, P. J., Renken, R. J., & Georgiadis, J. R. (2013). Disgust trait modulates frontal-posterior coupling as
a function of disgust domain. Social Cognitive and Affective Neuroscience, 8(3), 351–358. https://doi.org/10.1093/scan/
nss006
Borg, J. S., Lieberman, D., & Kiehl, K. A. (2008). Infection, incest, and iniquity: Investigating the neural correlates of
disgust and morality. Journal of Cognitive Neuroscience, 20(9), 1529–1546. https://doi.org/10.1162/jocn.2008.20109
Borison, H. L. (1989). Area postrema: Chemoreceptor circumventricular organ of the medulla oblongata. Progress in
Neurobiology, 32(5), 351–390. https://doi.org/10.1016/0301-0082(89)90028-2
Boulet, N. P., Cloutier, C. J., Ossenkopp, K. P.,  & Kavaliers, M. (2016). Oxytocin, social factors, and the expres-
sion of conditioned disgust (anticipatory nausea) in male rats. Behavioural Pharmacology, 27(8), 718–725. https://doi.
org/10.1097/fbp.0000000000000271
Buss, D. M., & Schmitt, D. P. (1993). Sexual strategies theory: An evolutionary perspective on human mating. Psychologi-
cal Review, 100(2), 204–232. https://doi.org/10.1037/0033-295X.100.2.204
Case, T. I., Repacholi, B. M., & Stevenson, R. J. (2006). My baby doesn’t smell as bad as yours: The plasticity of dis-
gust. Evolution and Human Behavior, 27(5), 357–365. https://doi.org/10.1016/j.evolhumbehav.2006.03.003
Chapman, H. A., & Anderson, A. K. (2014). Trait physical disgust is related to moral judgments outside of the purity
domain. Emotion, 14(2), 341–348. https://doi.org/10.1037/a0035120

81
Klebl, Donner and Bishnoi

Chapman, H. A., Kim, D. A., Susskind, J. M., & Anderson, A. K. (2009). In bad taste: Evidence for the oral origins of
moral disgust. Science, 323(5918), 1222–1226. https://doi.org/10.1126/science.1165565
Ciaramelli, E., Sperotto, R. G., Mattioli, F., & di Pellegrino, G. (2013). Damage to the ventromedial prefrontal cortex
reduces interpersonal disgust. Social Cognitive and Affective Neuroscience, 8(2), 171–180. https://doi.org/10.1093/scan/
nss087
Coburn, A., Kardan, O., Kotabe, H., Steinberg, J., Hout, M. C., Robbins, A., MacDonald, J., Hayn-Leichsenring, G., &
Berman, M. G., . . . & Berman. (2019). Psychological responses to natural patterns in architecture. Journal of Environ-
mental Psychology, 62, 133–145. https://doi.org/10.1016/j.jenvp.2019.02.007
Confer, J. C., Easton, J. A., Fleischman, D. S., Goetz, C. D., Lewis, D. M., Perilloux, C.,  & Buss, D. M. (2010).
Evolutionary psychology: Controversies, questions, prospects, and limitations. American Psychologist, 65(2), 110–126.
https://doi.org/10.1037/a0018413
Cotter, K. N., Silvia, P. J., Bertamini, M., Palumbo, L.,  & Vartanian, O. (2017). Curve appeal: Exploring individ-
ual differences in preference for curved versus angular objects.  I-Perception,  8(2), 2041669517693023. https://doi.
org/10.1177/2041669517693023
Critchley, H. D., & Rolls, E. T. (1996). Hunger and satiety modify the responses of olfactory and visual neurons in the pri-
mate orbitofrontal cortex. Journal of Neurophysiology, 75(4), 1673–1686. https://doi.org/10.1152/jn.1996.75.4.1673
Croley, J. A., Reese, V., & Wagner, R. F. (2017). Dermatologic features of classic movie villains: The face of evil. JAMA
Dermatology, 153(6), 559–564. https://doi.org/10.1001/jamadermatol.2016.5979
Curtis, V., De Barra, M., & Aunger, R. (2011). Disgust as an adaptive system for disease avoidance behaviour. Philosophi-
cal Transactions of the Royal Society of London. Series B, Biological Sciences, 366(1563), 389–401. https://doi.org/10.1098/
rstb.2010.0117
Deaner, R. O., Khera, A. V., & Platt, M. L. (2005). Monkeys pay per view: Adaptive valuation of social images by rhesus
macaques. Current Biology, 15(6), 543–548. https://doi.org/10.1016/j.cub.2005.01.044
DeBruine, L. M., Jones, B. C., Tybur, J. M., Lieberman, D., & Griskevicius, V. (2010). Women’s preferences for mas-
culinity in male faces are predicted by pathogen disgust, but not by moral or sexual disgust. Evolution and Human
Behavior, 31(1), 69–74. https://doi.org/10.1016/j.evolhumbehav.2009.09.003
Dempsey, E. W. (1973). Neural and vascular ultrastructure of the area postrema in the rat. Journal of Comparative Neurology,
150(2), 177–199. https://doi.org/10.1002/cne.901500206
Dion, K., Berscheid, E., & Walster, E. (1972). What is beautiful is good. Journal of Personality and Social Psychology, 24(3),
285–290. https://doi.org/10.1037/h0033731
Dolensek, N., Gehrlach, D. A., Klein, A. S., & Gogolla, N. (2020). Facial expressions of emotion states and their neuronal
correlates in mice. Science, 368(6486), 89–94. https://doi.org/10.1126/science.aaz9468
Donner, M. R., Azaad, S., Warren, G. A., & Laham, S. M. (2020). Specificity versus generality: A meta-analytic review
of the association between trait disgust sensitivity and moral judgment. Manuscript submitted for publication.
Dutton, D. (2009). The art instinct: Beauty, pleasure, & human evolution. Oxford University Press.
Eckel, L. A., & Ossenkopp, K. P. (1993). Novel diet consumption and body weight gain are reduced in rats chroni-
cally infused with lithium chloride: Mediation by the chemosensitive area postrema. Brain Research Bulletin, 31(5),
613–619. https://doi.org/10.1016/0361-9230(93)90130-4
Efran, M. G. (1974). The effect of physical appearance on the judgment of guilt, interpersonal attraction, and sever-
ity of recommended punishment in a simulated jury task. Journal of Research in Personality, 8(1), 45–54. https://doi.
org/10.1016/0092-6566(74)90044-0
Elliott, R., Friston, K. J., & Dolan, R. J. (2000). Dissociable neural responses in human reward systems. Journal of Neuro-
science, 20(16), 6159–6165. https://doi.org/10.1523/JNEUROSCI.20-16-06159.2000
Eskine, K. J., Kacinik, N. A., & Prinz, J. J. (2011). A bad taste in the mouth: Gustatory disgust influences moral judg-
ment. Psychological Science, 22(3), 295–299. https://doi.org/10.1177/0956797611398497
Eskine, K. J., Kacinik, N. A., & Webster, G. D. (2012). The bitter truth about morality: Virtue, not vice, makes a bland
beverage taste nice. PLOS ONE, 7(7), e41159. https://doi.org/10.1371/journal.pone.0041159
Faulkner, J., Schaller, M., Park, J. H.,  & Duncan, L. A. (2004). Evolved disease-avoidance mechanisms and
contemporary xenophobic attitudes.  Group Processes and Intergroup Relations,  7(4), 333–353. https://doi.
org/10.1177/1368430204046142
Faust, N. T., Chatterjee, A., & Christopoulos, G. I. (2019). Beauty in the eyes and the hand of the beholder: Eye and
hand movement’s differential responses to facial attractiveness. Journal of Experimental Social Psychology, 85, 103884.
https://doi.org/10.1016/j.jesp.2019.103884
Fingerhut, J., & Prinz, J. J. (2018). Wonder, appreciation, and the value of art. In J. F. Christensen & A. Gomila (Eds.),
Progress in brain research, 237. https://doi.org/10.1016/bs.pbr.2018.03.004. Elsevier.

82
Disliking

Fleischman, D. S. (2014). Women’s disgust adaptations. In T. K. Shackelford & V. A. Weekes-Shackelford (Eds.), Evolu-


tionary perspectives on human sexual psychology and behaviour. Springer.
Francis, S., Rolls, E. T., Bowtell, R., McGlone, F., O’Doherty, J., Browning, A., Clare, S., & Smith, E., . . . Smith.
(1999). The representation of pleasant touch in the brain and its relationship with taste and olfactory areas. NeuroRe-
port, 10(3), 453–459. https://doi.org/10.1097/00001756-199902250-00003
Fumagalli, M., & Priori, A. (2012). Functional and clinical neuroanatomy of morality. Brain, 135(7), 2006–2021. https://
doi.org/10.1093/brain/awr334
Gallagher, M.,  & Chiba, A. A. (1996). The amygdala and emotion. Current Opinion in Neurobiology, 6(2), 221–227.
https://doi.org/10.1016/S0959-4388(96)80076-6
Gangopadhyay, P., Chawla, M., Dal Monte, O., & Chang, S. W. C. (2021). Prefrontal-amygdala circuits in social decision-
making. Nature Neuroscience, 24(1), 5–18. https://doi.org/10.1038/s41593-020-00738-9
Garve, R., Garve, M., Türp, J. C., Fobil, J. N., & Meyer, C. G. (2017). Scarification in sub-Saharan Africa: Social skin,
remedy and medical import.  Tropical Medicine and International Health,  22(6), 708–715. https://doi.org/10.1111/
tmi.12878
Giner-Sorolla, R., Kupfer, T., & Sabo, J. (2018). What makes moral disgust special? An integrative functional review.
In. Advances in experimental social psychology (Vol. 57, pp.  223–289). Academic Press. https://doi.org/10.1016/
bs.aesp.2017.10.001
Goffman, E. (1963). Stigma: Notes on the management of spoiled identity. Prentice Hall.
Goodson, J. L. (2013). Deconstructing sociality, social evolution and relevant nonapeptide functions. Psychoneuroendocri-
nology, 38(4), 465–478. https://doi.org/10.1016/j.psyneuen.2012.12.005
Graham, J., Haidt, J., Koleva, S., Motyl, M., Iyer, R., Wojcik, S. P., & Ditto, P. H. (2013). Moral foundations theory:
The pragmatic validity of moral pluralism. In Advances in Experimental Social Psychology, 47, 55–130. https://doi.
org/10.1016/B978-0-12-407236-7.00002-4
Graham, J., Nosek, B. A., Haidt, J., Iyer, R., Koleva, S., & Ditto, P. H. (2011). Mapping the moral domain. Journal
of Personality and Social Psychology, 101(2), 366–385. https://psycnet.apa.org/doi/10.1037/a0021847. https://doi.
org/10.1037/a0021847
Grammer, K., Fink, B., Møller, A. P.,  & Thornhill, R. (2003). Darwinian aesthetics: Sexual selection and the biol-
ogy of beauty. Biological Reviews of the Cambridge Philosophical Society, 78(3), 385–407. https://doi.org/10.1017/
s1464793102006085
Haidt, J. (2001). The emotional dog and its rational tail: A  social intuitionist approach to moral judgment. Psycho-
logical Review, 108(4), 814–834. https://psycnet.apa.org/doi/10.1037/0033-295X.108.4.814. https://doi.
org/10.1037/0033-295x.108.4.814
Halberstadt, J., & Rhodes, G. (2000). The attractiveness of nonface averages: Implications for an evolutionary expla-
nation of the attractiveness of average faces.  Psychological Science,  11(4), 285–289. https://doi.org/10.1111
%2F/1467-9280.00257
Harris, L. T.,  & Fiske, S. T. (2006). Dehumanizing the lowest of the low: Neuroimaging responses to extreme out-
groups. Psychological Science, 17(10), 847–853. https://doi.org/10.1111/j.1467-9280.2006.01793.x
Harris, L. T., & Fiske, S. T. (2007). Social groups that elicit disgust are differentially processed in mPFC. Social Cognitive
and Affective Neuroscience, 2(1), 45–51. https://doi.org/10.1093/scan/nsl037
Haselton, M. G., & Buss, D. M. (2000). Error management theory: A new perspective on biases in cross-sex mind read-
ing. Journal of Personality and Social Psychology, 78(1), 81–91. https://doi.org/10.1037//0022-3514.78.1.81
Horberg, E. J., Oveis, C., Keltner, D., & Cohen, A. B. (2009). Disgust and the moralization of purity. Journal of Personal-
ity and Social Psychology, 97(6), 963–976. https://psycnet.apa.org/doi/10.1037/a0017423. https://doi.org/10.1037/
a0017423
Hosoda, M., Stone-Romero, E. F., & Coats, G. (2003). The effects of physical attractiveness on job-related outcomes:
A meta-analysis of experimental studies. Personnel Psychology, 56(2), 431–462. https://doi.org/10.1111/j.1744-6570.
2003.tb00157.x
Hutcherson, C. A., & Gross, J. J. (2011). The moral emotions: A social-functionalist account of anger, disgust, and con-
tempt. Journal of Personality and Social Psychology, 100(4), 719–737. https://doi.org/10.1037/a0022408
Inbar, Y., Pizarro, D. A., Iyer, R., & Haidt, J. (2012). Disgust sensitivity, political conservatism, and voting. Social Psycho-
logical and Personality Science, 3(5), 537–544. https://doi.org/10.1177%2F/1948550611429024
Inbar, Y., Pizarro, D. A., Knobe, J., & Bloom, P. (2009). Disgust sensitivity predicts intuitive disapproval of gays. Emo-
tion, 9(3), 435–439. https://doi.org/10.1037/a0015960
Ishizu, T., & Zeki, S. (2011). Toward a brain-based theory of beauty. PLOS ONE, 6(7), e21852. https://doi.org/10.1371/
journal.pone.0021852

83
Klebl, Donner and Bishnoi

Janeway, C. A., Travers, P., Walport, M., & Shlomchik, M. J., Jr. (2001). Immunobiology (5th ed). Garland Science.
Johnson, Z. V., Walum, H., Xiao, Y., Riefkohl, P. C.,  & Young, L. J. (2017). Oxytocin receptors modulate a social
salience neural network in male prairie voles. Hormones and Behaviour, 87, 16–24. https://doi.org/10.1016/j.
yhbeh.2016.10.009
Jones, A., & Fitness, J. (2008). Moral hypervigilance: The influence of disgust sensitivity in the moral domain. Emotion,
8(5), 613–627. https://doi.org/10.1037/a0013435
Jones, B. C., Feinberg, D. R., Watkins, C. D., Fincher, C. L., Little, A. C., & DeBruine, L. M. (2013). Pathogen disgust
predicts women’s preferences for masculinity in men’s voices, faces, and bodies. Behavioural Ecology, 24(2), 373–379.
https://doi.org/10.1093/beheco/ars173
Juniper, A. (2011). Wabi sabi: The Japanese art of impermanence. Tuttle Publishing.
Kandel, E. R. (2012). The age of insight: The quest to understand the unconscious in art, mind and Brain. Random
House.
Kapp-Simon, K. A., & McGuire, D. E. (1997). Observed social interaction patterns in adolescents with and without
craniofacial conditions. Cleft Palate-Craniofacial Journal, 34(5), 380–384. https://doi.org/10.1597%2F/1545-1569_
1997_034_0380_osipia_2.3.co_2
Karinen, A. K., Molho, C., Kupfer, T. R., & Tybur, J. M. (2019). Disgust sensitivity and opposition to immigration:
Does contact avoidance or resistance to foreign norms explain the relationship? Journal of Experimental Social Psychol-
ogy, 84, 103817. https://doi.org/10.1016/j.jesp.2019.103817
Kavaliers, M., Bishnoi, I. R., Ossenkopp, K. P.,  & Choleris, E. (2021). Differential effects of progesterone on social
recognition and the avoidance of pathogen threat by female mice. Hormones and Behavior, 127, 104873. https://doi.
org/10.1016/j.yhbeh.2020.104873
Kavaliers, M.,  & Choleris, E. (2018). The role of social cognition in parasite and pathogen avoidance. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 373(1751), 20170206. https://doi.org/10.1098/
rstb.2017.0206
Kavaliers, M., Choleris, E., Agmo, A., Braun, W. J., Colwell, D. D., Muglia, L. J., Ogawa, S., & Pfaff, D. W., . . . Pfaff.
(2006). Inadvertent social information and the avoidance of parasitized male mice: A role for oxytocin. Proceedings
of the National Academy of Sciences of the United States of America, 103(11), 4293–4298. https://doi.org/10.1073/
pnas.0600410103
Kavaliers, M., Choleris, E., Ågmo, A., Muglia, L. J., Ogawa, S., & Pfaff, D. W. (2005). Involvement of the oxytocin gene
in the recognition and avoidance of parasitized males by female mice. Animal Behaviour, 70(3), 693–702. https://doi.
org/10.1016/j.anbehav.2004.12.016
Kavaliers, M., Choleris, E., & Colwell, D. D. (2001). Brief exposure to female odors “emboldens” male mice by reduc-
ing predator-induced behavioural and hormonal responses. Hormones and Behavior, 40(4), 497–509. https://doi.
org/10.1006/hbeh.2001.1714
Kavaliers, M., Colwell, D. D., & Choleris, E. (2005). Kinship, familiarity and social status modulate social learning about
“micropredators” (biting flies) in deer mice. Behavioral Ecology and Sociobiology, 58(1), 60–71. http://www.jstor.org/
stable/25063586. https://doi.org/10.1007/s00265-004-0896-0
Kavaliers, M., Colwell, D. D., Choleris, E., Agmo, A., Muglia, L. J., Ogawa, S., & Pfaff, D. W. (2003). Impaired discrimi-
nation of and aversion to parasitized male odors by female oxytocin knockout mice. Genes, Brain, and Behavior, 2(4),
220–230. https://doi.org/10.1034/j.1601-183x.2003.00021.x
Kavaliers, M., Colwell, D. D., Cloutier, C. J., Ossenkopp, K.-P., & Choleris, E. (2014). Pathogen threat and unfamiliar
males rapidly bias the social responses of female mice. Animal Behaviour, 97, 105–111. https://doi.org/10.1016/j.
anbehav.2014.09.006
Kavaliers, M., Ossenkopp, K. P., & Choleris, E. (2019). Social neuroscience of disgust. Genes, Brain, and Behavior, 18(1),
e12508. https://doi.org/10.1111/gbb.12508
Kayyal, H., Yiannakas, A., Kolatt Chandran, S., Khamaisy, M., Sharma, V., & Rosenblum, K. (2019). Activity of insula to
basolateral amygdala projecting neurons is necessary and sufficient for taste valence representation. Journal of Neurosci-
ence, 39(47), 9369–9382. https://doi.org/10.1523/JNEUROSCI.0752-19.2019
Kiefer, S. W., & Orr, M. R. (1992). Taste avoidance, but not aversion, learning in rats lacking gustatory cortex. Behav-
ioural Neuroscience, 106(1), 140–146. https://doi.org/10.1037/0735-7044.106.1.140
Kivikangas, J. M., Fernández-Castilla, B., Järvelä, S., Ravaja, N.,  & Lönnqvist, J. E. (2020). Moral foundations and
political orientation: Systematic review and meta-analysis. Psychological Bulletin [Advance online publication], 147(1),
55–94. https://doi.org/10.1037/bul0000308
Klebl, C., Greenaway, K. H., Rhee, J. J. S., & Bastian, B. (2021). Ugliness judgments alert us to cues of pathogen pres-
ence. Social Psychological and Personality Science, 12(5), 617–628. https://doi.org/10.1177/1948550620931655

84
Disliking

Koenigs, M., Young, L., Adolphs, R., Tranel, D., Cushman, F., Hauser, M.,  & Damasio, A. (2007). Damage to the
prefrontal cortex increases utilitarian moral judgements. Nature, 446(7138), 908–911. https://doi.org/10.1038/
nature05631
Krendl, A. C., Macrae, C. N., Kelley, W. M., Fugelsang, J. A., & Heatherton, T. F. (2006). The good, the bad, and the
ugly: An fMRI investigation of the functional anatomic correlates of stigma. Social Neuroscience, 1(1), 5–15. https://
doi.org/10.1080/17470910600670579
Krolak-Salmon, P., Hénaff, M. A., Isnard, J., Tallon-Baudry, C., Guénot, M., Vighetto, A., Bertrand, O., & Mauguière,
F., . . . Mauguière. (2003). An attention modulated response to disgust in human ventral anterior insula. Annals of
Neurology, 53(4), 446–453. https://doi.org/10.1002/ana.10502
Kupfer, T. R., & Giner-Sorolla, R. (2016). Communicating moral motives: The social signalling function of disgust.
Social Psychological and Personality Science, 8, 632–640. https://doi.org/10.1177%2F1948550616679236
Kurzban, R., Dukes, A., & Weeden, J. (2010). Sex, drugs and moral goals: Reproductive strategies and views about
recreational drugs. Proceedings of the Royal Society Series B: Biological Sciences, 277(1699), 3501–3508. https://doi.
org/10.1098/rspb.2010.0608
Kurzban, R., & Leary, M. R. (2001). Evolutionary origins of stigmatization: The functions of social exclusion. Psycho-
logical Bulletin, 127(2), 187–208. https://doi.org/10.1037/0033-2909.127.2.187
LaBar, K. S. (2016). Fear and anxiety. In L. F. Barrett, M. Lewis & J. M. Haviland-Jones (Eds.), Handbook of emotions
(pp. 751–773). Guildford Press.
Landy, J. F., & Goodwin, G. P. (2015). Does incidental disgust amplify moral judgment? A meta-analytic review of exper-
imental evidence. Perspectives on Psychological Science, 10(4), 518–536. https://doi.org/10.1177/1745691615583128
Lieberman, D., Tooby, J., & Cosmides, L. (2007). The architecture of human kin detection. Nature, 445(7129), 727–731.
https://doi.org/10.1038/nature05510
Limebeer, C. L., Rock, E. M., Sharkey, K. A., & Parker, L. A. (2018). Nausea-induced 5-HT release in the interoceptive
insular cortex and regulation by monoacylglycerol lipase (MAGL) inhibition and cannabidiol. Neuro, 5(4). https://
doi.org/10.1523/ENEURO.0256-18.2018
Liu, Y., Lin, W., Xu, P., Zhang, D., & Luo, Y. (2015). Neural basis of disgust perception in racial prejudice. Human Brain
Mapping, 36(12), 5275–5286. https://doi.org/10.1002/hbm.23010
Maren, S. (1999). Long-term potentiation in the amygdala: A mechanism for emotional learning and memory. Trends in
Neurosciences, 22(12), 561–567. https://doi.org/10.1016/S0166-2236(99)01465-4
Marlin, B. J., & Froemke, R. C. (2017). Oxytocin modulation of neural circuits for social behaviour. Developmental Neu-
robiology, 77(2), 169–189. https://doi.org/10.1002/dneu.22452
Martín-Loeches, M., Hernández-Tamames, J. A., Martín, A., & Urrutia, M. (2014). Beauty and ugliness in the bod-
ies and faces of others: An fMRI study of person esthetic judgement. Neuroscience, 277, 486–497. https://doi.
org/10.1016/j.neuroscience.2014.07.040
Miller, A. D., & Leslie, R. A. (1994). The area postrema and vomiting. Frontiers in Neuroendocrinology, 15(4), 301–320.
https://doi.org/10.1006/frne.1994.1012
Mitchell, J. P., Macrae, C. N., & Banaji, M. R. (2006). Dissociable medial prefrontal contributions to judgments of simi-
lar and dissimilar others. Neuron, 50(4), 655–663. https://doi.org/10.1016/j.neuron.2006.03.040
Mitre, M., Marlin, B. J., Schiavo, J. K., Morina, E., Norden, S. E., Hackett, T. A., Aoki, C. J., Chao, M. V., & Froemke,
R. C. (2016). A distributed network for social cognition enriched for oxytocin receptors. Journal of Neuroscience, 36(8),
2517–2535. https://doi.org/10.1523/JNEUROSCI.2409-15.2016
Nabi, R. L. (2002). The theoretical versus the lay meaning of disgust: Implications for emotion research. Cognition and
Emotion, 16(5), 695–703. https://doi.org/10.1080/02699930143000437
Navarrete, C. D.,  & Fessler, D. M. T. (2006). Disease avoidance and ethnocentrism: The effects of disease vulner-
ability and disgust sensitivity on intergroup attitudes. Evolution and Human Behavior, 27(4), 270–282. https://doi.
org/10.1016/j.evolhumbehav.2005.12.001
Navarrete, C. D., Fessler, D. M. T., & Eng, S. J. (2007). Elevated ethnocentrism in the first trimester of pregnancy. Evolu-
tion and Human Behavior, 28(1), 60–65. https://doi.org/10.1016/j.evolhumbehav.2006.06.002
Newman, S. W. (1999). The medial extended amygdala in male reproductive behavior. A node in the mammalian social
behavior network. Annals of the New York Academy of Sciences, 877, 242–257. https://doi.org/10.1111/j.1749-6632.1999.
tb09271.x
Nowak, M. A., & Sigmund, K. (2005). Evolution of indirect reciprocity. Nature, 437(7063), 1291–1298. https://doi.
org/10.1038/nature04131
Oaten, M., Stevenson, R. J., & Case, T. I. (2009). Disgust as a disease-avoidance mechanism. Psychological Bulletin, 135(2),
303–321. https://doi.org/10.1037/a0014823

85
Klebl, Donner and Bishnoi

O’Connell, L. A., & Hofmann, H. A. (2011). The Vertebrate mesolimbic reward system and social behaviour network:
A comparative synthesis. Journal of Comparative Neurology, 519(18), 3599–3639. https://doi.org/10.1002/cne.22735
O’Doherty, J., Kringelbach, M. L., Rolls, E. T., Hornak, J., & Andrews, C. (2001). Abstract reward and punishment
representations in the human orbitofrontal cortex. Nature Neuroscience, 4(1), 95–102. https://doi.org/10.1038/82959
O’Doherty, J., Winston, J., Critchley, H., Perrett, D., Burt, D. M., & Dolan, R. J. (2003). Beauty in a smile: The role
of medial orbitofrontal cortex in facial attractiveness. Neuropsychologia, 41(2), 147–155. https://doi.org/10.1016/
S0028-3932(02)00145-8
Olatunji, B. O., Adams, T., Ciesielski, B., David, B., Sarawgi, S., & Broman-Fulks, J. (2012). The three domains of dis-
gust scale: Factor structure, psychometric properties, and conceptual limitations. Assessment, 19(2), 205–225. https://
doi.org/10.1177%2F/1073191111432881
Olatunji, B. O., Williams, N. L., Tolin, D. F., Abramowitz, J. S., Sawchuk, C. N., Lohr, J. M., & Elwood, L. S. (2007).
The disgust scale: Item analysis, factor structure, and suggestions for refinement. Psychological Assessment, 19(3), 281–
297. https://doi.org/10.1037/1040-3590.19.3.281
Ossenkopp, K. P.,  & Eckel, L. A. (1995). Toxin-induced conditioned changes in taste reactivity and the role
of the chemosensitive area postrema. Neuroscience and Biobehavioral Reviews,  19(1), 99–108. https://doi.
org/10.1016/0149-7634(94)00024-U
Ossenkopp, K. P., Hirst, M., & Rapley, W. A. (1994). Deoxynivalenol (vomitoxin)-induced conditioned taste aversions
in rats are mediated by the chemosensitive area postrema. Pharmacology, Biochemistry, and Behaviour, 47(2), 363–367.
https://doi.org/10.1016/0091-3057(94)90024-8
Park, J. H., Faulkner, J., & Schaller, M. (2003). Evolved disease-avoidance processes and contemporary antisocial behav-
iour: Prejudicial attitudes and avoidance of people with physical disabilities.  Journal of Nonverbal Behavior,  27(2),
65–87. https://doi.org/10.1177/1368430204046142
Park, J. H., Van Leeuwen, F., & Chochorelou, Y. (2013). Disease-avoidance processes and stigmatization: Cues of sub-
standard health arouse heightened discomfort with physical contact. Journal of Social Psychology, 153(2), 212–228.
https://doi.org/10.1080/00224545.2012.721812
Penfield, W., & Faulk, M. E. (1955). The insula: Further observations on its function. Brain: a Journal of Neurology, 78(4),
445–470. https://doi.org/10.1093/brain/78.4.445
Pizarro, D. A., Inbar, Y., & Helion, C. (2011). On disgust and moral judgment. Emotion Review, 3(3), 267–268. https://
doi.org/10.1177%2F/1754073911402394
Poirotte, C., & Kappeler, P. M. (2019). Hygienic personalities in wild grey mouse lemurs vary adaptively with sex. Pro-
ceedings of the Royal Society Series B: Biological Sciences, 286(1908), 20190863. https://doi.org/10.1098/rspb.2019.0863
Regenbogen, C., Axelsson, J., Lasselin, J., Porada, D. K., Sundelin, T., Peter, M. G., Lekander, M., Lundström, J. N., &
Olsson, M. J. (2017). Behavioural and neural correlates to multisensory detection of sick humans. Proceedings of the
National Academy of Sciences, 114(24), 6400–6405. https://doi.org/10.1073/pnas.1617357114
Rolls, E. T., & Baylis, L. L. (1994). Gustatory, olfactory, and visual convergence within the primate orbitofrontal cortex.
Journal of Neuroscience, 14(9), 5437–5452. https://doi.org/10.1523/jneurosci.14-09-05437.1994
Royzman, E., & Kurzban, R. (2011). Minding the metaphor: The elusive character of moral disgust. Emotion Review,
3(3), 269–271. https://doi.org/10.1177/1754073911402371
Rozin, P., & Fallon, A. E. (1987). A perspective on disgust. Psychological Review, 94(1), 23–41. https://psycnet.apa.org/
doi/10.1037/0033-295X.94.1.23. https://doi.org/10.1037/0033-295X.94.1.23
Rozin, P., Haidt, J., & McCauley, C. R. (2008). Disgust. In M. Lewis, J. M. Haviland-Jones & L. F. Barrett (Eds.), Hand-
book of emotions (3rd ed., pp. 757–776). Guilford Press.
Rozin, P., Lowery, L., & Ebert, R. (1994). Varieties of disgust faces and the structure of disgust. Journal of Personality and
Social Psychology, 66(5), 870–881. https://doi.org/10.1037/0022-3514.66.5.870
Rozin, P., Lowery, L., Imada, S.,  & Haidt, J. (1999). The CAD triad hypothesis: A  mapping between three moral
emotions (contempt, anger, disgust) and three moral codes (community, autonomy, divinity). Journal of Personal-
ity and Social Psychology, 76(4), 574–586. https://psycnet.apa.org/doi/10.1037/0022-3514.76.4.574. https://doi.
org/10.1037//0022-3514.76.4.574
Rudebeck, P. H., Buckley, M. J., Walton, M. E., & Rushworth, M. F. S. (2006). A role for the macaque anterior cingu-
late gyrus in social valuation. Science, 313(5791), 1310–1312. https://doi.org/10.1126/science.1128197
Rumsey, N., Bull, R., & Gahagan, D. (1982). The effect of facial disfigurement on the proxemic behavior of the general
public. Journal of Applied Social Psychology, 12(2), 137–150. https://doi.org/10.1111/j.1559-1816.1982.tb00855.x
Ryan, S., Oaten, M., Stevenson, R. J., & Case, T. I. (2012). Facial disfigurement is treated like an infectious disease.
Evolution and Human Behaviour, 33(6), 639–646. https://doi.org/10.1016/j.evolhumbehav.2012.04.001
Schaller, M., & Park, J. H. (2011). The behavioural immune system (and why it matters). Current Directions in Psychological
Science, 20(2), 99–103. https://doi.org/10.1177%2F/0963721411402596

86
Disliking

Schnall, S., Haidt, J., Clore, G. L., & Jordan, A. H. (2008). Disgust as embodied moral judgment. Personality and Social
Psychology Bulletin, 34(8), 1096–1109. https://doi.org/10.1177%2F/0146167208317771
Sevinc, G., & Spreng, R. N. (2014). Contextual and perceptual brain processes underlying moral cognition: A quanti-
tative meta-analysis of moral reasoning and moral emotions. PLOS ONE, 9(2), e87427. https://doi.org/10.1371/
journal.pone.0087427
Smith, K. B., Oxley, D., Hibbing, M. V., Alford, J. R.,  & Hibbing, J. R. (2011). Disgust sensitivity and the neuro-
physiology of left–right political orientations. PLOS ONE, 6(10), e25552. https://doi.org/10.1371/journal.pone.00
25552
Stark, R., Schienle, A., Girod, C., Walter, B., Kirsch, P., Blecker, C., Ott, U., Schäfer, A., Sammer, G., Zimmermann,
M., & Vaitl, D. (2005). Erotic and disgust-inducing pictures—Differences in the hemodynamic responses of the brain.
Biological Psychology, 70(1), 19–29. https://doi.org/10.1016/j.biopsycho.2004.11.014
Stark, R., Walter, B., Schienle, A.,  & Vaitl, D. (2005). Psychophysiological correlates of disgust and disgust sensitiv-
ity. Journal of Psychophysiology, 19(1), 50–60. https://doi.org/10.1027/0269-8803.19.1.50
Steinman, M. Q., Duque-Wilckens, N., & Trainor, B. C. (2019). Complementary neural circuits for divergent effects
of oxytocin: Social approach versus social anxiety. Biological Psychiatry, 85(10), 792–801. https://doi.org/10.1016/j.
biopsych.2018.10.008
Stevenson, R. J., Case, T. I., Oaten, M. J., Stafford, L., & Saluja, S. (2019). A Proximal Perspective on Disgust. Emotion
Review, 11(3), 209–225. https://doi.org/10.1177/1754073919853355
Symons, D. (1979). The evolution of human sexuality. Oxford University Press. https://doi.org/10.1016/0169-5347(96)81051-2
Terrizzi, J. A., Shook, N. J., & McDaniel, M. A. (2013). The behavioural immune system and social conservatism: A meta-
analysis. Evolution and Human Behaviour, 34(2), 99–108. https://doi.org/10.1016/j.evolhumbehav.2012.10.003
Thornhill, R.,  & Gangestad, S. W. (2006). Facial sexual dimorphism, developmental stability, and susceptibil-
ity to disease in men and women. Evolution and Human Behaviour, 27(2), 131–144. https://doi.org/10.1016/j.
evolhumbehav.2005.06.001
Trivers, R. (1972). Parental investment and sexual selection. In B. Campbell (Ed.), Sexual selection and the descent of man,
1871–1971 (pp. 136–179). Aldine-Atherton.
Tuerke, K. J., Limebeer, C. L., Fletcher, P. J., & Parker, L. A. (2012). Double dissociation between regulation of con-
ditioned disgust and taste avoidance by serotonin availability at the 5-HT(3) receptor in the posterior and anterior
insular cortex. Journal of Neuroscience, 32(40), 13709–13717. https://doi.org/10.1523/JNEUROSCI.2042-12.2012
Turton, D. (2004). Lip-plates and ‘the people who take photographs’: Uneasy encounters between Mursi and tourists in
southern Ethiopia. Anthropology Today, 20(3), 3–8. https://doi.org/10.1111/j.0268-540X.2004.00266.x
Tybur, J. M., Bryan, A. D., Lieberman, D., Caldwell Hooper, A. E., & Merriman, L. A. (2011). Sex differences and
sex similarities in disgust sensitivity. Personality and Individual Differences, 51(3), 343–348. https://doi.org/10.1016/j.
paid.2011.04.003
Tybur, J. M., Inbar, Y., Aarøe, L., Barclay, P., Barlow, F. K., de Barra, M., Becker, D. V., Borovoi, L., Choi, I., Choi,
J. A., Consedine, N. S., Conway, A., Conway, J. R., Conway, P., Adoric, V. C., Demirci, D. E., Fernández, A. M.,
Ferreira, D. C., Ishii, K., . . Žeželj, I. (2016). Parasite stress and pathogen avoidance relate to distinct dimensions of
political ideology across 30 nations. Proceedings of the National Academy of Sciences of the United States of America, 113(44),
12408–12413. https://doi.org/10.1073/pnas.1607398113
Tybur, J. M., Inbar, Y., Güler, E.,  & Molho, C. (2015). Is the relationship between pathogen avoidance and ideo-
logical conservatism explained by sexual strategies? Evolution and Human Behaviour, 36(6), 489–497. https://doi.
org/10.1016/j.evolhumbehav.2015.01.006
Tybur, J. M., & Karinen, A. K. (2018). Measurement and theory in disgust sensitivity. In V. ZeiglerHill & T. K. Shackel-
ford (Eds.), The SAGE handbook of personality and individual differences: Volume III: Applications of personality and individual
differences (pp. 159–179). https://doi.org/10.4135/9781526451248.n7. SAGE.
Tybur, J. M., Lieberman, D., Fan, L., Kupfer, T. R., & De Vries, R. E. (2020). Behavioral Immune Trade-Offs: Inter-
personal Value Relaxes Social Pathogen Avoidance. Psychological Science, 31(10), 1211–1221. https://doi.org/10.117
7%2F/0956797620960011
Tybur, J. M., Lieberman, D., & Griskevicius, V. (2009). Microbes, mating, and morality: Individual differences in three
functional domains of disgust. Journal of Personality and Social Psychology, 97(1), 103–122. https://doi.org/10.1037/
a0015474
Tybur, J. M., Lieberman, D., Kurzban, R., & DeScioli, P. (2013). Disgust: Evolved function and structure. Psychological
Review, 120(1), 65–84. https://doi.org/10.1037/a0030778
Ugazio, G., Lamm, C., & Singer, T. (2012). The role of emotions for moral judgments depends on the type of emo-
tion and moral scenario. Emotion, 12(3), 579–590. https://psycnet.apa.org/doi/10.1037/a0024611. https://doi.
org/10.1037/a0024611

87
Klebl, Donner and Bishnoi

Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Rostrup, N., Skov, M., Corradi,
G., & Nadal, M. (2019). Preference for curvilinear contour in interior architectural spaces: Evidence from experts
and nonexperts. Psychology of Aesthetics, Creativity, and the Arts, 13(1), 110–116. https://doi.org/10.1037/aca0000150
Vessel, E. A., & Rubin, N. (2010). Beauty and the beholder: Highly individual taste for abstract, but not real-world
images. Journal of Vision, 10(2), 18. https://doi.org/10.1167/10.2.18
Vicario, C. M., Rafal, R. D., Martino, D., & Avenanti, A. (2017). Core, social and moral disgust are bounded: A review
on behavioural and neural bases of repugnance in clinical disorders. Neuroscience and Biobehavioural Reviews, 80, 185–
200. https://doi.org/10.1016/j.neubiorev.2017.05.008
Weeden, J., Cohen, A. B., & Kenrick, D. T. (2008). Religious attendance as reproductive support. Evolution and Human
Behaviour, 29(5), 327–334. https://doi.org/10.1016/j.evolhumbehav.2008.03.004
Wicker, B., Keysers, C., Plailly, J., Royet, J. P., Gallese, V.,  & Rizzolatti, G. (2003). Both of us disgusted in my
insula: The common neural basis of seeing and feeling disgust. Neuron, 40(3), 655–664. https://doi.org/10.1016/
S0896-6273(03)00679-2
Wolfe, N. D., Dunavan, C. P., & Diamond, J. (2007). Origins of major human infectious diseases. Nature, 447(7142),
279–283. https://doi.org/10.1038/nature05775
Wright, P., He, G., Shapira, N. A., Goodman, W. K., & Liu, Y. (2004). Disgust and the insula: fMRI responses to pictures of
mutilation and contamination. NeuroReport, 15(15), 2347–2351. https://doi.org/10.1097/00001756-200410250-00009
Ying, X., Luo, J., Chiu, C. Y., Wu, Y., Xu, Y., & Fan, J. (2018). Functional dissociation of the posterior and anterior
insula in moral disgust. Frontiers in Psychology, 9, 860. https://doi.org/10.3389/fpsyg.2018.00860
Zhang, M., Hu, S., Xu, L., Wang, Q., Xu, X., Wei, E., Yan, L., Hu, J., Wei, N., Zhou, W., Huang, M., & Xu, Y. (2011).
Neural circuits of disgust induced by sexual stimuli in homosexual and heterosexual men: An fMRI study. European
Journal of Radiology, 80(2), 418–425. https://doi.org/10.1016/j.ejrad.2010.05.021

88
5
THE INFLUENCE OF
INTEROCEPTIVE SIGNALS ON
THE PROCESSING OF EXTERNAL
SENSORY STIMULI
Alejandro Galvez-Pol, Enric Munar and James M. Kilner

Interoception refers to the set of physiological and cognitive processes involved in sensing, interpreting, and
integrating information that arises inside the body, providing a continuous mapping of our ever-fluctuating
internal milieu across conscious and unconscious levels (Khalsa et al., 2018). It can be distinguished from
exteroception (sensation of the environment) and proprioception (sensation of the body in space). While
far more research has focused on how external stimuli are represented by the brain, research on interocep-
tion focuses on the effect of ever-fluctuating afferent bodily signals on brain processes. Importantly, these
so-called interoceptive signals not only inform the brain about the state of the body but also influence how
we relate to our environment; that is, they influence our perception of the world. In this chapter, we focus
on this latter facet, with special emphasis on the way the cardiovascular system modulates the processing of
external stimuli. First, we outline the historical roots of interoception. Second, we describe how changes
inside the body are consciously perceived and such signals influence perception of external stimuli. Third, we
describe the physiological pathway of the heart–brain axis and its impact on stimuli processing. Fourth, we
review the link between afferent bodily signals and the neural encoding of subjective values and discuss what
is known about the way interoception affects hedonic coding of sensory objects. Last, we consider current
challenges of the field and how these can be overcome.

A brief history of the study of interoceptive signals on cognition


“The stability of the organism’s internal landscape (milieu intérieur) is the condition for the free and independ-
ent life” (Bernard, 1878). Although 150 years old, these words resonate well with our current understand-
ing of the way the integrity of living organisms rests upon upholding a homeostatic equilibrium. Some
years after Bernard, William James (1884) and Carl Lange (1885) proposed that our emotional experiences
originate from responses in the body that accompany the perception of external events. In contrast, Walter
Cannon (1927) and Philip Bard (1928) proposed that stimulating neurons in the central nervous system was
sufficient to elicit feelings and physical reactions in a simultaneous manner. This debate regarding the cause
of emotions—body states vs brain activity—has continued to this day.
By the turn of the 20th century, Charles Sherrington (1906) coined the term interoceptor to describe
the presence of an internal bodily surface dedicated to the monitoring of changes within the body. By the
time Sherrington and Edgar Adrian received the Nobel Prize for their discoveries regarding the functions

DOI: 10.4324/9781003008675-6 89
Galvez-Pol, Munar and Kilner

of neurons, Cannon further elaborated Bernard’s notion of a milieu intérieur in what he termed homeostasis,
intended to describe self-regulating processes that promote survival by maintaining the stability of the organ-
ism (Cannon, 1939). During this first half of the 20th century, most studies in interoception were conducted
in the Soviet Union, with Pavlov’s work on learnt reflexes and interoceptive processes in preparation for
digestion the most famous. However, some of these works remained unnoticed until later decades, when
Soviet psychophysiology became more appreciated by international audiences.
The advent of operant conditioning and the discovery of various types of interoceptors laid the founda-
tion for our present understanding of interoception. Novel paradigms in cardioception such as the heartbeat
detection task allowed scientists to estimate how much interoceptive information reaches awareness (Schandry,
1981). Later, the development of neuroimaging techniques allowed neuroscientists to map the brain activity
correlated with these tasks (Craig, 2002), and new theories such as the “Somatic Marker Hypothesis” stressed
how subjective value computations integrate bodily representations to form decision values (Damasio, 1999,
1994). Since then, the field of interoception has grown, matured, and diversified. The number of studies
has grown exponentially and our understanding of the interaction between interoceptive mechanisms and
external perception has improved significantly. The current chapter reviews this body of work and moves
beyond the recognized role of interoception for homeostasis, providing insight into the role that afferent
interoceptive signals play in the computation of perception, valuation, and reasoning.

Conscious interoception mediates stimulus processing


Interoception is multifaceted; it comprises distinguishable dimensions and different physiological systems
(cardiovascular, gastrointestinal, hormonal, circulatory) acting across conscious and unconscious levels
(Khalsa et al., 2018; Quigley et al., 2021). Yet the primary physiological focus of most interoception studies
is the cardiovascular system. This is likely due to the emergence of early evidence for a modulatory effect of
the carotid sinus on central and autonomous nervous processes (Koch, 1932; Kreindler, 1946) and the meth-
odological ease of monitoring discrete regular events (i.e., heartbeats) that can be recorded via noninvasive
tools, such as electrocardiogram (ECG), pulse oximeters, or wearable heart rate monitors.
Most interoceptive processes, such as the monitoring of one’s psychophysiological state, unfold at the
unconscious level and often reach awareness only when the system is compromised (e.g., pain, thirst). The
study of this facet of interoception usually focuses on the effect of afferent bodily signals upon the processing
of external stimuli and brain function. At the conscious level, numerous studies have examined the ability
to notice bodily states and fluctuations, mostly in the cardioception domain. Three dimensions of conscious
interoception have dominated this research: (i) interoceptive sensibility, which refers to the subjective experi-
ence of internal sensations as measured by self-reports; (ii) interoceptive sensitivity, which refers to the objective
measure of interoception and involves testing of the capability to perceive bodily changes in behavioural
tasks; and (iii) interoceptive awareness, which refers to metacognitive insights about the two former dimensions
and is most often operationalized as the distance between one’s beliefs and the person’s actual ability to per-
ceive inner body states (see Garfinkel et al., 2015; Murphy et al., 2019). In the following, we briefly review
findings that this body of work has produced with respect to the role these dimensions play in the processing
of external stimuli.

Interoceptive sensibility
One way to examine the conscious processing of interoceptive signals is by using self-report question-
naires that allow assessing participants’ experience with bodily sensations (interoceptive sensibility), as well
as how this sensitivity is related to other cognitive domains. The choice of a self-report questionnaire
depends largely on the research question, for example, noticing vs body listening or dissociating adaptive

90
Interoception and sensory processing

vs maladaptive interoception (see the Body Perception Questionnaire [Porges, 1993], the Multidimensional
Assessment of Interoceptive Awareness [Mehling et al., 2018, 2012], or the Interoceptive Sensory Question-
naire [Fiene et al., 2018]).
Self-report measures have shown that participants with a heightened interoceptive sensibility might expe-
rience more anxiety, and this is more likely to occur when they have difficulties in identifying and describing
emotions. The lack of attributing interoceptive signals to emotional states predisposes participants to anxiety-
related characteristics (Palser et al., 2018). This in turn may intensify the attribution of negative valence to
stimuli, as well as exacerbating a negative narrative when processing external events. In this way, self-reports
allow us to quickly obtain information about the interoceptive-mediated valuation of external stimuli. Relat-
edly, Paulus et al. (2019) highlighted two scenarios in which the negative appraisal of stimuli can be medi-
ated by maladaptive interoceptive mechanisms. In the first scenario, people who are exposed to threatening
stimuli experience high levels of arousal. If the situation is positively resolved, future related events should
be perceived as less hazardous, decreasing sympathetic engagement. However, if people fail to update their
beliefs, similar scenarios will continue to elicit heightened visceral responses. In the second scenario, people
extrapolate physiological responses from the original context to other situations. Both scenarios involve
persistent non-veridical perception that negatively affects the regular function of the viscera, which eventu-
ally feeds back to the central nervous system, where it affects the assessment of external stimuli. In sum, if
one’s set point (i.e., optimal bodily state) is missattuned, the representation of interoceptive signals becomes
imprecise, and the valuation of external stimuli does not necessarily meet the most appropriate behaviour
(Linson et al., 2020; Paulus et al., 2019).

Interoceptive sensitivity
Another way to examine the conscious processing of interoceptive signals is by using lab-based tasks that
allow obtaining more objective measures. This usually involves asking participants to notice bodily changes
while their physiological rhythms are being recorded (e.g., counting heartbeats while the ECG is recorded).
Then, the difference between participants’ subjective reports and the objective quantification of their bod-
ily changes are compared against each other. The difference between these measures provides an objective
estimate of interoceptive sensitivity. This allows “profiling” participants, ranking them on a continuum from
poor to good interoceptors, and relating this individual variance to other measures and tasks. For example,
good interoceptors display larger electrodermal responses to unfair offers in the context of the ultimatum
game (Dunn et al., 2012), have better memory recall for words encoded during the systolic phase of the
cardiac cycle (i.e., hearts’ contraction pumping the blood, Garfinkel et al., 2013), exhibit greater sympathetic
reactivity during mental stress and subjective arousal during emotional picture viewing (Herbert et al., 2010),
and display higher reinforced learning of emotional faces (Pfeifer et al., 2017). Good cardioception has also
been associated with stronger expectancy for unconditioned stimuli (Zaman et al., 2016), better learning
with the corresponding modulation of hippocampal activity (Stevenson et al., 2018), and higher sensitivity
to negative affect but lower accuracy in recognizing faces depicting fear and sadness (Georgiou et al., 2018).
A recent study, combining both objective and subjective measures of interoception, showed that participants
with better cardiac interoceptive awareness and insight are able to withhold actions and respond more slowly
in a Go/NoGo task, while the opposite pattern was observed for participants with poorer interoception (Rae
et al., 2020). This suggests that precise afferent input may support sensorimotor decisions. In contrast, noisier
signalling could prompt hasty responses to external stimuli.
Taken all together, current findings suggest that cardiac sensitivity is related to greater perception and
memory encoding of emotional stimuli. However, these results should be considered with caution. More
consistent research is needed, and some studies have also shown that interoceptive signals can inhibit stimulus
processing (see, e.g., Park et al., 2014; Salomon et al., 2018). Overall, the question of the relation between

91
Galvez-Pol, Munar and Kilner

interoceptive sensitivity and everyday well-being needs further work. Furthermore, research on atypical
interoception, a common denominator in all mental disorders, suggests that either low or amplified function-
ing in interoceptive sensibility, sensitivity, or awareness could be linked to a maladaptive valuation of stimulus
processing (Khalsa et al., 2018).

The influence of afferent bodily signals on the processing of stimuli


As noted previously, the primary focus of studies on interoception has been the influence of the cardiac cycle
on cognition. To understand the influence of the heart on stimulus perception, in the following we summa-
rize what is currently known about neural signalling from the heart to the brain and how these projections
modulate cognition.

The heart–brain axis


One of the distinctive features of the heart is that it is endowed with pacemaker properties; that is, the heart
can generate its own intrinsic oscillatory electrical activity. In a single heartbeat, two phases are observed:
in the systole phase, the heart contracts and ejects the blood, whereas in the diastole phase, the heart expands
and fills. Both phases constitute a cardiac cycle, with the R-peak (peak in ECG depicting the contraction
at systole) denoting the beginning of a new cycle. During the systolic phase, pressure sensors located in the
carotid sinus, coronary arteries, and aortic arch (i.e., baroreceptors) detect changes in blood pressure due to
the ejection of the blood from the left ventricle. Baroreceptors convey information to the brain about the
strength and timing of the heartbeats during the systolic phase while being quiescent during the diastole
phase of the cardiac cycle (Critchley & Harrison, 2013). Many studies have observed that neural and behav-
ioural responses to external stimuli vary according to the phase of the cardiac cycle during which they occur
(see, e.g., Azevedo, Badoud et al., 2017; Leganes-Fonteneau et al., 2021). This is usually demonstrated by
meticulously locking the presentation of stimuli to the systolic or diastolic phase of the cardiac cycle. Many
studies have associated the variation in participants’ responses along the cardiac cycle with the phasic firing
of the baroreceptors. Although this is under investigation, it is clear that heartbeats selectively modulate the
processing of external stimuli by constantly facilitating, competing with, or inhibiting information processing
(see Figure 5.1 and the following section).
As an intrinsic oscillator, the heart has an inherent nervous system composed of interconnecting, efferent,
and afferent ganglionated nerve plexi (a branching network of intersecting nerves). These project through
the spinal cord and the vagus nerve to the nucleus of the solitary tract (NTS) and other autonomic nuclei of
the brainstem, which in turn allow for dynamic regulation of efferent1 sympathetic and parasympathetic car-
diomotor activity. Interestingly, approximately 80% of the fibres of the vagus nerve are afferent, which makes
it more of a listener than a storyteller (as revisited by Wolpert et al., 2020). In the NTS, the convergence of
signals from different bodily systems (e.g., cardiac, gastric) projects to viscero-sensitive brain regions such as
the thalamus, hypothalamus, amygdala, cingulate cortex, and insula. The insula is considered a major hub
for interoceptive information (Craig, 2009). The posterior insula receives inputs from bodily systems and
the anterior part re-represents this information with emotional, cognitive, and subjective states. Informa-
tion from the environment and interoceptive signals seem to be assimilated across this posterior-to-anterior
insular gradient (Namkung et al., 2017).

Effect of cardiac phases on stimulus perception


Many studies have examined the iterative influence of the cardiac cycle upon the perception of stimuli
by presenting these during systolic or diastolic phases (Figure 5.1C). The idea is to exploit these naturally

92
Interoception and sensory processing

Figure 5.1 Interoceptive processes and their influence on stimulus processing. (A) Bodily organs endowed with pace-
maker properties interact with the brain via afferent and efferent signals. The activity of the stomach, heart,
and brain unfolds at various frequencies. Changes in such a complex oscillatory activity affects one another
and influence cognition (Azzalini et al., 2019). (B) The brain’s responses to sensory inputs (e.g., external
stimuli) not only depend on the stimuli’s properties but also on its own internal state at the time when
the stimuli are processed. Conscious perception and awareness of one’s inner bodily states moderates the
perception of stimuli and corresponding physiological correlates (interoceptive sensibility and predictions;
Murphy et al., 2019). (C) Studies examining the influence of the cardiac cycle on cognition often measure
participants’ ECG and present stimuli during time windows phase-locked to systole or diastole. In principle,
stimulus presentation is shifted in time to account for the time needed for afferent baroreceptor activity and
other physiological changes to reach the brain (~300 ms). (D) An overall trend in studies using cardiac phase-
locking shows that motor responses are more prone in the early period of the cycle (systole) whereas sensory
processing of stimuli seems enhanced in the later quiescent physiological period (diastole). Icons in all panels
are under Creative Commons license CC0 from Pixabay.com.

occurring fluctuations to understand how subsequent variations (e.g., bodily arousal, firing of baroreceptors
at systole) affect stimulus processing at the neural and behavioural levels. These studies are usually focused
on inspecting the processing of emotional or non-emotional stimuli presented at near or suprathreshold
perceptual levels.
Depending on the task and stimuli, cardiac phases can both selectively facilitate and inhibit stimulus pro-
cessing. Systolic modulation of sensory processing has been observed for subjective pain perception and sen-
sitivity to tactile stimuli, which are attenuated during this phase compared to diastole (Wilkinson et al., 2013;
Al et al., 2020; Motyka et al., 2019; see also concurrent effects of respiration in Grund et al., 2021). Similarly,
the startle reflex, an unconscious defensive response that induces an immediate eyeblink response to sudden
or threatening stimuli such as sudden noises or sharp movements, is attenuated by systolic afferent signals
(Larra et al., 2020). Startle responses are also modulated by phase respiratory and gastric rhythms (Schulz
et al., 2017, 2016). Conversely, enhanced processing at systole has often been linked to negative emotional
stimuli such as the detection of fearful faces and memories (Garfinkel et al., 2020; Garfinkel & Critchley,
2016). Likewise, the processing of threatening stimuli associated with racial stereotypes is heightened during

93
Galvez-Pol, Munar and Kilner

systole (Azevedo, Garfinkel et  al., 2017). Specifically, participants in a first-person shooter task produced
more errors (“shooting” un-armed Black vs. White targets) during systole. These results may be associated
with a cardiac modulation of error monitoring (Bury et al., 2019), and motor inhibition might be more
prone to fail during systole (Makowski et al., 2020; but see Rae et al., 2018).
Overall, current research indicates that there is a putative effect of afferent cardiac signalling on stimulus
processing (Critchley & Garfinkel, 2018). However, results seem to vary significantly across experimental
manipulations. While more consistent research is needed, it is clear that cardiac afferent signals and concur-
rent physiological changes moderate cognitive processes by competing for the allocation of attentional and
representational resources. This in turn can reduce or amplify the sensory processing of stimuli. Sensory
processes that are concurrent with the systolic phase of the cardiac cycle (the noisy period of the cycle) seem
to be reduced, whereas motor behaviour seems to be facilitated (see the later section “Active Sensing”).

Cardiac-related effects upon the processing of face stimuli


Influential theories of emotion highlight that certain emotions are likely coupled with particular bodily
states. For instance, disgust is closely coupled with parasympathetic responses, whereas feelings of anxiety or
fear are associated with heightened cardiovascular arousal caused by sympathetic activation. In this context,
physiological arousal is often understood as a consequence of top-down processes rather than as a cause of
emotional experience. Yet several studies have shown that detection and appraisal of facial emotional expres-
sions fluctuate according to the effect of short-term baroreceptors’ activity, that is, a transient increase of
visceral arousal.
The modulatory effect of the heart upon the processing of face stimuli varies according to both the emo-
tion displayed and the task employed. In a forewarned reaction time task, facial expressions of disgust, but not
sad, happy, or neutral expressions, were judged as more intense when presented in systole. Furthermore, the
processing of disgust and happy faces resulted in a more pronounced deacceleration of subsequent heartbeats,
a mechanism proposed to facilitate perception and appraisal (Gray et al., 2012). In a rapid serial presentation
task, the detection and intensity ratings of fearful faces were found to be higher when these were presented
in systole, suggesting that heartbeats might facilitate conscious processing of briefly presented and emotion-
ally strong stimuli (Garfinkel et al., 2014). The cardiac interaction between processing of fearful faces and
cardiac phase correlated with neural activations in several brain regions with the most prominent found in
the amygdala, a structure associated with threat processing and the integration of physiological and affective
information. Interestingly, regardless of the emotion, the overall effect of the cardiac cycle on emotional pro-
cessing was found in the anterior insula. In an emotional spatial cueing task, a systolic effect on attentional
engagement to fearful faces was found at different spatial frequency ranges. The systolic phase enhanced the
processing of fearful facial expressions at low, but not broad or high, spatial frequencies (Azevedo, Badoud
et al., 2017), implying that afferent bodily signals modulate the processing of faces by distinctly influencing
the magnocellular and parvocellular pathways at the early stages of the visual processing.
More recently, Leganes-Fonteneau et al. (2021) used an emotional visual search task where participants
saw a target emotional face on the screen (fearful, happy, sad, or disgust) surrounded by five neutral dis-
tractors. An interesting point of this task is that it allows capturing attentional processes allocated to the
scanning of faces in a crowd, as well as differentiating between the correct detection of the target stimuli
(accuracy in the visual search task) and the correct identification of the emotion presented. Accuracy in the
visual search was higher for disgust and happy faces presented during systole, whereas the opposite effect
was found for fearful faces. The identification of fearful and sad faces was higher when presented in diastole.
Overall, these studies highlight that detection and appraisal of facial emotional expressions are the result of
body–brain interactions. The role of cardiac interoceptive signals goes beyond the processing of fearful faces
and depends on the core task and emotional expression. Presumably, these effects have been explained as a

94
Interoception and sensory processing

consequence of afferent baroreceptor signals conveyed through the brainstem via the vagus nerve. However,
other physiological changes are concurrent to heartbeats (Davos et al., 2002), and these may instigate the
previous effects by directly or concurrently affecting key neural substrates for face processing.

The self and the heart–brain axis support subjective preferences


Research in cognitive neuroscience and psychology have long focused on the significance of the body as the
ground of the self—the person as the object of its own reflective consciousness. More recently, the study
of the self has focused on the importance of the body from within. It has been proposed that regardless of
the bodily state, organs endowed with pacemaker properties such as the heart and the stomach could work
as constant transmitters that signal the presence of a body to the brain (Azzalini et al., 2019; Tallon-Baudry
et al., 2018). Then, the central nervous system would make use of this information to generate a first-person
perspective. Compelling evidence for this mechanism comes from observing that the amplitude of brain
responses to heartbeats (heartbeat-evoked responses; HERs) correlates with the self-relatedness of thoughts
in the ventromedial prefrontal cortex (vmPFC; Babo-Rebelo et  al., 2016; Tallon-Baudry et  al., 2018).
Relatedly, the amplitude of the HER has been linked to the conscious perception of stimuli presented at
near-threshold detection (Al et al., 2020; Park et al., 2014; Park & Blanke, 2019). This suggests that brain
responses to heartbeats might regulate the perception of near-threshold stimuli by moderating one’s self-
consciousness during the perceptual experience.
Following this line of work, it has been proposed that this relationship, between the self and the heart–
brain axis, supports the valuation of what we like or dislike. Specifically, it has been exposed that preference-
based decisions about external stimuli such as cultural goods are subjected to the self. For instance, “Do you
prefer Forrest Gump or The Matrix? Only you know which movie you like best” (Azzalini et al., 2021, p. 1).
These authors examined participants’ brain activity while they were presented with pairs of movie titles. The
participants either indicated which movie they preferred or had to discriminate between versions of the title
written with different levels of contrast. The results of the study showed that when choosing the partici-
pants’ preferred movie, HERs signalled the recruitment of self-reflective processes in vmPFC. Conversely,
this association was not observed for the contrast discrimination task. Moreover, the interaction between
HERs and subjective value encoding reflected the inter-individual variability in choice consistency and the
trial-by-trial fluctuations within participants. These results indicate that the neural monitoring of cardiac
signals and the neural encoding of subjective values are related to each other. Considering these findings,
the sensory valuation of stimuli appears to depend on the novelty of the stimuli and the use/absence of a
subjective approach, including possible re-enactments of autobiographical and bodily memories (Galvez-Pol
et al., 2020a; Riva, 2018).

The gut–brain axis


The gastrointestinal tract has received increasing attention in recent research on interoception. Similar to the
heart, the gut generates its own intrinsic oscillatory electrical activity. The gastrointestinal tract has a rhythm
that unfolds in the form of a continuous slow electrical pulse (one cycle every 20 seconds, ~0.05 Hz). This
rhythm results from the activity of interstitial cells of Cajal, pacemaker cells that generate rings of electrical
waveforms. These cells mediate between the autonomic nervous system and the muscle layers of the stom-
ach, whereby mechanical changes in smooth muscles can be detected by making contact with vagal sensory
neurons along with intramuscular arrays (Powley et al., 2016).
The gastric rhythm can be non-invasively measured via electrogastrography (EGG), which reflects a com-
bination of the gastric rhythm caused by interstitial cells of Cajal and of gastric smooth muscle contractions
(Wolpert et al., 2020). Importantly, the recording of spontaneous brain activity (non–task-related fluctuating

95
Galvez-Pol, Munar and Kilner

neural activity) has revealed that several brain regions are linked to gastric function (Rebollo et al., 2018).
In this so-called gastric network, the gastric phase seems to modulate the neural activity of viscero-sensitive
brain areas (somatosensory cortices and parieto-occipital regions). The presence of this gastric network sug-
gests that automatic regulation of basic processes for life such as digestion are linked to complex patterns of
brain activity that affect how people could perceive external stimuli. Evidence for this has been shown in
a behavioural task where participants were given a dose of domperidone (an antiemetic moderating gastric
rhythm) or a placebo while their oculomotor responses to neutral and disgust images were recorded. The
study showed that domperidone did not change the subjective ratings of disgust but decreased oculomotor
avoidance following incentivized exposure to disgusting stimuli (Nord et al., 2021). Future work is needed
to examine the link between gastrointestinal afferent signals and the brain and subsequent behaviour, as well
as to study other aspects of the gut–brain axis that are likely to mediate this link (e.g., microbiota).

Active sensing in interoception: beyond the phase-locked presentation of stimuli


Many of the findings that we have presented here were obtained by deliberately locking the presentation of
stimuli to the different phases of the cardiac cycle. However, in our everyday life, the environment does not
normally provide us with sensory information phase-locked to our physiological cycles. Instead, we freely
and actively sense the stimuli at hand. Interoception studies in active sensing examine whether participants
are naturally prone to sample the stimuli in the environment in a particular phase of the bodily cycles. These
types of studies do not impose temporal constraints and allow participants to access the stimuli at their own
pace. Simultaneously, participants’ behaviour and their physiological rhythms are coregistered (e.g., ECG,
breathing, and participants’ responses). After data collection, researchers work “backwards” in the data by
situating each of the participants access to the stimuli along with the recorded continuous rhythms. Then,
they analyse whether participants are more prone to sample the stimuli, for instance, in the systolic or dias-
tolic phase of the cardiac cycle (Figure 5.2A, B).
A significant coupling between the active sampling of stimuli and cardiovascular oscillations has been
found in various tasks. For instance, in a visual search task comparable to a “spot the difference” task, par-
ticipants’ oculomotor behaviour and ECG were recorded while they searched for differences between two
bilateral arrays continuously displayed on the screen (Figure 5.2C; Galvez-Pol et al., 2020b). Across three
different analyses, the results showed a significant coupling of saccades, subsequent fixations, and blinks
with the cardiac cycle. More eye movements to sample the arrays were generated during systole, which has
been reported as the period of the maximal firing of the baroreceptors. Conversely, more ocular fixations
were found during the diastole phase (quiescent baroreceptors). Last, more blinks were generated in the
later period of the systolic phase. These results show that (1) in a more ecological setting, interoceptive and
exteroceptive processes adjust to each other, and (2) the active sampling of external stimuli might occur when
more computational resources are available, that is, during quiescent periods of the inner body (Galvez-Pol
et al., 2020b). While this latter hypothesis needs further examination, recent research seems to support this
idea (see, e.g., Kluger et al., 2021).
Beyond ocular sensory sampling, Kunzendorf et  al. (2019) found that participants freely generated a
keypress leading to the onset of images in a memory task more often during the systolic phase, though it
did not influence memory performance. Similarly, Perl et al. (2019) reported that participants preferred to
self-initiate the onset of non-olfactory cognitive tests to coincide with the beginning of the inspiratory phase.
Also, recently it has been shown that participants tend to initiate actions during the expiration phase of the
breathing cycle. Moreover, the neural marker of these self-initiated movements seems to be modulated by
the respiratory phase (Park et al., 2020). Overall, active sensing is entrained by cardiorespiratory fluctuations,
which indicates the constant incorporation of bodily signals into one’s engagement with the stimuli in the
environment. However, the behavioural relevance of this entrainment remains unclear.

96
Interoception and sensory processing
97

Figure 5.2 Active sensing of stimuli in interoception. (A) ECG depicting one cardiac cycle. Here, two events (e.g., saccades to sample stimuli, keypresses leading
to stimuli onset, etc.) occur red at the early phase of the cardiac cycle and one at the later phase. (B) Schematic of analysis; from left to r ight, the num-
ber of events in this cycle are depicted in phase i) as deg rees of each event relative to the concur rent heartbeat, ii) as counted events binned into time
windows along the cycle, and iii) as counted events in the systolic or diastolic phases. After n tr ials, it is possible to compute the frequency of events as
a function of phase, the event changing ratio, and their phasic occur rence. (C) Active visual sampling task and results. Participants reported the number
of boxes differ ing in colouration between two bilateral ar rays by compar ing each box in the left ar ray with the homologous box in the r ight Ar ray. (D)
The succeeding results showed that saccade onset, mean time point of fixations, and blink onset occur red significantly more often in the early, mid, and
later per iod of the cardiac cycle, respectively. Adapted from Galvez-Pol et al. (2020b).
Galvez-Pol, Munar and Kilner

Future research: a more ecological approach to interoceptive processing


Many of the studies reviewed in this chapter show that perception, reasoning, and emotional experiences
vary according to changes in the body (e.g., heartbeats). Yet most of this research has been conducted by:
(1) meticulously locking the presentation of brief stimuli to distinct phases of our bodily cycles, (2) adopting an
idiosyncratic perspective: physiological bodily signals serve “one’s purposes,” or (3) somewhat overlooking that
humans might relate to stimuli in their environment according to the processing of their physiological signals.
Active sensing paradigms tackle the first limitation (see the previous section). Meanwhile, very recent
work has shown that dynamic changes in our bodies can be inferred by others (hence, unfolding a dialecti-
cal perspective; Galvez-Pol et al., 2021a). Also, recent work has shown that our sense and interpretation of
inner bodily signals correlate with how we seek and experience our surrounding environment (Galvez-Pol
et al., 2021b). We believe that these lines of research will allow for a better understanding of the physiological
mechanisms underpinning sensory processing in real-life scenarios, including work, family life, entertain-
ment, or even art experiences.

Concluding remarks
In recent years, the field of interoception has grown, matured, and expanded. Exponential growth in the
number of studies has produced a better grasp of the relationship between the perception of the outside
world and mechanisms inside the body. These studies have shown that our responses to external stimuli not
only depend on the stimuli’s properties but also on our internal bodily state at the time when the stimuli
are processed. Internal bodily changes (e.g., afferent signalling from the heart and stomach) moderate cog-
nitive processes by competing for the allocation of attentional and representational resources. This in turn
might dampen, enhance, or modify the processing of stimuli. While the field is still in development, more
consistent methods, paradigms, and integrated theories are needed. Likewise, the consideration of various
physiological systems (beyond the cardiac system, e.g., gastrointestinal, hormonal, respiratory) is a promis-
ing avenue for developing next-generation studies. Furthermore, we believe that research in interoception
should advance towards a more ecological understanding of how humans function in the real world, that
is, the ecological niche in which the brain has evolved. In this setting, including interoceptive signals in the
study of sensory acquisition in active sensing is a fundamental step towards a more ecological understanding
of exteroceptive and interoceptive processes.

Acknowledgements
This work was supported by the Autonomous Community of the Balearic Islands (CAIB). Postdoctoral
grant Margalida Comas to Alejandro Galvez-Pol, Ref PD/036/2019.

Note
1 Efferent neurons carry signals from the central nervous system (i.e., the brain) to the body’s muscles, glands, and organs.
Afferent neurons project signals from sensory receptors and the autonomous nervous system (i.e., the body) to the
central nervous system.

References
Al, E., Iliopoulos, F., Forschack, N., Nierhaus, T., Grund, M., Motyka, P., Gaebler, M., Nikulin, V. V., & Villringer, A.
(2020). Heart—Brain interactions shape somatosensory perception and evoked potentials. Proceedings of the National
Academy of Sciences of the United States of America, 117(19), 10575–10584. https://doi.org/10.1073/pnas.1915629117

98
Interoception and sensory processing

Azevedo, R. T., Badoud, D., & Tsakiris, M. (2017). Afferent cardiac signals modulate attentional engagement to low
spatial frequency fearful faces. Cortex, 1–9. https://doi.org/10.1016/j.cortex.2017.06.016
Azevedo, R. T., Garfinkel, S. N., Critchley, H. D., & Tsakiris, M. (2017). Cardiac afferent activity modulates the expres-
sion of racial stereotypes. Nature Communications, 8, 13854. https://doi.org/10.1038/ncomms13854
Azzalini, D., Buot, A., Palminteri, S., & Tallon-Baudry, C. (2021). Responses to heartbeats in ventromedial prefrontal
cortex contribute to subjective preference-based decisions. Journal of Neuroscience, 41(23), 5102–5114. https://doi.
org/10.1523/JNEUROSCI.1932-20.2021
Azzalini, D., Rebollo, I., & Tallon-Baudry, C. (2019). Visceral signals shape brain dynamics and cognition. Trends in
Cognitive Sciences, 23(6), 488–509. https://doi.org/10.1016/j.tics.2019.03.007
Babo-Rebelo, M., Richter, C. G.,  & Tallon-Baudry, C. (2016). Neural responses to heartbeats in the default net-
work encode the self in spontaneous thoughts. Journal of Neuroscience, 36(30), 7829–7840. https://doi.org/10.1523/
JNEUROSCI.0262-16.2016
Bard, P. (1928). A diencephalic mechanism for the expression of rage with special reference to the sympathetic nervous
system. American Journal of Physiology-Legacy Content, 84(3), 490–515. https://doi.org/10.1152/ajplegacy.1928.84.
3.490
Bernard, C. (1878). Les phenom enes de la vie. Bailliere, Paris, 879.
Bury, G., García-Huéscar, M., Bhattacharya, J., & Ruiz, M. H. (2019). Cardiac afferent activity modulates early neu-
ral signature of error detection during skilled performance. Neuroimage, 199, 704–717. https://doi.org/10.1016/j.
neuroimage.2019.04.043
Cannon, W. B. (1927). The James-Lange theory of emotions: A critical examination and an alternative theory. American
Journal of Psychology, 100(3–4), 567–586. https://doi.org/10.2307/1422695
Cannon, W. B. (1939). The wisdom of the body. W.W. Norton & Company.
Craig, A. D. (2002). How do you feel? Interoception: The sense of the physiological condition of the body. Nature
Reviews. Neuroscience, 3(8), 655–666. https://doi.org/10.1038/nrn894
Craig, A. D. B. (2009). How do you feel—Now? The anterior insula and human awareness. Nature Reviews Neuroscience,
10(1), 59–70. https://doi.org/10.1038/nrn2555
Critchley, H. D.,  & Garfinkel, S. N. (2018). The influence of physiological signals on cognition. Current Opinion in
Behavioral Sciences, 19, 13–18. https://doi.org/10.1016/j.cobeha.2017.08.014
Critchley, H. D., & Harrison, N. A. (2013). Visceral influences on brain and behavior. Neuron, 77(4), 624–638. https://
doi.org/10.1016/j.neuron.2013.02.008
Damasio, A. R. (1994). Descartes’ error: Emotion, rationality and the human brain. Avon Books.
Damasio, A. R. (1999). The feeling of what happens: Body and emotion in the making of consciousness. Harcourt Publishers.
Davos, C. H., Davies, L. C., & Piepoli, M. (2002). The effect of baroreceptor activity on cardiovascular regulation. Hel-
lenic Journal of Cardiology, 43, 145–155.
Dunn, B. D., Evans, D., Makarova, D., White, J., & Clark, L. (2012). Gut feelings and the reaction to perceived ineq-
uity: The interplay between bodily responses, regulation, and perception shapes the rejection of unfair offers on
the ultimatum game. Cognitive, Affective and Behavioral Neuroscience, 12(3), 419–429. https://doi.org/10.3758/
s13415-012-0092-z
Fiene, L., Ireland, M. J., & Brownlow, C. (2018). The Interoception Sensory Questionnaire (ISQ): A scale to meas-
ure interoceptive challenges in adults. Journal of Autism and Developmental Disorders, 48(10), 3354–3366. https://doi.
org/10.1007/s10803-018-3600-3
Galvez-Pol, A, Antoine, S., Li, C., & Kilner, J. M. (2021a). Humans can infer the heart rate of others when looking at their face.
PsyArxiv 1–12. https://doi.org/10.1016/j.cortex.2022.03.003
Galvez-Pol, A., Forster, B., & Calvo-Merino, B. (2020a). Beyond action observation: Neurobehavioral mechanisms of
memory for visually perceived bodies and actions. Neuroscience and Biobehavioral Reviews, 116, 508–518. https://doi.
org/10.1016/j.neubiorev.2020.06.014
Galvez-Pol, A., McConnell, R., & Kilner, J. M. (2020b). Active sampling in visual search is coupled to the cardiac cycle.
Cognition, 196, 104149. https://doi.org/10.1016/j.cognition.2019.104149
Galvez-Pol, A., Nadal, M.,  & Kilner, J. M. (2021b). Emotional representations of space vary as a function of peo-
ples’ affect and interoceptive sensibility. Scientific Reports, 11(1), 16150. https://doi.org/10.1038/s41598-021-95
081-9
Garfinkel, S. N., Barrett, A. B., Minati, L., Dolan, R. J., Seth, A. K., & Critchley, H. D. (2013). What the heart forgets:
Cardiac timing influences memory for words and is modulated by metacognition and interoceptive sensitivity. Psycho-
physiology, 50(6), 505–512. https://doi.org/10.1111/psyp.12039
Garfinkel, S. N., & Critchley, H. D. (2016). Threat and the body: How the heart supports fear processing. Trends in Cogni-
tive Sciences, 20(1), 34–46. https://doi.org/10.1016/j.tics.2015.10.005

99
Galvez-Pol, Munar and Kilner

Garfinkel, S. N., Gould van Praag, C. D., Engels, M., Watson, D., Silva, M., Evans, S. L., Duka, T., & Critchley, H. D.
(2020). Interoceptive cardiac signals selectively enhance fear memories. Journal of Experimental Psychology: General, 44,
1–39. https://doi.org/10.1037/xge0000967
Garfinkel, S. N., Minati, L., Gray, M. A., Seth, A. K., Dolan, R. J., & Critchley, H. D. (2014). Fear from the heart:
Sensitivity to fear stimuli depends on individual heartbeats. Journal of Neuroscience, 34(19), 6573–6582. https://doi.
org/10.1523/JNEUROSCI.3507-13.2014
Garfinkel, S. N., Seth, A. K., Barrett, A. B., Suzuki, K., & Critchley, H. D. (2015). Knowing your own heart: Distinguish-
ing interoceptive accuracy from interoceptive awareness. Biological Psychology, 104, 65–74. https://doi.org/10.1016/j.
biopsycho.2014.11.004
Georgiou, E., Mai, S., Fernandez, K. C., & Pollatos, O. (2018). I see neither your fear, nor your sadness—interoception
in adolescents. Consciousness and Cognition, 60, 52–61. https://doi.org/10.1016/j.concog.2018.02.011
Gray, M. A., Beacher, F. D., Minati, L., Nagai, Y., Kemp, A. H., Harrison, N. A., & Critchley, H. D. (2012). Emo-
tional appraisal is influenced by cardiac afferent information. Emotion, 12(1), 180–191. https://doi.org/10.1037/a00
25083
Grund, M., Al, E., Pabst, M., Dabbagh, A., Stephani, T., & Villringer, A. (2021).03.22.436396. Respiration, heartbeat,
and conscious tactile perception. bioRxiv
Herbert, B. M., Pollatos, O., Flor, H., Enck, P.,  & Schandry, R. (2010). Cardiac awareness and autonomic cardiac
reactivity during emotional picture viewing and mental stress. Psychophysiology, 47(2), 342–354. https://doi.
org/10.1111/j.1469-8986.2009.00931.x
James, W. (1884). What is an emotion? Mind, 9(34), 188–205.
Khalsa, S. S., Adolphs, R., Cameron, O. G., Critchley, H. D., Davenport, P. W., Feinstein, J. S., Feusner, J. D., Garfinkel,
S. N., Lane, R. D., Mehling, W. E., Meuret, A. E., Nemeroff, C. B., Oppenheimer, S., Petzschner, F. H., Pollatos,
O., Rhudy, J. L., Schramm, L. P., Simmons, W. K., Stein, M. B.,  .  .  . Interoception Summit 2016 participants.
(2018). Interoception and mental health: A roadmap. Biological Psychiatry. Cognitive Neuroscience and Neuroimaging, 3(6),
501–513. https://doi.org/10.1016/j.bpsc.2017.12.004
Kluger, D. S., Balestrieri, E., Busch, N. A., & Gross, J. (2021). Respiration aligns perception with neural excitability. bioRxiv
Prepr. 1–28. https://doi.org/10.1101/2021.03.25.436938
Koch, E. (1932). Die irradiation der pressoreceptorischen kreislaufreflexe. Klinische Wochenschrift, 11, 225–227.
Kreindler, A. (1946). Recherches experimentales sur les relations entre le sinus carotidien et le systeme nerveux central.
Bulletin de la Section Scientifique de l’Académie Roumaine, 28, 448–481.
Kunzendorf, S., Klotzsche, F., Akbal, M., Villringer, A., Ohl, S., & Gaebler, M. (2019). Active information sampling
varies across the cardiac cycle. Psychophysiology, 56(5), 1–16. https://doi.org/10.1111/psyp.13322
Lange, C. G. (1885). Om sindsbevaegelser; et psyko-fysiologisk studie. Lund, India.
Larra, M. F., Finke, J. B., Wascher, E., & Schächinger, H. (2020). Disentangling sensorimotor and cognitive cardioafferent
effects: A cardiac-cycle-time study on spatial stimulus-response compatibility. Scientific Reports, 10(1), 4059. https://
doi.org/10.1038/s41598-020-61068-1
Leganes-Fonteneau, M., Buckman, J. F., Suzuki, K., Pawlak, A., & Bates, M. E. (2021). More than meets the heart:
Systolic amplification of different emotional faces is task dependent. Cognition and Emotion, 35, 400–408. https://doi.
org/10.1080/02699931.2020.1832050
Linson, A., Parr, T., & Friston, K. J. (2020). Active inference, stressors, and psychological trauma: A neuroethological
model of (mal)adaptive explore-exploit dynamics in ecological context. Behavioural Brain Research, 380, 112421.
https://doi.org/10.1016/j.bbr.2019.112421
Makowski, D., Sperduti, M., Blondé, P., Nicolas, S., & Piolino, P. (2020). The heart of cognitive control: Cardiac phase
modulates processing speed and inhibition. Psychophysiology, 57(3), e13490. https://doi.org/10.1111/psyp.13490
Mehling, W. E., Acree, M., Stewart, A., Silas, J., & Jones, A. (2018). The multidimensional assessment of interoceptive
awareness, version 2 (MAIA-2). PLOS ONE, 13(12), 1–12. https://doi.org/10.1371/journal.pone.0208034
Mehling, W. E., Price, C., Daubenmier, J. J., Acree, M., Bartmess, E.,  & Stewart, A. (2012). The multidimensional
assessment of interoceptive awareness (MAIA). PLOS ONE, 7(11), e48230. https://doi.org/10.1371/journal.
pone.0048230
Motyka, P., Grund, M., Forschack, N., Al, E., Villringer, A., & Gaebler, M. (2019). Interactions between cardiac activity
and conscious somatosensory perception. Psychophysiology, 56(10), 1–13. https://doi.org/10.1111/psyp.13424
Murphy, J., Catmur, C., & Bird, G. (2019). Classifying individual differences in interoception: Implications for the meas-
urement of interoceptive awareness. Psychonomic Bulletin and Review, 26(5), 1467–1471. https://doi.org/10.3758/
s13423-019-01632-7
Namkung, H., Kim, S. H., & Sawa, A. (2017). The insula: An underestimated brain area in clinical neuroscience, psy-
chiatry, and neurology. Trends in Neurosciences, 40(4), 200–207. https://doi.org/10.1016/j.tins.2017.02.002

100
Interoception and sensory processing

Nord, C. L., Dalmaijer, E. S., Armstrong, T., Baker, K., Nord, C. L., Dalmaijer, E. S., Armstrong, T., Baker, K., & Dal-
gleish, T. (2021). Report a causal role for gastric rhythm in human disgust avoidance a causal role for gastric rhythm
in human disgust avoidance. Current Biology, 1–6. https://doi.org/10.1016/j.cub.2020.10.087
Palser, E. R., Palmer, C. E., Galvez-Pol, A., Hannah, R., Fotopoulou, A., & Kilner, J. M. (2018). Alexithymia medi-
ates the relationship between interoceptive sensibility and anxiety. PLOS ONE, 13(9), e0203212. https://doi.
org/10.1371/journal.pone.0203212
Park, H. D., Barnoud, C., Trang, H., Kannape, O. A., Schaller, K., & Blanke, O. (2020). Breathing is coupled with
voluntary action and the cortical readiness potential. Nature Communications, 11(1), 289. https://doi.org/10.1038/
s41467-019-13967-9
Park, H. D., & Blanke, O. (2019). Heartbeat-evoked cortical responses: Underlying mechanisms, functional roles, and
methodological considerations. Neuroimage, 197, 502–511. https://doi.org/10.1016/j.neuroimage.2019.04.081
Park, H. D., Correia, S., Ducorps, A., & Tallon-Baudry, C. (2014). Spontaneous fluctuations in neural responses to heart-
beats predict visual detection. Nature Neuroscience, 17(4), 612–618. https://doi.org/10.1038/nn.3671
Paulus, M. P., Feinstein, J. S., & Khalsa, S. S. (2019). An active inference approach to interoceptive psychopathology.
Annual Review of Clinical Psychology, 15, 97–122. https://doi.org/10.1146/annurev-clinpsy-050718-095617
Perl, O., Ravia, A., Rubinson, M., Eisen, A., Soroka, T., Mor, N., Secundo, L.,  & Sobel, N. (2019). Human non-
olfactory cognition phase-locked with inhalation. Nature Human Behaviour, 3(5), 501–512. https://doi.org/10.1038/
s41562-019-0556-z
Pfeifer, G., Garfinkel, S. N., Gould van Praag, C. D., Sahota, K., Betka, S., & Critchley, H. D. (2017). Feedback from the
heart: Emotional learning and memory is controlled by cardiac cycle, interoceptive accuracy and personality. Biological
Psychology, 126, 19–29. https://doi.org/10.1016/j.biopsycho.2017.04.001
Porges, S. (1993). Body perception questionnaire. Lab. Dev. Assessment. Universidad Maryl.
Powley, T. L., Hudson, C. N., McAdams, J. L., Baronowsky, E. A., & Phillips, R. J. (2016). Vagal intramuscular arrays:
The specialized mechanoreceptor arbors that innervate the smooth muscle layers of the stomach examined in the rat.
Journal of Comparative Neurology, 524(4), 713–737. https://doi.org/10.1002/cne.23892
Quigley, K. S., Kanoski, S., Grill, W. M., Barrett, L. F.,  & Tsakiris, M. (2021). Functions of interoception: From
energy regulation to experience of the self. Trends in Neurosciences, 44(1), 29–38. https://doi.org/10.1016/j.tins.2020.
09.008
Rae, C. L., Ahmad, A., Larsson, D. E. O., Silva, M., van Praag, C. D. G. V., Garfinkel, S. N., & Critchley, H. D. (2020).
Impact of cardiac interoception cues and confidence on voluntary decisions to make or withhold action in an inten-
tional inhibition task. Scientific Reports, 10(1), 4184. https://doi.org/10.1038/s41598-020-60405-8
Rae, C. L., Botan, V. E., Gould Van Praag, C. D., Herman, A. M., Nyyssönen, J. A. K., Watson, D. R., Duka, T., Garfin-
kel, S. N., & Critchley, H. D. (2018). Response inhibition on the stop signal task improves during cardiac contraction.
Scientific Reports, 8(1), 9136. https://doi.org/10.1038/s41598-018-27513-y
Rebollo, I., Devauchelle, A. D., Béranger, B., & Tallon-Baudry, C. (2018). Stomach-brain synchrony reveals a novel,
delayed-connectivity resting-state network in humans. eLife, 7, 1–25. https://doi.org/10.7554/eLife.33321
Riva, G. (2018). The neuroscience of body memory: From the self through the space to the others. Cortex; a Jour-
nal Devoted to the Study of the Nervous System and Behavior, 104, 241–260. https://doi.org/10.1016/j.cortex.2017.
07.013
Salomon, R., Ronchi, R., Dönz, J., Bello-Ruiz, J., Herbelin, B., Faivre, N., Schaller, K., & Blanke, O. (2018). Insula
mediates heartbeat related effects on visual consciousness. Cortex; a Journal Devoted to the Study of the Nervous System
and Behavior, 101, 87–95. https://doi.org/10.1016/j.cortex.2018.01.005
Schandry, R. (1981). Heart beat perception and emotional experience. Psychophysiology, 18(4), 483–488. https://doi.
org/10.1111/j.1469-8986.1981.tb02486.x
Schulz, A., Schilling, T. M., Vögele, C., Larra, M. F., & Schächinger, H. (2016). Respiratory modulation of startle eye
blink: A new approach to assess afferent signals from the respiratory system. Philosophical Transactions of the Royal Society
of London. Series B, Biological Sciences, 371(1708). https://doi.org/10.1098/rstb.2016.0019
Schulz, A., van Dyck, Z., Lutz, A. P. C., Rost, S., & Vögele, C. (2017). Gastric modulation of startle eye blink. Biological
Psychology, 127, 25–33. https://doi.org/10.1016/j.biopsycho.2017.05.004
Sherrington, C. (1906). The integrative action of the nervous system. Yale University Press.
Stevenson, R. J., Francis, H. M., Oaten, M. J., & Schilt, R. (2018). Hippocampal dependent neuropsychological tests
and their relationship to measures of cardiac and self-report interoception. Brain and Cognition, 123, 23–29. https://
doi.org/10.1016/j.bandc.2018.02.008
Tallon-Baudry, C., Campana, F., Park, H. D., & Babo-Rebelo, M. (2018). The neural monitoring of visceral inputs,
rather than attention, accounts for first-person perspective in conscious vision. Cortex; a Journal Devoted to the Study of
the Nervous System and Behavior, 102, 139–149. https://doi.org/10.1016/j.cortex.2017.05.019

101
Galvez-Pol, Munar and Kilner

Wilkinson, M., McIntyre, D., & Edwards, L. (2013). Electrocutaneous pain thresholds are higher during systole than
diastole. Biological Psychology, 94(1), 71–73. https://doi.org/10.1016/j.biopsycho.2013.05.002
Wolpert, N., Rebollo, I., & Tallon-Baudry, C. (2020). Electrogastrography for psychophysiological research: Practical
considerations, analysis pipeline, and normative data in a large sample. Psychophysiology, 57(9), e13599. https://doi.
org/10.1111/psyp.13599
Zaman, J., De Peuter, S., Van Diest, I., Van den Bergh, O., & Vlaeyen, J. W. S. (2016). Interoceptive cues predicting extero-
ceptive events. International Journal of Psychophysiology, 109, 100–106. https://doi.org/10.1016/j.ijpsycho.2016.09.003

102
6
NEURAL CORRELATES OF VISUAL
AESTHETIC APPEAL
Edward A. Vessel, Tomohiro Ishizu and Giacomo Bignardi

The question of how visual experiences acquire aesthetic value is an old one. Many philosophers and poets
have mused about what makes a face attractive, what makes a mountain vista beautiful, or how a painting can
move us. Not only was the question of visual beauty of prime importance for early work in philosophical
and empirical aesthetics (from Kant [1790/1987] to Fechner [1876], Arnheim [1954], and Berlyne [1971]),
it was also one of the earliest domains of aesthetic experiences to be systematically explored using the mod-
ern tools of cognitive neuroscience. In part, this was due to the fact that in the late 1990s, knowledge of
the human visual system was a fair amount ahead of an understanding of other sensory systems, and novel
brain imaging techniques and analyses, particularly those for functional magnetic resonance imaging (fMRI),
were often pioneered in the visual domain. The existence of theoretical sketches of human visual processes
(e.g., Goodale & Milner, 1992; Ungerleider & Mishkin, 1982; Van Essen et al., 1992) served as a springboard
for early ideas about how visual experiences might come to be experienced as pleasurable or even beautiful,
and fMRI methods allowed for these ideas to be tested.
There are a variety of definitions of what constitutes an aesthetic experience (see Anglada-Tort & Skov,
2020). Here, we will adopt a rather wide definition. Visual aesthetic experiences clearly involve perception
yet also engage sense making (comprehension). They are evaluative, affectively absorbing, and linked to spe-
cific appraisals ( judgments) of a stimulus, such as whether they are pleasing, beautiful, moving, or attractive
(Pearce et al., 2016; Vessel, 2020). Visual aesthetic appraisals are a source of hedonic pleasure—a hedonic
value that derives from processing the perceptual and conceptual aspects of a visual experience itself (and apart
from any value associated with what an image or object may stand for, e.g., the value of a hundred-dollar
bill or the sweetness of a particular strawberry). Such appraisals are not confined to encounters with visual
art but can also occur during encounters with the natural world, with people, with objects, and even with
thoughts and concepts.
In order to clarify what is known about visual aesthetic appreciation, we begin this chapter with an over-
view of relevant behavioural work. Much of the empirical work on visual aesthetics has focused on under-
standing whether visual aesthetic appraisals, for example, feeling beauty or being moved, can be predicted
from properties of an image. More recently, an increasing focus has also been on understanding the highly
idiosyncratic nature of visual aesthetic tastes and how differences in context, or in a perceiver’s personality
and past experience, interact with a visual stimulus to make for an appealing aesthetic experience.
We then turn to focus on what is known about the neural mechanisms that support aesthetic appreciation.
Much recent work in visual neuroaesthetics has focused on understanding how visual aesthetic appraisals are

DOI: 10.4324/9781003008675-7 103


Vessel, Ishizu and Bignardi

represented in the brain. What neural systems contribute to visual aesthetic appeal? Does the brain represent
aesthetic appeal in a domain-specific or domain-general fashion? Are aesthetic appraisals automatic, or do
they require explicit judgment?
Finally, we briefly outline some initial thoughts on perhaps the most challenging question in the field—
why do we like what we like? Unfortunately, there is very little direct evidence that bears on the question of
why humans find certain visual experiences aesthetically appealing; why they can be so impactful; and what
evolutionary benefit, if any, such a mechanism might afford. We therefore present a necessarily selective and
tentative account that is informed by recent empirical and theoretical work in empirical aesthetics both in
vision and beyond (see also the discussions of this issue in Chapters 2 and 11).

Measuring visual aesthetic appreciation


Before diving deeper into the neural correlates of visual appreciation, we first summarize a body of behav-
ioural research in order to better clarify the relevant constructs. We begin by describing types of measure-
ment that have been employed by researchers to assess the neural correlates of visual aesthetic appreciation to
highlight sources of variation that might affect interpretation and synthesis of findings. Figure 6.1 shows the
two main types of measurement that have been employed to assess appraisal in neuroaesthetic studies, direct
and indirect, with the first being used more frequently than the second. Other measures, such as production
or adjustment tasks, are also possible, but given the minor role they have played in neuroimaging studies, they
will not be reviewed here (the curious reader can refer to Palmer et al., 2013).
Direct measurements quantify appraisal that is directly reported by participants. They can be collected
when participants are appraising images by means of dichotomous (e.g., “no,” “yes”; Jacobsen et al., 2006),
ordinal (e.g., from “ugly, “indifferent,” “beautiful”; Ishizu  & Zeki, 2011), interval, or continuous rating
scales (e.g., aesthetically “moving,” from low to high; Vessel et al., 2019) and when participants are express-
ing their preferences by choosing between two images (e.g., Kim et al., 2007) or through ratings obtained
continuously over time (e.g., Belfi et al., 2019; Isik & Vessel, 2021). Indirect measures, on the other hand, are
collected using a variety of alternative tools that are intended to measure an appraisal indirectly, with the goal
of measuring appeal in a manner that is unmediated, or at least only partially mediated, by conscious reflec-
tion and labelling of experience. Examples include measurements of viewing time (e.g., with or without a
requirement to expend “effort,” Aharon et al., 2001), reaction time (e.g., Lopez-Persem et al., 2020), arousal
(e.g., as measured by skin conductance changes, Salimpoor et al., 2011), pupil dilation (e.g., Laeng et al.,
2016), heart rate (e.g., Tschacher et al., 2012), and “willingness to pay” (e.g., Smith et al., 2010). It is worth
noting, however, that while for some aesthetic domains, there is evidence that indirect measures can be used
to assess liking (e.g., skin conductance response to asses pleasure from music; see Fleurian & Pearce, 2020),
there is so far little consensus regarding which indirect measures are reliable indicators of visual appraisal.
Further, an additional source of variation between studies is whether the task involves explicit appraisal, for
example, whether participants are asked to express their appraisal while viewing the stimuli (and thus also
during brain imaging). Alternatively, researchers study implicit appraisal by not imposing a rating task during
brain imaging, relying instead on pre- or post-session ratings or ratings from other observers. While explicit
appraisal is important to establish neural correlates of appreciation, it is important to investigate whether the
neural correlates of implicit appraisal differ from the ones of explicit appraisal, even when individuals are
focused on a non-aesthetic task such as identity recognition (Chatterjee et al., 2009) or perceptual judgments
(Kim et al., 2007). To make things more complicated, not only do studies employ different measures and/
or tasks, but they also introduce other sources of methodological variation by asking a variety of questions.
For example, when asked to judge images, participants can be asked questions ranging from how attrac-
tive (Chatterjee et al., 2009) or how liked images of faces are (Lopez-Persem et al., 2020) to how beautiful
(Ishizu & Zeki, 2011) or how moved participants feel when viewing a painting (Vessel et al., 2012).

104
Neural correlates of visual appeal
105

Figure 6.1 Methodological “sources of variation.” Different types of questions and measurements (direct and/or indirect) are used to assess visual aesthetic appeal.
To establish the neural correlates of visual appreciation, such measures can be employed during (explicit) or before or after (implicit) neuroimaging
experiments. L = low; H = high; t = time.
Vessel, Ishizu and Bignardi

Finally, even when participants are asked the same question, such as how attractive a face is, the instruc-
tions are sometimes different, with some implying personal attraction and potential mating desire for the
person depicted (Kocsor et al., 2013) and others ignoring it (Chatterjee et al., 2009). While it is the case that
different types of measures are strongly related, such as interval ratings being predictive of choice (Lopez-
Persem et al., 2020), it is possible that one can influence the other (see Ariely & Norton, 2008; Izuma et al.,
2010), and even when different questions lead to highly correlated ratings, such as pleasure and beauty
(Brielmann & Pelli, 2019), the relationship is rarely perfect. All in all, given that only a few studies have
investigated such differences, it is important to take these considerations in mind when interpreting results
from different studies.

Aesthetics from below


A central question in visual aesthetics is how representations of visual features and content map onto repre-
sentations of value or appreciation. To address this question, it is helpful to recall how visual perception and
representation is structured. Vision is generally divided into low-, mid-, and high-level processes. Low-level
vision includes the computation of local features such as contrast, orientation, colour, and motion. Mid-level
vision is conceptualized as organizational processes such as element grouping, contour completion, surface
segmentation, and depth assignment. Finally, high-level vision includes processes such as object and scene
recognition that require making contact with stored representations.
In empirical aesthetics, there is a strong tradition of an “aesthetics from below” approach that seeks to
explain aesthetic preferences for a visual experience as the combination of preferences for a variety of objec-
tively measurable low-, mid-, and high-level features. Yet while there do indeed appear to be a number of
stimulus dimensions that are hedonically “marked,” this approach has yielded mixed results.
Two of the most studied features in this regard are symmetry and angular versus curved contours. Across
a wide variety of stimuli from abstract geometrical patterns ( Jacobsen & Höfel, 2002) to faces (see Rhodes,
2006), symmetry positively predicts average preferences (Bertamini et al., 2019). Similarly, smoothly curved
contours tend to be preferred over angular and/or jagged contours for a variety of visual stimuli, including
images of objects and abstract patterns (Bar & Neta, 2006), closed shapes (Bertamini et al., 2016), product
packaging (Westerman et al., 2012), and architectural interiors (Vartanian et al., 2013). In addition, a number
of other low- and mid-level features have also been claimed to predict preferences, such as contrast and clar-
ity (Reber et al., 1998; Tinio & Leder, 2009), the presence of certain colours (McManus et al., 1981), and
aspect ratio (McManus, 1980).
Beyond localized features, a number of global features of images, such as anisotropy (inhomogeneity of
orientation, such as the predominance of vertical and horizontal features in city scenes) and self-similarity
(similar structure at multiple scales, such as fractals) have been explored under the hypothesis that such fea-
tures reflect key aspects of the statistics of natural images to which the visual system is sensitive (e.g., Mallon
et al., 2014; Redies, 2007; Spehar et al., 2016). Researchers have also investigated configural or holistic fea-
tures that could be subsumed under the label of “good composition,” such as balance (Hübner & Fillinger,
2019) and positioning in a frame (Sammartino & Palmer, 2012).
Yet this approach has some clear limitations. Often, such features fail to show a clear relationship to aes-
thetic preferences, and within a single study, the effects are often quite small, with a large degree of variance
from person to person (e.g., McManus et al., 2010; Spehar et al., 2016).

High-level aesthetic domains and shared versus individual taste


Research that specifically measures the degree of “shared” versus “individual” taste across a set of observ-
ers has found that people often don’t agree in what they find visually appealing. Vessel and Rubin (2010)

106
Neural correlates of visual appeal

reported a study in which observers made preference judgments on a set of photographs of real-world scenes
and, separately, a set of abstract images that contained variation in low- and mid-level features but lacked any
shared semantic interpretations. While the observers’ judgments revealed a strong degree of shared taste for
real-world scenes, the same observers were much more idiosyncratic in their ratings of the abstract images.
That the presence of visual features alone did not lead to consistent ratings across observers highlights the
importance of higher-level, semantic meanings and associations for appreciation of scenes. A study of simple
colour preferences reached a similar conclusion, finding that differences in colour preferences were well pre-
dicted by participants’ liking of objects associated with those colours (e.g., positive associations of sky with
blue but negative associations of excrement with yellow-brown; Palmer & Schloss, 2010).
This issue becomes even more apparent when one turns to a defining feature of high-level vision—the
specialized processing of object and scene features in category-selective visual regions. The human ventral
occipitotemporal cortex (VOT) contains distinct regions that respond selectively to faces, bodies, scene
layout and identity, objects, and symbols (see Grill-Spector & Weiner, 2014 for a review). Domains of aes-
thetic evaluation (e.g., faces, natural landscape, dance, fashion, architecture, visual artwork, design objects)
do not neatly map onto these visual categories in a one-to-one manner (for example, architectural images
may contain aspects of scene layout but also object-like properties). Yet it has become increasingly clear that
appreciation of visual images that fall under different domains, such as faces and paintings, results in different
levels of agreement across individuals (Bignardi et al., 2021; Leder et al., 2016; Vessel et al., 2018).
Specifically, individuals tend to agree more on what they like when they are appraising images of natural
kinds such as faces and landscapes than when they are judging cultural artefacts such as architecture and
paintings (see Figure 6.2; variance component analysis of data adapted from Vessel et al., 2018; Martinez
et  al., 2020). What leads to these differences is still a matter of debate. One possible explanation is that
inherited versus acquired brain concepts play a role when appraising visual images of biological versus non-
biological kinds (Zeki, 2011; Zeki & Chén, 2020; see Zeki et al., 2020 for a definition of brain concepts).
An alternative suggestion is that judgments of natural kinds have greater behavioural relevance for everyday
decisions compared to cultural artefacts (Vessel et al., 2018); over many years of interaction, this could lead
different people to base their appraisals on similar information.
These differences in shared taste provide a way to organize findings from studies of visual appreciation in
a manner that links to the organizational principles of high-level vision. While this approach has limitations
(e.g., scenes may contain a mixture of naturalistic and human-made content), we will use this taxonomy
from naturalistic (people, natural landscapes) to artefactual images (human-created objects) to describe the
findings from visual appreciation.
Among naturalistic visual aesthetic domains, appreciation of faces has the highest level of agreement across
individuals (see Figure 6.2; Vessel et al., 2018). There is a strong degree of shared taste for evaluations of
faces, and individuals have a good estimation of imagined ratings of others (Bignardi et al., 2021; Hönekopp,
2006; Leder et al., 2016; Vessel et al., 2018). Moreover, individuals from different cultures tend to agree on
which faces are on average considered more attractive, regardless of the ethnicity of the faces being rated, and
infants tend to look more at faces that are indeed rated more attractive by adults (Langlois et al., 2000; Slater
et al., 1998). Several features of faces, including degree of symmetry, averageness, and statistical typicality
more generally, can account for some of this agreement across observers (see Rhodes, 2006 for a review; see
Ryali et al., 2020 for details for statistical typicality).
Although visual scenes (or “places”) are often considered a category from the perspective of visual percep-
tion, this may be too general from the perspective of aesthetics, especially as it cuts across the natural versus
artefactual distinction. Rather, we will discuss natural landscapes separate from humanmade architectural
settings (exterior and interior).
Natural landscapes are another category for which agreement is substantially shared, though to a lesser
extent than for images of faces (see Figure 6.2, Vessel et al., 2018). Within the environmental assessment

107
Vessel, Ishizu and Bignardi

Figure 6.2 Amount of “shared taste” across observers varies by visual aesthetic domain. Variance component analysis
(upper panel) of aesthetic ratings for images of faces, places, exterior and interior architectural places, and
paintings. By using a multilevel intercept-only null model (Martinez et al., 2020), it is possible to partition
the percentage of variance in (1) appraisal that is common across individuals (image, here represented in yel-
low, that is variance accounted for by item level characteristics, or better by the shared experiences evoked
in the individuals by such characteristics), (2) appraisal that is unique to the individual (participant, here rep-
resented in green, that is the overall individual participant appraisal averaged across images within a domain),
(3) appraisal that emerges as an interaction from the individual appraising and the image appraised (interac-
tion, here represented in purple), and (4) error (residual, here represented in gray). CI represents the confi-
dence interval, obtained following a bootstrapping procedure (2000 bootstrap samples, following Sutherland
et al. [2020]). The lower panel shows the proportion of variance expressed only over non-residual variance
which is accounted for by shared (“shared taste,” yellow) and unique (blue) components. “Shared taste”
decreases along the axis that goes from natural to artefactual (faces to paintings), while “idiosyncratic taste”
tends to follow an inverse trend. Reanalysis of visual appreciation ratings from Vessel et al.(2018). Data at:
https://edmond.mpdl.mpg.de/imeji/collection/dMlhGcI642YmIIF2.

literature, there is a strong tradition of investigating the presence of natural versus humanmade intrusions
(Biederman & Vessel, 2006; S. Kaplan et al., 1972), the presence of refuge or an expansive view (Appleton,
1975), the likelihood of new information emerging (“mystery”; Kaplan, 1992), and the presence of specific
biomes (see Kaplan & Kaplan, 1995; see also Chapter 10). More recently, some of these measures have been
related to spectral measures of scene shape (e.g., openness, Oliva & Torralba, 2001; Pegors & Epstein, 2011).
Yet while aesthetic ratings for naturalistic images show higher agreement than artefactual ones, higher
doesn’t mean universal. For example, even for faces, only about half of the variation in appraisal can be
explained by shared preferences (Germine et al., 2015; Hönekopp, 2006; Vessel et al., 2018). That is, while
it seems that agreement on visual appraisal follows a gradient from shared to idiosyncratic—from naturalistic
to artefactual, respectively—idiosyncrasies still tend to emerge. As Hönekopp (2006, p. 208) said “(almost) all
people are good-looking—at least to some.” This is also another descriptive finding from the field of empiri-
cal aesthetics, namely that variation in appraisal is the norm, not the exception.
In contrast to natural landscapes, images of architecture produce quite idiosyncratic ratings of appeal
across individuals (both interior and exterior; see Figure 6.2; Vessel et al., 2018), despite being photographs
of real-world “places.” Factors such as angular (rectilinear) versus curved interiors, open versus closed spaces,
and ceiling height have been found to have an impact on average aesthetic judgments of architectural interi-
ors (Coburn et al., 2020; Vartanian et al., 2013) and also to affect approach-avoid judgments in non-experts

108
Neural correlates of visual appeal

(Vartanian et al., 2017). A recent study has proposed that the aesthetic experience of interior architecture can
be organized along dimensions of “fascination,” “coherence” (ease of understanding the organization of a
scene), and “hominess” (Coburn et al., 2020; though it should be noted that hominess was highly correlated
with naturalness; see Chapter 10 for more information).
Finally, aesthetic judgments of images of visual artwork (paintings) tend to produce highly individual rat-
ings; it is typical that less than 10% of repeatable variance can be accounted for by a shared factor (Leder et al.,
2016; see Figure 6.2, Vessel et al., 2018). Of course, artworks are a very heterogeneous category of stimuli.
Yet even within artworks, representational works produce higher agreement than abstract works (Schepman
et al., 2015). This agreement has been related to similarity in liking for associated semantics (Schepman et al.,
2015) and for objects contained in paintings (Levitan et al., 2019). Ambiguity and multiple-interpretability
can also play a central role (Muth et al., 2015), lending artworks multiple levels of meaning.
Taken together, this body of work suggests that especially for human artefacts, a purely stimulus-driven
approach that seeks to compute liking based solely on an analysis of visual features is limited in what it can
achieve. Not only is it unable to capture the wide degree of individual differences observed in aesthetic pref-
erences, it may not even be able to account for a majority of variance in average ratings for some aesthetic
domains.

The interactionist view


A second approach emphasizes the role of the idiosyncratic experience of the perceiver in aesthetic appraisal:
How does a visual stimulus interact with an existing mind/brain, whose goal is to make sense of the world?
This approach can be characterized as essentially about information processing and recasts aesthetic apprecia-
tion as pleasure from understanding.
This approach also has a long history. Constructs such as novelty, familiarity, complexity, and order
have received quite a bit of attention, particularly by Zajonc (1968; the mere exposure effect), Berlyne
(1958, 1970, 1974), and their contemporaries in the mid-20th century.1 In the environmental aesthet-
ics literature, Kaplan (1992) summarized a variety of findings on landscape preferences in a theory that
emphasized the roles of exploration and understanding of one’s environment. More recently, processing
fluency (Reber et al., 2004), which holds that perceptual and cognitive experiences that are processed
more easily are preferred, has received widespread attention, particularly as it is capable of incorporating
both preferences for previously exposed stimuli and also preferences for a number of specific stimulus
dimensions (e.g., higher contrast, prototypical shape). We note, though, that processing fluency has dif-
ficulty accounting for the fact that people often exhibit novelty preferences or in fact seek out, and get
more pleasure from, complex, challenging, or informationally rich experiences (e.g., Muth et al., 2015;
see also Graf & Landwehr, 2015).
Another line of work has emphasized the role of meaning and semantic associations (Biederman & Vessel,
2006; Levitan et al., 2019; Martindale, 1984; Palmer & Schloss, 2010; Vessel & Rubin, 2010) and measures
that more directly relate to information processing, such as expectations, surprise, uncertainty, entropy, com-
pressibility, and learnability (Biederman & Vessel, 2006; Schmidhuber, 2010; Schoeller & Perlovsky, 2016;
Silvia, 2005; see Chapter 26). Although details vary across several formulations, the central theme is that the
brain is constantly engaging in prediction of its sensory world (Clark, 2018; Friston, 2010), and violations of
those expectations (surprise) engage mechanisms that attempt to make sense of the unexpected input. Both
the initial surprise (Loewenstein, 1994; triggering interest/curiosity; Silvia, 2008) and the potential subse-
quent “click of comprehension” from successful sense-making (Biederman & Vessel, 2006; Muth & Carbon,
2013) are experienced as pleasurable.
This approach has gained significant momentum and support, particularly in the music domain
(e.g., Cheung et al., 2019; Gold et al., 2019; see also Chapters 7 and 16). Two key insights of this class of

109
Vessel, Ishizu and Bignardi

theories are that (1) higher-order (semantic, conceptual, personal) information matters more than lower-
order information, reflecting the hierarchy of sensory processing, and (2) that people get the most pleasure
from experiences that provide the greatest opportunity for learning, because they offer novel information
that is richly interpretable to the perceiver.
While some of these measures may reflect common aspects of human experience and could therefore be
potentially computed solely from a stimulus (or statistics of occurrence over a set of stimuli), the majority of
these measures, and the finer aspects of all of them, can only really be computed by taking into account the
interaction of a particular stimulus with the specifics of an observer’s past history (exposure), associations,
beliefs and expertise. Indeed, even low- and mid-level predictors like contrast, symmetry, and angularity can
likely be recast in terms of access to information content. Visual appreciation is less about specific features
and more about one’s ability to predict, make sense of, and learn about our world.

How the brain supports visual aesthetic appreciation


How does the brain get from a neural representation of what we see to a representation of what we like? One
of the primary topics tackled by neuroaesthetics research is the question of whether there is a specific pat-
tern of brain activity associated with the psychological state of finding something aesthetically appealing. As
reviewed in the previous section, it is clear that these states can be elicited by different experiences for differ-
ent people. Many experiments in neuroaesthetics have used fMRI to identify brain regions that are activated
in response to the subjective experience of high aesthetic appeal, variously characterized as the feeling of
beauty, “being moved,” or liking. These experiments aim to detect correlations between brain responses and
subjective states by computing statistical contrasts between average patterns of blood oxygen level–depend-
ent (BOLD) signal for different experimental conditions (e.g., “beautiful” versus “ugly” or “highly moving”
versus “not moving”) or by measuring an association between BOLD signal and a continuously varying
parameter (e.g., rated enjoyment).
While less often the focus, several studies also have sought to identify brain systems that support explicit
judgments of aesthetic appeal by comparing aesthetic versus non-aesthetic judgment tasks. More recently,
the question of how the brain transforms a neural representation of what we see into a representation of
what we find appealing has also come into sharper focus, though at this stage, it is fair to say that an answer
remains far off.
Based upon comparisons of neural activation for high versus low aesthetic appeal, as well as studies
that compare aesthetic versus non-aesthetic tasks, there is evidence that several large-scale brain systems
contribute to visual aesthetic appreciation. In addition to the visual system and subcortical brain systems
for valuation and reward, there is also evidence for engagement of medial and inferior lateral prefrontal
cortical areas, the default-mode network, and in some cases somato-motor systems. In the following,
we will discuss findings related to specific brain systems, followed by a discussion of integration across
networks.

The visual system


It is clear that visual experiences with aesthetic material can engage most, if not all, of the visual system,
depending on the type of stimulus (static/dynamic, scenes versus objects or bodies, moving through space,
etc.) What is less clear is how a neural representation of what we see, as is thought to exist in feature- or
category-selective regions of the ventral visual pathway (Grill-Spector & Weiner, 2014; Kanwisher, 2010;
Pitcher et al., 2009), is related to representations of aesthetic or affective dimensions. Is aesthetic value rep-
resented only in core “valuation” regions such as the striatum or orbitofrontal cortex? Or are there already
signals present in visual regions reflecting aesthetic value?

110
Neural correlates of visual appeal

Early visual processing


As discussed previously, there is evidence that at least some visual feature dimensions correlate with aesthetic
appeal, though with the caveat that such evidence is often seen at the level of average preference, and often
with small effect sizes. However, there do not exist many neural studies that directly link such preferences to
neural activation. For example, higher-contrast stimuli lead to higher responses in early visual regions (Boyn-
ton et al., 1999), and this sensitivity is reduced in higher-order regions such as the lateral occipitotemporal
cortex (LOC; Avidan et al., 2002), but no link has been made between these findings and preferences for
high-contrast images. Symmetry responses first appear in V3, with strongest responses in area VO1 (Keefe
et al., 2018; Sasaki et al., 2005). However, an experiment that specifically looked at aesthetic judgments of
symmetrical versus asymmetrical stimuli found no differences for either symmetry nor aesthetic appeal of
symmetric stimuli in visual regions ( Jacobsen et al., 2006), though these authors did report higher activity
in early visual regions for more complex stimuli. On the other hand, contour curvature strongly modulates
neural responses in V4 (Pasupathy & Connor, 2002), and angular (versus curved) stimuli lead to greater acti-
vation in a number of early visual areas (Bar & Neta, 2007). Within the domain of motion, a study of kinetic

Figure 6.3 The visual system and aesthetic appreciation. On both (A) lateral and (B) ventral surfaces of the brain, a
variety of feature- and category-selective visual regions have been identified. These include the fusiform face
area (FFA), parahippocampal place area (PPA), lateral occipital complex (LOC), occipital face area (OFA),
and extrastriate body area (EBA). In addition, a number of retinotopic visual regions (V1–V4) as well as
V5/MT also show selectivity for a variety of low- and mid-level visual features. A number of studies have
reported modulations of these regions by aesthetic appeal. Also shown are approximate locations of regions
within ventral occipital and inferior temporal areas, such as the ventral occipital temporal cortex (VOT),
inferior temporal cortex (IT), fusiform gyrus (FG), lingual gyrus (LG), and parahippocampal gyrus (PHG).
Regions can also be organized into broader visual streams (dotted arrows), including the ventral (“what”)
pathway, the dorsal (“where/how”) pathway, and a recently identified middle pathway subserving social
vision including the posterior superior temporal sulcus (pSTS). To date, most work on visual aesthetics has
focused on the ventral pathway. The regions consisting of the action observation network (AON), including
the parietal cortex, premotor cortex, and inferior frontal gyrus (IFG), are thought to be engaged by aesthetic
appeal of dance and body movements. The upper and lower arrows denote the dorsal and ventral streams,
respectively, and the middle arrow, through V1/V2, EBA/V5, and pSTS, denotes the recently proposed
“social perception” pathway (Pitcher & Ungerleider, 2021).

111
Vessel, Ishizu and Bignardi

dot patterns reported greater activity for aesthetically preferred patterns in an area of visual cortex likely
overlapping with motion-sensitive regions (V5/middle temporal (MT) area; Zeki & Stutters, 2012), though
the relationship between observers’ preferences and measures of motion energy was not entirely clear (for an
in-depth discussion of the neural correlates of motion appreciation, see Chapter 9). In contrast to motion, an
fMRI study on harmonious colours found no significant difference in activity within V4 between aestheti-
cally preferred colour patches and non-preferred ones (Ikeda et al., 2015).
Overall, there is very little brain imaging work that attempts to link low- and mid-level features, and their
fMRI correlates, directly to aesthetic appeal, and what work there is does not provide a coherent picture.
One possible interpretation of the literature is that more activity in visual regions is generally preferred. With
the potential exception of angularity, where angular stimuli were associated with greater activity (Bar  &
Neta, 2007), most of the existing work is not inconsistent with this hypothesis. A more parsimonious sum-
mary of the existing evidence is that activity in early visual cortices does not generally bear a strong, consist-
ent relation to aesthetic preferences.

Higher-level visual processing


Across a number of visual aesthetic domains, there is evidence for activity changes in higher-level por-
tions of ventral, lateral, and medial occipito-temporal cortex correlated with aesthetic appeal. The exact
location of these modulations depends on the stimulus domain, though there has generally been a greater
focus on parts of the ventral visual pathway, including the ventral occipitotemporal cortex, thought to
support object and scene recognition (e.g., the “what” pathway; Goodale & Milner, 1992; Ungerleider &
Mishkin, 1982).
For faces, the most well-studied domain of visual aesthetic appraisal, there is evidence that the intensity
of aesthetic appeal is correlated with activity in portions of a face-selective network that is composed of a
series of patches in the occipital and the temporal lobes. Two of these regions, the occipital face area (OFA;
Gauthier et al., 2000) in the inferior occipital gyrus and the fusiform face area (FFA; Kanwisher et al., 1997)
in the fusiform gyrus (FG), are part of the ventral visual pathway, while a third, the superior temporal sul-
cus (STS; Haxby et al., 2002), is part of a recently suggested “social perception” visual pathway (Pitcher &
Ungerleider, 2021). BOLD activity in the FG and/or the STS correlates with the attractiveness of faces as
reported by participants (Kranz & Ishai, 2006; Pegors et al., 2015; Winston et al., 2007). For example, Kranz
and Ishai (2006) reported enhanced activation in all of the face-specific brain network areas for attractive
faces when compared to neutral or non-attractive faces. Moreover, a correlation with attractiveness was
observed in the FFA even when participants were not explicitly rating the attractiveness of the face, suggest-
ing that the beauty of a face is accessed by the human brain automatically (Chatterjee et al., 2009; Kocsor
et al., 2013). Several more advanced methods have also been used to probe facial attractiveness representa-
tions in the visual system. For example, a study using multivariate pattern analysis (MVPA), a technique that
assesses spatial patterns of BOLD signal rather than activation per se, found that patterns of BOLD signal in
FFA encode the perceived beauty of faces (Yang et al., 2021). Further, using representational similarity analy-
sis (RSA) with electroencephalographic (EEG) data, Kaiser and Nyga (2020) found that facial attractiveness
judgments are reflected in signals recorded as early as 150 ms after stimulus onset.
However, other studies, including meta-analytic studies, did not find any relationship between activity in
face-selective regions and the appraisal of faces (Bzdok et al., 2011; Chuan-Peng et al., 2020). In addition,
the strong link between stimulus features of a face and attractiveness ratings (and thus high “shared taste”
across observers) makes it difficult to assess the degree to which these reported modulations reflect stimulus
differences or true sensitivity to aesthetic appeal.
Similar to faces, a number of studies using visual scenes have also found modulations of ventral visual
regions correlated with aesthetic appeal. In one study, a mixture of indoor and outdoor scenes rated as

112
Neural correlates of visual appeal

pleasing were found to correlate with higher activity in a region of the parahippocampal gyrus (PHG; Yue
et al., 2007) that shows selectivity for scenes (the parahippocampal “place” area; PPA; Epstein & Kanwisher,
1998). Yet given the important distinction between natural kinds and cultural artefacts, it likely makes sense
to separate out effects observed with natural landscapes versus those with humanmade environments and
architecture. Within studies that have specifically focused on natural landscapes, one study that asked observ-
ers to rate how “sublime” the landscapes were found that more sublime images were correlated with more
activity in an extensive portion of VOT stretching from the FG to PHG and the hippocampus (Ishizu &
Zeki, 2014). Another study also found that attractive landscapes were correlated with greater activity in the
PHG, but the activity differences only survived a stringent threshold in an object-selective region (lateral
occipitotemporal cortex) and not in the PPA per se (Pegors et al., 2015). Overall, the precise relationship of
appeal-related activity to scene-selective regions remains unclear. In a study using movies of natural land-
scapes rather than images, Isik and Vessel (2021) found that aesthetic appeal was associated with greater activ-
ity in regions of PHG but also with the middle occipital sulcus (MOS) and posterior middle temporal gyrus
(pMTG) on the lateral surface. These activations were immediately adjacent to scene-selective occipital place
area (OPA) and motion-sensitive MT, but these regions themselves (identified using independent localizers)
were not significantly modulated by appeal.
In contrast to natural landscapes, architecture is the aesthetic domain that generally encompasses human-
made places such as building exteriors or interior spaces, and ratings of aesthetic appeal of architecture tend
to be characterized by strong individual taste (e.g., Figure 6.2; Vessel et al., 2018). In a study using images
of building exteriors, Kirk et al. (2009) did not report any modulation by aesthetic appeal in early visual
regions. However, Vartanian et  al. (2013) found that activity in the middle occipital gyrus (MOG) was
correlated with beauty ratings of interior spaces and also that curvilinear spaces (which are preferred) led
to greater activation than rectilinear spaces in early visual regions (calcarine sulcus, lingual gyrus). Using
the same data, Coburn et al. (2020) additionally found that ratings of fascination, coherence (the degree to
which a scene could be easily understood), and “hominess” (which correlated strongly with naturalness),
factors that reflect peoples’ assessments of interior architectural spaces, were correlated with differences in
activity in early visual regions.
Turning to artworks, another visual aesthetic domain that shows a very low degree of shared taste across
observers (Leder et al., 2016; e.g., Figure 6.2; Vessel et al., 2018), there are again a number of reports of mod-
ulations in visual regions by aesthetic appeal. Correlations with aesthetic appeal have been reported in the
inferotemporal sulcus (ITS; Vessel et al., 2012), FG (Vartanian & Goel, 2004), and PHG (Vessel et al., 2012).
The heterogeneity of the types of artworks used (from representational paintings of landscapes, people, and
still lifes to surrealist or abstract works with few to no identifiable objects) makes it difficult to identify a sys-
tematic relationship between the features of an artwork and where one might expect modulation by appeal.
On the other hand, the low degree of agreement in Vessel et al. (2012) on which particular paintings people
found aesthetically moving makes it more likely that the observed activations reflected aesthetic appeal per se
and not visual features of the artworks.
In a study comparing several visual aesthetic domains (natural landscapes, architecture, visual art-
work), Vessel et al. (2019) found that category-selective regions in VOT tended to be more active for
appealing images of all three domains. In addition, multivoxel patterns across VOT contained a small but
significant degree of information about aesthetic appeal for all three domains. However, these patterns
were only weakly predictive of appeal and highly domain specific: the spatial pattern of activity that pre-
dicted appeal for architecture, for example, was not predictive of appeal for artwork or landscapes (and
vice versa).
Watching and enjoying a dance performance critically depends on seeing and interpreting bodies and
body movements. As such, it is not surprising that the extrastriate body area (EBA), a part of the action-
observation network (AON), is engaged by watching dance and is also modulated by rated appeal of short

113
Vessel, Ishizu and Bignardi

dance movements (Calvo-Merino et al., 2008; Cross et al., 2011). As the growing body of research on the
neuroaesthetics of dance is covered in a separate chapter (Chapter 16), it will not be dealt with in depth here.
There have been several meta-analysis studies which attempt to isolate brain systems that are reliably acti-
vated across neuroimaging studies in response to visual aesthetic appreciation regardless of task differences
(e.g., Vartanian & Skov, 2014; Boccia et al., 2016). A recent study applied activation likelihood estimation
(ALE) in order to identify the neural correlates of reactions to specific visual categories of artworks (portraits,
scenes, abstract paintings, body sculptures). It showed different content-dependent areas of the VOT are
involved in aesthetic appreciation; for example, the FFA is involved in evaluating beauty of a portrait, and a
few additional frontal and subcortical areas are also involved (Boccia et al., 2016). But other ALE studies pri-
marily emphasize the engagement of frontal and subcortical regions with aesthetic appreciation (e.g., Brown
et al., 2011; Chuan-Peng et al., 2020; Kühn & Gallinat, 2012; see subsequently).
Taken together, this body of literature suggests that aesthetic appeal can lead to modulations of portions of
the visual system in a manner that appears dependent on the precise nature of the stimulus being considered.
However, it remains unclear how these activations relate to the well-documented feature and category selec-
tivity of the higher-level visual cortex and what exactly is being computed or represented in these regions
modulated by aesthetic appeal.

Subcortical reward circuitry


Reward processing, particularly hedonic pleasure but likely also reward prediction and learning, are core
aspects of visual aesthetic appreciation. Like other abstract rewards such as money (Delgado et al., 2000) and
music (Blood & Zatorre, 2001), appealing visual images can activate regions of the basal ganglia associated
with reward processing (Aharon et al., 2001; Vartanian & Goel, 2004; Yue et al., 2007). While the poten-
tially rewarding properties of an attractive face or a resource-rich location may make it seem obvious that
such stimuli would engage the brain’s reward circuitry, it is in fact quite significant that visual stimuli such as
artwork, which have no a priori survival value associated with them, can also engage this same system (Lacey
et al., 2011).
Within the basal ganglia, modulations by aesthetic appeal appear most often in portions of the striatum.
The main components of the striatum are its dorsal part, composed of portions of the caudate and the puta-
men, and its ventral part, primarily the nucleus accumbens (NAcc). This subcortical structure is one of the
major dopamine-containing areas in the brain, receiving most of the dopaminergic projections from the
substantia nigra in its dorsal part and from the ventral tegmental area in its ventral part (Purves et al., 2018).
It is also one of the brain regions with the highest density of mu-opioid receptors, especially in its ventral
section (Meier et al., 2021).
Across a variety of natural and artefactual visual aesthetic domains, BOLD activity in the striatum has
been found to be modulated by aesthetic appeal. In several studies, activation in the striatum has been found
to correlate linearly with the attractiveness of faces (Aharon et al., 2001; Cloutier et al., 2008; Kim et al.,
2007; T. Wang et al., 2015), and there is some evidence that activation of the ventral striatum occurs early
and in a relatively automatic fashion (Kim et al., 2007). Similarly, pleasing scenes (Yue et al., 2007), sublime
natural landscapes (Ishizu & Zeki, 2014), and aesthetically appealing movies of natural landscapes (Isik &
Vessel, 2021) have also been found to modulate striatal activity. For artefactual categories, the aesthetic appeal
of artworks has been repeatedly associated with activity in the striatum (Belfi et al., 2019; Vartanian & Goel,
2004; Vessel et al., 2012), and the beauty of interior architecture has been reported to modulate activity in
the globus pallidus (Vartanian et al., 2013), an adjacent region of the basal ganglia. It has also been claimed
that interactions with artwork engage the striatum compared to non-art images regardless of aesthetic appeal,
reflecting the rewarding nature of engaging with artwork (Lacey et al., 2011). There is generally a lack of

114
Neural correlates of visual appeal

reports of aesthetically appealing dance sequences leading to striatal activation, though it is the case that
higher NAcc activity was observed for liked dance sequences before individuals were trained on the dance
moves (but not after; Kirsch et al., 2015).
However, the precise location of activations within the basal ganglia, particularly with respect to ventral
(NAcc) versus middle or dorsal (putamen, caudate) striatum, is less clear. The NAcc has been tightly linked
to reward processing, together with the medial orbitofrontal cortex (mOFC)/ventromedial prefrontal cortex
(vmPFC), and is considered a key node of the reward circuit, along with the interconnected medial, dorso-
lateral prefrontal, and orbitofrontal areas; the amygdala; and dopaminergic midbrain nuclei (Heekeren et al.,
2007; O’Doherty, 2004; Yacubian et al., 2007). BOLD increases in NAcc have been reported both for pre-
ferred faces (Kim et al., 2007) and for moving artwork (Belfi et al., 2019). Yet for visual images, activations
in the dorsal striatum are perhaps even more common, appearing for preferred faces (T. Wang et al., 2015),
scenes (Yue et al., 2007), and artwork (Vessel et al., 2012, 2019).
In the decision-making literature, NAcc has been identified with “critic” functions by representing actual
reward and reward-prediction errors (Schultz et al., 1992; Setlow et al., 2003; Wan & Peoples, 2006; though
see Knutson & Greer, 2008), whereas dorsal activations have been identified with “actor” functions of learn-
ing and habit implementation (Maia, 2009) and reward expectation (Delgado et al., 2000, 2003). Within
the music literature, it has been suggested that NAcc activation reflects peak moments of (consummatory)
hedonic pleasure, while dorsal striatal activation reflects pleasure from anticipation (e.g., prediction) of such
moments (Blood  & Zatorre, 2001; Salimpoor et  al., 2011). Yet the results of studies with visual stimuli
would argue against such a strict interpretation: aesthetically appealing artworks and landscapes have been
more directly associated with modulation in dorsal striatum, and the pleasure derived from visual aesthetic
experiences cannot be considered purely anticipatory. In addition, recent imaging work using poetry found
activation in NAcc preceding the reported onset of chills (Wassiliwizky et al., 2017). The precise function of
these different structures in aesthetically appealing experiences thus remains an open question.
Another question that remains empirically unanswered is whether the neurotransmitters that play a sig-
nificant role in striatal and especially NAcc functioning (i.e., dopamine, GABA, and opioids) also play a role
in visual aesthetic appraisal (Spee et al., 2018). On the one hand, while dopaminergic signalling has been
shown to broadly play a role in reward-based associative learning (Schultz, 1998) and has already been linked
to appealing music (Ferreri et al., 2019; Salimpoor et al., 2011), its link with visual aesthetic appraisal is yet
unknown.
On the other hand, there is at least some evidence that the endogenous opioid system may play a role in
visual appraisal. In a study where males were asked to rate female faces, blocking opioid activity (by naltrex-
one) decreased liking and wanting to highly attractive faces, while increasing opioid activity (by morphine)
led to an increased liking for the most attractive faces and an increase in exerted effort to both see highly
attractive and avoid highly unattractive faces (Chelnokova et al., 2014; though we note that the effects were
modest; see Meier et al., 2021). However, it is yet unclear whether opioidergic modulation of neural activ-
ity impacts visual aesthetic appreciation by exerting a direct effect in the ventral striatum or if it also impacts
activity in the sensory cortices, as has been previously suggested (Biederman & Vessel, 2006) based on their
distribution in these areas (Wise & Herkenham, 1982).
Another subcortical region that seems to encode information relevant for aesthetic appeal is the amyg-
dala. In primates, the amygdala has been shown to track the reward value associated with abstract visual
stimuli (Paton et al., 2006). Yet reports of amygdalar activation positively associated with aesthetic appeal in
humans are more rare. BOLD activity in the amygdala has been related to differential activity for images of
angular (versus curved) objects (Bar & Neta, 2007), to disharmonious (versus harmonious) colour combina-
tions (Ikeda et al., 2015), to beautiful versus ugly sculptures (Di Dio et al., 2007), and to facial attractiveness
(though in a nonlinear manner potentially reflecting arousal; Bzdok et al., 2011).

115
Vessel, Ishizu and Bignardi

Figure 6.4 Prefrontal and subcortical structures implicated in visual aesthetic appeal. (A) On the medial surface of
the cortex, the divisions of the medial prefrontal cortex are visible, including the anterior cingulate cortex
(ACC); medial prefrontal cortex (mPFC); and, below the gyrus rectus, the ventromedial prefrontal cortex
(vmPFC), which is contiguous with the medial orbitofrontal cortex (mOFC). (B) In a coronal section, the
substructures of the basal ganglia are visible, including the caudate, putamen, and nucleus accumbens (NAcc;
one of the structures of the ventral striatum). (C) On the ventral surface, the medial (mOFC) and lateral
(lOFC) divisions of the orbitofrontal cortex are visible.

Prefrontal cortex
Across a variety of sensory modalities, several regions of the prefrontal cortex are thought to mediate impor-
tant aspects of aesthetic experiences (e.g., Brown et al., 2011; Vessel, 2020). Yet, while visual appreciation
is no exception, as will be discussed in the following, it is also the case that there is a remarkable degree of
heterogeneity in reported engagement of the prefrontal cortex for aesthetically appealing images.
Primary among the prefrontal regions implicated by aesthetic processing are those thought to support
valuation, representation of conscious feeling/emotion, and integration of valuation with ongoing state and
goals. In particular, the ventral portion of the medial prefrontal cortex, including the ventral and anterior
medial prefrontal cortex (vMPFC, aMPFC), anterior cingulate (ACC), especially the ventral subgenual
portion, and medial orbitofrontal cortex, have been repeatedly identified as central for aesthetic appeal. We
note that summarizing fMRI findings from mOFC and vMPFC can be challenging. Not only are labelling
conventions for this part of the brain inconsistent across different subfields and labs (Ishizu, 2019; Vessel,
2020), the orbital and medial prefrontal cortex exhibit large anatomical variability across individuals (e.g.,
Fornito et al., 2008). To make matters worse, these regions are highly susceptible to spatial distortions during
fMRI due to their location above the eyes and sinus cavities; these distortions can lead to mislocalization of
an activation by up to several centimetres.
Over a decade of research suggests a close relationship between aesthetic appeal and activity within the
mOFC and vMPFC using a wide range of stimuli. Modulations of fMRI activity by aesthetic appeal have
been reported below the superior rostral sulcus (SRS) for faces (Kim et al., 2007; O’Doherty et al., 2003),
artwork (Kawabata & Zeki, 2004; Lacey et al., 2011) scenes containing central objects (Kirk, 2008), and
buildings (Kirk et al., 2009). In turn, modulations by aesthetic appeal have been reported in or above the
SRS for faces (O’Doherty, 2004; Smith et al., 2010); artwork (Lacey et al., 2011; Vartanian & Goel, 2004;
Vessel et al., 2012); abstract patterns ( Jacobsen et al., 2006); art and music (Ishizu & Zeki, 2011); bodies

116
Neural correlates of visual appeal

(Martín-Loeches et al., 2014); architecture (Vartanian et al., 2013); motion stimuli (Zeki & Stutters, 2012);
harmonious versus disharmonious colours (Ikeda et  al., 2015); and non-visual domains, including math-
ematical beauty (Zeki et al., 2014). A more dorsal portion of the MPFC has also been reported to correlate
with aesthetic preferences, at least for paintings (Vessel et al., 2012). The relationship between rated appeal
and mOFC or MPFC activity has generally been reported to be positive and monotonic, though the precise
shape is less clear.
There is a notable lack of reports of modulation by aesthetic appeal in the MPFC/mOFC for sev-
eral visual aesthetic domains, notably natural landscapes (Isik & Vessel, 2021; Yue et al., 2007) and dance
stimuli. Yet given the well-documented role of this region in valuation more generally (Bartra et al., 2013;
O’Doherty et al., 2003), it would be surprising if it were to play no role for certain types of aesthetic valua-
tion. One potential explanation for this discrepancy may be the spatial scale at which aesthetic appeal is rep-
resented, making it difficult to detect using activation methods. Another possibility is that the lack of MPFC/
mOFC activation for some domains stems from methodological or analytical differences, such as how BOLD
responses to temporally extended stimuli (e.g., dance, music) are modelled.
Studies using multivariate analysis methods, which generally do not average signal over neighbouring
voxels but rather assess whether spatial patterns across a region of cortex can be related to a psychologi-
cal state, have also identified the mOFC and vMPFC as important for a variety of sensory valuations.
Multivoxel patterns across the OFC and vMPFC were able to predict valence judgments (pleasant-to-
unpleasant) for both visual and gustatory stimuli, whereas other regions (ventral temporal cortex and
anterior insula) contained patterns that were specific to either vision or taste, respectively (Chikazoe
et al., 2014). A study of faces and natural landscapes found that a multivoxel spatial correlation meas-
ure across the vMPFC was predictive of attractiveness ratings for both domains (though there was also
a degree of domain specificity; Pegors et al., 2015), and a study of natural landscapes, architecture, and
visual artworks also found that multivoxel patterns in parts of the aMPFC and dMPFC (belonging to
the default-mode network; see the following) represented aesthetic appeal in a domain-general manner
(Vessel et al., 2019).
A recent study using intracranial EEG found that high gamma band activity in the vMPFC and OFC—a
measure of high frequency neuronal oscillations thought to reflect coordinated firing of local neural ensembles
(Lachaux et al., 2012)—was related in a domain-general fashion to the likability that participants expressed
for images of faces, paintings, and food (Lopez-Persem et al., 2020). Critically, neuronal activity was related
to rated appreciation even when individuals were not explicitly rating appeal, supporting the hypothesis that
assessments of visual appreciation occur implicitly, even when not required. Interestingly, the authors also
reported that baseline activity in the vMPFC and OFC before image presentation was significantly related to
rated appreciation, potentially reflecting the influence of an observers’ internal state on visual appreciation.
While it is difficult to obtain strong causal information on brain function in humans, two studies have
used brain stimulation to explore the potential causal role of the mOFC/vMPFC in aesthetic appeal. They
report that transcranial direct current stimulation (tDCS) designed to enhance the excitability of the MPFC
increased aesthetic ratings of paintings (Cattaneo et al., 2019; Nakamura & Kawabata, 2015). Given the dif-
ficulty in modelling how tDCS current flows through the cortex, a note of caution is warranted in interpret-
ing these results.
Another approach to studying causal relationships between cognition and brain function are lesion stud-
ies. Although there are no neuroaesthetics studies exploring change in aesthetic evaluation by patients with
an mOFC/vmPFC lesion, one study reported that patients with injuries to the ventral PFC (including the
mOFC/vmPFC) reportedly suffered impairments when making moral judgments (Young et  al., 2010).
Considering the fact that this region is engaged when judging moral beauty (Tsukiura & Cabeza, 2011), it
would be interesting to test whether such a patient population would show alteration or impairment in other
aesthetic domains.

117
Vessel, Ishizu and Bignardi

Together, this breadth of research links many diverse and highly subjective experiences of feeling beauty,
pleasure, and being moved to activity in the mOFC/vMPFC and suggests that this part of the brain con-
tains a representation that can function as a “common currency” in hedonic appraisals independently of the
source (Bartra et al., 2013; Chikazoe et al., 2014; Levy & Glimcher, 2011; O’Doherty, 2004; Pegors et al.,
2015). This common neurological basis for valuations may allow for comparison of all manner of experiences
derived from different sources. Yet there is also evidence that signals in parts of the OFC and MPFC can
represent differential aspects of various flavours of hedonically rewarding experiences (Grabenhorst & Rolls,
2011; Sescousse et al., 2013), such as those derived from primary rewards (e.g., sucrose), through association
(e.g., money), or through engaging in sense making (e.g., aesthetic experiences). How these different rep-
resentations interact to support both subjective experience and the coordinated control of behaviour during
aesthetic experiences remains to be seen.
Beyond the medial and orbital surfaces, several other prefrontal regions also appear to play important roles
in visual aesthetic appraisals. On the lateral surface, the inferior frontal gyrus (IFG), particularly the pars trian-
gularis or opercularis segments, as well as the insula, have received quite a bit of attention. As in the vMPFC/
mOFC, anatomical localization and labelling are an issue for these regions: the IFG lies on top of the insula,
and activations often appear to straddle both cortical surfaces, a problem that is exacerbated by the moder-
ate to high degree of spatial blurring that is often applied before averaging across observers. Modulation by
aesthetic appeal has been reported here for artwork (Vessel et al., 2012), abstract patterns ( Jacobsen et al.,
2006), faces (Chatterjee et al., 2009; Kim et al., 2007; O’Doherty et al., 2003), human bodies (Di Dio et al.,
2011), and abnormal scenes (containing an out-of-context object; Kirk, 2008). Interestingly, a meta-analysis
of aesthetically appealing stimuli across four modalities (vision, audition, gustation, olfaction) and 93 studies
identified the anterior insula, rather than the mOFC/MPFC, as the region most consistently engaged by
aesthetic appeal (Brown et al., 2011; though we note that many of the included studies used stimuli that are
not typically evaluated in an aesthetic manner, such as glucose, water, and sexually arousing images). While
it is unclear whether these findings reflect engagement of the anterior insula, IFG, or both, the status of the
anterior insula as a hub of the “salience network” (or ventral attention network; Seeley et al., 2007; Thomas
Yeo et al., 2011) and also as important for interoception (Critchley et al., 2004; see also Chapter 5) makes it
an important target for future study.
Also on the lateral surface, activations in the lateral OFC have been reported for aesthetically appeal-
ing faces (O’Doherty et al., 2003; Pegors et al., 2015), as well as for artworks (Vessel et al., 2012). One
magnetoencephalography (MEG) study (Cela-Conde et al., 2004) and a few brain-stimulation studies (see
Cattaneo, 2020 for a review) have suggested that the dorsolateral PFC (dlPFC) plays a role in aesthetic pro-
cessing, though few other reports have corroborated such a link (though see a recent study on musical pleas-
ure; Mas-Herrero et al., 2021). The frontopolar and frontomedian cortex have been found to respond more
to scenes containing an out-of-context object (Kirk, 2008) and to curvilinear versus rectilinear architecture
(Vartanian et al., 2013) and also to contain multivoxel information correlated with aesthetic appeal for art and
architecture (Vessel et al., 2019). The superior frontal gyrus was found to be modulated by aesthetic appeal
for artworks (Vessel et al., 2012) and architecture (Vartanian et al., 2013).

Subcortical vs. cortical representations of hedonic pleasure


Both the striatum and OFC/MPFC are key nodes of the brain’s reward circuitry. However, their functions
within that network appear to differ. In one fMRI study, both expert architects and non-experts were asked
to aesthetically evaluate images of buildings (Kirk et al., 2009). While both NAcc and OFC were modu-
lated by appeal, only cortical regions, namely the OFC and ventral ACC, were modulated by expertise.
A study by Wang et al. (2015) also found differential profiles in the striatum and OFC. They compared
qualitatively different types of beauty judgments, facial beauty (physical attractiveness), and moral beauty

118
Neural correlates of visual appeal

(abstract and higher-order beauty) and found that both types of beauty judgments commonly engaged the
mOFC/vMPFC, consistent with previous neuroaesthetics research. However, only facial beauty judgment
engaged the striatum. Activity in the striatum and OFC may also occur at temporally dissociable stages.
Kim and colleagues showed that when participants evaluate the attractiveness of a face, increases in activa-
tion in the NAcc were detectable during initial brief exposures to the image, whereas OFC activation was
more dominant for later exposures (2007). These findings are generally consistent with the view that the
ventral striatum functions as a “hedonic hotspot” supporting core pleasure (along with the ventral pallidum),
whereas the OFC supports conscious representations of affect that can influence decisions (Berridge  &
Kringelbach, 2013), potentially in concert with interoceptive information coming from the anterior insula
(Brown et al., 2011).
Interestingly, it has also been suggested that different dopamine pathways, which play an important role
in the experience of pleasure, may differentially affect activity within the striatum and OFC. The OFC and
vmPFC are targets of both the mesolimbic and mesocortical pathways, whereas the striatum is strongly mod-
ulated by the mesolimbic pathway alone. The mesolimbic pathway projects to the ventral striatum (primarily
NAcc) and hippocampus, as well as the OFC and vmPFC, whereas the mesocortical pathway transmits dopa-
mine from the ventral tegmental area to the PFC, especially the dlPFC, and eventually through the parietal,
occipital, and temporal areas. These different projections of the dopaminergic system may contribute to the
temporal dissociation between the striatum and OFC (Spee et al., 2018).

The default-mode network and strongly moving aesthetic experiences


Beyond activations of localized brain regions, strongly moving aesthetic experiences with visual art involve
activation of a brain system known as the default-mode network (DMN). Nodes of the DMN are typically
suppressed during tasks requiring external focus, such as looking at images, and are engaged by tasks that
require internally directed or self-generated thought such as autobiographical memory, prospective thinking,
or judgments of self-relevance (Andrews-Hanna et al., 2010; Axelrod et al., 2017; Fox et al., 2005; Raichle
et al., 2001). Whereas disliked artworks and artworks rated as pleasant are accompanied by suppression of the
DMN, viewing of artworks rated as strongly moving results in a release from this suppression and engage-
ment of the DMN by ongoing visual processing (Belfi et al., 2019; Vessel et al., 2012, 2013). Evidence for
likely DMN engagement has also been found using MEG: a pattern of coherence across brain regions con-
sistent with the DMN emerges 1000–1500 ms after onset of an image rated as beautiful (Cela-Conde et al.,
2013). While the precise role played by the DMN in moving experiences remains unsettled, its central role in
self-generated and internally directed thought suggests that it may become engaged by aesthetic experiences
that trigger an assessment of self-relevance or that resonate with aspects of self identity and autobiographical
memory (Vessel et al., 2013).
Interestingly, while modulation of DMN regions by aesthetic appeal have also been reported for other
visual stimuli such as faces (O’Doherty et  al., 2003), abstract visual patterns ( Jacobsen et  al., 2006), and
architecture (Vartanian et al., 2013), DMN activation has not been robustly reported for appealing landscapes
(e.g., Isik & Vessel, 2021) or dance and was reported to be deactivated by awe-inducing landscape movies
(van Elk et al., 2019). It remains an open question whether the DMN is engaged by all moving aesthetic
experiences or only in a restricted set of circumstances. On the other hand, it has been shown using multi-
variate methods that the DMN contains a representation of aesthetic appeal that generalizes across multiple
visual aesthetic domains, even if it is not, on average, more active. Multivoxel activity patterns in DMN can
be used to predict observers’ aesthetic responses, and these patterns are consistent across visual artworks,
architecture, and natural landscapes (Vessel et al., 2019). In contrast, while multivoxel patterns in higher-level
visual regions can also predict a degree of aesthetic appeal, the predictive patterns in these regions are unique
to each domain.

119
Vessel, Ishizu and Bignardi

Figure 6.5 The default-mode network (DMN) is a network of highly interconnected brain regions that are thought
to support self-referential and inwardly directed thought and are typically suppressed by tasks that require
external focus (Andrews-Hanna et al., 2010; Fox et al., 2005; Raichle et al., 2001). However, aesthetically
moving visual artworks engage DMN regions, which have also been shown to contain multivoxel patterns
that can be used to decode aesthetic appeal across multiple visual domains (Vessel et al., 2012; Vessel et al.,
2019). The core regions of the DMN include the medial prefrontal cortex (mPFC) and posterior cingulate
cortex (PCC)/precuneus on the cortical midline. The DMN also includes regions in the inferior parietal
lobule (IPL) lateral temporal cortex (LTC), and hippocampus (not shown).

Coordination across different brain systems


Given their highly integrative nature, it is not surprising that visual aesthetic experiences engage multi-
ple brain systems. Several models have tried to make sense of interactions amongst these many systems.
One popular model highlights an “aesthetic triad” of processes engaged during aesthetic experiences
(Chatterjee & Vartanian, 2014; see also Chapter 10): sensory-motor (sensation, perception, motor sys-
tem), emotion-valuation (reward, emotion, wanting/liking), and knowledge-meaning processes (sense-
making, context, culture, expertise). We would note that while a fair amount is known about the brain
systems that mediate the first two sets of processes, much less is known about knowledge-meaning neural
processes during aesthetic experiences. A second emerging model proposes a difference between ordinary
pleasures that engage sensory and reward networks and more intense or contemplative aesthetic experi-
ences that cross a threshold to become strongly moving (Brielmann & Pelli, 2018; Pelowski et al., 2017;
Vessel et al., 2013). Whereas externally focused (sensory) and internally focused (DMN) brain systems are
typically anticorrelated (Fox et al., 2005), strongly moving experiences may result from change in these
large-scale network dynamics, bringing these two systems into closer coordination (Belfi et  al., 2019;
Vessel et al., 2013). Such coordinated activity would provide a pathway by which representations of an
external stimulus could mutually interact with internally generated states, generating insight and pleas-
ure not just from an analysis of the stimulus but also from the interplay between a stimulus and internal
thought.
Although visual neuroaesthetics research has heavily relied on regional/inter-regional BOLD measure-
ments, it is also important to consider time frames of information processing to understand such coor-
dination across large systems. Recent studies argue that aesthetic appraisal may rely on the activation of

120
Neural correlates of visual appeal

two different networks—an initial aesthetic network and a delayed network—engaged within distinct time
frames (e.g., Cela-Conde et al., 2013). This study suggests that the DMN may correspond mainly to the
delayed aesthetic network processing with more cognitive analysis of a stimulus in a comparatively late stage
(1000–1500 ms). This can be compared with the mOFC/vmPFC, which may have an important role in
the initial aesthetic network occurring in the earlier stage where a fast aesthetic appreciative perception is
formed. It could be interesting to combine fMRI and M/EEG with network/connectivity analyses in order
to investigate directional influences of one brain system on another.
Given the degree of variability seen in neural responses for different visual aesthetic domains, it may be
that the degree of cross-network integration may differ depending on the domain. For example, a visual
domain like natural landscapes may be more driven by exogenous, stimulus-derived factors and thus require
less input from self-generated or self-referential processes (Isik & Vessel, 2021). Such differences in the bal-
ance between bottom-up and top-down activation may additionally be related to differences in the state of
mind evoked by a particular aesthetic task or context (e.g., Herz et al., 2020)

Why do we like what we like?


Why we like what we like is a hard question and is one that may not have any single answer. Given the cur-
rent “cartographic” state of knowledge in the field of visual neuroaesthetics (and more generally in cognitive
neuroscience; Poeppel, 2008), it is difficult to say much about why certain visual experiences come to be felt
as beautiful or moving. Keeping this in mind, we will outline some of the explanations put forward for why
humans like what they like and put forward some explanations of our own as well.
From an evolutionary perspective, it is clear that there must be some adaptive benefit for our ability
to make appraisals of a diverse array of visual stimuli, from the face of a person we see on the street, to
a remote mountain meadow, to a piece of contemporary art. Classical theories have often focused on
why humans have preferences for specific features within specific classes of objects. For example, it has
been suggested that preferences for certain facial features may reflect implicit knowledge about markers
for health or desirable genetic qualities (e.g., Thornhill & Gangestad, 1999; see also Chapter 11). Similar
claims have been made about the potential advantage of a preference for curved over angular visual fea-
tures, as an implicit mechanism for avoiding sharp objects or dangerous places (Bar & Neta, 2007). Within
environmental aesthetics, preferences for certain types of landscapes, such as ones that afford a view or that
signal resource availability (e.g., water, flora and fauna), have been touted as conferring a potential survival
advantage (Orians & Heerwagen, 1992). The aesthetic appraisal in this sense seems to play an essential
role as a deciding factor in the decisions we face in various situations. In Plato’s theory of ideas, truth,
goodness, and beauty are universal values that we pursue as ideals. According to this view, humans have a
cognitive bias that links the good with the beautiful and the true with the beautiful, which is pervasive in
human societies (Dion et al., 1972). Here, one could say that aesthetic appraisal has the function of bring-
ing in a kind of emotional information that allows us to judge the “rightness” or “goodness” of things or
situations. This determinant can be applied to a variety of entities, not only faces or landscapes but also
morality or arts.
However, this approach often suffers from equivocal empirical evidence and extensive post hoc reasoning
and largely ignores the fact, highlighted in this chapter, that people often don’t agree on what they like. Even
for faces, one of the visual domains that elicits a high degree of shared taste, evolutionary predictions are
sometimes far off from the observed data. For example, sexual facial dimorphic traits seem to be preferred
in a culturally specific fashion, suggesting cultural, and not evolutionary, pressure to influence shared liking
(Scott et al., 2014). Although there is some evidence suggesting that variation in preferences for specific
features can be accounted for by genetic variance (e.g., sexual dimorphism in faces; Zietsch et al., 2015), the

121
Vessel, Ishizu and Bignardi

assumption that evolutionary pressures acted on the formation of preferences for very specific features has so
far received little support. If anything, we now know that preferences are influenced mainly by environmen-
tal, and not genetic, factors, as studies on the heritability and development of preferences suggest (Germine
et al., 2015; Rodway et al., 2016; Sutherland et al., 2020). Further, even when individuals tend to agree on
what they like across cultures, the empirical evidence for a link between such shared appraisals and signalling
properties of what is appraised is contradictory. For example, the link between rated attractiveness of faces (or
the trait predictor of attractiveness, e.g., symmetry) and actual health has been recently disputed (Foo et al.,
2017; Pound et al., 2014).
Here we would argue that a set of alternative views, some of which have been proposed to account
for non-human animal appraisal, could potentially explain aspects of why individuals like what they like
(Ryan & Cummings, 2013). First, it is possible that sensory biases, as in the non-human animal kingdom,
could account for some of the shared appraisals in humans. For example, the female tungara frog auditory
system tends to be biased toward conspecific calls with a lower-than-average frequency range (Ryan et al.,
2019). These sets of biases constitute part of the experiential world of non-human animals (their umwelt
or environment; Ryan, 2011). Being animals, it is thus possible that similar biases constituting the human
environment could potentially and partially explain why individuals tend to show some shared preferences.
However, here it is important to note that, as for animals, these biases would constitute only an initial set of
pressures from which environmental exposure and contextual information would then operate (Ryan et al.,
2019), making aesthetic evaluations a more flexible and time-dependent (Nadal & Chatterjee, 2019), though
not entirely unconstrained, mechanism (Bignardi et al., 2020).
More generally, we would argue that preferences for specific features may not be at the right level of
granularity for linking to evolutionary pressures and that evolution may instead have acted upon more
general-purpose mechanisms. A  mechanism that promotes learning about, and modelling (Brielmann  &
Dayan, 2021), one’s environment is a promising potential target for such evolutionary pressures. The goal of
such a mechanism would be to enable better predictions in the future.
According to this view, curiosity about our world, and the pleasure we derive from making sense of it,
act as a motivational drive that pushes humans to keep exploring, keep learning, and keep developing new
ways of parsing our world (Biederman & Vessel, 2006; Koelsch et al., 2018; Loewenstein, 1994; Schoeller &
Perlovsky, 2016). The consequence is that humans can get pleasure from understanding and also that what
we find pleasurable can change as we learn. First, optimal interest is directed to potential sources of learn-
ing, often corresponding to moments of high surprise or high uncertainty. Then, the subsequent decrease
in entropy that comes from making sense (“aha” moments associated with resolution, the creation of new
categories, paradigm shifts, etc.) is experienced as pleasurable and signals that something has potentially been
learned (Van de Cruys et al., 2021). Although such information-seeking behaviours, which incur the cost
of receiving information that will not directly result in receiving rewards, can also be observed in animals
(Wang & Hayden, 2019), they are likely most elaborated in humans.
Such explanations, including learning-based theories for visual aesthetic appeal, can potentially explain
appraisals that happen at many different levels of representation, interacting with different types of envi-
ronmental exposures over the lifespan. An infant may get pleasure from learning about simple shapes and
their parent’s faces. A child may actively seek certain environments over others, exposing themselves to new
sources of visual input. A teenager may find that expressionist art, subversive graffiti, or Japanese manga reso-
nates with their rapidly changing view of their place in the world. A neuroscientist captivated by complexity
may find that two-photon microscopy captures neural signalling in a way they had not previously considered.
While it may not be the case that every experience of beauty is also an “aha” moment, relating aesthetics to
information processing is a promising direction for future research.
Our proposed explanations are far from complete. Testing whether such explanations carry true explana-
tory power will require neuroaesthetics to develop new paradigms and increasingly quantitative models. The

122
Neural correlates of visual appeal

task of understanding “why we like what we like” is a complicated puzzle that still requires many pieces to
be completed.

Some concluding remarks and open questions


Interest in the neural basis of visual aesthetic appreciation has increased dramatically over the past 25 years,
and while the field is still relatively small, much has been learned. Despite the highly individual and subjec-
tive nature of aesthetic experiences, a host of tools and methods have been developed that are facilitating
systematic study. The psychological nature of aesthetic appeal has itself been much better clarified, and there
exists a general outline of the brain systems that are involved. Yet it is clear that there is still very much to
learn. One crucial limitation, highlighted here at the beginning of the chapter, is represented by the same
strength that facilitated the rise of the systematic study of aesthetic appeal: “methodological variation.” In
particular, the array of different questions and tasks that researchers have introduced, and the clear lack of
a priori agreement on the terminology used, puts constraints on the possibility of interpreting results from
different studies. In spite of such weaknesses, we now have a general map of where in the brain one can find
modulations by aesthetic appeal. Yet we still understand very little about the underlying computations and
representations that can be found in these regions.
In the following, we highlight some key open questions in visual neuroaesthetics:

• How do representations of what we see inform representations of aesthetic appeal?


• How does activation in visual pathways relate to aesthetic appeal? Is appeal computed or represented in
these local regions, or is appeal computed elsewhere?
• How much of the variability in reported activation of the mOFC/vMPFC and IFG/insula is due to
methodological concerns versus reflective of true differences in mechanisms?
• Are the MPFC and/or the DMN engaged for all aesthetically appealing visual experiences, or for only
for a subset?
• What are the precise roles of various brain regions (MPFC, OFC, IFG/insula) in computing and repre-
senting aesthetic appeal?
• What brain structures support knowledge-meaning processes during aesthetic experiences?
• How might we expect neural engagement to change when moving from images to more ecologically
valid aesthetic experiences?
• How are differences in brain structure and function linked to individual differences in aesthetic
experiences?
• What are the etiological (i.e., genetic and environmental) sources of variation in aesthetic appeal?
• How can empirical findings from neuroaesthetics be translated into humanities and philosophy to con-
tribute to our understanding of human’s aesthetic and artistic activities?
• What are the differences, if any, between the neural systems supporting experiences of beauty, pleasure,
and reward?
• Is it possible to study aesthetic appraisals/beauty judgements in infants and animals?
• How can neuroaesthetics benefit from increasing incorporation of neuropsychopharmacological and
causal (lesion, TMS) approaches?

Acknowledgements
We thank Margherita Bignardi for help with Figure 6.1 and MacKenzie Trupp for thoughtful discussions
and for editing an earlier version of this manuscript. This work was partially supported by the BMBF and
Max Planck Society to G.B.

123
Vessel, Ishizu and Bignardi

Note
1 We note that complexity and order have been investigated using both objective and subjective frameworks. The for-
mer approach seeks to quantify complexity and order from a stimulus, while the latter posits that these constructs can
only be understood by taking the perceiver’s subjective judgment into account.

References
Aharon, I., Etcoff, N., Ariely, D., Chabris, C. F., O’Connor, E., & Breiter, H. C. (2001). Beautiful faces have variable reward
value: FMRI and behavioral evidence. Neuron, 32(3), 537–551. https://doi.org/10.1016/S0896-6273(01)00491-3
Andrews-Hanna, J. R., Reidler, J. S., Sepulcre, J., Poulin, R., & Buckner, R. L. (2010). Functional-anatomic fractiona-
tion of the brain’s default network. Neuron, 65(4), 550–562. https://doi.org/10.1016/j.neuron.2010.02.005
Anglada-Tort, M., & Skov, M. (2020). What counts as aesthetics in science? A bibliometric analysis and visualization of
the scientific literature from 1970 to 2018. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/
aca0000350
Appleton, J. (1975). The experience of landscape. John Wiley.
Ariely, D., & Norton, M. I. (2008). How actions create—Not just reveal—Preferences. Trends in Cognitive Sciences, 12(1),
13–16. https://doi.org/10.1016/j.tics.2007.10.008
Arnheim, R. (1954). Art and visual perception: A psychology of the creative eye (pp. xii, 408). University of California Press.
Avidan, G., Harel, M., Hendler, T., Ben-Bashat, D., Zohary, E., & Malach, R. (2002). Contrast sensitivity in human
visual areas and its relationship to object recognition. Journal of Neurophysiology, 87(6), 3102–3116. https://doi.
org/10.1152/jn.2002.87.6.3102
Axelrod, V., Rees, G., & Bar, M. (2017, December). The default network and the combination of cognitive processes
that mediate self-generated thought. Nature Human Behaviour, 1, 896–910. https://doi.org/10.1038/s41562-017-
0244-9
Bar, M.,  & Neta, M. (2006). Humans prefer curved visual objects. Psychological Science, 17(8), 645–648. https://doi.
org/10.1111/j.1467-9280.2006.01759.x
Bar, M., & Neta, M. (2007). Visual elements of subjective preference modulate amygdala activation. Neuropsychologia,
45(10), 2191–2200. https://doi.org/10.1016/j.neuropsychologia.2007.03.008
Bartra, O., McGuire, J. T., & Kable, J. W. (2013). The valuation system: A coordinate-based meta-analysis of BOLD fMRI
experiments examining neural correlates of subjective value. NeuroImage, 76, 412–427. https://doi.org/10.1016/j.
neuroimage.2013.02.063
Belfi, A. M., Vessel, E. A., Brielmann, A. A., Isik, A. I., Chatterjee, A., Leder, H., Pelli, D. G., & Starr, G. G. (2019).
Dynamics of aesthetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.
org/10.1016/j.neuroimage.2018.12.017
Berlyne, D. E. (1958). The influence of complexity and novelty in visual figures on orienting responses. Journal of Experi-
mental Psychology, 55(3), 289–296. https://doi.org/10.1037/h0043555
Berlyne, D. E. (1970). Novelty, complexity, and hedonic value. Perception and Psychophysics, 8(5), 279–286. https://doi.
org/10.3758/BF03212593
Berlyne, D. E. (1971). Aesthetics and psychobiology (pp. xiv, 336). Appleton-Century-Crofts.
Berlyne, D. E. (1974). The new experimental aesthetics. In D. E. Berlyne (Ed.), Studies in the new experimental aesthetics:
Steps toward an objective psychology of aesthetic appreciation (pp. 1–25). Hemisphere Publishing Corporation.
Berridge, K. C., & Kringelbach, M. L. (2013). Neuroscience of affect: Brain mechanisms of pleasure and displeasure.
Current Opinion in Neurobiology, 23(3), 294–303. https://doi.org/10.1016/j.conb.2013.01.017
Bertamini, M., Palumbo, L., Gheorghes, T. N., & Galatsidas, M. (2016). Do observers like curvature or do they dislike
angularity? British Journal of Psychology, 107(1), 154–178. https://doi.org/10.1111/bjop.12132
Bertamini, M., Rampone, G., Makin, A. D. J., & Jessop, A. (2019). Symmetry preference in shapes, faces, flowers and
landscapes. PeerJ, 7, e7078. https://doi.org/10.7717/peerj.7078
Biederman, I., & Vessel, E. A. (2006). Perceptual pleasure and the brain: A novel theory explains why the brain craves
information and seeks it through the senses. American Scientist, 94(3), 247–253. JSTOR.
Bignardi, G., Ishizu, T., & Zeki, S. (2021). The differential power of extraneous influences to modify aesthetic judgments
of biological and artifactual stimuli. PsyCh Journal, 10(2), 190–199. https://doi.org/10.1002/pchj.415
Bignardi, G., Ticini, L. F., Smit, D.,  & Polderman, T. J. (2020). Domain-specific and domain-general genetic and
environmental effects on the intensity of visual aesthetic appraisal. Psychiatry, Ar.Xiv. https://doi.org/10.31234/osf.
io/79nbq

124
Neural correlates of visual appeal

Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Boccia, M., Barbetti, S., Piccardi, L., Guariglia, C., Ferlazzo, F., Giannini, A. M., & Zaidel, D. W. (2016). Where does
brain neural activation in aesthetic responses to visual art occur? Meta-analytic evidence from neuroimaging studies.
Neuroscience and Biobehavioral Reviews, 60, 65–71. https://doi.org/10.1016/j.neubiorev.2015.09.009
Boynton, G. M., Demb, J. B., Glover, G. H., & Heeger, D. J. (1999). Neuronal basis of contrast discrimination. Vision
Research, 39(2), 257–269. https://doi.org/10.1016/s0042-6989(98)00113-8
Brielmann, A. A., & Dayan, P. (2021). Introducing a computational model of aesthetic value. Psychiatry, Ar.Xiv. https://
doi.org/10.31234/osf.io/eaqkc
Brielmann, A. A., & Pelli, D. G. (2018). Aesthetics. Current Biology, 28(16), R859–R863. https://doi.org/10.1016/j.
cub.2018.06.004
Brielmann, A. A., & Pelli, D. G. (2019). Intense beauty requires intense pleasure. Frontiers in Psychology, 10, 2420. https://
doi.org/10.3389/fpsyg.2019.02420
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. NeuroImage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Bzdok, D., Langner, R., Caspers, S., Kurth, F., Habel, U., Zilles, K., Laird, A., & Eickhoff, S. B. (2011). ALE meta-
analysis on facial judgments of trustworthiness and attractiveness. Brain Structure and Function, 215(3–4), 209–223.
https://doi.org/10.1007/s00429-010-0287-4
Calvo-Merino, B., Jola, C., Glaser, D. E., & Haggard, P. (2008). Towards a sensorimotor aesthetics of performing art.
Consciousness and Cognition, 17(3), 911–922. https://doi.org/10.1016/j.concog.2007.11.003
Cattaneo, Z. (2020). Neural correlates of visual aesthetic appreciation: Insights from non-invasive brain stimulation.
Experimental Brain Research, 238(1), 1–16. https://doi.org/10.1007/s00221-019-05685-x
Cattaneo, Z., Ferrari, C., Schiavi, S., Alekseichuk, I., Antal, A., & Nadal, M. (2019). Medial prefrontal cortex involvement
in aesthetic appreciation of paintings: A tDCS study. Cognitive processing, 89, 01234567. https://doi.org/10.1007/
s10339-019-00936-9
Cela-Conde, C. J., García-Prieto, J., Ramasco, J. J., Mirasso, C. R., Bajo, R., Munar, E., Flexas, A., Del-Pozo, F., &
Maestú, F. (2013). Dynamics of brain networks in the aesthetic appreciation. Proceedings of the National Academy of Sci-
ences of the United States of America, 110(Suppl. 2), 10454–10461. https://doi.org/10.1073/pnas.1302855110
Cela-Conde, C. J., Marty, G., Maestú, F., Ortiz, T., Munar, E., Fernández, A., Roca, M., Rosselló, J., & Quesney, F.
(2004). Activation of the prefrontal cortex in the human visual aesthetic perception. Proceedings of the National Academy
of Sciences of the United States of America, 101(16), 6321–6325. https://doi.org/10.1073/pnas.0401427101
Chatterjee, A., Thomas, A., Smith, S. E., & Aguirre, G. K. (2009). The neural response to facial attractiveness. Neuropsy-
chology, 23(2), 135–143. https://doi.org/10.1037/a0014430
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Chelnokova, O., Laeng, B., Eikemo, M., Riegels, J., Løseth, G., Maurud, H., Willoch, F., & Leknes, S. (2014). Rewards
of beauty: The opioid system mediates social motivation in humans. Molecular Psychiatry, 19(7), 746–747. https://doi.
org/10.1038/mp.2014.1
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology:
CB, 29(23), 4084–4092.e4. https://doi.org/10.1016/j.cub.2019.09.067
Chikazoe, J., Lee, D. H., Kriegeskorte, N., & Anderson, A. K. (2014). Population coding of affect across stimuli, modali-
ties and individuals. Nature Neuroscience, 17(8), 1114–1122. https://doi.org/10.1038/nn.3749
Chuan-Peng, H., Huang, Y., Eickhoff, S. B., Peng, K.,  & Sui, J. (2020). Seeking the “beauty center” in the brain:
A meta-analysis of fMRI studies of beautiful human faces and visual art. Cognitive, Affective and Behavioral Neuroscience,
20(6), 1200–1215. https://doi.org/10.3758/s13415-020-00827-z
Clark, A. (2018). A nice surprise? Predictive processing and the active pursuit of novelty. Phenomenology and the Cognitive
Sciences, 17(3), 521–534. https://doi.org/10.1007/s11097-017-9525-z
Cloutier, J., Heatherton, T. F., Whalen, P. J., & Kelley, W. M. (2008). Are attractive people rewarding? Sex differences in
the neural substrates of facial attractiveness. Journal of Cognitive Neuroscience, 20(6), 941–951. https://doi.org/10.1162/
jocn.2008.20062
Coburn, A., Vartanian, O., Kenett, Y. N., Nadal, M., Hartung, F., Hayn-Leichsenring, G., Navarrete, G., González-
Mora, J. L., & Chatterjee, A. (2020). Psychological and neural responses to architectural interiors. Cortex; a Journal
Devoted to the Study of the Nervous System and Behavior, 126, 217–241. https://doi.org/10.1016/j.cortex.2020.01.009

125
Vessel, Ishizu and Bignardi

Critchley, H. D., Wiens, S., Rotshtein, P., Öhman, A., & Dolan, R. J. (2004). Neural systems supporting interoceptive
awareness. Nature Neuroscience, 7(2), 189–195. https://doi.org/10.1038/nn1176
Cross, E. S., Kirsch, L., Ticini, L. F.,  & Schütz-Bosbach, S. (2011, September). The impact of aesthetic evalua-
tion and physical ability on dance perception. Frontiers in Human Neuroscience, 5, 1–10. https://doi.org/10.3389/
fnhum.2011.00102
Delgado, M. R., Locke, H. M., Stenger, V. A., & Fiez, J. A. (2003). Dorsal striatum responses to reward and punishment:
Effects of valence and magnitude manipulations. Cognitive, Affective and Behavioral Neuroscience, 3(1), 27–38. https://
doi.org/10.3758/CABN.3.1.27
Delgado, M. R., Nystrom, L. E., Fissell, C., Noll, D. C., & Fiez, J. A. (2000). Tracking the hemodynamic responses
to reward and punishment in the striatum. Journal of Neurophysiology, 84(6), 3072–3077. https://doi.org/10.1152/
jn.2000.84.6.3072
Di Dio, C., Canessa, N., Cappa, S. F., & Rizzolatti, G. (2011). Specificity of esthetic experience for artworks: An fMRI
study. Frontiers in Human Neuroscience, 5, 139. https://doi.org/10.3389/fnhum.2011.00139
Di Dio, C., Macaluso, E., & Rizzolatti, G. (2007). The golden beauty: Brain response to Classical and Renaissance sculp-
tures. PlOS ONE, 2(11), e1201—e1201. https://doi.org/10.1371/journal.pone.0001201
Dion, K., Berscheid, E., & Walster, E. (1972). What is beautiful is good. Journal of Personality and Social Psychology, 24(3),
285–290. https://doi.org/10.1037/h0033731
Epstein, R.,  & Kanwisher, N. (1998). A  cortical representation of the local visual environment. Nature, 392(6676),
598–601. https://doi.org/10.1038/33402
Fechner, G. T. (1876). Vorschule der Aesthetik. http://archive.org/details/vorschulederaest12fechuoft. Breitkopf.
Ferreri, L., Mas-Herrero, E., Zatorre, R. J., Ripollés, P., Gomez-Andres, A., Alicart, H., Olivé, G., Marco-Pallarés, J.,
Antonijoan, R. M., Valle, M., Riba, J., & Rodriguez-Fornells, A. (2019). Dopamine modulates the reward experi-
ences elicited by music. Proceedings of the National Academy of Sciences of the United States of America, 116(9), 3793–3798.
https://doi.org/10.1073/pnas.1811878116
Fleurian, de R., & Pearce, M. (2020). Chills in music: A systematic review. Psychiatry, Ar.Xiv. https://doi.org/10.31234/
osf.io/yc6d8
Foo, Y. Z., Simmons, L. W., & Rhodes, G. (2017). Predictors of facial attractiveness and health in humans. Scientific
Reports, 7(1), 39731. https://doi.org/10.1038/srep39731
Fornito, A., Wood, S. J., Whittle, S., Fuller, J., Adamson, C., Saling, M. M., Velakoulis, D., Pantelis, C., & Yücel, M.
(2008). Variability of the paracingulate sulcus and morphometry of the medial frontal cortex: Associations with corti-
cal thickness, surface area, volume, and sulcal depth. Human Brain Mapping, 29(2), 222–236. https://doi.org/10.1002/
hbm.20381
Fox, M. D., Snyder, A. Z., Vincent, J. L., Corbetta, M., Van Essen, D. C., & Raichle, M. E. (2005). The human brain is
intrinsically organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences
of the United States of America, 102(27), 9673–9678. https://doi.org/10.1073/pnas.0504136102
Friston, K. (2010). The free-energy principle: A  unified brain theory? Nature Reviews. Neuroscience, 11(2), 127–138.
https://doi.org/10.1038/nrn2787
Gauthier, I., Tarr, M. J., Moylan, J., Skudlarski, P., Gore, J. C., & Anderson, A. W. (2000). The fusiform “face area” is
part of a network that processes faces at the individual level. Journal of Cognitive Neuroscience, 12(3), 495–504. https://
doi.org/10.1162/089892900562165
Germine, L., Russell, R., Bronstad, P. M., Blokland, G. A. M., Smoller, J. W., Kwok, H., Anthony, S. E., Nakayama, K.,
Rhodes, G., & Wilmer, J. B. (2015). Individual aesthetic preferences for faces are shaped mostly by environments, not
genes. Current Biology: CB, 25(20), 2684–2689. https://doi.org/10.1016/j.cub.2015.08.048
Gold, B. P., Mas-Herrero, E., Zeighami, Y., Benovoy, M., Dagher, A., & Zatorre, R. J. (2019). Musical reward prediction
errors engage the nucleus accumbens and motivate learning. Proceedings of the National Academy of Sciences of the United
States of America, 116(8), 3310–3315. https://doi.org/10.1073/pnas.1809855116
Goodale, M. A., & Milner, A. D. (1992). Separate visual pathways for perception and action. Trends in Neurosciences, 15(1),
20–25. https://doi.org/10.1016/0166-2236(92)90344-8
Grabenhorst, F., & Rolls, E. T. (2011). Value, pleasure and choice in the ventral prefrontal cortex. Trends in Cognitive Sci-
ences, 15(2), 56–67. https://doi.org/10.1016/j.tics.2010.12.004
Graf, L. K. M.,  & Landwehr, J. R. (2015). A  dual-process perspective on fluency-based aesthetics: The pleasure-
interest model of aesthetic liking. Personality and Social Psychology Review, 19(4), 395–410. https://doi.org/10.1177/
1088868315574978
Grill-Spector, K., & Weiner, K. S. (2014). The functional architecture of the ventral temporal cortex and its role in cat-
egorization. Nature Reviews Neuroscience, 15(8), 536–548. https://doi.org/10.1038/nrn3747

126
Neural correlates of visual appeal

Haxby, J. V., Hoffman, E. A., & Gobbini, M. I. (2002). Human neural systems for face recognition and social commu-
nication. Society of biological psychiatry, 51(1), 59–67.
Heekeren, H. R., Wartenburger, I., Marschner, A., Mell, T., Villringer, A.,  & Reischies, F. M. (2007). Role of
ventral striatum in reward-based decision making. NeuroReport, 18(10), 951–955. https://doi.org/10.1097/
WNR.0b013e3281532bd7
Herz, N., Baror, S., & Bar, M. (2020). Overarching states of mind. Trends in Cognitive Sciences, 24(3), 184–199. https://
doi.org/10.1016/j.tics.2019.12.015
Hönekopp, J. (2006). Once more: Is beauty in the eye of the beholder? Relative contributions of private and shared
taste to judgments of facial attractiveness. Journal of Experimental Psychology. Human Perception and Performance, 32(2),
199–209. https://doi.org/10.1037/0096-1523.32.2.199
Hübner, R., & Fillinger, M. G. (2019). Perceptual balance, stability, and aesthetic appreciation: Their relations depend on
the picture type. i-Perception, 10(3), 2041669519856040. https://doi.org/10.1177/2041669519856040
Ikeda, T., Matsuyoshi, D., Sawamoto, N., Fukuyama, H., & Osaka, N. (2015). Color harmony represented by activ-
ity in the medial orbitofrontal cortex and amygdala. Frontiers in Human Neuroscience, 9( July), 382–382. https://doi.
org/10.3389/fnhum.2015.00382
Ishizu, T. (2019, August 12). Functional neuroimaging in empirical aesthetics and neuroaesthetics. In The Oxford hand-
book of empirical aesthetics. https://doi.org/10.1093/oxfordhb/9780198824350.013.14.
Ishizu, T., & Zeki, S. (2011). Toward A brain-based theory of beauty. PLOS ONE, 6(7), e21852. https://doi.org/10.1371/
journal.pone.0021852
Ishizu, T., & Zeki, S. (2014). A neurobiological enquiry into the origins of our experience of the sublime and beautiful.
Frontiers in Human Neuroscience, 8, 891. https://doi.org/10.3389/fnhum.2014.00891
Isik, A. I., & Vessel, E. A. (2021). From visual perception to aesthetic appeal: Brain responses to aesthetically appeal-
ing natural landscape movies. Frontiers in Human Neuroscience, 15, 676032. https://doi.org/10.3389/fnhum.2021.
676032
Izuma, K., Matsumoto, M., Murayama, K., Samejima, K., Sadato, N., & Matsumoto, K. (2010). Neural correlates of
cognitive dissonance and choice-induced preference change. Proceedings of the National Academy of Sciences of the United
States of America, 107(51), 22014–22019. https://doi.org/10.1073/pnas.1011879108
Jacobsen, T., & Höfel, L. (2002). Aesthetic judgments of novel graphic patterns: Analyses of individual judgments. Per-
ceptual and Motor Skills, 95(3 Pt. 1), 755–766. https://doi.org/10.2466/pms.2002.95.3.755
Jacobsen, T., Schubotz, R. I., Höfel, L., & Cramon, D. Y. V. (2006). Brain correlates of aesthetic judgment of beauty.
NeuroImage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010
Kaiser, D., & Nyga, K. (2020). Tracking cortical representations of facial attractiveness using time-resolved representa-
tional similarity analysis. Scientific Reports, 10(1), 16852. https://doi.org/10.1038/s41598-020-74009-9
Kant, I. (1790/1987). Critique of judgment. Hackett Publishing.
Kanwisher, N. (2010). Functional specificity in the human brain: A  window into the functional architecture of the
mind. Proceedings of the National Academy of Sciences of the United States of America, 107(25), 11163–11170. https://doi.
org/10.1073/pnas.1005062107
Kanwisher, N., McDermott, J.,  & Chun, M. M. (1997). The fusiform face area: A  module in human extrastriate
cortex specialized for face perception. Journal of Neuroscience, 17(11), 4302–4311. https://doi.org/10.1523/
JNEUROSCI.17-11-04302.1997
Kaplan, R., & Kaplan, S. (1995). The experience of nature: A psychological perspective. Ulrich’s bookstore.
Kaplan, S. (1992). Environmental preference in a knowledge-seeking, knowledge-using organism. In J. H. Barkow
Cosmides, L. & J. Tooby (Eds.), The adapted mind: Evolutionary psychology and the generation of culture (pp. 581–598).
Oxford University Press.
Kaplan, S., Kaplan, R., & Wendt, J. S. (1972). Rated preference and complexity for natural and urban visual material.
Perception and Psychophysics, 12(4), 354–356. https://doi.org/10.3758/BF03207221
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Keefe, B. D., Gouws, A. D., Sheldon, A. A., Vernon, R. J. W., Lawrence, S. J. D., McKeefry, D. J., Wade, A. R., &
Morland, A. B. (2018). Emergence of symmetry selectivity in the visual areas of the human brain: FMRI responses
to symmetry presented in both frontoparallel and slanted planes. Human Brain Mapping, 39(10), 3813–3826. https://
doi.org/10.1002/hbm.24211
Kim, H., Adolphs, R., O’Doherty, J. P., & Shimojo, S. (2007). Temporal isolation of neural processes underlying face
preference decisions. Proceedings of the National Academy of Sciences of the United States of America, 104(46), 18253–
18258. https://doi.org/10.1073/pnas.0703101104

127
Vessel, Ishizu and Bignardi

Kirk, U. (2008). The neural basis of object-context relationships on aesthetic judgment. PlOS ONE, 3(11), e3754.
https://doi.org/10.1371/journal.pone.0003754
Kirk, U., Skov, M., Christensen, M. S., & Nygaard, N. (2009). Brain correlates of aesthetic expertise: A parametric fMRI
study. Brain and Cognition, 69(2), 306–315. https://doi.org/10.1016/j.bandc.2008.08.004
Kirsch, L. P., Dawson, K., & Cross, E. S. (2015). Dance experience sculpts aesthetic perception and related brain circuits.
Annals of the New York Academy of Sciences, 1337(1), 130–139. https://doi.org/10.1111/nyas.12634
Knutson, B.,  & Greer, S. M. (2008) [Review]. Anticipatory affect: Neural correlates and consequences for choice.
Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 363(1511), 3771–3786. https://doi.
org/10.1098/rstb.2008.0155
Kocsor, F., Feldmann, A., Bereczkei, T., & Kállai, J. (2013). Assessing facial attractiveness: Individual decisions and evo-
lutionary constraints. Socioaffective Neuroscience and Psychology, 3(1), 21432. https://doi.org/10.3402/snp.v3i0.21432
Koelsch, S., Vuust, P., & Friston, K. (2018). Predictive processes and the peculiar case of music. Trends in Cognitive Sciences,
23(1), 63–77. https://doi.org/10.1016/j.tics.2018.10.006
Kranz, F., & Ishai, A. (2006). Face perception is modulated by sexual preference. Current Biology, 16(1), 63–68. https://
doi.org/10.1016/j.cub.2005.10.070
Kühn, S., & Gallinat, J. (2012). The neural correlates of subjective pleasantness. NeuroImage, 61(1), 289–294. https://doi.
org/10.1016/j.neuroimage.2012.02.065
Lacey, S., Hagtvedt, H., Patrick, V. M., Anderson, A., Stilla, R., Deshpande, G., Hu, X., Sato, J. R., Reddy, S.,  &
Sathian, K. (2011). Art for reward’s sake: Visual art recruits the ventral striatum. NeuroImage, 55(1), 420–433. https://
doi.org/10.1016/j.neuroimage.2010.11.027
Lachaux, J. P., Axmacher, N., Mormann, F., Halgren, E., & Crone, N. E. (2012). High-frequency neural activity and
human cognition: Past, present and possible future of intracranial EEG research. Progress in Neurobiology, 98(3), 279–
301. https://doi.org/10.1016/j.pneurobio.2012.06.008
Laeng, B., Eidet, L. M., Sulutvedt, U., & Panksepp, J. (2016). Music chills: The eye pupil as a mirror to music’s soul.
Consciousness and Cognition, 44, 161–178. https://doi.org/10.1016/j.concog.2016.07.009
Langlois, J. H., Kalakanis, L., Rubenstein, A. J., Larson, A., Hallam, M.,  & Smoot, M. (2000). Maxims or myths
of beauty? A  meta-analytic and theoretical review. Psychological Bulletin, 126(3), 390–423. https://doi.
org/10.1037/0033-2909.126.3.390
Leder, H., Goller, J., Rigotti, T., & Forster, M. (2016). Private and shared taste in art and face appreciation. Frontiers in
Human Neuroscience, 10, 155. https://doi.org/10.3389/fnhum.2016.00155
Levitan, C. A., Winfield, E. C., & Sherman, A. (2019). Grumpy toddlers and dead pheasants: Visual art preferences
are predicted by preferences for the depicted objects. Psychology of Aesthetics, Creativity, and the Arts, 14(2), 155–161.
https://doi.org/10.1037/aca0000240
Levy, D. J., & Glimcher, P. W. (2011). Comparing apples and oranges: Using reward-specific and reward-general sub-
jective value representation in the brain. Journal of Neuroscience, 31(41), 14693–14707. https://doi.org/10.1523/
JNEUROSCI.2218-11.2011
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 75–98.
https://doi.org/10.1037/0033-2909.116.1.75
Lopez-Persem, A., Bastin, J., Petton, M., Abitbol, R., Lehongre, K., Adam, C., Navarro, V., Rheims, S., Kahane, P.,
Domenech, P., & Pessiglione, M. (2020). Four core properties of the human brain valuation system demonstrated in
intracranial signals. Nature Neuroscience, 23(5), 664–675. https://doi.org/10.1038/s41593-020-0615-9
Maia, T. V. (2009). Reinforcement learning, conditioning, and the brain: Successes and challenges. Cognitive, Affective and
Behavioral Neuroscience, 9(4), 343–364. https://doi.org/10.3758/CABN.9.4.343
Mallon, B., Redies, C., & Hayn-Leichsenring, G. U. (2014). Beauty in abstract paintings: Perceptual contrast and statisti-
cal properties. Frontiers in Human Neuroscience, 8(161), 161. https://doi.org/10.3389/fnhum.2014.00161
Martindale, C. (1984). The pleasures of thought: A theory of cognitive hedonics. Journal of Mind and Behavior, 5(1).
Martinez, J. E., Funk, F., & Todorov, A. (2020). Quantifying idiosyncratic and shared contributions to judgment. Behavior
Research Methods, 52(4), 1428–1444. https://doi.org/10.3758/s13428-019-01323-0
Martín-Loeches, M., Hernández-Tamames, J. A., Martín, A., & Urrutia, M. (2014). Beauty and ugliness in the bod-
ies and faces of others: An fMRI study of person esthetic judgement. Neuroscience, 277, 486–497. https://doi.
org/10.1016/j.neuroscience.2014.07.040
Mas-Herrero, E., Dagher, A., Farrés-Franch, M., & Zatorre, R. J. (2021). Unraveling the temporal dynamics of reward
signals in music-induced pleasure with TMS. Journal of Neuroscience, 41(17), 3889–3899. https://doi.org/10.1523/
JNEUROSCI.0727-20.2020
McManus, I. C. (1980). The aesthetics of simple figures. British Journal of Psychology, 71(4), 505–524. https://doi.
org/10.1111/j.2044-8295.1980.tb01763.x

128
Neural correlates of visual appeal

McManus, I. C., Cook, R., & Hunt, A. (2010). Beyond the Golden Section and normative aesthetics: Why do individu-
als differ so much in their aesthetic preferences for rectangles? Psychology of Aesthetics, Creativity, and the Arts, 4(2),
113–126. https://doi.org/10.1037/a0017316
McManus, I. C., Jones, A. L., & Cottrell, J. (1981). The aesthetics of colour. Perception, 10(6), 651–666. https://doi.
org/10.1068/p100651
Meier, I. M., Eikemo, M., & Leknes, S. (2021). The role of mu-opioids for reward and threat processing in humans:
Bridging the gap from preclinical to clinical opioid drug studies. Current Addiction Reports, 8(2), 306–318. https://doi.
org/10.1007/s40429-021-00366-8
Muth, C., & Carbon, C. C. (2013). The aesthetic aha: On the pleasure of having insights into gestalt. Acta Psychologica,
144(1), 25–30. https://doi.org/10.1016/j.actpsy.2013.05.001
Muth, C., Hesslinger, V. M., & Carbon, C.-C. (2015). The appeal of challenge in the perception of art: How ambiguity,
solvability of ambiguity, and the opportunity for insight affect appreciation. Psychology of Aesthetics, Creativity, and the
Arts, 9(3), 206–216. https://doi.org/10.1037/a0038814
Nadal, M., & Chatterjee, A. (2019). Neuroaesthetics and art’s diversity and universality. Wiley Interdisciplinary Reviews.
Cognitive Science, 10(3), e1487. https://doi.org/10.1002/wcs.1487
Nakamura, K., & Kawabata, H. (2015). Transcranial direct current stimulation over the medial prefrontal cortex and left
primary motor cortex (mPFC-lPMC) affects subjective beauty but not ugliness. Frontiers in Human Neuroscience. DEC,
9, 654–654. https://doi.org/10.3389/fnhum.2015.00654
O’Doherty, J. P. (2004). Reward representations and reward-related learning in the human brain: Insights from neuroim-
aging. Current Opinion in Neurobiology, 14(6), 769–776. https://doi.org/10.1016/j.conb.2004.10.016
O’Doherty, J. P., Winston, J., Critchley, H. D., Perrett, D., Burt, D. M., & Dolan, R. J. (2003). Beauty in a smile: The
role of medial orbitofrontal cortex in facial attractiveness. Neuropsychologia, 41(2), 147–155. https://doi.org/10.1016/
s0028-3932(02)00145-8
Oliva, A., & Torralba, A. (2001). Modeling the shape of the scene: A holistic representation of the spatial envelope. Inter-
national Journal of Computer Vision, 42(3), 145–175. https://doi.org/10.1023/A:1011139631724
Orians, G. H., & Heerwagen, J. H. (1992). Evolved responses to landscapes. In J. H. Barkow, L. Cosmides, & J. Tooby
(Eds.), The adapted mind: Evolutionary psychology and the generation of culture (pp. 555–579). Oxford University Press.
Palmer, S. E., & Schloss, K. B. (2010). An ecological valence theory of human color preference. Proceedings of the National
Academy of Sciences of the United States of America, 107(19), 8877–8882. https://doi.org/10.1073/pnas.0906172107
Palmer, S. E., Schloss, K. B., & Sammartino, J. (2013). Visual aesthetics and human preference. Annual Review of Psychol-
ogy, 64(1), 77–107. https://doi.org/10.1146/annurev-psych-120710-100504
Pasupathy, A., & Connor, C. E. (2002). Population coding of shape in area V4. Nature Neuroscience, 5(12), 1332–1338.
https://doi.org/10.1038/nn972
Paton, J. J., Belova, M. A., Morrison, S. E., & Salzman, C. D. (2006). The primate amygdala represents the positive and
negative value of visual stimuli during learning. Nature, 439(7078), 865–870. https://doi.org/10.1038/nature04490
Pearce, M. T., Zaidel, D. W., Vartanian, O., Skov, M., Leder, H., Chatterjee, A., & Nadal, M. (2016). Neuroaesthetics:
The cognitive neuroscience of aesthetic experience. Perspectives on Psychological Science: A Journal of the Association for
Psychological Science, 11(2), 265–279. https://doi.org/10.1177/1745691615621274
Pegors, T. K., & Epstein, R. A. (2011). Sensitivity to the aesthetic value of scenes in the parahippocampal place area.
Journal of Vision, 11(11), 1119–1119. https://doi.org/10.1167/11.11.1119
Pegors, T. K., Kable, J. W., Chatterjee, A., & Epstein, R. A. (2015). Common and unique representations in pFC for
face and place attractiveness. Journal of Cognitive Neuroscience, 27(5), 959–973. https://doi.org/10.1162/jocn_a_00777
Pelowski, M., Markey, P. S., Forster, M., Gerger, G., & Leder, H. (2017). Move me, astonish me . . . delight my eyes
and brain: The Vienna integrated model of top-down and bottom-up processes in art perception (VIMAP) and cor-
responding affective, evaluative, and neurophysiological correlates. Physics of Life Reviews, 21, 80–125. https://doi.
org/10.1016/j.plrev.2017.02.003
Pitcher, D., Charles, L., Devlin, J. T., Walsh, V., & Duchaine, B. (2009). Triple dissociation of faces, bodies, and objects
in extrastriate cortex. Current Biology, 19(4), 319–324. https://doi.org/10.1016/j.cub.2009.01.007
Pitcher, D., & Ungerleider, L. G. (2021). Evidence for a third visual pathway specialized for social perception. Trends in
Cognitive Sciences, 25(2), 100–110. https://doi.org/10.1016/j.tics.2020.11.006
Poeppel, D. (2008). The cartographic imperative: Confusing localization and explanation in human brain map-
ping. Bildwelten des Wissen: Ikonographie des Gehirns. https://nyuscholars.nyu.edu/en/publications/
the-cartographic-imperative-confusing-localization-and-explanatio
Pound, N., Lawson, D. W., Toma, A. M., Richmond, S., Zhurov, A. I., & Penton-Voak, I. S. (2014). Facial fluctuating
asymmetry is not associated with childhood ill-health in a large British cohort study. Proceedings. Biological Sciences,
281(1792), 20141639. https://doi.org/10.1098/rspb.2014.1639

129
Vessel, Ishizu and Bignardi

Purves, D., Augustine, G. J., Fitzpatrick, D., Hall, W. C., LaMantia, A. S., Mooney, R. D., White, L. E., & Platt, M. L.
(2018). Neuroscience. Oxford University Press. https://books.google.at/books?id=R__8uwEACAAJ.
Raichle, M. E., MacLeod, A. M., Snyder, A. Z., Powers, W. J., Gusnard, D. A., & Shulman, G. L. (2001). A default
mode of brain function. Proceedings of the National Academy of Sciences of the United States of America, 98(2), 676–682.
https://doi.org/10.1073/pnas.98.2.676
Reber, R., Schwarz, N., & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the perceiver’s
processing experience? Personality and Social Psychology Review: An Official Journal of the Society for Personality and Social
Psychology, Inc, 8(4), 364–382. https://doi.org/10.1207/s15327957pspr0804_3
Reber, R., Winkielman, P., & Schwarz, N. (1998). Effects of perceptual fluency on affective judgments. Psychological
Science, 9(1), 45–48. https://doi.org/10.1111/1467-9280.00008
Redies, C. (2007). A universal model of esthetic perception based on the sensory coding of natural stimuli. Spatial Vision,
21(1–2), 97–117. https://doi.org/10.1163/156856807782753886
Rhodes, G. (2006). The evolutionary psychology of facial beauty. Annual Review of Psychology, 57(1), 199–226. https://
doi.org/10.1146/annurev.psych.57.102904.190208
Rodway, P., Kirkham, J., Schepman, A., Lambert, J., & Locke, A. (2016). The development of shared liking of represen-
tational but not abstract art in primary school children and their justifications for liking. Frontiers in Human Neurosci-
ence, 10, 21. https://doi.org/10.3389/fnhum.2016.00021
Ryali, C. K., Goffin, S., Winkielman, P., & Yu, A. J. (2020). From likely to likable: The role of statistical typicality in
human social assessment of faces. Proceedings of the National Academy of Sciences of the United States of America, 117(47),
29371–29380. https://doi.org/10.1073/pnas.1912343117
Ryan, M. J. (2011). The brain as a source of selection on the social niche: Examples from the psychophysics of mate
choice in túngara frogs. Integrative and Comparative Biology, 51(5), 756–770. https://doi.org/10.1093/icb/icr065
Ryan, M. J., & Cummings, M. E. (2013). Perceptual biases and mate choice. Annual Review of Ecology, Evolution, and
Systematics, 44(1), 437–459. https://doi.org/10.1146/annurev-ecolsys-110512-135901
Ryan, M. J., Page, R. A., Hunter, K. L., & Taylor, R. C. (2019). ‘Crazy love’: Nonlinearity and irrationality in mate
choice. Animal Behaviour, 147, 189–198. https://doi.org/10.1016/j.anbehav.2018.04.004
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Sammartino, J., & Palmer, S. E. (2012). Aesthetic issues in spatial composition: Effects of vertical position and perspective
on framing single objects. Journal of Experimental Psychology. Human Perception and Performance, 38(4), 865–879. https://
doi.org/10.1037/a0027736
Sasaki, Y., Vanduffel, W., Knutsen, T., Tyler, C., & Tootell, R. (2005). Symmetry activates extrastriate visual cortex in
human and nonhuman primates. Proceedings of the National Academy of Sciences of the United States of America, 102(8),
3159–3163. https://doi.org/10.1073/pnas.0500319102
Schepman, A., Rodway, P., Pullen, S. J., & Kirkham, J. (2015). Shared liking and association valence for representational
art but not abstract art. Journal of Vision, 15(5), 11. https://doi.org/10.1167/15.5.11
Schmidhuber, J. (2010). Formal theory of creativity and intrinsic motivation (1990–2010). IEEE Transactions on Autono-
mous Mental Development, 2(3), 1–18.
Schoeller, F., & Perlovsky, L. (2016). Aesthetic chills: Knowledge-acquisition, meaning-making, and aesthetic emotions.
Frontiers in Psychology, 7(AUG), 1–16. https://doi.org/10.3389/fpsyg.2016.01093
Schultz, W. (1998). Predictive reward signal of dopamine neurons. Journal of Neurophysiology, 80(1), 1–27. https://doi.
org/10.1152/jn.1998.80.1.1
Schultz, W., Apicella, P., Scarnati, E.,  & Ljungberg, T. (1992). Neuronal activity in monkey ventral striatum
related to the expectation of reward. Journal of Neuroscience, 12(12), 4595–4610. https://doi.org/10.1523/
JNEUROSCI.12-12-04595.1992
Scott, I. M., Clark, A. P., Josephson, S. C., Boyette, A. H., Cuthill, I. C., Fried, R. L., Gibson, M. A., Hewlett, B. S.,
Jamieson, M., Jankowiak, W., Honey, P. L., Huang, Z., Liebert, M. A., Purzycki, B. G., Shaver, J. H., Snodgrass, J.
J., Sosis, R., Sugiyama, L. S., Swami, V., . . . Penton-Voak, I. S. (2014). Human preferences for sexually dimorphic
faces may be evolutionarily novel. Proceedings of the National Academy of Sciences of the United States of America, 111(40),
14388–14393. https://doi.org/10.1073/pnas.1409643111
Seeley, W. W., Menon, V., Schatzberg, A. F., Keller, J., Glover, G. H., Kenna, H., Reiss, A. L., & Greicius, M. D. (2007).
Dissociable intrinsic connectivity networks for salience processing and executive control. Journal of Neuroscience, 27(9),
2349–2356. https://doi.org/10.1523/JNEUROSCI.5587-06.2007

130
Neural correlates of visual appeal

Sescousse, G., Caldú, X., Segura, B., & Dreher, J. C. (2013). Processing of primary and secondary rewards: A quantitative
meta-analysis and review of human functional neuroimaging studies. In Neuroscience and Biobehavioral Reviews, 37(4),
681–696. https://doi.org/10.1016/j.neubiorev.2013.02.002
Setlow, B., Schoenbaum, G., & Gallagher, M. (2003). Neural encoding in ventral striatum during olfactory discrimina-
tion learning. Neuron, 38(4), 625–636. https://doi.org/10.1016/S0896-6273(03)00264-2
Silvia, P. J. (2005). What is interesting?—Exploring the appraisal structure of interest. Emotion, 5(1), 89–102. https://doi.
org/10.1037/1528-3542.5.1.89
Silvia, P. J. (2008). Interest—The curious emotion. Current Directions in Psychological Science, 17(1), 57–60. https://doi.
org/10.1111/j.1467-8721.2008.00548.x
Slater, A., Von der Schulenburg, C., Brown, E., Badenoch, M., Butterworth, G., Parsons, S., & Samuels, C. (1998).
Newborn infants prefer attractive faces. Infant Behavior and Development, 21(2), 345–354. https://doi.org/10.1016/
S0163-6383(98)90011-X
Smith, D. V., Hayden, B. Y., Truong, T. K., Song, A. W., Platt, M. L., & Huettel, S. A. (2010). Distinct value sig-
nals in anterior and posterior ventromedial prefrontal cortex. Journal of Neuroscience, 30(7), 2490–2495. https://doi.
org/10.1523/JNEUROSCI.3319-09.2010
Spee, B., Ishizu, T., Leder, H., Mikuni, J., Kawabata, H., & Pelowski, M. (2018). Neuropsychopharmacological aesthet-
ics: A theoretical consideration of pharmacological approaches to causative brain study in aesthetics and art. Progress
in Brain Research, 237, 343–372. https://doi.org/10.1016/bs.pbr.2018.03.021
Spehar, B., Walker, N., & Taylor, R. P. (2016). Taxonomy of individual variations in aesthetic responses to fractal pat-
terns. Frontiers in Human Neuroscience, 10( July), 350. https://doi.org/10.3389/fnhum.2016.00350
Sutherland, C. A. M., Burton, N. S., Wilmer, J. B., Blokland, G. A. M., Germine, L., Palermo, R., Collova, J. R., &
Rhodes, G. (2020). Individual differences in trust evaluations are shaped mostly by environments, not genes. Proceed-
ings of the National Academy of Sciences of the United States of America, 117(19), 10218–10224. https://doi.org/10.1073/
pnas.1920131117
Thornhill, R., & Gangestad, S. W. (1999). Facial attractiveness. Trends in Cognitive Sciences, 3(12), 452–460. https://doi.
org/10.1016/s1364-6613(99)01403-5
Tinio, P. P. L., & Leder, H. (2009). Natural scenes are indeed preferred, but image quality might have the last word.
Psychology of Aesthetics, Creativity, and the Arts, 3(1), 52–56. https://doi.org/10.1037/a0014835
Tschacher, W., Greenwood, S., Kirchberg, V., Wintzerith, S., van den Berg, K., & Tröndle, M. (2012). Physiological cor-
relates of aesthetic perception of artworks in a museum. Psychology of Aesthetics, Creativity, and the Arts, 6(1), 96–103.
https://doi.org/10.1037/a0023845
Tsukiura, T.,  & Cabeza, R. (2011). Shared brain activity for aesthetic and moral judgments: Implications for the
beauty-is-good stereotype. Social Cognitive and Affective Neuroscience, 6(1), 138–148. https://doi.org/10.1093/scan/
nsq025
Ungerleider, L. G., & Mishkin, M. (1982). Two cortical visual systems. In M. A. G. D. J. Ingle & R. J. W. Mansfield
(Eds.), Analysis of visual behavior (pp. 549–586). MIT Press.
Van de Cruys, S., Damiano, C., Boddez, Y., Król, M., Goetschalckx, L., & Wagemans, J. (2021). Visual affects: Linking
curiosity, aha-erlebnis, and memory through information gain. Cognition, 212, 104698. https://doi.org/10.1016/j.
cognition.2021.104698
van Elk, M., Arciniegas Gomez, M. A., van der Zwaag, W., van Schie, H. T., & Sauter, D. (2019). The neural correlates
of the awe experience: Reduced default mode network activity during feelings of awe. Human Brain Mapping, 40(12),
3561–3574. https://doi.org/10.1002/hbm.24616
Van Essen, D. C., Anderson, C. H., & Felleman, D. J. (1992). Information processing in the primate visual system: An
integrated systems perspective. Science, 255(5043), 419–423. https://doi.org/10.1126/science.1734518
Vartanian, O., & Goel, V. (2004). Neuroanatomical correlates of aesthetic preference for paintings. NeuroReport, 15(5),
893–897. https://doi.org/10.1097/01.wnr.00001
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov, M.
(2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceedings of the
National Academy of Sciences of the United States of America, 110(Suppl. 2), 10446–10453. https://doi.org/10.1073/
pnas.1301227110
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Rostrup, N., Skov, M., Corradi,
G., & Nadal, M. (2017). Preference for curvilinear contour in interior architectural spaces: Evidence from experts
and nonexperts. Psychology of Aesthetics, Creativity, and the Arts, 13(1), 110–116. https://doi.org/10.1037/aca0
000150

131
Vessel, Ishizu and Bignardi

Vartanian, O.,  & Skov, M. (2014). Neural correlates of viewing paintings: Evidence from a quantitative meta-
analysis of functional magnetic resonance imaging data. Brain and Cognition, 87(1), 52–56. https://doi.org/10.1016/j.
bandc.2014.03.004
Vessel, E. A. (2020). Neuroaesthetics. In Reference module in neuroscience and biobehavioral psychology. Elsevier. https://doi.
org/10.1016/B978-0-12-809324-5.24104-7
Vessel, E. A., Isik, A. I., Belfi, A. M., Stahl, J. L., & Starr, G. G. (2019). The default-mode network represents aesthetic
appeal that generalizes across visual domains. Proceedings of the National Academy of Sciences of the United States of America,
116(38), 19155–19164. https://doi.org/10.1073/pnas.1902650116
Vessel, E. A., Maurer, N., Denker, A. H., & Starr, G. G. (2018). Stronger shared taste for natural aesthetic domains than
for artifacts of human culture. Cognition, 179, 121–131. https://doi.org/10.1016/j.cognition.2018.06.009
Vessel, E. A., & Rubin, N. (2010). Beauty and the beholder: Highly individual taste for abstract, but not real-world
images. Journal of Vision, 10(2), 18.1–1814. https://doi.org/10.1167/10.2.18
Vessel, E. A., Starr, G. G., & Rubin, N. (2012). The brain on art: Intense aesthetic experience activates the default mode
network. Frontiers in Human Neuroscience, 6, 66. https://doi.org/10.3389/fnhum.2012.00066
Vessel, E. A., Starr, G. G., & Rubin, N. (2013). Art reaches within: Aesthetic experience, the self and the default mode
network. Frontiers in Neuroscience, 7, 258. https://doi.org/10.3389/fnins.2013.00258
Wan, X., & Peoples, L. L. (2006). Firing patterns of accumbal neurons during a Pavlovian-conditioned approach task.
Journal of Neurophysiology, 96(2), 652–660. https://doi.org/10.1152/jn.00068.2006
Wang, M. Z.,  & Hayden, B. Y. (2019). Monkeys are curious about counterfactual outcomes. Cognition, 189, 1–10.
https://doi.org/10.1016/j.cognition.2019.03.009
Wang, T., Mo, L., Mo, C., Tan, L. H., Cant, J. S., Zhong, L., & Cupchik, G. (2015). Is moral beauty different from
facial beauty? Evidence from an fMRI study. Social Cognitive and Affective Neuroscience, 10(6), 814–823. https://doi.
org/10.1093/scan/nsu123
Wassiliwizky, E., Koelsch, S., Wagner, V., Jacobsen, T., & Menninghaus, W. (2017). The emotional power of poetry:
Neural circuitry, psychophysiology and compositional principles. Social Cognitive and Affective Neuroscience, 12(8),
1229–1240. https://doi.org/10.1093/scan/nsx069
Westerman, S. J., Gardner, P. H., Sutherland, E. J., White, T., Jordan, K., Watts, D.,  & Wells, S. (2012). Product
design: Preference for rounded versus angular design elements. Psychology and Marketing, 29(8), 595–605. https://doi.
org/10.1002/mar.20546
Winston, J. S., O’Doherty, J. P., Kilner, J. M., Perrett, D. I., & Dolan, R. J. (2007). Brain systems for assessing facial
attractiveness. Neuropsychologia, 45(1), 195–206. https://doi.org/10.1016/j.neuropsychologia.2006.05.009
Wise, S. P., & Herkenham, M. (1982). Opiate receptor distribution in the cerebral cortex of the rhesus monkey. Science,
218(22), 387–389. https://doi.org/10.1126/science.6289441
Yacubian, J., Sommer, T., Schroeder, K., Gläscher, J., Braus, D. F.,  & Büchel, C. (2007). Subregions of the ventral
striatum show preferential coding of reward magnitude and probability. NeuroImage, 38(3), 557–563. https://doi.
org/10.1016/j.neuroimage.2007.08.007
Yang, T., Formuli, A., Paolini, M., & Zeki, S. (2021). The neural determinants of beauty. European Journal of Neuroscience,
55(1), 91–106. https://doi.org/10.1111/ejn.15543
Yeo, B. T., Krienen, F. M., Sepulcre, J., Sabuncu, M. R., Lashkari, D., Hollinshead, M., Roffman, J. L., Smoller, J.
W., Zöllei, L., Polimeni, J. R., Fischl, B., Liu, H., & Buckner, R. L. (2011). The organization of the human cer-
ebral cortex estimated by intrinsic functional connectivity. Journal of Neurophysiology, 106(3), 1125–1165. https://doi.
org/10.1152/jn.00338.2011
Young, L., Bechara, A., Tranel, D., Damasio, H., Hauser, M., & Damasio, A. (2010). Damage to ventromedial prefron-
tal cortex impairs judgment of harmful intent. Neuron, 65(6), 845–851. https://doi.org/10.1016/j.neuron.2010.
03.003
Yue, X., Vessel, E. A.,  & Biederman, I. (2007). The neural basis of scene preferences. NeuroReport, 18(6), 525–529.
https://doi.org/10.1097/WNR.0b013e328091c1f9
Zajonc, R. B. (1968). Attitudinal effects of mere exposure. Journal of Experimental Social Psychology, 9(2), 1–28. https://
doi.org/10.1016/0022-1031(71)90078-3
Zeki, S. (2011). Splendors and miseries of the brain: Love, creativity, and the quest for human happiness. John Wiley & Sons.
Zeki, S.,  & Chén, O. Y. (2020). The Bayesian-Laplacian brain. European Journal of Neuroscience, 51(6), 1441–1462.
https://doi.org/10.1111/ejn.14540
Zeki, S., Javier, A., & Mylonas, D. (2020). The biological basis of the experience and categorization of colour. European
Journal of Neuroscience, 51(2), 670–680. https://doi.org/10.1111/ejn.14557

132
Neural correlates of visual appeal

Zeki, S., Romaya, J. P., Benincasa, D. M. T., & Atiyah, M. F. (2014). The experience of mathematical beauty and its
neural correlates. Frontiers in Human Neuroscience, 8(1 FEB), 68–68. https://doi.org/10.3389/fnhum.2014.00068
Zeki, S., & Stutters, J. (2012). A brain-derived metric for preferred kinetic stimuli. Open Biology, 2(2), 120001. https://
doi.org/10.1098/rsob.120001
Zietsch, B. P., Lee, A. J., Sherlock, J. M., & Jern, P. (2015). Variation in women’s preferences regarding male facial mas-
culinity is better explained by genetic differences than by previously identified context-dependent effects. Psychological
Science, 26(9), 1440–1448. https://doi.org/10.1177/0956797615591770

133
7
AUDITORY PLEASURE ELICITED BY
MUSIC
Ernest Mas-Herrero

Sounds warn us of potential dangers (e.g., a storm approaching, a snow avalanche, or the presence of a preda-
tor) and help us detect possible resources (e.g., a stream of water or prey). In addition, our ability to produce
a large variety of sounds and our capacity to extract and learn regularities has given rise to language: crucial
for social interaction, communication of complex thoughts, and, ultimately, essential for the spectacular suc-
cess of our species.
Besides its direct and pragmatic survival utility, sounds may also be the source of intense pleasurable aes-
thetic responses. For example, natural environmental sounds not only alert us of potential threats or benefits
but may also be found relaxing and pleasant. Who has not been hypnotized by the sound of waves breaking
along the shore, the leaves rustling in the wind, or the rain falling on the roof ? Similarly, we may find some
voices particularly appealing and relaxing (e.g., the voice of our favourite radio/podcast host or our favourite
actor). However, if there is one auditory aesthetic response that stands out above all the others, it is, without
doubt, the one elicited by music.
Nowadays, it is almost impossible to spend a day without listening to music, either at home, on TV, shop-
ping, or while having drinks at a bar. Music constantly surrounds us. However, the presence of music in our
societies is not new. Indeed, human beings have been enjoying music for a long time. The earliest musical
instruments found in archaeological excavations (a bone flute and two ivory flutes) date back to the Upper
Palaeolithic period, more than 40,000 years ago (Conard et al., 2009; see also Chapter 29). These findings
demonstrate that humans had already developed a musical tradition by that time, one that has endured until
this day.
Despite music’s ubiquity and universality, the reason music evolved and prospered remains a matter of
debate, one that has intrigued many scientists and philosophers. Some authors have suggested that music
appeared as a prelinguistic form of communication (Fitch, 2006). Others hold that music has a sexual-selec-
tion basis similar to songbirds’ songs (Miller, 2000; Ravignani, 2018). And others posit that music is nothing
more than a consequence of the adaptations required for language and other high-order cognitive functions,
devoid of evolutionary purpose in and of itself (Pinker, 1997).
However, what is indubitable is that music strongly impacts us. Indeed, music is generally ranked as
one of the most pleasurable experiences in life (Dubé & Bel, 2003; Mehr et al., 2019). But how does the
human brain translate a structured sequence of sounds, with no apparent biological advantage, into pleasure?
The investigation of music from a neuroscientific perspective is indeed a challenging enterprise. Because of
the lack of animal models, the neural mechanisms of music-induced pleasure have remained elusive until the

134 DOI: 10.4324/9781003008675-8


Auditory pleasure

development of modern neuroimaging techniques that allowed for the non-invasive measurement of brain
activity in humans. Although this field is still in its early stages (at least compared to the neuroscience of
primary pleasures such as food), a reasonable number of studies have addressed this question over the last
20 years. In the following sections, I review the main findings to emerge from this body of work.

Music perception
Music appreciation begins with the decoding of acoustic features and encoding them into neural firing pat-
terns (for a detailed description of this process, see Chapter 15). This process is essential to the representation
of sounds in our brains, categorizing and identifying the tones we are listening to, establishing temporal and
structural patterns, and so on. Sounds are nothing more than vibrations propagating as waves through the
air or other media (e.g., water). And musical sounds are not that different from other sounds at the stage of
early perception, sharing most of the initial processing mechanisms with other auditory stimuli. Striking the
strings of a guitar makes them vibrate and generate musical sound waves that propagate through the air and
ultimately hit the listeners’ cochlea. There, through a complex cascade of mechanical reactions, air vibra-
tions are transformed into neural (electrical) signals. This information then travels via the auditory nerve to
the central nervous system, where it is progressively transformed in the brainstem and the thalamus until it
reaches the auditory cortex (located in the superior temporal gyrus; STG). This auditory-specific cortical
region is located in temporal cortices and responsible for the more fine-grained analysis of acoustic features.
Notably, contrary to speech, music is preferentially processed in the right hemisphere (Zatorre et al., 2002).
This distinction may be driven by complementary specialization of the two hemispheres for fine acoustic
modulations in the temporal vs. the spectral domain, with the right side showing the highest sensitivity to
spectral information crucial for pitch-based aspects of music (Albouy et al., 2020; Zatorre & Belin, 2001).
It is important to note that music, unlike other sounds, relies not only on the processing of discrete fea-
tures (e.g., pitch, time, or timbre) but also on identifying temporal and structural patterns. To do so, the audi-
tory cortex does not work in isolation but through collaboration with many different brain regions located
in the cortex, particularly the posterior parietal cortex, the inferior frontal gyrus, and motor regions. Com-
munication among this set of cortical areas is crucial for identifying temporal and structural patterns of pitch
and rhythm to ultimately create musical percepts. For instance, the interplay between the STG and inferior
frontal gyrus (IFG) is involved in the recognition of musical patterns, melodic expectations, and auditory
working memory (Albouy et al., 2019; Herholz et al., 2012). Likewise, interactions with parietal, frontal,
and motor regions support the perception of rhythm and rhythmic expectations and predictions (Chen et al.,
2008; Thaut et al., 2014).
A great model for understanding the neurobiology of music perception is provided by individuals suffer-
ing from amusia or tone-deafness. Congenital amusia is a lifelong musical disorder characterized by pith per-
ception deficits, with neuro-genetic underpinnings, that affects 1.5% of the population, according to recent
reports (Peretz & Vuvan, 2017). Individuals suffering from amusia have difficulties identifying out-of-tune
and off-key notes and distinguishing familiar melodies without the aid of lyrics. They just do not get music,
no matter how much time they invest in musical training (Lebrun et al., 2012; Wilbiks et al., 2016). Yet they
may correctly recognize voices or environmental sounds and have no apparent deficits in language processing
and speech. These deficits are associated with a reduced communication between frontotemporal cortices,
including the IFG and the STG, particularly in the right hemisphere, and likely occur because of abnormal
developmental processes (Loui et al., 2009).
Impaired music perception in the amusical brain may limit the appreciation of music as an aesthetic expe-
rience. Indeed, amusical individuals are less interested in musical activities and are less likely to enjoy music
than control-matched individuals (Mcdonald & Stewart, 2008; Omigie et al., 2012). Other evidence comes
from patients developing amusia after brain damage (also known as acquired amusia; Sihvonen et al., 2019),

135
Ernest Mas-Herrero

which may be concurrent with a sudden inability to experience pleasure from music (Hirel et al., 2014;
Mazzucchi et al., 1982; Satoh et al., 2016). However, there may still be some room for musical pleasure to
occur in these individuals since some amusical individuals do report experiencing considerable pleasure from
exposure to music (Omigie et al., 2012; Wilbiks et al., 2016). Particularly astonishing is the case of Tim
Falconer, a unique case in Wilbiks and colleagues’ (2016) study of a self-described “musicophile” despite suf-
fering from amusia. His motivation was such that he followed an 18-month program of formal vocal training
to overcome his condition. Unfortunately, despite his efforts, Tim did not show substantial improvement
in rhythm and meter perception after his training (Wilbiks et al., 2016). The mechanisms through which
individuals like Tim may drive pleasure from music despite their striking musical perceptual deficits are still
unclear but point to at least a partial dissociation between music perception and appreciation.

Neural circuits involved in musical pleasure

The common reward circuitry


The cortical network previously described plays a key role in auditory cognition and perception. It helps
us understand and make sense of music, but this is only one part of the story. What circuits are behind
the rewarding feelings of music? Psychologists and neuroscientists have been particularly intrigued by how
rewards are processed in the brain, especially in response to primary sensory pleasures (e.g., food) that
(1) possess clear evolutionary advantages and (2) are shared with other organisms. A great deal of research in
both humans and nonhuman animals supports the existence of a core reward system in the brain involved in
many facets of reward, from motivation to pleasure and from pleasure to learning (Berridge & Kringelbach,
2015; see Chapter 3). The identification of an anatomically identifiable reward circuit in the brain dates back
to 1954 when James Olds and Peter Milner reported one of the most important studies in the neuroscience
of reward. They surgically implanted electrodes into rats’ brains through which electrical impulses were
delivered. The rats were placed in a cage where a lever controlled the delivery of current to the electrodes
so that rats could self-stimulate their brains by simply pressing that lever. By testing different parts of the
brain, Olds and Milner observed that at specific locations, a rat would press the lever “to stimulate itself in
these places frequently and regularly for long periods of time if permitted to do so” (Olds & Milner, 1954).
Subsequent studies have shown the rewarding effects of stimulating various subcortical and cortical
sites in several species, including humans (Olds, 1969). Intracranial recordings in nonhuman primates and
human neuroimaging studies investigating brain responses to both primary (e.g., food and sex) and second-
ary rewards (e.g., money) have further shown that these different reward experiences, irrespective of their
nature, engage a common set of brain regions, constituting what is known as the reward circuit (Bartra et al.,
2013; Sescousse et al., 2013). The reward circuit is highly consistent across species, phylogenetically old, and
responsible for the feelings of pleasure and the reinforcement of biologically relevant behaviours such as eat-
ing or sex. This circuitry includes several brain regions, but the striatum (particularly the nucleus accumbens;
NAcc), the orbitofrontal cortex (OFC), and dopaminergic neurons located in the ventral tegmental area/
substantia nigra appear especially important (Haber & Knutson, 2010; Kelley & Berridge, 2002).
Does this phylogenetically ancient reward circuitry respond to high-order pleasures such as music? The
first study investigating this question was conducted by Blood and colleagues in 1999. The authors used
positron emission tomography (PET) to examine regional cerebral blood flow changes—considered a proxy
of brain activation—elicited by affective responses to music. Ten participants listened to a novel melody and
five versions of that melody in which different chords, differing in their degree of dissonance, were used as
accompaniment. Using this approach, the authors aimed to modulate the pleasantness of the melody by add-
ing more or less dissonance to the melody’s harmonic structure. Their results revealed that changes of activa-
tion in the OFC were associated with the subjective degree of pleasure reported by the participants. Thus,

136
Auditory pleasure

the study showed, for the first time, that despite its complexity and sophistication, the functional mechanisms
employed by music reward were comparable to that of other, adaptive sensory pleasures, tapping into a com-
mon neural circuitry for computing pleasure and reward.
In this first attempt, the authors elegantly manipulated the stimuli’s unpleasantness by manipulating dis-
sonance. Their approach, however, was not optimized to maximize the individual participant’s pleasure but
rather their displeasure. So, although individuals reported liking more the consonant condition than the
different dissonant versions, the stimuli used hardly evoked intense feelings of pleasure in the listeners. One
way to solve this problem is to use real music known to elicit strong emotional reactions. Yet individuals vary
greatly in their musical preferences: a piece of music can be highly pleasant for one individual but unpalat-
able to another. To overcome this limitation, Blood and Zatorre conducted a second study two years later
(2001). They now asked participants to bring their own favourite music, known to evoke intense feelings of
pleasure, into the lab. Specifically, the researchers asked for music that evoked chills, goosebumps, or shivers
down the spine. Because chills and goosebumps are clear, discrete, and highly reproducible events that are
generally associated with particularly intense pleasurable music excerpts, they are often used as an index of
musical pleasure experiences (Grewe et al., 2005; Pelofi et al., 2021; Salimpoor et al., 2009). As a control,
each participant also listened to one of the other participants’ pieces which they knew would not elicit chills
in them. That way, the authors ensured that differences in acoustic features that could characterize chill-
inducing music did not drive the neural responses associated with chills; instead, those would be directly
related to each individual’s subjective experience. The results of the study revealed that the occurrence of
chills was associated with the engagement of a distributed network that involved cortical and subcortical
brain regions engaged in the computation of reward and emotion, such as the insula, the anterior cingulate
cortex (ACC), and the orbitofrontal cortex, as well as the ventral striatum.
Since these first seminal studies, a large body of neuroimaging research has investigated the neural cor-
relates of music-induced pleasure. In a recent meta-analysis, we showed that independently of the experi-
mental design used, intense feelings of pleasure to music are consistently associated with the engagement of
the core reward circuitry, particularly the insula, striatum (specifically the NAcc), and orbitofrontal cortex
(Mas-Herrero, Maini et al., 2021). Notably, these structures were also engaged in studies investigating food-
induced pleasure, consistent with the existence of a common set of brain regions underlying the rewarding
feelings of any rewarding event and experience—from primary to aesthetic rewards.
Another piece of evidence supporting the role of the phylogenetically ancient reward circuitry in musi-
cal reward comes from an investigation by Salimpoor and colleagues in 2011. Here, the authors aimed to
investigate whether music could lead to the release of dopamine, a neurotransmitter well known for its role
in reward processing. The different regions of the reward system communicate through dopamine, and
dopamine is also responsible for the addictive properties of drugs (Volkow et al., 2017). Would dopamine
also mediate the rewarding properties of music? To test this hypothesis, Salimpoor and colleagues used PET
to non-invasively measure dopamine release in a group of volunteers while listening to either pleasant or
non-pleasant music, using a similar approach as that pioneered by Blood and Zatorre (2001). Their results
elegantly revealed that highly pleasurable pieces of music induced greater dopamine release in both ventral
and dorsal portions of the striatum compared to the neutral control music. Notably, dopaminergic release
levels associated with music listening predicted the number and intensity of participants’ chills: the greater
the release of striatal dopamine, the more pleasure participants experienced. This study further supported
the hypothesis that musical reward relies on similar brain mechanisms as those identified in studies of food
or drug pleasures.
We have recently extended these findings by showing that dopamine is not only correlated with musical
reward, as shown by Salimpoor and colleagues, but that dopamine release is causally related to music enjoy-
ment. Neuroimaging methods are correlational, and thus they are not capable of teasing apart those brain
regions that are directly involved in generating the hedonic experience of music from those that are only

137
Ernest Mas-Herrero

modulated by this experience (e.g., responding as a consequence of pleasure rather than being its cause). To
disentangle whether dopaminergic striatal circuits are the cause or consequence of music reward experiences,
it is necessary to modulate the functioning of this circuitry and investigate if that leads to changes in reward-
related measures. Nowadays, it is possible to modulate human brain activity non-invasively by applying
transcranial magnetic stimulation (TMS). TMS is a non-invasive brain stimulation technique that uses mag-
netic fields to cause temporal electric current changes in specific brain areas. It is basically a big magnet that
generates magnetic pulses, which will go through the scalp, hit the target region, and there cause a change in
electric current. Yet TMS has a limited range of action and is only effective to stimulate cortical areas, and
therefore, we cannot directly stimulate subcortical structures such as the striatum.
Nevertheless, because the striatum is highly connected with cortical brain regions (Haber & Knutson,
2010), striatal functioning can be indirectly modulated by applying TMS over certain cortical regions, specifi-
cally over the left dorsolateral prefrontal cortex (DLPFC; Strafella et al., 2001). Stimulation of the DLPFC
may propagate until it reaches subcortical striatal regions via white matter paths connecting both, result-
ing in alterations of these regions as well. Taking advantage of this procedure, we recently tested whether
modulation of striatal circuits through excitatory and inhibitory TMS stimulation over the left DLPFC could
actually enhance or disrupt music-induced pleasure, respectively (Mas-Herrero, Dagher et al., 2021; Mas-
Herrero et al., 2018a). In two independent studies, over three separate days, two samples of 17 participants
received either excitatory, inhibitory, or sham stimulation (as a control) over the left DLPFC, counterbalanc-
ing for the order. Following each stimulation, the participants listened to a set of musical excerpts while pro-
viding continuous real-time ratings of experienced pleasure. Stimuli consisted of five self-selected favourite
songs and 10 experimenter-selected songs that were likely to be experienced as positive (since they matched
participants’ musical taste). In addition, the participants had the opportunity to purchase our music selection
through an auction paradigm procedure in which participants could bet their own money to buy any of the
musical items they would find appealing. Finally, we recorded neuroimaging data and objective physiological
measures of emotional arousal, such as electrodermal activity (EDA). Previous studies have shown that EDA
increases as a function of perceived pleasure, with a maximum during the occurrence of chills, particularly
in those cases where pleasure is experienced as highly intense (Benedek & Kaernbach, 2010; Mas-Herrero
et al., 2014; Salimpoor et al., 2009). Our results showed that exciting fronto-striatal pathways via TMS sig-
nificantly increased subjective reports of pleasure, psychophysiological measures of emotion, and participants’
motivation to purchase our music selection. In contrast, inhibition of this system reduced all these measures.
Notably, neuroimaging results indicated that the NAcc was the key brain structure underlying these modula-
tions of pleasure and motivation. Changes in the functioning of the NAcc (but not other brain regions) due
to TMS stimulation predicted TMS-induced changes in participants’ pleasure and motivation.
These findings were further replicated by directly modulating systemic dopamine levels through a phar-
macological intervention (Ferreri et al., 2019). In a double-blind, within-subject pharmacological design,
27 healthy participants were engaged in a similar music listening paradigm as used in our previous TMS
experiment. Yet now participants either received a dopamine precursor (levodopa), a dopamine antagonist
(risperidone), or a placebo (lactose) on three different days. Consistent with the previous TMS findings,
participants reported experiencing more pleasure, and more willingness to spend their own money buy-
ing music, following the administration of the dopamine precursor levodopa. In contrast, the dopamine
antagonist risperidone reduced the number of chills experienced and participants’ motivation to buy music
from our musical list. These two studies represent the first direct demonstration that affective reactions to
music can be manipulated by the stimulation of dopaminergic striatal circuits, providing strong support for
the hypothesis that this circuitry mediates feelings of pleasure induced by music. Interestingly, these findings
seem to contradict previous evidence from nonhuman animal studies using primary rewards such as gusta-
tory sweetness showing that dopamine’s role is limited to the motivational aspects of reward but does not
affect pleasure itself (Berridge & Kringelbach, 2015; see also Chapter 3). Dopaminergic manipulations do

138
Auditory pleasure

not change liking responses in rodents (the pleasure remains intact, no matter the dopamine levels). Still,
it influences the amount of work they are willing to exert to get more sweet pellets (they are more moti-
vated when dopamine levels are up). Therefore, together, these findings may suggest a possible dissociation
between primary and abstract pleasures, with the latter requiring dopaminergic transmission for pleasure to
happen—at least in humans.

Music-specific pathways
If the pleasure we experience in relation to music is driven by an ancient circuitry that we share with many
other organisms, how is it that it seems a uniquely human trait? If other animals possess this “hardware,” why
is musical pleasure only evident in humans? Indeed, studies performed in nonhuman primates indicate that
they generally prefer silence to music (McDermott & Hauser, 2007), even though they may be responsive
to aspects of sound relevant to music (Izumi, 2000; Wright et al., 2000). How does the human brain turn
music into pleasure? To answer this question, we have to consider that the reward circuitry does not work in
isolation but is highly interconnected with the rest of the brain. Notably, as animals become more evolved,
the connections between this circuitry and the rest of the brain, particularly the cortex, gain complexity and
are more numerous (Kaas, 2000, 2005). Specifically, humans show the largest interconnections with tempo-
ral and frontal cortices (Balsters et al., n.d.; Gordon et al., 2021), which, in turn, are particularly evolved in
humans (Dick & Tremblay, 2012; Friederici, 2017; Smaers et al., 2011). These cortico-striatal connections
may have provided alternative ways to trigger reward-related signals in humans, expanding the range of
events and experiences we may find pleasurable.
Evidence for the existence of music-specific reward pathways involving cortical brain regions comes from
patients who develop alterations in their aesthetic appreciation of music after brain damage. The most intrigu-
ing example of this is found in patients who, after brain damage, lose the capacity to experience pleasure in
connection with music listening despite a preserved ability to perceive music (i.e., they are not amusical) and
form reward-related responses to other pleasant stimuli. This condition is known as acquired specific musical
anhedonia. Remarkably, most patients with musical anhedonia exhibit lesions in cortical areas—including
temporal, frontal, and parietal cortices—but not in reward-related structures. This acquired condition is
actually rare (Belfi et al., 2017). The first case reported was a 24-year-old male who, after a haemorrhage in
the right tempo-parietal lobe, lost the capacity to experience aesthetic pleasure from music (Mazzoni et al.,
1993). Notably, before the vascular accident he was an amateur guitarist. Yet, on the day of the accident he
realized that he had difficulties responding to music in the way he was used to. As he would report later:
“my perception is changed . . . it’s flat, it’s no longer 3-dimensional; it’s only on two planes . . . there’s no
emotion.” Interestingly, his musical abilities were intact: he could perfectly perceive rhythm, melody, and
harmony, yet the emotional feelings were gone. The second case study concerned a 52-year-old radio
announcer reported by Griffiths and colleagues (2004). Following an infarction of the left insula extending
into the frontal lobe and amygdala, he experienced a general loss of pleasure to music. However, he had
preserved music perceptual abilities and hedonic reactions to other rewarding activities. Finally, Satoh and
colleagues reported a 71-year-old retired teacher who suffered an infarction of the right tempo-parietal lobe
(2011). After this injury, he was unable to enjoy music, including the music of his favourite artists. However,
as in the previous cases, he was perfectly capable of driving pleasure from other activities and did not present
musical perceptual deficits.
Notably, other patients may develop a specific craving for music, known as musicophilia. Indeed, this seems
to be a common phenotype in some patients suffering from frontotemporal lobe dementia (FTLD; Fletcher
et al., 2013, 2015). A fascinating case was reported by Geroldi et al. (2000). The patient was a 73-year-old
woman suffering from FTLD who never really enjoyed music in her life. Yet, one year after the onset of
FTLD, she developed a strong interest in Italian pop bands and singers. As reported by the authors, “she

139
Ernest Mas-Herrero

listened to Italian pop bands and singers, commenting that they had beautiful voices and played good music
with nice rhythms.” Although she developed apathy and lost interest in other activities due to her dementia,
she suddenly started enjoying music. This particular musicophilia phenotype in FTLD has been associated
with structural changes to a widespread network, including the right temporal cortex, frontal areas, and
striatum (Fletcher et al., 2013, 2015).
Another example of musicophila was reported by Rohrer and colleagues in 2006. EEG of a 65-year-
old woman treated for partial epilepsy revealed focal discharges in temporal cortices, particularly in right
frontotemporal regions (Rohrer et al., 2006). After being diagnosed with partial epilepsy, she was treated
with lamotrigine, a medication generally prescribed to treat certain types of epileptic seizures. Before the
treatment, she barely listened to music and never showed any interest in music-related activities. Indeed,
she seemed to actively avoid music. As the author reported: “she would shut the door to avoid hearing her
husband playing piano music, and found choral singing irritating.” Yet, after some weeks of treatment, she
suddenly developed a craving for music. She would now spend many hours listening to music on the radio
and TV and developed a strong preference for classical music. Music was not irritating anymore; instead, she
would now be extremely annoyed by people talking in a concert. Notably, this change in behaviour turned
out to be specific to music appreciation, since she and her family did not notice any other evident changes in
behaviour or personality, and she did not lose or gain interest in any of her previous hobbies either.
Many of these case studies share the fact that disruptions or increases in sensitivity to music were driven
not by alterations to reward-related structures but rather involved damages to frontal and temporal cortices,
particularly in the right hemisphere. As we have seen in the first section of this chapter, the right superior
temporal gyrus and inferior frontal gyrus play key roles in music perception and musical pattern recogni-
tion. Yet the lesions did not affect music perception, suggesting that these structures were functioning well.
Nor was dysfunction to the reward circuitry observed, with affective reactions to other rewarding stimuli
preserved. Thus, one possibility is that the patients’ lesions left unaltered the local functioning of cortical and
reward-related regions but impaired the coupling between both systems.
To directly test the role of cortico-striatal interactions in music-induced pleasure, Salimpoor and col-
leagues investigated the connectivity patterns between the nucleus accumbens and cortical regions during
music listening (2013). Participants were scanned while listening to a list of novel music that they had not
heard before but matched their musical preferences. After listening to each piece, participants were given
the chance to purchase the music in an auction paradigm. The authors used participants’ bids to measure the
reward value associated with each piece of music. In accord with previous studies, the results showed that
while listening to those songs on which participants would later bet higher, the participants’ striatum, espe-
cially the NAcc, was more activated than during those songs towards which they did not show any interest.
Yet, the most relevant finding was that highly rewarding musical pieces were also associated with an increased
coupling of the STG and frontal areas (including the IFG), on the one hand, and the NAacc, on the other.
Altogether, accumulating evidence from neuroimaging and lesion studies indicates that music-induced pleas-
ure relies not only on the engagement of the common reward circuitry but also on the crosstalk between this
circuitry and higher-order cortical regions involved in auditory cognition and learning, which are phyloge-
netically newer and especially well developed in humans. Indeed, these cortico-striatal interactions appear
to be a hallmark of musical reward compared to other reward types and provide a biological basis for why
music is found pleasurable in humans but not in other species with whom we share a similar reward circuitry.

Do we all love music?


The fact that music is present in all cultures and societies and plays an essential part in most people’s lives has
led to the general and implicit assumption that everybody loves music. Yet, as we saw in the previous section,
there are indeed individual differences in how rewarding music is experienced to be, even in the healthy

140
Auditory pleasure

population. Some do not find music appealing at all, while others spend most of their time and money
engaging in music-related activities. In addition, musical pleasure is a complex, multifactorial construct since
it provides different pleasure sources: from the mere fact of passively enjoying a highly emotional musical
excerpts to dancing or social bonding in concert settings.
In 2013, we developed a psychometric instrument (the Barcelona Music Reward Questionnaire; BMRQ)
to provide a fine-grained description of the different factors that contribute to music reward and allow for
an assessment of individual differences in music reward sensitivity (Mas-Herrero et al., 2013). Using factor
analysis on data collected from around 1600 participants, we identified five main factors that critically con-
tributed to music reward. We named these factors Music Seeking, Emotion Evocation, Mood Regulation, Sen-
sorimotor, and Social Reward. Music Seeking refers to the experience of pleasure associated with discovering
novel compositions and bands/singers, or music information seeking. Not surprisingly, musicians generally
score higher on this factor than non-musicians. The Emotion Evocation factor measures music’s emotional
impact, that is, the degree to which individuals get emotional listening to music, experience chills, or even
cry when listening to certain pieces. The Mood Regulation factor captures the rewarding feelings elicited
by music when it efficiently regulates our mood by relaxing and comforting us when stressed, anxious, or
depressed. The Sensorimotor factor refers to the pleasure of spontaneously synchronizing and coordinating
our body to music by either tapping, humming, or dancing. Finally, the Social Reward facet measures the
reward of social bonding through music, either by attending concerts, sharing music, or just sharing discus-
sions of music-related topics.
In this first study, we also administered standardized scales related to individual differences in general
reward processing in order to measure reward sensitivity to other rewarding stimuli (e.g., food, sex, sports, or
social activities, among others). In accordance with the idea that rewards are processed by a common reward
circuitry, we found a strong correlation between these measures (Mas-Herrero et al., 2013). Individuals who
were more sensitive to reward in general were also more likely to display a high sensitivity to music reward.
However, although these different measures were significantly correlated across the population studied, some
individuals did not follow this rule. Concretely, about 5% of the studied population presented average scores
on sensitivity to non-music rewarding stimuli yet presented low scores on the BMRQ. That is, 5% of the
studied population reported not enjoying music despite being perfectly capable of experiencing pleasure in
response to other rewarding stimuli. These preliminary findings suggested the existence of specific musical
anhedonic individuals in the healthy population. Yet we could not rule out that these individuals were not
actually suffering from amusia, which could explain their specific low sensitivity to music. In addition, this
initial finding was completely dependent on self-reported questionnaires (which might be subject to bias or
even malingering). It was then necessary to provide further validation with behavioural and physiological
responses to music and other rewarding stimuli to ultimately conclude that healthy individuals with specific
musical anhedonia existed.
To overcome these limitations, we conducted a second study in which we initially administrated the
BMRQ, other scales of reward sensitivity, and tests of music perceptual abilities in a large sample (Mas-
Herrero et al., 2014). Based on participants’ scores on this battery of tests, we selected three groups of indi-
viduals (10 participants each) who reported high (music hyperhedonics; HHDN), average (music hedonics;
HDN), and low (music anhedonics; ANH) sensitivity to music but similar sensitivity to other rewards and no
music perceptual difficulties. Therefore, the differences in sensitivity to musical reward among groups could
not be attributed to a general lack of sensitivity to reward (general anhedonia or inability to feel pleasure) or
deficits in music perception. In other words, their lack of sensitivity to music could not be driven by a general
dysfunction of either the reward circuitry or auditory cortical regions. Next, the three groups of participants
performed two tasks: a music task in which they had to indicate in real time the degree of pleasure they
experienced during music listening and a monetary task in which the participants had to respond as quickly
as possible to a target to either win or avoid losing money. For the music task, and following the procedure

141
Ernest Mas-Herrero

introduced by Blood and Zatorre, we asked the participants to bring in their three favourite musical excerpts,
those evoking the most intense feelings of pleasure. However, because, by definition, specific musical anhe-
donics do not find music appealing in general, we anticipated that selecting highly pleasurable music excerpts
would be challenging for this group of individuals. Therefore, we performed a survey in more than 200
individuals of similar age and socioeconomic status in which we asked them to list those musical excerpts that
they found most pleasurable. Next, we selected the 20 musical pieces that were most frequently reported by
our sample. Finally, an independent sample of 45 participants rated the degree of pleasure experienced while
listening to the 20 selected excerpts. We then selected those excerpts that were consistently rated as pleasur-
able by most participants, leading to a final selection of 13 musical pieces that were “universally pleasing.”
The three groups of participants, therefore, in addition to self-selected excerpts, listened to this set of music
pieces. To assess whether the differences among groups also translated into objective physiological responses
associated with emotional intensity, we measured participants’ EDA response and heart rate while participants
performed both tasks.
As we expected, the group with low sensitivity to musical reward had trouble selecting their favourite
music. Indeed, one participant of the ANH group was unable to provide any piece of music, two participants
only provided one, and two more asked for help from their family or friends since they reported not having a
strong preference for any music in particular. Furthermore, the ANH group rated all the music as less pleas-
urable and reported chills less frequently and of lower intensity than the other two groups (regardless of music
selection). These differences were also reflected in their physiological responses. While HDN and HHDN
groups displayed the expected increase in EDA and heart rate as a function of the pleasure experienced, the
ANH group did not show any significant changes in these measures throughout the task. Thus, although
some ANH participants reported experiencing chills while listening to the music, those responses were not
accompanied by changes in physiological responses as expected. Notably, and in contrast to the differences
found among groups in the music task, the three groups performed similarly when playing for monetary
rewards, and physiologically, the three groups exhibited similar EDA and heart rate responses to monetary
reward-predicting cues. Therefore, although the ANH groups did show lower hedonic reactions and attenu-
ated physiological responses to music, they showed average performance and physiological responses when
receiving monetary rewards.
In a second session, conducted a year later, we replicated the behavioural results obtained in the music
task, suggesting that the effects were stable over time. Furthermore, we verified that the observed effects
were not explained by differences in familiarity with our music list (all groups showed the same familiarity
with the melodies presented). Finally, we verified that the participants with low sensitivity to musical reward
correctly recognized the emotions conveyed by the melodies ( joy, sadness, etc.), even when they did not
experience such feelings. Altogether, these results demonstrated that the lack of musical pleasure reported
by these people was not driven by a general inability to feel pleasure, music perceptual deficits, differences
in exposure, or inability to recognize emotions from music. In subsequent studies, we have also shown that
this condition does not reflect a more general deficit affecting either aesthetic rewards in general or other
types of emotional sounds (Mas-Herrero et al., 2018b): Specific musical anhedonics demonstrated average
emotional reactions to pictorial art (e.g., the Guernica of Picasso) and emotional sounds (e.g., baby crying
and sexual moans).
What, then, is going on? Based on previous neuroimaging and lesion studies, we hypothesized that this
particular lack of sensitivity to music rewards could be driven by altered connectivity between the nucleus
accumbens and frontal and temporal cortices. To test this hypothesis, we selected three new groups of partici-
pants with high, average, or low sensitivity to music but with similar sensitivity to other rewarding stimuli and
no music perceptual deficits, using the same battery of test used before (Martínez-Molina et al., 2016). These
three groups of participants also performed a music and a monetary task but now inside a magnetic resonance
imaging scanner to investigate the neural correlates of specific musical anhedonia. Besides replicating our

142
Auditory pleasure

previous behavioural and physiological findings, we also found that those with specific musical anhedonia
exhibited a reduced activity in the NAcc while listening to pleasant music. Notably, this reduction was not
related to a general dysfunction of this brain region. When playing for money, activations of the nucleus
accumbens did not differ from the other groups. Why, then, did the nucleus accumbens fail to respond to
music in the members of the ANH group? In answer to this question, we observed decreased functional con-
nectivity between the right STG and the NAcc during music listening in the ANH participants compared to
the other two groups. It is important to highlight that the pattern of activity in the STG while listening to
music was similar among the three groups, consistent with the intact music perceptual abilities of the ANH
group. Therefore, the problem does not reside in auditory regions but in how these structures communicated
with the NAcc. These findings fit very well with the idea that music-induced pleasure may be mediated via
cortical pathways involved in auditory cognition in interaction with reward-related structures.
However, although these findings suggest that communication between temporal cortices and the NAcc
is necessary for music reward to emerge, it is important to notice that these two regions are only sparsely
connected through direct anatomical pathways. It remains possible that music information processed by
temporal cortices accesses the NAcc through a third region. Here, the orbitofrontal cortex represents an
ideal candidate for the mediation of this interaction. The OFC is a key structure in the reward circuitry that
receives inputs from all sensory modalities and projects to the ventral striatum, including the NAcc (Haber &
Knutson, 2010). In this way, the OFC acts as a multisensory hub, integrating signals from sensory and per-
ceptual regions into the reward circuitry. To investigate the mediating role of the OFC in the interaction
between the STG and the NAcc, we used diffusion tensor imaging (DTI)—an MRI technique that allows
for measuring the integrity of white matter fibre tracts—in a subset of participants from our previous fMRI
study (Martínez-Molina et al., 2019). Specifically, we studied the relationship between music reward sensitiv-
ity (measured by the BMRQ) and the white matter microstructure of the pathways connecting the STG and
the NAcc via the OFC (STG–OFC and OFC–NAcc connections). The results revealed that white matter
microstructure in the right STG–OFC and the OFC–NAcc connectivity predicted individual differences in
music reward sensitivity. Remarkably, the integrity of the entire pathway connecting the STG and NAcc via
the OFC (STG–OFC–NAcc) explained individual differences in NAcc activation during music listening.
These results extended the previous findings by suggesting that the OFC plays a pivotal role in the interac-
tion between high-order cortical brain regions and the NAcc in music-induced pleasure. Overall, research
on specific musical anhedonia has provided significant insights into the brain mechanism involved in musical
pleasure, validating the idea that music-induced pleasure arises through a complex interplay between percep-
tual, integrative, and reward systems.

Major challenges, goals, and suggestions


Despite the significant advances in understanding the neural basis of music-induced pleasure during the last
20 years, there are still many open questions requiring further research. First, although the literature reviewed
here indicates that the interactions between auditory cortical regions and the reward system play a crucial
role in music reward, it is still unclear what the directionality and the dynamics of these interactions are. Is
the reward system controlled by auditory cortical signals, or is this coupling only reflecting greater auditory
processing because of the pleasure experienced? The implementation of new connectivity methods to esti-
mate directed functional (causal) connectivity, such as Granger causality analysis, as well as new techniques
to improve the temporal resolution of MRI data acquisition, may provide important insights into this topic.
Second, it remains to be seen if specific musical anhedonia is a heritable trait. Investigating the genetic
underpinnings of this condition may provide important advances in understanding the evolutionary trajec-
tory of music reward sensitivity. Twin or family studies may be the first step to determining whether the
predisposition to developing musical anhedonia is genetic. If so, genome-wide association studies (GWA or

143
Ernest Mas-Herrero

GWAS) may enable us to map the genetic loci for musical anhedonia and, through comparative genetics,
investigate evolutionary trajectories.
Third, as we have seen in this chapter, we now have a pretty good picture of the brain circuits involved
in musical pleasure. Yet we do not fully understand what it is about music that triggers these circuitries (see
also Chapter 15). Why do we like some music and detest other music? Why are there such large individual
differences in musical tastes and preferences, even within the same culture and society? Theoretical models
have long posited that the key may rely on music’s power to generate predictions and expectations through
its temporal and structural patterns. According to these models, either the fulfilment or violation of our
expectations may give rise to pleasure peaks. However, large individual differences are present in musical
preferences indicating that the balance between predictability and surprise may differ across individuals. For
instance, some people might enjoy music characterized by repetitive patterns (pop, techno, hip hop, etc.). In
contrast, others may prefer musical genres involving improvisation and surprise ( jazz, experimental music,
among others). Are these differences in musical tastes a matter of exposure? Or do they reflect individual
differences in auditory cognition and perception shaping our preferences for more or less music complex-
ity? Studying the development of musical tastes and the brain changes associated with it as a function of
individuals’ auditory abilities may provide an excellent model to understand how music-induced pleasure
arises. In addition, the implementation of computational models to estimate music’s ongoing expectations
and surprises may be an excellent tool to assess music complexity objectively and thus investigate the role of
surprise, uncertainty, and predictability in music-induced pleasure.

References
Albouy, P., Benjamin, L., Morillon, B.,  & Zatorre, R. J. (2020). Distinct sensitivity to spectrotemporal modulation
supports brain asymmetry for speech and melody. Science, 367(6481), 1043–1047. https://doi.org/10.1126/science.
aaz3468
Albouy, P., Peretz, I., Bermudez, P., Zatorre, R. J., Tillmann, B., & Caclin, A. (2019). Specialized neural dynamics for
verbal and tonal memory: FMRI evidence in congenital amusia. Human Brain Mapping, 40(3), 855–867. https://doi.
org/10.1002/hbm.24416
Balsters, J. H., Zerbi, V., Sallet, J., Wenderoth, N., & Mars, R. B. (n.d.). Primate homologs of mouse cortico-striatal
circuits. eLife, 9. https://doi.org/10.7554/eLife.53680
Bartra, O., McGuire, J. T., & Kable, J. W. (2013). The valuation system: A coordinate-based meta-analysis of BOLD fMRI
experiments examining neural correlates of subjective value. NeuroImage, 76, 412–427. https://doi.org/10.1016/j.
neuroimage.2013.02.063
Belfi, A. M., Evans, E., Heskje, J., Bruss, J., & Tranel, D. (2017). Musical anhedonia after focal brain damage. Neuropsy-
chologia, 97, 29–37. https://doi.org/10.1016/j.neuropsychologia.2017.01.030
Benedek, M., & Kaernbach, C. (2010). A continuous measure of phasic electrodermal activity. Journal of Neuroscience
Methods, 190(1–5), 80–91. https://doi.org/10.1016/j.jneumeth.2010.04.028
Berridge, K. C.,  & Kringelbach, M. L. (2015). Pleasure systems in the brain. Neuron, 86(3), 646–664. https://doi.
org/10.1016/j.neuron.2015.02.018
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Blood, A. J., Zatorre, R. J., Bermudez, P., & Evans, A. C. (1999). Emotional responses to pleasant and unpleasant music
correlate with activity in paralimbic brain regions. Nature Neuroscience, 2(4), 382–387. https://doi.org/10.1038/7299
Chen, J. L., Penhune, V. B., & Zatorre, R. J. (2008). Listening to musical rhythms recruits motor regions of the brain.
Cerebral Cortex, 18(12), 2844–2854. https://doi.org/10.1093/cercor/bhn042
Conard, N. J., Malina, M., & Münzel, S. C. (2009). New flutes document the earliest musical tradition in southwestern
Germany. Nature, 460(7256), 737–740. https://doi.org/10.1038/nature08169
Dick, A. S.,  & Tremblay, P. (2012). Beyond the arcuate fasciculus: Consensus and controversy in the connectional
anatomy of language. Brain, 135(12), 3529–3550. https://doi.org/10.1093/brain/aws222
Dubé, L., & Bel, J. L. (2003). The content and structure of laypeople’s concept of pleasure. Cognition and Emotion, 17(2),
263–295. https://doi.org/10.1080/02699930302295

144
Auditory pleasure

Ferreri, L., Mas-Herrero, E., Zatorre, R. J., Ripollés, P., Gomez-Andres, A., Alicart, H., Olivé, G., Marco-Pallarés, J.,
Antonijoan, R. M., Valle, M., Riba, J., & Rodriguez-Fornells, A. (2019). Dopamine modulates the reward experi-
ences elicited by music. Proceedings of the National Academy of Sciences of the United States of America, 116(9), 3793–3798.
https://doi.org/10.1073/pnas.1811878116
Fitch, W. T. (2006). The biology and evolution of music: A comparative perspective. Cognition, 100(1), 173–215. https://
doi.org/10.1016/j.cognition.2005.11.009
Fletcher, P. D., Downey, L. E., Golden, H. L., Clark, C. N., Slattery, C. F., Paterson, R. W., Schott, J. M., Rohrer, J.
D., Rossor, M. N., & Warren, J. D. (2015). Auditory hedonic phenotypes in dementia: A behavioural and neuro-
anatomical analysis. Cortex; a Journal Devoted to the Study of the Nervous System and Behavior, 67, 95–105. https://doi.
org/10.1016/j.cortex.2015.03.021
Fletcher, P. D., Downey, L. E., Witoonpanich, P., & Warren, J. D. (2013). The brain basis of musicophilia: Evidence
from frontotemporal lobar degeneration. Frontiers in Psychology, 4, 347. https://doi.org/10.3389/fpsyg.2013.
00347
Friederici, A. D. (2017). Evolution of the neural language network. Psychonomic Bulletin and Review, 24(1), 41–47.
https://doi.org/10.3758/s13423-016-1090-x
Geroldi, C., Metitieri, T., Binetti, G., Zanetti, O., Trabucchi, M., & Frisoni, G. B. (2000). Pop music and frontotemporal
dementia. Neurology, 55(12), 1935–1936. https://doi.org/10.1212/WNL.55.12.1935
Gordon, E. M., Laumann, T. O., Marek, S., Newbold, D. J., Hampton, J. M., Seider, N. A., Montez, D. F., Nielsen, A.
M., Van, A. N., Zheng, A., Miller, R., Siegel, J. S., Kay, B. P., Snyder, A. Z., Greene, D. J., Schlaggar, B. L., Petersen,
S. E., Nelson, S. M., & Dosenbach, N. U. F. (2021). Human fronto-striatal connectivity is organized into discrete
functional subnetworks. bioRxiv. https://doi.org/10.1101/2021.04.12.439415
Grewe, O., Nagel, F., Kopiez, R., & Altenmüller, E. (2005). How does music arouse “chills”? Investigating strong emo-
tions, combining psychological, physiological, and psychoacoustical methods. Annals of the New York Academy of Sci-
ences, 1060, 446–449. https://doi.org/10.1196/annals.1360.041
Griffiths, T. D., Warren, J. D., Dean, J. L., & Howard, D. (2004). ‘When the feeling’s gone’: A selective loss of musi-
cal emotion. Journal of Neurology, Neurosurgery, and Psychiatry, 75(2), 344–345. https://doi.org/10.1136/jnnp.2003.
015586
Haber, S. N., & Knutson, B. (2010). The reward circuit: Linking primate anatomy and human imaging. Neuropsychophar-
macology: Official Publication of the American College of Neuropsychopharmacology, 35(1), 4–26. https://doi.org/10.1038/
npp.2009.129
Herholz, S. C., Halpern, A. R., & Zatorre, R. J. (2012). Neuronal correlates of perception, imagery, and memory for
familiar tunes. Journal of Cognitive Neuroscience, 24(6), 1382–1397. https://doi.org/10.1162/jocn_a_00216
Hirel, C., Lévêque, Y., Deiana, G., Richard, N., Cho, T. H., Mechtouff, L., Derex, L., Tillmann, B., Caclin, A., &
Nighoghossian, N. (2014). [Acquired amusia and musical anhedonia]. Revue neurologique, 170(8–9), 536–540. https://
doi.org/10.1016/j.neurol.2014.03.015
Izumi, A. (2000). Japanese monkeys perceive sensory consonance of chords. Journal of the Acoustical Society of America,
108(6), 3073–3078. https://doi.org/10.1121/1.1323461
Kaas, J. H. (2000). Why is brain size so important: Design problems and solutions as neocortex gets bigger or smaller.
Brain and Mind, 1(1), 7–23. https://doi.org/10.1023/A:1010028405318
Kaas, J. H. (2005). From mice to men: The evolution of the large, complex human brain. Journal of Biosciences, 30(2),
155–165. https://doi.org/10.1007/BF02703695
Kelley, A. E., & Berridge, K. C. (2002). The neuroscience of natural rewards: Relevance to addictive drugs. Journal of
Neuroscience, 22(9), 3306–3311. https://doi.org/20026361
Lebrun, M. A., Moreau, P., McNally-Gagnon, A., Mignault Goulet, G., & Peretz, I. (2012). Congenital amusia in child-
hood: A case study. Cortex; a Journal Devoted to the Study of the Nervous System and Behavior, 48(6), 683–688. https://
doi.org/10.1016/j.cortex.2011.02.018
Loui, P., Alsop, D., & Schlaug, G. (2009). Tone deafness: A new disconnection syndrome? Journal of Neuroscience, 29(33),
10215–10220. https://doi.org/10.1523/JNEUROSCI.1701-09.2009
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J.,  & Marco-Pallarés, J. (2016). Neural
correlates of specific musical anhedonia. Proceedings of the National Academy of Sciences of the United States of America,
113(46), E7337–E7345. https://doi.org/10.1073/pnas.1611211113
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J., & Marco-Pallarés, J. (2019). White mat-
ter microstructure reflects individual differences in music reward sensitivity. Journal of Neuroscience, 39(25), 5018–5027.
https://doi.org/10.1523/JNEUROSCI.2020-18.2019
Mas-Herrero, E., Dagher, A., & Zatorre, R. J. (2018a). Modulating musical reward sensitivity up and down with tran-
scranial magnetic stimulation. Nature Human Behaviour, 2(1), 27–32. https://doi.org/10.1038/s41562-017-0241-z

145
Ernest Mas-Herrero

Mas-Herrero, E., Dagher, A., Farrés-Franch, M., & Zatorre, R. J. (2021). Unraveling the temporal dynamics of reward
signals in music-induced pleasure with TMS. Journal of Neuroscience, 41(17), 3889–3899. https://doi.org/10.1523/
JNEUROSCI.0727-20.2020
Mas-Herrero, E., Karhulahti, M., Marco-Pallares, J., Zatorre, R. J., & Rodriguez-Fornells, A. (2018b). The impact of
visual art and emotional sounds in specific musical anhedonia. Progress in Brain Research, 237, 399–413. https://doi.
org/10.1016/bs.pbr.2018.03.017
Mas-Herrero, E., Maini, L., Sescousse, G., & Zatorre, R. J. (2021). Common and distinct neural correlates of music
and food-induced pleasure: A coordinate-based meta-analysis of neuroimaging studies. Neuroscience and Biobehavioral
Reviews, 123, 61–71. https://doi.org/10.1016/j.neubiorev.2020.12.008
Mas-Herrero, E., Marco-Pallares, J., Lorenzo-Seva, U., Zatorre, R. J., & Rodriguez-Fornells, A. (2013). Individual dif-
ferences in music reward experiences. Music Perception, 31(2), 118–138. https://doi.org/10.1525/mp.2013.31.2.118
Mas-Herrero, E., Zatorre, R. J., Rodriguez-Fornells, A., & Marco-Pallarés, J. (2014). Dissociation between musical and
monetary reward responses in specific musical anhedonia. Current Biology, 24(6), 699–704. https://doi.org/10.1016/j.
cub.2014.01.068
Mazzoni, M., Moretti, P., Pardossi, L., Vista, M., Muratorio, A., & Puglioli, M. (1993). A case of music imperception.
Journal of Neurology, Neurosurgery, and Psychiatry, 56(3), 322. https://doi.org/10.1136/jnnp.56.3.322
Mazzucchi, A., Marchini, C., Budai, R., & Parma, M. (1982). A case of receptive amusia with prominent timbre percep-
tion defect. Journal of Neurology, Neurosurgery, and Psychiatry, 45(7), 644–647. https://doi.org/10.1136/jnnp.45.7.644
McDermott, J., & Hauser, M. D. (2007). Nonhuman primates prefer slow tempos but dislike music overall. Cognition,
104(3), 654–668. https://doi.org/10.1016/j.cognition.2006.07.011
Mcdonald, C., & Stewart, L. (2008). Uses and functions of music in congenital amusia. Music Perception, 25(4), 345–355.
https://doi.org/10.1525/mp.2008.25.4.345
Mehr, S. A., Singh, M., Knox, D., Ketter, D. M., Pickens-Jones, D., Atwood, S., Lucas, C., Jacoby, N., Egner, A.
A., Hopkins, E. J., Howard, R. M., Hartshorne, J. K., Jennings, M. V., Simson, J., Bainbridge, C. M., Pinker,
S., O’Donnell, T. J., Krasnow, M. M., & Glowacki, L. (2019). Universality and diversity in human song. Science,
366(6468). https://doi.org/10.1126/science.aax0868
Miller, G. (2000). Evolution of human music through sexual selection. In The origins of music (pp. 329–360). MIT Press.
Olds, J. (1969). The central nervous system and the reinforcement of behavior. American Psychologist, 24(2), 114–132.
https://doi.org/10.1037/h0027145
Olds, J., & Milner, P. (1954). Positive reinforcement produced by electrical stimulation of septal area and other regions
of rat brain. Journal of Comparative and Physiological Psychology, 47(6), 419–427. https://doi.org/10.1037/h0058775
Omigie, D., Müllensiefen, D., & Stewart, L. (2012). The experience of music in congenital amusia. Music Perception,
30(1), 1–18. https://doi.org/10.1525/mp.2012.30.1.1
Pelofi, C., Goldstein, M., Bevilacqua, D., McPhee, M., Abrams, E., & Ripollés, P. (2021). Chiller: A computer human
interface for the live labeling of emotional responses. https://nime.pubpub.org/pub/kdahf9fq/release/1. International Con-
ference on New Interfaces for Musical Expression.
Peretz, I., & Vuvan, D. T. (2017). Prevalence of congenital amusia. European Journal of Human Genetics, 25(5), 625–630.
https://doi.org/10.1038/ejhg.2017.15
Pinker, S. (1997). How the mind works. W. W. Norton, and Co.
Ravignani, A. (2018). Darwin, sexual selection, and the origins of music. Trends in Ecology and Evolution, 33(10), 716–
719. https://doi.org/10.1016/j.tree.2018.07.006
Rohrer, J. D., Smith, S. J., & Warren, J. D. (2006). Craving for music after treatment for partial epilepsy. Epilepsia, 47(5),
939–940. https://doi.org/10.1111/j.1528-1167.2006.00565.x
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Salimpoor, V. N., Benovoy, M., Longo, G., Cooperstock, J. R., & Zatorre, R. J. (2009). The rewarding aspects of music
listening are related to degree of emotional arousal. PLOS ONE, 4(10), e7487. https://doi.org/10.1371/journal.
pone.0007487
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059
Satoh, M., Kato, N., Tabei, K. I., Nakano, C., Abe, M., Fujita, R., Kida, H., Tomimoto, H., & Kondo, K. (2016). A case
of musical anhedonia due to right putaminal hemorrhage: A disconnection syndrome between the auditory cortex
and insula. Neurocase, 22(6), 518–525. https://doi.org/10.1080/13554794.2016.1264609

146
Auditory pleasure

Satoh, M., Nakase, T., Nagata, K., & Tomimoto, H. (2011). Musical anhedonia: Selective loss of emotional experience
in listening to music. Neurocase, 17(5), 410–417. https://doi.org/10.1080/13554794.2010.532139
Sescousse, G., Caldú, X., Segura, B., & Dreher, J. C. (2013). Processing of primary and secondary rewards: A quantita-
tive meta-analysis and review of human functional neuroimaging studies. Neuroscience and Biobehavioral Reviews, 37(4),
681–696. https://doi.org/10.1016/j.neubiorev.2013.02.002
Sihvonen, A. J., Särkämö, T., Rodríguez-Fornells, A., Ripollés, P., Münte, T. F., & Soinila, S. (2019). Neural archi-
tectures of music—Insights from acquired amusia. Neuroscience and Biobehavioral Reviews, 107, 104–114. https://doi.
org/10.1016/j.neubiorev.2019.08.023
Smaers, J. B., Steele, J., Case, C. R., Cowper, A., Amunts, K., & Zilles, K. (2011). Primate prefrontal cortex evolution:
Human brains are the extreme of a lateralized ape trend. Brain, Behavior and Evolution, 77(2), 67–78. https://doi.
org/10.1159/000323671
Strafella, A. P., Paus, T., Barrett, J., & Dagher, A. (2001). Repetitive transcranial magnetic stimulation of the human pre-
frontal cortex induces dopamine release in the caudate nucleus. Journal of Neuroscience: The Official Journal of the Society
for Neuroscience, 21(15), RC157. https://doi.org/10.1523/JNEUROSCI.21-15-j0003.2001
Thaut, M. H., Trimarchi, P. D., & Parsons, L. M. (2014). Human brain basis of musical rhythm perception: Common
and distinct neural substrates for meter, tempo, and pattern. Brain Sciences, 4(2), 428–452. https://doi.org/10.3390/
brainsci4020428
Volkow, N. D., Wise, R. A., & Baler, R. (2017). The dopamine motive system: Implications for drug and food addiction.
Nature Reviews. Neuroscience, 18(12), 741–752. https://doi.org/10.1038/nrn.2017.130
Wilbiks, J. M. P., Vuvan, D. T., Girard, P. Y., Peretz, I., & Russo, F. A. (2016). Effects of vocal training in a musicophile
with congenital amusia. Neurocase, 22(6), 526–537. https://doi.org/10.1080/13554794.2016.1263339
Wright, A. A., Rivera, J. J., Hulse, S. H., Shyan, M.,  & Neiworth, J. J. (2000). Music perception and octave gen-
eralization in rhesus monkeys. Journal of Experimental Psychology. General, 129(3), 291–307. https://doi.
org/10.1037//0096-3445.129.3.291
Zatorre, R. J., & Belin, P. (2001). Spectral and temporal processing in human auditory cortex. Cerebral Cortex, 11(10),
946–953. https://doi.org/10.1093/cercor/11.10.946
Zatorre, R. J., Belin, P., & Penhune, V. B. (2002). Structure and function of auditory cortex: Music and speech. Trends
in Cognitive Sciences, 6(1), 37–46. https://doi.org/10.1016/S1364-6613(00)01816-7

147
8
ODOUR AESTHETICS
Hedonic perception of olfactory stimuli

Gulce Nazli Dikecligil and Jay A. Gottfried

Olfaction is a key sensory modality that detects and interprets volatile chemical compounds to ultimately
inform, guide and shape an organism’s behaviour in response to its environment. Many species rely on their
sense of smell for finding and selecting food (Arenas & Farina, 2012; Laidre, 2009; Raghuram et al., 2009;
Stutz et al., 2016), avoiding predators (Amo et al., 2008; Amo et al., 2011; Ferrero et al., 2011; Kelley &
Magurran, 2003; J. J. Li et al., 2014; Roth et al., 2008; Sündermann et al., 2008), mating (Clarke et al., 2009;
M. Keller et al., 2009; Keverne, 2004; Mclennan, 2003), communicating with conspecifics (Arakawa et al.,
2008; Banks et al., 2016; Chauvin & Thierry, 2005; Kullmann et al., 2008; Logan et al., 2012; Ryan et al.,
2008), and spatial navigation (Buehlmann et al., 2015; Jacobs et al., 2015). In humans, the sense of smell has
historically been considered an archaic and weak sense that has relatively little influence on cognition and
behaviour and is deemed inferior to the olfactory abilities of other mammals (McGann, 2017), yet studies
in the past decades have consistently shown that humans, much like other mammals, have an acute sense of
smell (Bushdid et al., 2014) that informs a rich spectrum of behaviour from food and mate selection to spatial
navigation ( Jacobs et al., 2015; Shepherd, 2004; Stevenson, 2010). The importance of olfaction for human
behaviour has been evident in studies focusing on smell disorders: It has been shown that impoverished olfac-
tion negatively impacts daily functions such as cooking, eating, detecting hazardous chemicals, and assessing
foul odours (Croy et al., 2014; Miwa et al., 2001; Temmel et al., 2002). Furthermore, smell disorders can
decrease the pleasure and reward associated with eating, as they diminish the complex flavour experiences
associated with food consumption (Blomqvist et al., 2004; A. Keller & Malaspina, 2013; Nordin et al., 2011).
Collectively, the consequences of impoverished smell contribute to decreased quality of life (Blomqvist et al.,
2004; Miwa et al., 2001; Nordin et al., 2011) and correlate with increased likelihood of depression (Bonfils
et al., 2005; Croy et al., 2012; Frasnelli & Hummel, 2005; Neuland et al., 2011; Smeets et al., 2009; Temmel
et al., 2002). These findings highlight that the human sense of smell, in addition to its crucial role in sup-
porting survival functions via odour detection, discrimination, and identification, also contributes to human
well-being through its role in appreciation of the surrounding chemosensory landscape. This crucial role
of olfaction in human pleasure and reward and how it relates to the field of aesthetics is often overlooked
in favour of work that centres on audition and vision. Despite being relatively understudied as a subject of
aesthetics, olfactory pleasures (and displeasures) form the backbone of many industries ranging from culinary
arts and oenology to cosmetics.
Olfactory appreciation builds on affective assessments of scents, which is one of the key functions of
the olfactory system to inform approach or avoidance behaviours. A mouse smelling peanut butter might

148 DOI: 10.4324/9781003008675-9


Odour aesthetics

navigate towards the odour source to locate its meal for the day, or a human sitting next to foul-smelling
garbage might decide to relocate a different bench in the park. It is important to note that appetitive and
aversive odours can shape behaviour either because they act as a cue that is associated with a positive or nega-
tive outcome, respectively (e.g., peanut butter scent signals the presence of a higher-calorie meal), or because
they act as reinforcers in and of themselves (e.g., the unpleasant smell of sweaty socks can cause someone to
move away merely because it is unpleasant, and likewise the scent of an amber candle might cause someone to
linger in a room merely because they enjoy the scent itself ). Beyond approach and avoidance behaviours, our
affective assessments of odours can bias our perceptions and decisions even in non-olfactory domains; pleasant
odours have been shown to improve pain tolerance (Bartolo et al., 2013; Jo et al., 2020; Prescott & Wilkie,
2007) and to modulate interpretation of facial expressions (Cook et al., 2017; Leleu et al., 2015; Poncet et al.,
2020). These findings support the notion that odour valence perception is not only crucial for well-being but
also can shape cognition and decision making in non-olfactory domains such as vision and touch.
Aesthetics of olfaction, while interconnected with simple hedonic judgements of odours and learned
odour-outcome associations, extends beyond mere assessments of whether an odour smells good or bad or
whether it signals approach or avoidance behaviour. Olfactory aesthetic appreciation, much like aesthetic
preferences in other sensory modalities, is influenced by stimulus novelty; complexity; familiarity; unex-
pected features of the stimulus; interplay between elements of appetitive, neutral, and aversive components;
the context in which the odour is experienced; and importantly the identity of the odour object as well as
the social and cultural associations the odour quality evokes. Interestingly, aesthetic appreciation of odours
is not necessarily aligned with approach or avoidance behaviours either; for example, the smell of gasoline
may both elicit liking and avoidance, whereas the smell of blue cheese may elicit dislike and approach. In a
similar dissociation, odour hedonics judgments and odour aesthetic appreciation do not always necessarily
run parallel either. An odour that is disliked may still be aesthetically appreciated, whereas an odour that is
liked may be found to have low aesthetic value. In other words, olfactory aesthetics does not necessarily go
hand in hand with the perceived hedonic value (e.g., smells good or bad), utility of the odour object and its
source (e.g., this scent indicates there is food nearby), or appropriate odour response (e.g., this scent means
I should avoid this area) and can be affected by a multitude of cultural factors, personal experiences, stimulus
properties, and metacognitive processes.
While olfactory aesthetics is the overarching topic of this chapter, we will mainly discuss literature on
odour valence perception, and its neural basis, as a building block for understanding the mechanistic under-
pinnings of olfactory aesthetic appreciation, given that a much greater number of neuroscientific studies
focus on valence perception rather than the more complex and subjective topic of odour aesthetics per se.
We will begin the chapter with discussing the basics of odour perception and how odour valence perception
relates to odour quality perception. In the next section, we will give an overview of the olfactory system and
neural circuits of odour quality coding. We will then highlight factors such as odour stimulus parameters and
cognitive factors that are known to shape odour valence perception. We will finish the chapter with neural
mechanisms that underlie affective assessment of odours.

General principles of odour perception in humans


The human olfactory system is excellent at detecting the presence of odours even at concentrations as low
as 10 parts per billion (Cain et al., 2007). Once an odourant is detected, the olfactory system can extract
many attributes from an odour stimulus: how intense it is, how long it lasts, what perceptual qualities it has
(e.g., fruity, flowery, fishy, etc.), how familiar it is, and how pleasant it is. Of these attributes, assessing the
perceptual qualities and characteristics of an odourant (e.g., “this is a smoky odour”) and comparing the
odour-evoked percept to existing olfactory templates (e.g., “does this smell like burning rubber?”) perhaps
serves the most crucial function for odour-guided behaviour, allowing organisms to identify and in return

149
Gulce Nazli Dikecligil and Jay A. Gottfried

act appropriately to odour stimuli (to run away from fire or prepare for eating popcorn, depending on odour
identity). Not only can the human olfactory system identify and differentiate odours arising from distinct
odour sources, but it can even learn to discriminate between pairs of enantiomers that were initially indis-
tinguishable through association of one enantiomer with an aversive outcome (Li et al., 2008), suggesting
that odour perceptual qualities can be specific to even small molecular differences and yet malleable as new
associations are acquired. Despite highly sensitive odour detection and discrimination abilities, humans are
surprisingly limited in naming odours when these are presented without any contextual cues (i.e., sights
and sounds that may typically accompany the odour source). Identification performance increases when
subjects are presented with multiple-choice options instead of doing free recall (Cain, 1979; Cain et  al.,
1998; Lawless & Engen, 1977). While naming odourants is challenging, assessing odour pleasantness comes
easily; humans can rapidly report how much they like an odour and use odour pleasantness to categorize
odours (Berglund et al., 1973; Schiffman et al., 1977; Zarzo, 2008). These findings have led researchers to
reason that difficulties in odour-naming might reflect limitations in odour quality perception and propose
that odour valence instead may be the key feature that dominates odour perception (valence-centred approach)
(Yeshurun & Sobel, 2010). An alternative model proposes that difficulties in odour naming may reflect an
evolutionary and ecological disconnect between language and olfactory systems (Olofsson  & Gottfried,
2015) rather than a difficulty in odour quality perception per se and argues that odour perceptual qualities
are the primary defining axis of odour objects (object-centred approach). Specifically, object-centred approaches
suggest that odour quality coding and perception occur earlier in the olfactory processing stream, and the
emerging odour-object representations inform the downstream affective assessments of the odour object
(Gottfried, 2010; Olofsson et al., 2013; Olofsson et al., 2012). Experiments directly testing this hypothesis
of whether odour quality perception arises prior or after odour valence perception (Olofsson et al., 2013;
Olofsson et al., 2012) have shown that, in case of familiar odours, humans can identify odours faster than
they can assess odour valence. These studies suggest that while odour valence is a crucial dimension of odour
perception, it hinges on odour-quality representations that then inform affective assessments of odour quality.
From an evolutionary and behavioural standpoint, it is easy to see why odour-valence information, in the
absence of odour-quality information, would be insufficient for odour-guided behaviour; for example, while
both garlic bread and a jasmine-scented candle might smell equally pleasant, they smell pleasant in distinct
ways that inform the subject on how to interact with the odour source. While both pleasant odours may
generate an approach response, the organism uses the odour quality information to determine what specific
approach behaviour it will follow (e.g., initiate food consumption or light the scented candle).

Neural mechanisms of odour identity coding and perception


What are the neural mechanisms that underlie the translation of myriad volatile compounds arising from a
wide range of odour sources into complex olfactory percepts? This section gives a brief overview of the cur-
rent understanding of the neuroanatomy and neurophysiology of the olfactory system and how it encodes
odour information.
Volatile odorous compounds (“odourants”) can access the olfactory receptors and initiate odour pro-
cessing through two different routes, the orthonasal route, where odourants are inhaled into the nasal cavity
through the nostrils, and the retronasal route, where the odourants emitted by food and drinks in the mouth
travel through the back of the oral cavity and into the nose during exhalation. Furthermore, in rodents,
odourants reaching the nasal cavity can be processed through either the main olfactory system and/or the
accessory olfactory system, depending on the odour source and the volatility of the odourant. However, it
has been shown that the vomeronasal organ (VNO; the olfactory sensory organ of the accessory olfactory
system) is not present in primates and humans (Salazar & Quinteiro, 2009). Therefore, we will focus our
discussion on the main olfactory system.

150
Odour aesthetics

Figure 8.1 Illustration of two differing models of odour valence perception. Object-centred models propose that the
olfactory system identifies the odour object first and then uses this information to determine the hedonic
value of the odour object. Conversely, the valence-centred approach proposes that odour hedonic percep-
tion arises prior to odour object identification and is necessary for odour object identification. Figure modi-
fied with permission from Olofsson, J. K., Bowman, N. E., Khatibi, K. & Gottfried, J. A. “A Time-Based
Account of the Perception of Odor Objects and Valences.” Psychol. Sci. 23, 1224–1232 (2012).

Orthonasal olfactory processing in the main olfactory system begins with the inhalation of odourants into
the nasal cavity. Once inside the nasal cavity, odourants bind to olfactory receptors (Ors) expressed on the cell
membrane of olfactory sensory neurons (OSNs) residing in the olfactory epithelium (OE). The binding of
odourants to G-protein–coupled Ors initiates the electrochemical signalling in OSNs. Each OSN expresses
a single OR (Serizawa et al., 2004, 2005) from a repertoire of ~350 Ors in humans (Malnic et al., 2004)
or ~1000 Ors in mice (Godfrey et al., 2004) (the size of the olfactory gene repertoire varies across species
[Keller & Vosshall, 2008; Niimura et al., 2014]). However, a given OR can bind multiple monomolecular
odourants, and, likewise, a given monomolecular odourant can bind to multiple Ors, albeit with different
affinities (Malnic et al., 1999). Furthermore, odourants can also act as antagonists for Ors (Oka et al., 2004;
Reddy et al., 2018). The electrochemical signals initiated in OSNs travel down the axons of OSNs and target
the olfactory bulb (OB). In this first stage of olfactory processing, axons of OSNs expressing the same OR
converge onto a small number of spherical structures called glomeruli such that each glomerulus represents
one OR. Within the glomeruli, OSN axons synapse onto the dendrites of output neurons of the OB called
the mitral/tufted cells (M/T cells), which then project to downstream olfactory regions. In addition to OSN
inputs and mitral/tufted cell outputs, OB neural circuits contain local neurons (granule cells and juxtaglo-
merular cells); top-down projections from the cortex (Boyd et  al., 2015; Matsutani, 2010; Matsutani  &
Yamamoto, 2008; Rothermel & Wachowiak, 2014); and extrinsic neuromodulatory inputs from serotoner-
gic, cholinergic, and noradrenergic projections as well as intrinsic dopaminergic neurons (Linster & Cleland,
2016; McLean & Shipley, 1987; McLean et al., 1989; Petzold et al., 2009).

151
Gulce Nazli Dikecligil and Jay A. Gottfried

As a result of converging projections from OSNs expressing the same OR, each monomolecular odour-
ant binds to a small subset of Ors and consequently gives rise to a sparse pattern of glomerular activity
(Davison & Katz, 2007). However, animals and humans rarely encounter monomolecular odours in isolation
and instead interact with natural scents that are complex odour mixtures composed of a large number of
monomolecular odourants (e.g., hundreds of organic volatile compounds have been identified in relation to
the aroma of strawberry [Ulrich et al., 2018]). Such odour mixtures elicit complex spatiotemporal patterns
of glomerular activity (Chae et al., 2019; Gschwend et al., 2016; Vincis et al., 2012). Both the spatial and
temporal patterns of odour-evoked responses in the OB carry odour identity information (Uchida et al.,
2014). It is important to note that glomerular activity reflects the first stage of olfactory processing, where
the information relayed by the OSNs is shaped by local OB circuits (specifically, lateral inhibition plays an
important role in early odour processing [Cleland & Linster, 2012; Geramita et al., 2016; Najac et al., 2015;
Shao et al., 2012; Urban, 2002; Zavitz et al., 2020]), top-down projections [Boyd et al., 2015; Boyd et al.,
2012; Markopoulos et al., 2012; Matsutani, 2010; Matsutani & Yamamoto, 2008; Rothermel & Wachowiak,
2014], and neuromodulatory inputs [D’Souza & Vijayaraghavan, 2014; Escanilla et al., 2009; Guérin et al.,
2008; Linster & Cleland, 2016; Liu et al., 2012; Mandairon et al., 2006; McLean & Shipley, 1987; Petzold
et al., 2009; Suyama et al., 2021]).
Olfactory information diverges onto multiple downstream regions after the first stage of processing in
the OB; outputs of the OB, mitral/tufted cells, project to cortical and subcortical regions: anterior olfactory
nucleus (AON), tenia tecta (TT), olfactory tubercle (OT), amygdala, entorhinal cortex, and piriform cortex
(PCx). The direct monosynaptic projections from the OB to the PCx are a unique neuroanatomical feature
of the olfactory system, as ascending sensory information in all other sensory modalities is relayed to the
primary sensory cortices through the thalamus.
While neighbouring mitral/tufted cells in a given glomeruli receive converging information from the
same OR, this spatial organization of olfactory information in the OB does not hold for the next stage of
processing in the PCx. M/T cells receiving input from the same OR project their axons in a seemingly
random fashion to the PCx (Ghosh et  al., 2011; Igarashi et  al., 2012). As a result, nearby PCx neurons
receive input from a spatially distributed set of glomeruli (Miyamichi et al., 2011). Furthermore, a given PCx
neuron receives converging inputs from multiple M/T cells arising from distinct glomeruli, which suggests
that PCx neurons integrate information across distinct glomeruli and hence Ors (Apicella et al., 2010). This
neuroanatomical organization results in PCx odour responses lacking a clear topographical organization.
Electrophysiological and imaging studies in PCx of rodents (Illig & Haberly, 2003; Rennaker et al., 2007;
Roland et  al., 2017; Stettler  & Axel, 2009) and neuroimaging studies in humans (Gottfried, 2010) have
shown that odours activate distributed ensembles of neurons with unique spatiotemporal activity profiles.
While PCx neural activity patterns can be used to decode odour identity information, studies to date sug-
gest that PCx does not have a chemotopic map (the physical locations of the neural responses do not carry
information about the spatial, physiochemical, or perceptual features of the odour stimuli). The absence of
topographic organization in PCx is yet another feature that differentiates the olfactory system from the audi-
tory, visual, and somatosensory systems, where neurons responding to similar features of sensory stimuli are
spatially clustered together. However, similar to the piriform cortex, the primary gustatory cortex also lacks
a chemotopic organization, and single gustatory cortex neurons can respond to broad range of tastants (Chen
et al., 2021; Dikecligil et al., 2020; Fletcher et al., 2017), which suggests that the cortical representation of
chemical senses is not organized spatially according to the chemical or perceptual features of the stimuli. In
addition to the diffuse ascending projections from the OB, the PCx receives extensive local inputs from other
PCx neurons, the AON, and the orbitofrontal cortex (OFC), as well as the amygdala and entorhinal cortex
(Bekkers & Suzuki, 2013). Given this connectivity, the odour information arriving to the PCx is interpreted
and processed in context of these extensive top-down and auto-associative projections and hence does not
simply reflect the same information arriving from the OB.

152
Odour aesthetics

From the PCx, odour information further diverges; PCx axons project to the OB, AON, and other
olfactory regions, including those that receive direct OB input as well as the OFC, hypothalamus, amygdala,
entorhinal cortex, and mediodorsal thalamus. In addition to the monosynaptic connections listed previously,
the PCx has indirect connections with key regions involved in emotion, memory, hedonics, and cognitive
processing, including the insula, cingulate cortex, nucleus accumbens, hippocampus, and ventral putamen.
Many of these regions are reciprocally connected among themselves as well.
Among the higher-order regions that receive PCx input, OFC is considered the secondary olfactory
cortex, as it receives dense projections from PCx (Carmichael et al., 1994). This projection from the PCx
to OFC provides a pathway for olfactory information to reach neocortex without going through a thalamic
relay. In the past two decades, neuroimaging studies in humans have shown that odour stimuli consistently
evoke activity in the OFC, frequently in a bilateral fashion (Gottfried et al., 2002; Gottfried et al., 2003;
Small et al., 1997; Sobel et al., 1998; Zatorre et al., 1992). Electrophysiology studies in animals have shown
that OFC neurons not only encode odour identity but can respond to olfactory context cues and odour
cue-outcome contingencies (Rolls et al., 1996; Rolls et al., 1996). Current theories suggest that the OFC
plays an important role in integrating odour identity information with expected outcomes to guide odour-
related behaviour and decision making (Gottfried, 2007; Gottfried  & Zald, 2005; Howard et  al., 2015;
W. Li et al., 2010).
Overall, the olfactory system, while specializing in detecting and identifying a vast array of volatile
compounds, does so through highly interconnected olfactory and non-olfactory regions that reciprocally
exchange olfactory, contextual, and neuromodulatory information that ultimately leads to interpretation of

Figure 8.2 Neuroanatomy of the human olfactory regions. (A) A ventral view of the human brain in which the ante-
rior part of the right temporal lobe has been resected in the coronal plane to better visualize the underlying
olfactory and limbic regions. (B) Close-up of the outlined area in panel (A). The outputs of the olfactory
bulb (OB) pass through the lateral olfactory tract (LOT) and send parallel projections to many downstream
olfactory and limbic regions, including the olfactory tubercle (OT), piriform cortex (PCx; anterior portion
of PCx shown in purple and posterior portion of PCx shown in blue), amygdala (AM), and entorhinal cor-
tex (EM). Other key regions involved in processing higher-order olfactory information, the hippocampus
(HP) and olfactory portion of the orbitofrontal cortex (OFC), can also be seen in the ventral view. Figure
modified with permission from Gottfried, J. A. “Central mechanisms of odour object perception.” Nat. Rev.
Neurosci. 11, 628–641 (2010).

153
Gulce Nazli Dikecligil and Jay A. Gottfried

olfactory information in context of the organism’s current physiological, emotional, and cognitive states as
well as past experiences and memories.

Factors that influence odour valence perception


As discussed earlier in the chapter, odour valence perception is one of the key functions of the olfactory
system that complements odour quality perception to guide behaviour. While there are evolutionarily con-
served appetitive and aversive odours that are hardwired in many species (Ferrero et al., 2011; Hallem &
Sternberg, 2008; Kobayakawa et al., 2007; Li et al., 2013; Papes et al., 2010; Semmelhack & Wang, 2009;
Stowers & Logan, 2010; Suh et al., 2004), such as fox odour (Endres & Fendt, 2009; Saito et al., 2017)
or volatiles found in mouse urine (Li et  al., 2013) that promote avoidance and approach behaviours, a
much greater number of odourants acquire their hedonic value through learning (Chao et al., 2004; Li &
Liberles, 2015; Logan et al., 2012; Zhang et al., 2005). Furthermore, the hedonic value of an odourant is
not fixed for a given organism but can be further shaped by contextual cues, multisensory inputs, an organ-
ism’s physiological state, and the properties of the odour stimulus such as its duration and concentration
(Barkat et al., 2008; Bensafi et al., 2007; Chao et al., 2004; Linster & Cleland, 2016; Luisier et al., 2019;
O’Doherty et  al., 2000; Poncelet, Rinck, Bourgeat et  al., 2010; Poncelet, Rinck, Ziessel et  al., 2010;
Small et al., 2001). The olfactory system, in concert with neural circuits that mediate cognitive, limbic,
and homeostatic functions, integrates across these factors to determine the hedonic value of an odour for
that given moment, and hence, as the organism and the circumstances around the odourant change, so
does its potential hedonic interpretation. In the following, we will discuss some of the well-studied factors
that contribute to odour hedonic perception. These factors can be broadly grouped into three categories:
(1) properties of the odour stimulus itself, (2) environmental and contextual cues accompanying the odour,
and, finally, (3) the cognitive and physiological state of the organism that is interacting with the odourant.
We note that social and cultural factors also play a very important role in learned odour-outcome and
odour valence perceptions; however, this topic is beyond the scope of the chapter, and we direct readers
to recent articles on this topic (Coppin et al., 2016; Ferdenzi et al., 2017; Ferdenzi et al., 2013; Ferdenzi
et al., 2011).

Effects of odour stimulus properties on odour valence perception


One of the puzzling aspects of olfaction compared to other sensory systems is the lack of a clear relation-
ship between odour stimulus parameters and odour quality perception (Sell, 2006). In vision and audition,
wavelength of light and sound are the basic building blocks of sensory processing and resulting perceptions.
In contrast, there is no evident parameter of monomolecular odours (e.g., molecule size, functional groups,
carbon chain length, etc.) that can reliably predict aspects of the resulting percept such as odour quality and
pleasantness across individuals.
A handful of studies in recent years (Khan et al., 2007; Koulakov et al., 2011; Ravia et al., 2020; Snitz
et al., 2013) has aimed to assess if large data sets using hundreds of molecular and perceptual features could
unveil any relationship between groups of physiochemical and perceptual features such as odour valence.
One of the first studies using this approach asked subjects (perfumers and olfactory scientists) to rate how well
160 unique odourants matched each of the 146 available verbal descriptors (e.g., “fruity,” “floral,” “woody,”
etc.). After using dimensionality reduction approaches, the authors found that odour pleasantness explained
30% of the variance in the perceptual space (Khan et al., 2007). Additionally, the physiochemical features
that accounted for the largest variance in the molecular space were correlated with odour pleasantness. These
findings suggest that there might be groups of physiochemical properties partly associated with odours that
are found to be pleasant, although molecular properties alone cannot account for the other 70% of variance

154
Odour aesthetics

in odour perception. A similar study where subjects sampled and rated 476 odour molecules for their per-
ceived intensity, pleasantness, and semantic category showed that machine learning algorithms trained on
a subset of the molecule-percept relationships could predict, above chance level, the perceptual outcomes
of novel molecules that the algorithm had not seen before (Keller et al., 2017). Specifically, physiochemi-
cal properties could be used to predict the pleasantness of novel molecules using the algorithms trained on
the existing pairings. Overall, these studies do not necessarily suggest that physiochemical properties dictate
odour percepts per se but that there may be broad associations between groups of physiochemical proper-
ties and general odour quality perceptions. It is possible that these associations arise because odourants from
similar odour sources (e.g., flowers) tend to share molecular features (Dudareva et al., 2006; Dudareva et al.,
2004; Tholl et al., 2011) and are perceptually grouped together due to the similarities in the ecological func-
tion of the odour source (i.e., flowers eliciting approach behaviour). In other words, the olfactory system
might be grouping odour objects based on the functional and ecological relevance of the odour source rather
than the molecular features, but the underlying correlations in molecular features of odours coming from
similar sources might lead to predictable patterns between molecular features and odour quality and valence
percepts.

Odour stimulus intensity, duration, and frequency


Beyond the molecular composition of the volatiles entering the nose, odour stimuli are also defined by the
concentration of the molecules, the duration of the odour plume, and the frequency with which the same
odour object is encountered in a short amount of time. Concentration of odourants in the natural envi-
ronment is subject to change depending on proximity to the odour source, wind, and temperature. Such
changes in odour concentration typically alter the perceived intensity of a given odourant (although the
relationship between concentration and perceived intensity is not linear, and can vary widely across odour-
ants) (Cain, 1969; Moskowitz et al., 1974). Studies investigating whether changes in odour intensity alter
odour valence perception have reported conflicting findings, with some studies showing that increasing
odour intensity results in decreased odour pleasantness (Henion, 1971; Moskowitz et al., 1974; Moskowitz
et al., 1976) and others showing both increased and decreased odour pleasantness depending on the odour
identity being tested (Doty, 1975). Yet another study using food odours (sweet and savoury) reported that
higher concentrations of the same odour were rated to be more pleasant compared to lower concentrations,
suggesting that food odours that are already considered appetitive in low concentrations may become more
pleasant with increasing odour intensity (Howard et al., 2015). A possible explanation is that odour inten-
sity informs the organism about the relative proximity of the odour source and the appropriate behavioural
response that should be taken in response to the odour object depending on this proximity; therefore, the
potential changes in stimulus–outcome relationships as odour intensity changes may alter the hedonic value
of the odourant. For food odours, high odour intensity means that the food is close by, and this may increase
the motivation to seek and consume the food and can alter the hedonic value of the associated food odour.
Likewise, a low concentration predator odour may not be salient enough to generate an escape response,
as it signals that the predator is far away and may be considered hedonically neutral, whereas high-intensity
predator odour might trigger an avoidance response and consequently alter the hedonic value of the odour
(Ferrari et al., 2006; Hacquemand et al., 2013). It is important to note that once inside the nasal cavity,
many odorous molecules activate both the olfactory and the trigeminal nerves (e.g., acetone, ammonia) and
can lead to unpleasant non-olfactory sensations such as stinging or burning that can contribute to decreased
odour valence perception with increasing concentrations (Doty et al., 1978). Taken together, these findings
suggest that odour pleasantness may be altered by the changes in the intensity of the odourant in an odour
identity–dependent manner, with effects mediated by the simultaneous stimulation of the trigeminal nerve
by the odorous molecules.

155
Gulce Nazli Dikecligil and Jay A. Gottfried

In the natural environment, organisms often encounter odourants within complex, turbulent air streams
called odour plumes. Humans and animals can sample odour plumes repeatedly across multiple sniff cycles
to extract valuable information about the location and proximity of the odour source (Raithel & Gottfried,
2021). How, then, does repeated exposure to the same odourant, such as what might happen when a mouse
is sniffing the air for fox odour or a human smelling baked goods from a nearby pastry shop, affect the valence
of the odourant? One study in humans showed that an odour that is initially pleasant becomes less pleasant
with repeated exposure, whereas an odour that is unpleasant becomes less unpleasant (W. S. Cain & Johnson,
1978). This observation that the hedonic value of the odour becomes more neutral with repeated exposure
was found to be consistent with affective habituation, where repeated exposure diminishes the affective
response (Leventhal et al., 2007). It is possible that the observed decreases in odour valence occur as a result of
odour habituation. It has been shown that repetitions of the same odour stimuli evoke decreased responses in
the OSN, OB (Dalton & Wysocki, 1996; Kurahashi & Menini, 1997; Scott, 1977; Zufall & Leinders-Zufall,
2000), and PCx (Li et al., 2006; Sobel et al., 2000).

The effect of internal states and external cues on odour valence perception
While odour stimulus parameters might play a minor role in odour valence perception, a much greater vari-
ance in affective assessment of odourants arises from the cognitive and emotional state of subjects and the
external cues that provide contextual information within which odour stimuli are interpreted. Perception of
both odour quality and hedonic value are highly dependent on the surrounding sensory and contextual cues
that accompany the odour. Presenting the same odour mixture with different labels can influence familiar-
ity, intensity, edibility, and pleasantness ratings (Djordjevic et al., 2008; Herz, 2003; Herz & von Clef, 2001;
Manescu et al., 2014). In one study, subjects smelled the same set of odours under three different semantic
labels; for example, juniper berry odour was either labelled as green mango (a positive label), hospital disin-
fectant (a negative label), or the number “21” (a neutral label) (Djordjevic et al., 2008). Subjects reported the
odours to be more pleasant when they were paired with a positive label compared to a negative or neutral
label. Subjects also showed greater sniff volume and decreased skin conductance measurements for odours
that were labelled with positive names, suggesting that the odour labels affected both perceptual and physi-
ological responses. This effect generalized to 13 out of the 15 odours that varied across the pleasantness
spectrum, implying that the ability of labels to bias odour hedonic perception is not limited solely to pleasant
or unpleasant odours. In the following, we will highlight how contextual information arising from other
sensory modalities can influence odour valence perception.

Multisensory inputs
Sensory information arising from a given sensory modality does not lead to fixed percepts but is rather
interpreted within the context of a cacophony of multimodal sensory cues, giving rise to dynamic, context-
dependent percepts. For the olfactory system, the interpretation of an odour stimulus can be influenced by
the accompanying sights, sounds, and tastes (Dubose et al., 1980; Jadauji et al., 2012; Manesse et al., 2020;
Morrot et al., 2001; Osterbauer et al., 2005; Seo & Hummel, 2011; Seo et al., 2014; Spence et al., 2010;
Zellner et al., 1991). For example, a sour cheese odour can be aversive if one is walking by a dumpster in a
dark alley, or it could be pleasurable if one is having a cheese plate at their favourite restaurant. How does
information arriving from other sensory systems contribute to odour processing and specifically to the
hedonic assessment of odourants? What are the neural mechanisms that might contribute to this dynamic
interpretation of odour hedonics?
Interactions between taste and olfaction, sensory systems that specialize in translating chemical informa-
tion into actionable percepts, inform decisions about which foods to seek, consume, and return to. During

156
Odour aesthetics

food consumption, while tastants stimulate the taste receptor cells on the tongue, volatile molecules emit-
ted from food can travel through retronasal and orthonasal pathways and bind to olfactory receptors. This
simultaneous activation of the olfactory and gustatory systems (along with the oral somatosensory system)
forms the basis of a unitary flavour percept. The frequent pairing of taste–odour combinations during food
consumption (e.g., pairing of banana odour with sweet taste each time someone eats a banana) leads to
strong associations that can bias odour and taste perception when the taste or odour stimuli are presented
individually. For instance, presentation of just the banana odour can yield a “sweet” percept, as banana odour
has almost exclusively been paired with one sweet taste. Likewise, the hedonic value of one modality can
bias the hedonic perception, such as perceiving an otherwise neutral odour as pleasant when it has been
paired with a sweet taste. How quickly can such taste–odour pairings emerge and modulate odour hedonic
perception? In humans, neutral odours that were paired with a sweet solution were later reported to be more
pleasant in as few as three repetitions of the odour–sweet solution pairings (Barkat et al., 2008). A similar
study in rodents using an odour–taste association task (Blankenship et al., 2019) delivered odours either via
the orthonasal or retronasal route and paired them with either water (neutral taste) or sucrose solution (pleas-
ant taste). Following three days of training where the animals were presented with the pairings, the authors
tested rats’ odour preference. Odour preference was acquired faster through the retronasal route compared to
the orthonasal route, and preferences acquired through retronasal odour–taste pairings did not generalize to
the same odour delivered via the orthonasal route. Interestingly, it was shown that the inactivation of primary
gustatory cortex selectively impaired retronasal (but not orthonasal) preferences. These findings suggest that
the odours processed through the retronasal route might have a privileged association with taste stimuli due
to the frequent co-occurrence of retronasal odours and gustatory stimuli during food consumption. Not
surprisingly, electrophysiology studies in rodents have shown that gustatory stimuli can activate piriform
cortex neurons (Maier et al., 2012), and conversely, olfactory stimuli can activate gustatory cortex neural
activity (Samuelsen & Fontanini, 2017) suggesting that cross-modal chemosensory stimuli are represented at
the primary sensory cortices. Furthermore, olfactory cues predicting taste stimuli can be acquired faster than
visual, auditory, or somatosensory cues predicting taste stimuli and can be represented more extensively in
the primary gustatory cortex (Vincis & Fontanini, 2016). Collectively, these findings show that taste–odour
associations can form rapidly and influence the odour hedonic perception.
Olfactory encounters are often accompanied by visual cues that can shape expectations and subsequent
olfactory perceptions. A behavioural study has shown that artificially red-coloured white wine was described
just as if it was a red wine by a panel of wine professionals, suggesting that the colour cues could strongly
bias odour identity perception (Morrot et al., 2001). Other studies have also shown that the colour of the
odour source can influence odour identification and furthermore alter odour hedonic perception: orange-
coloured cherry drinks were reported to be orange flavoured (Dubose et al., 1980), and red-coloured straw-
berries were reported to be more pleasant compared to green-coloured strawberries (Zellner et al., 1991).
Given the strong behavioural influence of visual cues on odour perception, it is not surprising that early
neuroimaging studies have shown that olfactory tasks can activate visual cortices (Qureshy et  al., 2000;
Zatorre et al., 2000), suggesting that subjects may be forming a mental image of the odour source that may
aid in olfactory task performance. In a simple odour-detection task (Gottfried & Dolan, 2003), congruent
picture–odour pairing (e.g., displaying orange pictures while delivering orange odour) versus incongruent
picture–odour pairings (e.g., displaying lettuce picture while delivering orange odour) facilitated rapid odour
detection. Furthermore, congruent pairings activated the anterior hippocampus and OFC, suggesting that
the cross-modal interaction between the two senses might be facilitated by these regions. A different study
also provided complementary evidence where congruent bimodal stimuli (visual and olfactory) resulted in
increased activation in the OFC and insular cortex, key regions for multimodal integration of olfactory and
gustatory inputs to inform flavour perceptions (Osterbauer et al., 2005). A causal study investigating whether
visual cortex activity can influence odour perception ( Jadauji et al., 2012) reported that repetitive transcranial

157
Gulce Nazli Dikecligil and Jay A. Gottfried

magnetic stimulation (TMS) of the early visual cortex, but not auditory cortex, improved performance in an
olfactory discrimination task. Other studies investigated the role of olfactory cues in modulating visual per-
ception and found that odours can influence binocular rivalry (Zhou et al., 2010), visual attention (Seo et al.,
2010), and visual processing of faces (Leleu et al., 2015), raising the possibility that bidirectional interaction
between olfactory and visual systems can meaningfully inform multisensory perception.
Similar to gustatory and visual stimuli, auditory inputs have also been shown to influence olfactory
perception and hedonic assessment. In a study where odours were paired with either congruent (pairing of
Christmas songs with cinnamon smell) or incongruent sounds (pairing of a coffee advertisement with cin-
namon smell), the subjects reported the odours to be more pleasant when they were paired with congruent
sounds (Seo et al., 2014). In a similar experiment, subjects were presented with either a pleasant or unpleas-
ant sound prior to or during the odour presentation. Irrespective of the hedonic value of the odour stimuli,
the subjects’ hedonic judgements were modulated by how much they liked the sound, such that when pleas-
ant sounds were paired with unpleasant odours, the subjects reported the odour as being pleasant (Seo &
Hummel, 2011).

Neural circuits mediating odour hedonics


Neural mechanisms of rewarding and aversive stimuli have been extensively studied in both human and
animal models. While the nuances of how each brain region contributes to hedonic processing is an area of
active research, many regions, including the dorsal and ventral striatum, substantia nigra, orbitofrontal cor-
tex, insula, and anterior cingulate cortex, have consistently been shown to be involved in assessing and guid-
ing subjective value judgement and related decision making. Many of these regions are also involved in odour
hedonic processing, and in the following, we will discuss some of the well-studied brain regions in context
of odour valence coding. We will discuss findings from both human and animal studies; while experiments
in humans can provide greater insight about the subjective perceptual experience, animal studies can provide
finer details about the underlying neural circuits given the wider range of experimental techniques available
to study the brain. However, unlike human studies, where subjects can be asked to report their subjective
experience of an odourant, animal studies must rely on behavioural assays such as odour investigation time,
odour conditioned place preference, and other quantifiable approach and avoidance behaviours to interpret
whether an animal finds an odour appetitive or aversive. It is important to note that the results of these assays
may be confounded with other behavioural factors and psychological states that may drive approach and
avoidance behaviours such as curiosity, anxiety, hunger, exploration, or neo-phobia. Furthermore, animals
may spend more time investigating odours that they have a harder time identifying due to the speed–accuracy
trade-off (Rinberg et al., 2006), and hence odour investigation time may not correlate directly with how
much an animal likes an odour stimuli per se. These behavioural caveats should be considered carefully when
interpreting the results of animal studies.
In recent years, one region in the olfactory system has consistently been shown to be involved in odour
valence processing in rodent studies, the olfactory tubercle (OT). The OT receives direct projections from
the OB and is reciprocally connected with many regions in the olfactory network (OB, AON, PCx, OFC),
as well as regions associated with the reward circuitry, including the ventral tegmental area (VTA) (Z. Zhang,
Zhang et  al., 2017). The neuroanatomical connectivity of the OT supports integration of olfactory and
reward-related signals to influence behavioural output. Furthermore, modulation of OT activity can alter
motivated behaviours such as mating (Hitt et al., 1973) or preference for conspecific chemosignals (DiBen-
edictis et  al., 2015). Single neuron recordings from OT of awake rodents have found that OT neurons
preferentially encode odours that predict a reward outcome, and these reward-specific responses in the OT
emerge prior to motor movement (Gadziola et al., 2015). This finding suggests that OT plays an important
role in encoding odour–reward contingency and can be informing subsequent decision making. Another

158
Odour aesthetics

study using c-Fos activation to assess neural activity has shown that when mice were conditioned to associ-
ate odours with either a reward or an electric shock, odours that predicted reward versus aversive outcomes
activated distinct areas of the OT (Murata et al., 2015). These findings suggest that subregions of the OT
can encode odour-associated outcomes according to their hedonic value, independently from odour iden-
tity. A recent study investigated how learned odour–reward associations emerged in the OT and whether
the OT and posterior PC encoded odour–reward associations in a similar way (Millman & Murthy, 2020).
The authors found that mice could learn odour–reward contingencies within a single experimental session,
and OT neurons showed both odour identity–specific and reward-outcome–specific responses as early as 50
ms after the odour inhalation. In contrast, neurons in the posterior PC, while encoding odour identity, did
not show reward-specific responses.
Other studies have also investigated VTA to OT projection and its role in reward-related behaviour and
odour hedonics. Optogenetic activation of VTA to OT projections could elicit preference for an otherwise
neutral odour, and blocking OT dopaminergic receptors prevented the formation of preference for neutral
odours (Zhang, Liu et al., 2017). Interestingly, this pathway may be important for maintaining the learned
odour preference, as inactivation of dopaminergic input from VTA to OT eliminated the odour preference
that was already acquired.
Appetitive or aversive associative olfactory learning where organisms learn the relationship between a
neutral odour stimulus and a positive or negative outcome is a key mechanism that shapes hedonic value
of odours and odour responses (Barkat et al., 2008; Gottfried et al., 2002; Herz, 2005; Li, 2014; Li et al.,
2008). Animals can learn within a few trials that an odour may predict an electric shock response and display
avoidance response to the odourant (Herzog & Otto, 1997; Kucharski & Spear, 1984; Raineki et al., 2010).
Alternatively, an animal can rapidly learn that an odour predicts sucrose reward and display anticipatory
mouth movements in response to the odour (Vincis & Fontanini, 2016). Across different sensory stimuli and
paradigms, it has been shown that acquisition and retention of associations involves brain regions involved in
sensory, emotion, and memory networks (Fanselow & Poulos, 2005). While distinct experimental paradigms
differentially engage the previously mentioned networks, both the hippocampus and amygdala have been
shown to play a key role across a wide range of associative learning paradigms (Cousens & Otto, 1998; Davis
et al., 1994; Grace & Rosenkranz, 2002; Phillips & LeDoux, 1992).
The hippocampus receives olfactory information through the amygdala and entorhinal cortex, structures
that both receive direct input from the OB and PCx (Insausti et al., 2002; Keshavarzi et al., 2015; Schw-
erdtfeger et al., 1990; Uva & de Curtis, 2005). Hippocampal lesions in rodents impair olfactory memory
(Eichenbaum, 1998; Ergorul & Eichenbaum, 2004) and odour–place associative memory (Gilbert & Kesner,
2002). Studies in humans have shown that patients with bilateral hippocampal lesions can identify odours just
as well as control subjects but perform poorly in an odour–place association task where they need to learn
the pairing between an odour and a specific location (Goodrich-Hunsaker et al., 2009). An fMRI study
(Yeshurun et al., 2009) in humans investigated whether the temporal order and hedonic value of an odourant
affected strength of associative learning. The authors paired a sequence of pleasant or unpleasant olfactory
(or auditory stimuli) with visual stimuli and found that first pairings were better remembered as well as pair-
ing with unpleasant olfactory stimuli. Furthermore, the strength of activity in the hippocampus predicted
the recall of olfactory–visual associations (but not audio–visual associations). These studies highlight that the
hippocampus plays an important role in olfactory associative learning in both humans and animals, and the
valence of olfactory stimuli can affect the strength of the learned associations.
The amygdala, one of the most-studied brain regions involved in processing of emotional and salient
stimuli, receives direct projections from both the OB and PCx, and animal recordings have shown that amyg-
dala neurons respond to odours and these responses are modulated by learning (Cain & Bindra, 1972; Gore
et al., 2015; Grace & Rosenkranz, 2002; Schoenbaum et al., 1999; Sevelinges et al., 2004). Furthermore,
lesions of the lateral amygdala prevent olfactory associative learning (Cousens & Otto, 1998; Sevelinges et al.,

159
Gulce Nazli Dikecligil and Jay A. Gottfried

2009; Sevelinges et  al., 2004). Intracellular recordings from basolateral amygdala of anesthetized rodents
showed that when an odour is paired with a foot shock, the neural responses to the conditioned odour is
stronger and the repeated presentation of this odour does not lead to habituation as it does in unconditioned
odours. Interestingly, the administration of a dopamine antagonist prevents the acquisition of new odour–
foot shock conditioning but does not interfere with the previously acquired odour–shock pairings (Grace &
Rosenkranz, 2002; Rosenkranz & Grace, 2002) suggesting that dopamine-mediated basolateral amygdala
activity is crucial for acquiring new odour associations. A qualitative meta-analysis of human olfactory neu-
roimaging studies using functional magnetic resonance imaging (fMRI) and positron emission tomography
imaging (Patin & Pause, 2015) reports that a wide range of pleasant odours evoke bilateral amygdala activa-
tion, although in some studies, responses to pleasant odours were limited to either the left or right amygdala
only. Unpleasant odours, including mild- and moderate-strength stimuli, were also found to evoke bilateral
amygdala activation across many studies. Across studies, odours with higher concentrations and greater emo-
tional salience led to greater amygdala activation compared to lower concentration odours and hedonically
neutral odours. In one neuroimaging study investigating how odour valence is encoded in the human brain
and whether it is dissociated from odour intensity, the authors reported that amygdala activity is associated
with odour intensity, whereas orbitofrontal cortex activity is associated with odour valence (Anderson et al.,
2003). However, later studies have shown that the amygdala does respond to a range of odour valences, sug-
gesting that both regions may be involved in assessing and responding to odour valence ( Jin et al., 2015).
Similar to animal studies where hedonic representations in the amygdala have been shown to be malleable,
human studies also showed that top-down cognitive factors can influence olfactory hedonic judgement and
related amygdala activity. One study investigated whether subjects’ hedonic perception and corresponding
amygdala activation could be altered by presenting the same odour mixture under different labels (isovaleric
acid and cheddar cheese flavour mixture presented as either “body odour” or “cheddar cheese”) (de Araujo
et al., 2005). Subjects rated the odour mixture as being more unpleasant when it was paired with a “body
odour” label, whereas the level of amygdala activity (blood-oxygen-level dependent [BOLD] signal) was
positively correlated with pleasantness ratings. Collectively, these findings suggest that the amygdala plays an
important role in evaluation of olfactory stimuli in context of past experiences and present cognitive context.
The OFC, given its extensive inputs from the PCx and higher-order cortical regions, is critically involved
in integrating olfactory information with cognitive processing and value information. Early electrophysiol-
ogy studies in monkeys showed different subsets of OFC neurons respond to odour–identity and odour–taste
associations (Rolls, Critchley, Mason et al., 1996). A separate study showed that OFC neurons that respond
to odour–outcome associations on the basis of the outcome rather than odour identity can also change their
responses when associations are reversed (Rolls, Critchley, & Treves, 1996). Meta-analysis of human neuroim-
aging studies using pleasant odours reported consistent OFC activation across studies (pleasant versus neutral
or pleasant versus no odour contrasts), and the activity was greater when subjects were asked to report their
subjective experience compared to when they were passively smelling (Zou et al., 2016). This finding suggests
that conscious access to the olfactory hedonic value might depend on the OFC activity. In fact, a case study in a
patient with right OFC traumatic brain injury provides supporting evidence to this role of OFC (Li et al., 2010);
an otherwise healthy 33-year-old patient with no history of smell loss developed anosmia upon traumatic brain
injury limited to the right OFC. Surprisingly, the patient displayed differential physiological and neural (PCx)
responses to unpleasant odours in contrast to neutral odours as assessed by skin conductance measurements and
fMRI, but when asked, he was unable to detect presence of odours above chance level. These findings suggest
that even though the early stages of the olfactory system are intact and responsive to odourants, the right OFC
lesion interfered with the patient’s ability to consciously access and report their odour experience.
Similar to single neuron studies in animals showing plasticity of OFC reward value representations, human
neuroimaging studies have also shown that OFC can dynamically update reward contingency of odours.
Sensory-specific satiety studies (Gottfried et al., 2003; O’Doherty et al., 2000; Small et al., 2001), where the

160
Odour aesthetics

subjects consume a food item until it is no longer palatable, report that the odour associated with the satiated
food item decreases in pleasantness and, in conjunction, the OFC activation in response to the satiated food
odour is reduced. Does OFC activity merely encode information about the rewarding properties of odours
(affective value), or does OFC also encode information pertaining to what the stimulus is (e.g., is it chocolate
cake or pizza)? A human neuroimaging study investigated this question using a classic conditioning paradigm
where the subjects were presented with visual cues that predicted four distinct odour outcomes (sweet or
savoury food odours) at two different concentrations (low or high concentration, which were respectively
perceived as moderately or very pleasant) (Howard et al., 2015). The authors studied the patterns of OFC
activity evoked by visual cues in anticipation of the rewarding food odours and found that the OFC can
indeed encode odour identity information for odours that were rated to be equally rewarding. This finding,
in combination with the aforementioned single neuron recordings, suggests that the OFC simultaneously
represents both predicted reward value of odours and their identity.
One framework (Berridge et al., 2009; Pool et al., 2016) on reward-related behaviour proposes that reward
behaviour can be studied as two dissociable psychological and neurobiological constructs that are distinct yet
highly correlated: liking and wanting (see also Chapter 3). Liking corresponds to the subjective experience
associated with a reward (e.g., how likeable the chocolate flavour is when it is in the mouth), whereas want-
ing corresponds to the anticipatory motivational aspect that drives approach behaviour to expected rewards
(e.g., the motivation to seek and consume chocolate). In an fMRI experiment investigating whether wanting
and liking of food odours recruit distinct neural circuits ( Jiang et al., 2015), participants were asked to rate
the pleasantness of food odour (liking) as well as their desire to consume the food item corresponding to the
odour stimuli (wanting) either in a hungry or satiated state. The authors found that when subjects were rating
how much they liked the food odours, there was greater activity in nucleus accumbens (NAcc) compared to
the condition in which subjects were rating how much they wanted to consume the food item. Interestingly,
this effect was reversed when subjects were hungry, with the NAcc showing greater activity in response to
the ratings of wanting. Furthermore, when subjects were hungry, there was greater activity in the ventral
pallidum for the wanting condition compared to liking ratings, and this effect was not present when subjects
were satiated. These findings suggest that the neural circuits mediating odour liking and wanting recruit dif-
ferent areas in a motivational state–dependent manner.

Conclusion
The extent to which odours are found to be rewarding, aversive, or neutral varies greatly across individuals,
cultures, and even within an individual depending on the individual’s cognitive and emotional state and the
signals surrounding the odour stimuli. The olfactory system, in concert with emotion, memory, and cogni-
tive systems, integrates across a multitude of past and current factors in the context of an organism’s goals and
motivations to determine the hedonic value of the odours it encounters. The complexity and nonlinearity
of the factors that contribute to odour hedonics perhaps make it such that it may be nearly impossible to
predict the hedonic judgement of odourants in a deterministic fashion. We argue that this is a strength rather
than a weakness of studying odour hedonics; it opens a window into how the physical world is interpreted
and reinterpreted in the nose of the beholder, unfolding the mysteries of how our internal state, experiences,
biases, and goals shape our subjective experience and allowing for a scientific framework to investigate the
basis and complexity of aesthetic preferences.

References
Amo, L., Galvan, I., Tomas, G., & Sanz, J. J. (2008). Predator odour recognition and avoidance in a songbird. Functional
Ecology, 22(2), 289–293. https://doi.org/10.1111/j.1365-2435.2007.01361.x

161
Gulce Nazli Dikecligil and Jay A. Gottfried

Amo, L., Visser, M. E., & van Oers, K. (2011). Smelling out predators is innate in birds. Ardea, 99(2), 177–184. https://
doi.org/10.5253/078.099.0207
Anderson, A. K., Christoff, K., Stappen, I., Panitz, D., Ghahremani, D. G., Glover, G., Gabrieli, J. D., & Sobel, N.
(2003). Dissociated neural representations of intensity and valence in human olfaction. Nature Neuroscience, 6(2),
196–202. https://doi.org/10.1038/nn1001
Apicella, A., Yuan, Q., Scanziani, M., & Isaacson, J. S. (2010). Pyramidal cells in piriform cortex receive convergent
input from distinct olfactory bulb glomeruli. Journal of Neuroscience, 30(42), 14255–14260. https://doi.org/10.1523/
JNEUROSCI.2747-10.2010
Arakawa, H., Blanchard, D. C., Arakawa, K., Dunlap, C., & Blanchard, R. J. (2008). Scent marking behavior as an odor-
ant communication in mice. Neuroscience and Biobehavioral Reviews, 32(7), 1236–1248. https://doi.org/10.1016/j.
neubiorev.2008.05.012
Arenas, A., & Farina, W. M. (2012). Learned olfactory cues affect pollen-foraging preferences in honeybees, Apis mel-
lifera. Animal Behaviour, 83(4), 1023–1033. https://doi.org/10.1016/j.anbehav.2012.01.026
Banks, P. B., Daly, A., & Bytheway, J. P. (2016). Predator odours attract other predators, creating an olfactory web of infor-
mation. Biology Letters, 12(5). https://doi.org/10.1098/rsbl.2015.1053:ARTN 2015105310.1098/rsbl.2015.1053.
Barkat, S., Poncelet, J., Landis, B. N., Rouby, C., & Bensafi, M. (2008). Improved smell pleasantness after odor-taste
associative learning in humans. Neuroscience Letters, 434(1), 108–112. https://doi.org/10.1016/j.neulet.2008.01.037
Bartolo, M., Serrao, M., Gamgebeli, Z., Alpaidze, M., Perrotta, A., Padua, L., Pierelli, F., Nappi, G., & Sandrini, G.
(2013). Modulation of the human nociceptive flexion reflex by pleasant and unpleasant odors. Pain, 154(10), 2054–
2059. https://doi.org/10.1016/j.pain.2013.06.032
Bekkers, J. M., & Suzuki, N. (2013). Neurons and circuits for odor processing in the piriform cortex. Trends in Neuro-
sciences, 36(7), 429–438. https://doi.org/10.1016/j.tins.2013.04.005
Bensafi, M., Rinck, F., Schaal, B., & Rouby, C. (2007). Verbal cues modulate hedonic perception of odors in 5-year-old
children as well as in adults. Chemical Senses, 32(9), 855–862. https://doi.org/10.1093/chemse/bjm055
Berglund, B., Berglund, U., Engen, T., & Ekman, G. (1973). Multidimensional analysis of 21 odors. Scandinavian Journal
of Psychology, 14(2), 131–137. https://doi.org/10.1111/j.1467-9450.1973.tb00104.x
Berridge, K. C., Robinson, T. E., & Aldridge, J. W. (2009). Dissecting components of reward: “liking”, “wanting”, and
learning. Current Opinion in Pharmacology, 9(1), 65–73. https://doi.org/10.1016/j.coph.2008.12.014
Blankenship, M. L., Grigorova, M., Katz, D. B., & Maier, J. X. (2019). Retronasal odor perception requires taste cortex,
but orthonasal does not. Current Biology, 29(1), 62–69.e3. https://doi.org/10.1016/j.cub.2018.11.011
Blomqvist, E. H., Brämerson, A., Stjärne, P., & Nordin, S. (2004). Consequences of olfactory loss and adopted coping
strategies. Rhinology, 42(4), 189–194.
Bonfils, P., Avan, P., Faulcon, P.,  & Malinvaud, D. (2005). Distorted odorant perception: Analysis of a series of 56
patients with parosmia. Archives of Otolaryngology–Head and Neck Surgery, 131(2), 107–112. https://doi.org/10.1001/
archotol.131.2.107
Boyd, A. M., Kato, H. K., Komiyama, T., & Isaacson, J. S. (2015). Broadcasting of cortical activity to the olfactory bulb.
Cell Reports, 10(7), 1032–1039. https://doi.org/10.1016/j.celrep.2015.01.047
Boyd, A. M., Sturgill, J. F., Poo, C., & Isaacson, J. S. (2012). Cortical feedback control of olfactory bulb circuits. Neuron,
76(6), 1161–1174. https://doi.org/10.1016/j.neuron.2012.10.020
Buehlmann, C., Graham, P., Hansson, B. S., & Knaden, M. (2015). Desert ants use olfactory scenes for navigation. Ani-
mal Behaviour, 106, 99–105. https://doi.org/10.1016/j.anbehav.2015.04.029
Bushdid, C., Magnasco, M. O., Vosshall, L. B., & Keller, A. (2014). Humans can discriminate more than 1 trillion olfac-
tory stimuli. Science, 343(6177), 1370–1372. https://doi.org/10.1126/science.1249168
Cain, D. P., & Bindra, D. (1972). Responses of amygdala single units to odors in the rat. Experimental Neurology, 35(1),
98–110. https://doi.org/10.1016/0014-4886(72)90062-3
Cain, W. S. (1969). Odor intensity. Differences in exponent of psychophysical function. Perception and Psychophysics, 6(6)
(6a), 349–354. https://doi.org/ 10.3758/Bf03212789
Cain, W. S. (1979). To know with the nose: Keys to odor identification. Science, 203(4379), 467–470. https://doi.
org/10.1126/science.760202
Cain, W. S., de Wijk, R., Lulejian, C., Schiet, F., & See, L. C. (1998). Odor identification: Perceptual and semantic
dimensions. Chemical Senses, 23(3), 309–326. https://doi.org/10.1093/chemse/23.3.309
Cain, W. S., & Johnson, F., Jr. (1978). Lability of odor pleasantness: Influence of mere exposure. Perception, 7(4), 459–465.
https://doi.org/10.1068/p070459
Cain, W. S., Schmidt, R., & Wolkoff, P. (2007). Olfactory detection of ozone and D-limonene: Reactants in indoor
spaces. Indoor Air, 17(5), 337–347. https://doi.org/10.1111/j.1600-0668.2007.00476.x

162
Odour aesthetics

Carmichael, S. T., Clugnet, M. C., & Price, J. L. (1994). Central olfactory connections in the macaque monkey. Journal
of Comparative Neurology, 346(3), 403–434. https://doi.org/10.1002/cne.903460306
Chae, H., Kepple, D. R., Bast, W. G., Murthy, V. N., Koulakov, A. A., & Albeanu, D. F. (2019). Mosaic representations
of odors in the input and output layers of the mouse olfactory bulb. Nature Neuroscience, 22(8), 1306–1317. https://
doi.org/10.1038/s41593-019-0442-z
Chao, M. Y., Komatsu, H., Fukuto, H. S., Dionne, H. M., & Hart, A. C. (2004). Feeding status and serotonin rapidly
and reversibly modulate a Caenorhabditis elegans chemosensory circuit. Proceedings of the National Academy of Sciences of
the United States of America, 101(43), 15512–15517. https://doi.org/10.1073/pnas.0403369101
Chauvin, C., & Thierry, B. (2005). Tonkean macaques orient their food search from olfactory cues conveyed by conspe-
cifics. Ethology, 111(3), 301–310, https://doi.org/10.1111/j.1439-0310.2004.01066.x
Chen, K., Kogan, J. F., & Fontanini, A. (2021). Spatially distributed representation of taste quality in the gustatory insular
cortex of behaving mice. Current Biology, 31(2), 450. https://doi.org/10.1016/j.cub.2020.11.038
Clarke, P. M. R., Barrett, L., & Henzi, S. P. (2009). What role do olfactory cues play in chacma baboon mating? American
Journal of Primatology, 71(6), 493–502. https://doi.org/10.1002/ajp.20678
Cleland, T. A., & Linster, C. (2012). On-center/inhibitory-surround decorrelation via intraglomerular inhibition in the
olfactory bulb glomerular layer. Frontiers in Integrative Neuroscience, 6, 5. https://doi.org/10.3389/fnint.2012.00005
Cook, S., Kokmotou, K., Soto, V., Fallon, N., Tyson-Carr, J., Thomas, A., Giesbrecht, T., Field, M., & Stancak, A.
(2017). Pleasant and unpleasant odour-face combinations influence face and odour perception: An event-related
potential study. Behavioural Brain Research, 333, 304–313. https://doi.org/10.1016/j.bbr.2017.07.010
Coppin, G., Pool, E., Delplanque, S., Oud, B., Margot, C., Sander, D., & Van Bavel, J. J. (2016). Swiss identity smells like
chocolate: Social identity shapes olfactory judgments. Scientific Reports, 6, 34979. https://doi.org/10.1038/srep34979
Cousens, G., & Otto, T. (1998). Both pre- and posttraining excitotoxic lesions of the basolateral amygdala abolish the
expression of olfactory and contextual fear conditioning. Behavioral Neuroscience, 112(5), 1092–1103. https://doi.
org/10.1037//0735-7044.112.5.1092
Croy, I., Negoias, S., Novakova, L., Landis, B. N., & Hummel, T. (2012). Learning about the functions of the olfac-
tory system from people without a sense of smell. PLOS ONE, 7(3), e33365. https://doi.org/10.1371/journal.
pone.0033365
Croy, I., Nordin, S., & Hummel, T. (2014). Olfactory disorders and quality of life-an updated review. Chemical Senses,
39(3), 185–194. https://doi.org/10.1093/chemse/bjt072
Dalton, P., & Wysocki, C. J. (1996). The nature and duration of adaptation following long-term odor exposure. Perception
and Psychophysics, 58(5), 781–792. https://doi.org/10.3758/bf03213109
Davis, M., Rainnie, D., & Cassell, M. (1994). Neurotransmission in the rat amygdala related to fear and anxiety. Trends
in Neurosciences, 17(5), 208–214. https://doi.org/10.1016/0166-2236(94)90106-6
Davison, I. G., & Katz, L. C. (2007). Sparse and selective odor coding by mitral/tufted neurons in the main olfactory
bulb. Journal of Neuroscience, 27(8), 2091–2101. https://doi.org/10.1523/JNEUROSCI.3779-06.2007
de Araujo, I. E., Rolls, E. T., Velazco, M. I., Margot, C., & Cayeux, I. (2005). Cognitive modulation of olfactory process-
ing. Neuron, 46(4), 671–679. https://doi.org/10.1016/j.neuron.2005.04.021
DiBenedictis, B. T., Olugbemi, A. O., Baum, M. J., & Cherry, J. A. (2015). DREADD-induced silencing of the medial
olfactory tubercle disrupts the preference of female mice for opposite-sex chemosignals(1,2,3). eNeuro, 2(5). https://
doi.org/10.1523/ENEURO.0078-15.2015
Dikecligil, G. N., Graham, D. M., Park, I. M., & Fontanini, A. (2020). Layer- and cell type-specific response proper-
ties of gustatory cortex neurons in awake mice. Journal of Neuroscience, 40(50), 9676–9691. https://doi.org/10.1523/
JNEUROSCI.1579-19.2020
Djordjevic, J., Lundstrom, J. N., Clément, F., Boyle, J. A., Pouliot, S., & Jones-Gotman, M. (2008). A rose by any other
name: Would it smell as sweet? Journal of Neurophysiology, 99(1), 386–393. https://doi.org/10.1152/jn.00896.2007
Doty, R. L. (1975). Examination of relationships between pleasantness, intensity, and concentration of 10 odorous
stimuli. Perception and Psychophysics, 17(5), 492–496. https://doi.org/10.3758/BF03203300
Doty, R. L., Brugger, W. E., Jurs, P. C., Orndorff, M. A., Snyder, P. J., & Lowry, L. D. (1978). Intranasal trigeminal
stimulation from odorous volatiles: Psychometric responses from anosmic and normal humans. Physiology and Behav-
ior, 20(2), 175–185. https://doi.org/10.1016/0031-9384(78)90070-7
D’Souza, R. D.,  & Vijayaraghavan, S. (2014). Paying attention to smell: Cholinergic signaling in the olfactory bulb.
Frontiers in Synaptic Neuroscience, 6, 21. https://doi.org/10.3389/fnsyn.2014.00021
Dubose, C. N., Cardello, A. V., & Maller, O. (1980). Effects of colorants and flavorants on identification, perceived flavor
intensity, and hedonic quality of fruit-flavored beverages and cake. Journal of Food Science, 45(5), 1393–1399. https://
doi.org/10.1111/j.1365-2621.1980.tb06562.x

163
Gulce Nazli Dikecligil and Jay A. Gottfried

Dudareva, N., Negre, F., Nagegowda, D. A., & Orlova, I. (2006). Plant volatiles: Recent advances and future perspec-
tives. Critical Reviews in Plant Sciences, 25(5), 417–440. https://doi.org/10.1080/07352680600899973
Dudareva, N., Pichersky, E., & Gershenzon, J. (2004). Biochemistry of plant volatiles. Plant Physiology, 135(4), 1893–
1902. https://doi.org/10.1104/pp.104.049981
Eichenbaum, H. (1998). Using olfaction to study memory. Olfaction and taste Xii. Annals of the New York Academy of
Sciences, 855, 657–669. https://doi.org/10.1111/j.1749-6632.1998.tb10642.x
Endres, T., & Fendt, M. (2009). Aversion- vs fear-inducing properties of 2,4,5-trimethyl-3-thiazoline, a component of
fox odor, in comparison with those of butyric acid. Journal of Experimental Biology, 212(Pt 15), 2324–2327. https://
doi.org/10.1242/jeb.028498
Ergorul, C., & Eichenbaum, H. (2004). The hippocampus and memory for “what,” “where,” and “when”. Learning and
Memory, 11(4), 397–405. https://doi.org/10.1101/lm.73304
Escanilla, O., Yuhas, C., Marzan, D., & Linster, C. (2009). Dopaminergic modulation of olfactory bulb processing affects
odor discrimination learning in rats. Behavioral Neuroscience, 123(4), 828–833. https://doi.org/10.1037/a0015855
Fanselow, M. S., & Poulos, A. M. (2005). The neuroscience of mammalian associative learning. Annual Review of Psychol-
ogy, 56, 207–234. https://doi.org/10.1146/annurev.psych.56.091103.070213
Ferdenzi, C., Joussain, P., Digard, B., Luneau, L., Djordjevic, J., & Bensafi, M. (2017). Individual differences in verbal
and non-verbal affective responses to smells: Influence of odor label across cultures. Chemical Senses, 42(1), 37–46.
https://doi.org/10.1093/chemse/bjw098
Ferdenzi, C., Roberts, S. C., Schirmer, A., Delplanque, S., Cekic, S., Porcherot, C., Cayeux, I., Sander, D., & Grand-
jean, D. (2013). Variability of affective responses to odors: Culture, gender, and olfactory knowledge. Chemical Senses,
38(2), 175–186. https://doi.org/10.1093/chemse/bjs083
Ferdenzi, C., Schirmer, A., Roberts, S. C., Delplanque, S., Porcherot, C., Cayeux, I., Velazco, M. I., Sander, D., Scherer,
K. R., & Grandjean, D. (2011). Affective dimensions of odor perception: A comparison between Swiss, British, and
Singaporean populations. Emotion, 11(5), 1168–1181. https://doi.org/10.1037/a0022853
Ferrari, M. C. O., Capitania-Kwok, T.,  & Chivers, D. P. (2006). The role of learning in the acquisition of threat-
sensitive responses to predator odours. Behavioral Ecology and Sociobiology, 60(4), 522–527. https://doi.org/10.1007/
s00265-006-0195-z
Ferrero, D. M., Lemon, J. K., Fluegge, D., Pashkovski, S. L., Korzan, W. J., Datta, S. R., Spehr, M., Fendt, M., & Liber-
les, S. D. (2011). Detection and avoidance of a carnivore odor by prey. Proceedings of the National Academy of Sciences of
the United States of America, 108(27), 11235–11240. https://doi.org/10.1073/pnas.1103317108
Fletcher, M. L., Ogg, M. C., Lu, L., Ogg, R. J., & Boughter, J. D., Jr. (2017). Overlapping representation of primary
tastes in a defined region of the gustatory cortex. Journal of Neuroscience, 37(32), 7595–7605. https://doi.org/10.1523/
JNEUROSCI.0649-17.2017
Frasnelli, J.,  & Hummel, T. (2005). Olfactory dysfunction and daily life. European Archives of Oto-Rhino-Laryngology,
262(3), 231–235. https://doi.org/10.1007/s00405-004-0796-y
Gadziola, M. A., Tylicki, K. A., Christian, D. L., & Wesson, D. W. (2015). The olfactory tubercle encodes odor valence
in behaving mice. Journal of Neuroscience, 35(11), 4515–4527. https://doi.org/10.1523/JNEUROSCI.4750-14.2015
Geramita, M. A., Burton, S. D., & Urban, N. N. (2016). Distinct lateral inhibitory circuits drive parallel processing of
sensory information in the mammalian olfactory bulb. eLife, 5. https://doi.org/10.7554/eLife.16039
Ghosh, S., Larson, S. D., Hefzi, H., Marnoy, Z., Cutforth, T., Dokka, K., & Baldwin, K. K. (2011). Sensory maps in
the olfactory cortex defined by long-range viral tracing of single neurons. Nature, 472(7342), 217–220. https://doi.
org/10.1038/nature09945
Gilbert, P. E.,  & Kesner, R. P. (2002). Role of the rodent hippocampus in paired-associate learning involving asso-
ciations between a stimulus and a spatial location. Behavioral Neuroscience, 116(1), 63–71. https://doi.org/10.1037/
0735-7044.116.1.63
Godfrey, P. A., Malnic, B., & Buck, L. B. (2004). The mouse olfactory receptor gene family. Proceedings of the National
Academy of Sciences of the United States of America, 101(7), 2156–2161. https://doi.org/10.1073/pnas.0308051100
Goodrich-Hunsaker, N. J., Gilbert, P. E., & Hopkins, R. O. (2009). The role of the human hippocampus in odor-place
associative memory. Chemical Senses, 34(6), 513–521. https://doi.org/10.1093/chemse/bjp026
Gore, F., Schwartz, E. C., Brangers, B. C., Aladi, S., Stujenske, J. M., Likhtik, E., Russo, M. J., Gordon, J. A., Salzman,
C. D., & Axel, R. (2015). Neural representations of unconditioned stimuli in basolateral amygdala mediate innate and
learned responses. Cell, 162(1), 134–145. https://doi.org/10.1016/j.cell.2015.06.027
Gottfried, J. A. (2007). What can an orbitofrontal cortex-endowed animal do with smells? Annals of the New York Academy
of Sciences, 1121, 102–120. https://doi.org/10.1196/annals.1401.018
Gottfried, J. A. (2010). Central mechanisms of odour object perception. Nature Reviews. Neuroscience, 11(9), 628–641.
https://doi.org/10.1038/nrn2883

164
Odour aesthetics

Gottfried, J. A., Deichmann, R., Winston, J. S., & Dolan, R. J. (2002). Functional heterogeneity in human olfactory
cortex: An event-related functional magnetic resonance imaging study. Journal of Neuroscience, 22(24), 10819–10828.
https://doi.org/10.1523/JNEUROSCI.22-24-10819.2002
Gottfried, J. A., & Dolan, R. J. (2003). The nose smells what the eye sees: Crossmodal visual facilitation of human olfac-
tory perception. Neuron, 39(2), 375–386. https://doi.org/10.1016/s0896-6273(03)00392-1
Gottfried, J. A., O’Doherty, J.,  & Dolan, R. J. (2002). Appetitive and aversive olfactory learning in humans studied
using event-related functional magnetic resonance imaging. Journal of Neuroscience, 22(24), 10829–10837. https://doi.
org/10.1523/JNEUROSCI.22-24-10829.2002
Gottfried, J. A., O’Doherty, J., & Dolan, R. J. (2003). Encoding predictive reward value in human amygdala and orbito-
frontal cortex. Science, 301(5636), 1104–1107. https://doi.org/10.1126/science.1087919
Gottfried, J. A., & Zald, D. H. (2005). On the scent of human olfactory orbitofrontal cortex: Meta-analysis and com-
parison to non-human primates. Brain Research. Brain Research Reviews, 50(2), 287–304. https://doi.org/10.1016/j.
brainresrev.2005.08.004
Grace, A. A., & Rosenkranz, J. A. (2002). Regulation of conditioned responses of basolateral amygdala neurons. Physiol-
ogy and Behavior, 77(4–5), 489–493. https://doi.org/10.1016/s0031-9384(02)00909-5
Gschwend, O., Beroud, J., Vincis, R., Rodriguez, I., & Carleton, A. (2016). Dense encoding of natural odorants by
ensembles of sparsely activated neurons in the olfactory bulb. Scientific Reports, 6, 36514. https://doi.org/10.1038/
srep36514
Guérin, D., Peace, S. T., Didier, A., Linster, C., & Cleland, T. A. (2008). Noradrenergic neuromodulation in the olfac-
tory bulb modulates odor habituation and spontaneous discrimination. Behavioral Neuroscience, 122(4), 816–826.
https://doi.org/10.1037/a0012522
Hacquemand, R., Choffat, N., Jacquot, L.,  & Brand, G. (2013). Comparison between low doses of TMT and cat
odor exposure in anxiety- and fear-related behaviors in mice. Behavioural Brain Research, 238, 227–231. https://doi.
org/10.1016/j.bbr.2012.10.014
Hallem, E. A., & Sternberg, P. W. (2008). Acute carbon dioxide avoidance in Caenorhabditis elegans. Proceedings of the National
Academy of Sciences of the United States of America, 105(23), 8038–8043. https://doi.org/10.1073/pnas.0707469105
Henion, K. E. (1971). Odor pleasantness and intensity: A single dimension? Journal of Experimental Psychology, 90(2),
275–279. https://doi.org/10.1037/h0031549
Herz, R. S. (2003). The effect of verbal context on olfactory perception. Journal of Experimental Psychology. General,
132(4), 595–606. https://doi.org/10.1037/0096-3445.132.4.595
Herz, R. S. (2005). Odor-associative learning and emotion: Effects on perception and behavior. Chemical Senses,
30(Suppl. 1), i250–i251. https://doi.org/10.1093/chemse/bjh209
Herz, R. S., & von Clef, J. (2001). The influence of verbal labeling on the perception of odors: Evidence for olfactory
illusions? Perception, 30(3), 381–391. https://doi.org/10.1068/p3179
Herzog, C., & Otto, T. (1997). Odor-guided fear conditioning in rats: 2. Lesions of the anterior perirhinal cortex disrupt
fear conditioned to the explicit conditioned stimulus but not to the training context. Behavioral Neuroscience, 111(6),
1265–1272. https://doi.org/10.1037/0735-7044.111.6.1265
Hitt, J. C., Bryon, D. M., & Modianos, D. T. (1973). Effects of rostral medial forebrain bundle and olfactory tubercle
lesions upon sexual behavior of male rats. Journal of Comparative and Physiological Psychology, 82(1), 30–36. https://doi.
org/10.1037/h0033797
Howard, J. D., Gottfried, J. A., Tobler, P. N., & Kahnt, T. (2015). Identity-specific coding of future rewards in the human
orbitofrontal cortex. Proceedings of the National Academy of Sciences of the United States of America, 112(16), 5195–5200.
https://doi.org/10.1073/pnas.1503550112
Igarashi, K. M., Ieki, N., An, M., Yamaguchi, Y., Nagayama, S., Kobayakawa, K., Kobayakawa, R., Tanifuji, M.,
Sakano, H., Chen, W. R., & Mori, K. (2012). Parallel mitral and tufted cell pathways route distinct odor informa-
tion to different targets in the olfactory cortex. Journal of Neuroscience, 32(23), 7970–7985. https://doi.org/10.1523/
JNEUROSCI.0154-12.2012
Illig, K. R., & Haberly, L. B. (2003). Odor-evoked activity is spatially distributed in piriform cortex. Journal of Compara-
tive Neurology, 457(4), 361–373. https://doi.org/10.1002/cne.10557
Insausti, R., Marcos, P., Arroyo-Jiménez, M. M., Blaizot, X., & Martínez-Marcos, A. (2002). Comparative aspects of
the olfactory portion of the entorhinal cortex and its projection to the hippocampus in rodents, nonhuman primates,
and the human brain. Brain Research Bulletin, 57(3–4), 557–560. https://doi.org/10.1016/s0361-9230(01)00684-0
Jacobs, L. F., Arter, J., Cook, A., & Sulloway, F. J. (2015). Olfactory orientation and navigation in humans. PLOS ONE,
10(6), e0129387. https://doi.org/10.1371/journal.pone.0129387
Jadauji, J. B., Djordjevic, J., Lundström, J. N., & Pack, C. C. (2012). Modulation of olfactory perception by visual cortex
stimulation. Journal of Neuroscience, 32(9), 3095–3100. https://doi.org/10.1523/JNEUROSCI.6022-11.2012

165
Gulce Nazli Dikecligil and Jay A. Gottfried

Jiang, T., Soussignan, R., Schaal, B., & Royet, J. P. (2015). Reward for food odors: An fMRI study of liking and want-
ing as a function of metabolic state and BMI. Social Cognitive and Affective Neuroscience, 10(4), 561–568. https://doi.
org/10.1093/scan/nsu086
Jin, J., Zelano, C., Gottfried, J. A., & Mohanty, A. (2015). Human amygdala represents the complete spectrum of sub-
jective valence. Journal of Neuroscience, 35(45), 15145–15156. https://doi.org/10.1523/JNEUROSCI.2450-15.2015
Jo, H. G., Wudarczyk, O., Leclerc, M., Regenbogen, C., Lampert, A., Rothermel, M., & Habel, U. (2020). Effect of
odor pleasantness on heat-induced pain: An fMRI study. Brain Imaging and Behavior, 15(3), 1300–1312. https://doi.
org/10.1007/s11682-020-00328-0
Keller, A., Gerkin, R. C., Guan, Y., Dhurandhar, A., Turu, G., Szalai, B., Mainland, J. D., Ihara, Y., Yu, C. W., Wolfin-
ger, R., Vens, C., Schietgat, L., De Grave, K., Norel, R., DREAM Olfaction Prediction Consortium, Stolovitzky,
G., Cecchi, G. A., Vosshall, L. B., & Meyer, P. (2017). Predicting human olfactory perception from chemical features
of odor molecules. Science, 355(6327), 820–826. https://doi.org/10.1126/science.aal2014
Keller, A., & Malaspina, D. (2013). Hidden consequences of olfactory dysfunction: A patient report series. BMC Ear,
Nose, and Throat Disorders, 13(1), 8. https://doi.org/10.1186/1472-6815-13-8
Keller, A., & Vosshall, L. B. (2008). Better smelling through genetics: Mammalian odor perception. Current Opinion in
Neurobiology, 18(4), 364–369. https://doi.org/10.1016/j.conb.2008.09.020
Keller, M., Baum, M. J., Brock, O., Brennan, P. A., & Bakker, J. (2009). The main and the accessory olfactory systems
interact in the control of mate recognition and sexual behavior. Behavioural Brain Research, 200(2), 268–276. https://
doi.org/10.1016/j.bbr.2009.01.020
Kelley, J. L., & Magurran, A. E. (2003). Learned predator recognition and antipredator responses in fishes. Fish and Fisher-
ies, 4(3), 216–226, https://doi.org/10.1046/j.1467-2979.2003.00126.x
Keshavarzi, S., Power, J. M., Albers, E. H., Sullivan, R. K.,  & Sah, P. (2015). Dendritic organization of olfactory
inputs to medial amygdala neurons. Journal of Neuroscience, 35(38), 13020–13028. https://doi.org/10.1523/JNEU
ROSCI.0627-15.2015
Keverne, E. B. (2004). Importance of olfactory and vomeronasal systems for male sexual function. Physiology and Behavior,
83(2), 177–187. https://doi.org/10.1016/j.physbeh.2004.08.013
Khan, R. M., Luk, C. H., Flinker, A., Aggarwal, A., Lapid, H., Haddad, R.,  & Sobel, N. (2007). Predicting odor
pleasantness from odorant structure: Pleasantness as a reflection of the physical world. Journal of Neuroscience, 27(37),
10015–10023. https://doi.org/10.1523/JNEUROSCI.1158-07.2007
Kobayakawa, K., Kobayakawa, R., Matsumoto, H., Oka, Y., Imai, T., Ikawa, M., Okabe, M., Ikeda, T., Itohara, S., Kiku-
sui, T., Mori, K., & Sakano, H. (2007). Innate versus learned odour processing in the mouse olfactory bulb. Nature,
450(7169), 503–508. https://doi.org/10.1038/nature06281
Koulakov, A. A., Kolterman, B. E., Enikolopov, A. G., & Rinberg, D. (2011). In search of the structure of human olfac-
tory space. Frontiers in Systems Neuroscience, 5, 65. https://doi.org/10.3389/fnsys.2011.00065
Kucharski, D., & Spear, N. E. (1984). Conditioning of aversion to an odor paired with peripheral shock in the developing
rat. Developmental Psychobiology, 17(5), 465–479. https://doi.org/10.1002/dev.420170505
Kullmann, H., Thünken, T., Baldauf, S. A., Bakker, T. C., & Frommen, J. G. (2008). Fish odour triggers conspecific
attraction behaviour in an aquatic invertebrate. Biology Letters, 4(5), 458–460. https://doi.org/10.1098/rsbl.2008.0246
Kurahashi, T., & Menini, A. (1997). Mechanism of odorant adaptation in the olfactory receptor cell. Nature, 385(6618),
725–729. https://doi.org/10.1038/385725a0
Laidre, M. E. (2009). Informative breath: Olfactory cues sought during social foraging among Old World monkeys
(Mandrillus sphinx, M. leucophaeus, and Papio anubis). Journal of Comparative Psychology, 123(1), 34–44. https://doi.
org/10.1037/a0013129
Lawless, H., & Engen, T. (1977). Associations to odors: Interference, mnemonics, and verbal labeling. Journal of Experi-
mental Psychology. Human Learning and Memory, 3(1), 52–59. https://doi.org/10.1037/0278-7393.3.1.52
Leleu, A., Godard, O., Dollion, N., Durand, K., Schaal, B.,  & Baudouin, J. Y. (2015). Contextual odors modulate
the visual processing of emotional facial expressions: An ERP study. Neuropsychologia, 77, 366–379. https://doi.
org/10.1016/j.neuropsychologia.2015.09.014
Leventhal, A. M., Martin, R. L., Seals, R. W., Tapia, E., & Rehm, L. P. (2007). Investigating the dynamics of affect:
Psychological mechanisms of affective habituation to pleasurable stimuli. Motivation and Emotion, 31(2), 145–157.
https://doi.org/10.1007/s11031-007-9059-8
Li, J. J., Wang, Z. W., Tan, K., Qu, Y. F., & Nieh, J. C. (2014). Giant Asian honeybees use olfactory eavesdropping to
detect and avoid ant predators. Animal Behaviour, 97, 69–76. https://doi.org/10.1016/j.anbehav.2014.08.015
Li, Q., Korzan, W. J., Ferrero, D. M., Chang, R. B., Roy, D. S., Buchi, M., Lemon, J. K., Kaur, A. W., Stowers, L., Fendt,
M., & Liberles, S. D. (2013). Synchronous evolution of an odor biosynthesis pathway and behavioral response. Current
Biology, 23(1), 11–20. https://doi.org/10.1016/j.cub.2012.10.047

166
Odour aesthetics

Li, Q., & Liberles, S. D. (2015). Aversion and attraction through olfaction. Current Biology, 25(3), R120–R129. https://
doi.org/10.1016/j.cub.2014.11.044
Li, W. (2014). Learning to smell danger: Acquired associative representation of threat in the olfactory cortex. Frontiers in
Behavioral Neuroscience, 8, 98. https://doi.org/10.3389/fnbeh.2014.00098
Li, W., Howard, J. D., Parrish, T. B., & Gottfried, J. A. (2008). Aversive learning enhances perceptual and cortical dis-
crimination of indiscriminable odor cues. Science, 319(5871), 1842–1845. https://doi.org/10.1126/science.1152837
Li, W., Lopez, L., Osher, J., Howard, J. D., Parrish, T. B., & Gottfried, J. A. (2010). Right orbitofrontal cortex medi-
ates conscious olfactory perception. Psychological Science, 21(10), 1454–1463. https://doi.org/10.1177/095679761
0382121
Li, W., Luxenberg, E., Parrish, T., & Gottfried, J. A. (2006). Learning to smell the roses: Experience-dependent neu-
ral plasticity in human piriform and orbitofrontal cortices. Neuron, 52(6), 1097–1108. https://doi.org/10.1016/j.
neuron.2006.10.026
Linster, C., & Cleland, T. A. (2016). Neuromodulation of olfactory transformations. Current Opinion in Neurobiology, 40,
170–177. https://doi.org/10.1016/j.conb.2016.07.006
Liu, S., Aungst, J. L., Puche, A. C., & Shipley, M. T. (2012). Serotonin modulates the population activity profile of olfac-
tory bulb external tufted cells. Journal of Neurophysiology, 107(1), 473–483. https://doi.org/10.1152/jn.00741.2011
Logan, D. W., Brunet, L. J., Webb, W. R., Cutforth, T., Ngai, J., & Stowers, L. (2012). Learned recognition of mater-
nal signature odors mediates the first suckling episode in mice. Current Biology, 22(21), 1998–2007. https://doi.
org/10.1016/j.cub.2012.08.041
Luisier, A. C., Petitpierre, G., Clerc Bérod, A., Garcia-Burgos, D., & Bensafi, M. (2019). Effects of familiarization on
odor hedonic responses and food choices in children with autism spectrum disorders. Autism, 23(6), 1460–1471.
https://doi.org/10.1177/1362361318815252
Maier, J. X., Wachowiak, M., & Katz, D. B. (2012). Chemosensory convergence on primary olfactory cortex. Journal of
Neuroscience, 32(48), 17037–17047. https://doi.org/10.1523/JNEUROSCI.3540-12.2012
Malnic, B., Godfrey, P. A., & Buck, L. B. (2004). The human olfactory receptor gene family. Proceedings of the National
Academy of Sciences of the United States of America, 101(8), 2584–2589. https://doi.org/10.1073/pnas.0307882100
Malnic, B., Hirono, J., Sato, T., & Buck, L. B. (1999). Combinatorial receptor codes for odors. Cell, 96(5), 713–723.
https://doi.org/10.1016/s0092-8674(00)80581-4
Mandairon, N., Ferretti, C. J., Stack, C. M., Rubin, D. B., Cleland, T. A., & Linster, C. (2006). Cholinergic modulation
in the olfactory bulb influences spontaneous olfactory discrimination in adult rats. European Journal of Neuroscience,
24(11), 3234–3244. https://doi.org/10.1111/j.1460-9568.2006.05212.x
Manescu, S., Frasnelli, J., Lepore, F., & Djordjevic, J. (2014). Now you like me, now you don’t: Impact of labels on odor
perception. Chemical Senses, 39(2), 167–175. https://doi.org/10.1093/chemse/bjt066
Manesse, C., Fournel, A., Bensafi, M., & Ferdenzi, C. (2020). Visual priming influences olfactomotor response and per-
ceptual experience of smells. Chemical Senses, 45(3), 211–218. https://doi.org/10.1093/chemse/bjaa008
Markopoulos, F., Rokni, D., Gire, D. H., & Murthy, V. N. (2012). Functional properties of cortical feedback projections
to the olfactory bulb. Neuron, 76(6), 1175–1188. https://doi.org/10.1016/j.neuron.2012.10.028
Matsutani, S. (2010). Trajectory and terminal distribution of single centrifugal axons from olfactory cortical areas in the
rat olfactory bulb. Neuroscience, 169(1), 436–448. https://doi.org/10.1016/j.neuroscience.2010.05.001
Matsutani, S.,  & Yamamoto, N. (2008). Centrifugal innervation of the mammalian olfactory bulb. Anatomical Science
International, 83(4), 218–227. https://doi.org/10.1111/j.1447-073X.2007.00223.x
McGann, J. P. (2017). Poor human olfaction is a 19th-century myth. Science, 356(6338). ARTN eaam7263. https://doi.
org/10.1126/science.aam7263
McLean, J. H., & Shipley, M. T. (1987). Serotonergic afferents to the rat olfactory bulb: I. Origins and laminar speci-
ficity of serotonergic inputs in the adult rat. Journal of Neuroscience, 7(10), 3016–3028. https://doi.org/10.1523/
JNEUROSCI.07-10-03016.1987
McLean, J. H., Shipley, M. T., Nickell, W. T., Aston-Jones, G., & Reyher, C. K. (1989). Chemoanatomical organization
of the noradrenergic input from locus coeruleus to the olfactory bulb of the adult rat. Journal of Comparative Neurology,
285(3), 339–349. https://doi.org/10.1002/cne.902850305
Mclennan, D. A. (2003). The importance of olfactory signals in the gasterosteid mating system: Sticklebacks go multi-
modal. Biological Journal of the Linnean Society, 80(4), 555–572, https://doi.org/10.1111/j.1095-8312.2003.00254.x
Millman, D. J., & Murthy, V. N. (2020). Rapid learning of odor-value association in the olfactory striatum. Journal of
Neuroscience, 40(22), 4335–4347. https://doi.org/10.1523/JNEUROSCI.2604-19.2020
Miwa, T., Furukawa, M., Tsukatani, T., Costanzo, R. M., DiNardo, L. J., & Reiter, E. R. (2001). Impact of olfac-
tory impairment on quality of life and disability. Archives of Otolaryngology–Head and Neck Surgery, 127(5), 497–503.
https://doi.org/10.1001/archotol.127.5.497

167
Gulce Nazli Dikecligil and Jay A. Gottfried

Miyamichi, K., Amat, F., Moussavi, F., Wang, C., Wickersham, I., Wall, N. R., Taniguchi, H., Tasic, B., Huang, Z. J.,
He, Z., Callaway, E. M., Horowitz, M. A., & Luo, L. (2011). Cortical representations of olfactory input by trans-
synaptic tracing. Nature, 472(7342), 191–196. https://doi.org/10.1038/nature09714
Morrot, G., Brochet, F., & Dubourdieu, D. (2001). The color of odors. Brain and Language, 79(2), 309–320. https://doi.
org/10.1006/brln.2001.2493
Moskowitz, H. R., Dravnieks, A., & Gerbers, C. (1974). Odor intensity and pleasantness of butanol. Journal of Experi-
mental Psychology, 103(2), 216–223, https://doi.org/10.1037/h0036793
Moskowitz, H. R., Dravnieks, A., & Klarman, L. A. (1976). Odor intensity and pleasantness for a diverse set of odorants.
Perception and Psychophysics, 19(2), 122–128. https://doi.org/10.3758/BF03204218
Murata, K., Kanno, M., Ieki, N., Mori, K., & Yamaguchi, M. (2015). Mapping of learned odor-induced motivated
behaviors in the mouse olfactory tubercle. Journal of Neuroscience, 35(29), 10581–10599. https://doi.org/10.1523/
JNEUROSCI.0073-15.2015
Najac, M., Sanz Diez, A., Kumar, A., Benito, N., Charpak, S., & De Saint Jan, D. (2015). Intraglomerular lateral inhi-
bition promotes spike timing variability in principal neurons of the olfactory bulb. Journal of Neuroscience, 35(10),
4319–4331. https://doi.org/10.1523/JNEUROSCI.2181-14.2015
Neuland, C., Bitter, T., Marschner, H., Gudziol, H.,  & Guntinas-Lichius, O. (2011). Health-related and specific
olfaction-related quality of life in patients with chronic functional anosmia or severe hyposmia. Laryngoscope, 121(4),
867–872. https://doi.org/10.1002/lary.21387
Niimura, Y., Matsui, A., & Touhara, K. (2014). Extreme expansion of the olfactory receptor gene repertoire in African
elephants and evolutionary dynamics of orthologous gene groups in 13 placental mammals. Genome Research, 24(9),
1485–1496. https://doi.org/10.1101/gr.169532.113
Nordin, S., Blomqvist, E. H., Olsson, P., Stjärne, P., Ehnhage, A., & Group, N. S. S. (2011). Effects of smell loss on
daily life and adopted coping strategies in patients with nasal polyposis with asthma. Acta Oto-laryngologica, 131(8),
826–832. https://doi.org/10.3109/00016489.2010.539625
O'Doherty, J., Rolls, E. T., Francis, S., Bowtell, R., McGlone, F., Kobal, G., Renner, B., & Ahne, G., . . . Ahne. (2000).
Sensory-specific satiety-related olfactory activation of the human orbitofrontal cortex. NeuroReport, 11(4), 893–897.
https://doi.org/10.1097/00001756-200003200-00046
Oka, Y., Omura, M., Kataoka, H., & Touhara, K. (2004). Olfactory receptor antagonism between odorants. EMBO
Journal, 23(1), 120–126. https://doi.org/10.1038/sj.emboj.7600032
Olofsson, J. K., Bowman, N. E., & Gottfried, J. A. (2013). High and low roads to odor valence? A choice response-time
study. Journal of Experimental Psychology. Human Perception and Performance, 39(5), 1205–1211. https://doi.org/10.1037/
a0033682
Olofsson, J. K., Bowman, N. E., Khatibi, K., & Gottfried, J. A. (2012). A time-based account of the perception of odor
objects and valences. Psychological Science, 23(10), 1224–1232. https://doi.org/10.1177/0956797612441951
Olofsson, J. K., & Gottfried, J. A. (2015). The muted sense: Neurocognitive limitations of olfactory language. Trends in
Cognitive Sciences, 19(6), 314–321. https://doi.org/10.1016/j.tics.2015.04.007
Osterbauer, R. A., Matthews, P. M., Jenkinson, M., Beckmann, C. F., Hansen, P. C., & Calvert, G. A. (2005). Color of
scents: Chromatic stimuli modulate odor responses in the human brain. Journal of Neurophysiology, 93(6), 3434–3441.
https://doi.org/10.1152/jn.00555.2004
Papes, F., Logan, D. W., & Stowers, L. (2010). The vomeronasal organ mediates interspecies defensive behaviors through
detection of protein pheromone homologs. Cell, 141(4), 692–703. https://doi.org/10.1016/j.cell.2010.03.037
Patin, A.,  & Pause, B. M. (2015). Human amygdala activations during nasal chemoreception. Neuropsychologia, 78,
171–194. https://doi.org/10.1016/j.neuropsychologia.2015.10.009
Petzold, G. C., Hagiwara, A., & Murthy, V. N. (2009). Serotonergic modulation of odor input to the mammalian olfac-
tory bulb. Nature Neuroscience, 12(6), 784–791. https://doi.org/10.1038/nn.2335
Phillips, R. G., & LeDoux, J. E. (1992). Differential contribution of amygdala and hippocampus to cued and contextual
fear conditioning. Behavioral Neuroscience, 106(2), 274–285. https://doi.org/10.1037/0735-7044.106.2.274
Poncelet, J., Rinck, F., Bourgeat, F., Schaal, B., Rouby, C., Bensafi, M., & Hummel, T. (2010). The effect of early expe-
rience on odor perception in humans: Psychological and physiological correlates. Behavioural Brain Research, 208(2),
458–465. https://doi.org/10.1016/j.bbr.2009.12.011
Poncelet, J., Rinck, F., Ziessel, A., Joussain, P., Thévenet, M., Rouby, C.,  & Bensafi, M. (2010). Semantic knowl-
edge influences prewired hedonic responses to odors. PLOS ONE, 5(11), e13878. https://doi.org/10.1371/journal.
pone.0013878
Poncet, F., Leleu, A., Rekow, D., Damon, F., Durand, K., Schaal, B., & Baudouin, J. Y. (2020). Odor-evoked hedonic
contexts influence the discrimination of facial expressions in the human brain. Biological Psychology, 158, 108005.
https://doi.org/10.1016/j.biopsycho.2020.108005

168
Odour aesthetics

Pool, E., Sennwald, V., Delplanque, S., Brosch, T.,  & Sander, D. (2016). Measuring wanting and liking from ani-
mals to humans: A systematic review. Neuroscience and Biobehavioral Reviews, 63, 124–142. https://doi.org/10.1016/j.
neubiorev.2016.01.006
Prescott, J., & Wilkie, J. (2007). Pain tolerance selectively increased by a sweet-smelling odor. Psychological Science, 18(4),
308–311. https://doi.org/10.1111/j.1467-9280.2007.01894.x
Qureshy, A., Kawashima, R., Imran, M. B., Sugiura, M., Goto, R., Okada, K., Inoue, K., Itoh, M., Schormann, T.,
Zilles, K., & Fukuda, H. (2000). Functional mapping of human brain in olfactory processing: A PET study. Journal of
Neurophysiology, 84(3), 1656–1666. https://doi.org/10.1152/jn.2000.84.3.1656
Raghuram, H., Thangadurai, C., Gopukumar, N., Nathar, K., & Sripathi, K. (2009). The role of olfaction and vision in
the foraging behaviour of an echolocating megachiropteran fruit bat, Rousettus leschenaulti (Pteropodidae). Mammalian
Biology, 74(1), 9–14. https://doi.org/10.1016/j.mambio.2008.02.008
Raineki, C., Holman, P. J., Debiec, J., Bugg, M., Beasley, A.,  & Sullivan, R. M. (2010). Functional emergence of
the hippocampus in context fear learning in infant rats. Hippocampus, 20(9), 1037–1046. https://doi.org/10.1002/
hipo.20702
Raithel, C. U.,  & Gottfried, J. A. (2021). Using your nose to find your way: Ethological comparisons between
human and non-human species. Neuroscience and Biobehavioral Reviews, 128, 766–779. https://doi.org/10.1016/j.
neubiorev.2021.06.040
Ravia, A., Snitz, K., Honigstein, D., Finkel, M., Zirler, R., Perl, O., Secundo, L., Laudamiel, C., Harel, D., & Sobel,
N. (2020). A measure of smell enables the creation of olfactory metamers. Nature, 588(7836), 118–123. https://doi.
org/10.1038/s41586-020-2891-7
Reddy, G., Zak, J. D., Vergassola, M., & Murthy, V. N. (2018). Antagonism in olfactory receptor neurons and its implica-
tions for the perception of odor mixtures. eLife, 7. https://doi.org/10.7554/eLife.34958
Rennaker, R. L., Chen, C. F., Ruyle, A. M., Sloan, A. M., & Wilson, D. A. (2007). Spatial and temporal distribution of
odorant-evoked activity in the piriform cortex. Journal of Neuroscience, 27(7), 1534–1542. https://doi.org/10.1523/
JNEUROSCI.4072-06.2007
Rinberg, D., Koulakov, A., & Gelperin, A. (2006). Speed-accuracy tradeoff in olfaction. Neuron, 51(3), 351–358. https://
doi.org/10.1016/j.neuron.2006.07.013
Roland, B., Deneux, T., Franks, K. M., Bathellier, B., & Fleischmann, A. (2017). Odor identity coding by distributed
ensembles of neurons in the mouse olfactory cortex. eLife, 6. https://doi.org/10.7554/eLife.26337
Rolls, E. T., Critchley, H. D., & Treves, A. (1996). Representation of olfactory information in the primate orbitofrontal
cortex. Journal of Neurophysiology, 75(5), 1982–1996. https://doi.org/10.1152/jn.1996.75.5.1982
Rolls, E. T., Critchley, H. D., Mason, R., & Wakeman, E. A. (1996). Orbitofrontal cortex neurons: Role in olfactory and
visual association learning. Journal of Neurophysiology, 75(5), 1970–1981. https://doi.org/10.1152/jn.1996.75.5.1970
Rosenkranz, J. A., & Grace, A. A. (2002). Dopamine-mediated modulation of odour-evoked amygdala potentials during
Pavlovian conditioning. Nature, 417(6886), 282–287. https://doi.org/10.1038/417282a
Roth, T. C., Cox, J. G., & Lima, S. L. (2008). Can foraging birds assess predation risk by scent? Animal Behaviour, 76(6),
2021–2027. https://doi.org/10.1016/j.anbehav.2008.08.022
Rothermel, M., & Wachowiak, M. (2014). Functional imaging of cortical feedback projections to the olfactory bulb.
Frontiers in Neural Circuits, 8, 73. https://doi.org/10.3389/fncir.2014.00073
Ryan, B. C., Young, N. B., Moy, S. S., & Crawley, J. N. (2008). Olfactory cues are sufficient to elicit social approach
behaviors but not social transmission of food preference in C57BL/6J mice. Behavioural Brain Research, 193(2), 235–
242. https://doi.org/10.1016/j.bbr.2008.06.002
Saito, H., Nishizumi, H., Suzuki, S., Matsumoto, H., Ieki, N., Abe, T., Kiyonari, H., Morita, M., Yokota, H., Hirayama,
N., Yamazaki, T., Kikusui, T., Mori, K.,  & Sakano, H. (2017). Immobility responses are induced by photoacti-
vation of single glomerular species responsive to fox odour TMT. Nature Communications, 8, 16011. https://doi.
org/10.1038/ncomms16011
Salazar, I., & Quinteiro, P. S. (2009). The risk of extrapolation in neuroanatomy: The case of the mammalian vomeronasal
system. Frontiers in Neuroanatomy, 3, 22. https://doi.org/10.3389/neuro.05.022.2009
Samuelsen, C. L., & Fontanini, A. (2017). Processing of intraoral olfactory and gustatory signals in the gustatory cortex
of awake rats. Journal of Neuroscience, 37(2), 244–257. https://doi.org/10.1523/JNEUROSCI.1926-16.2016
Schiffman, S., Robinson, D. E.,  & Erickson, R. P. (1977). Multidimensional-scaling of odorants—Examination of
psychological and physicochemical dimensions. Chemical Senses, 2(3), 375–390, https://doi.org/10.1093/chemse/
2.3.375
Schoenbaum, G., Chiba, A. A., & Gallagher, M. (1999). Neural encoding in orbitofrontal cortex and basolateral amyg-
dala during olfactory discrimination learning. Journal of Neuroscience, 19(5), 1876–1884. https://doi.org/10.1523/
JNEUROSCI.19-05-01876.1999

169
Gulce Nazli Dikecligil and Jay A. Gottfried

Schwerdtfeger, W. K., Buhl, E. H., & Germroth, P. (1990). Disynaptic olfactory input to the hippocampus mediated
by stellate cells in the entorhinal cortex. Journal of Comparative Neurology, 292(2), 163–177. https://doi.org/10.1002/
cne.902920202
Scott, J. W. (1977). A measure of extracellular unit responses to repeated stimulation applied to observations of the time
course of olfactory responses. Brain Research, 132(2), 247–258. https://doi.org/10.1016/0006-8993(77)90419-x
Sell, C. S. (2006). On the unpredictability of odor. Angewandte Chemie International Edition, 45(38), 6254–6261. https://
doi.org/10.1002/anie.200600782
Semmelhack, J. L., & Wang, J. W. (2009). Select Drosophila glomeruli mediate innate olfactory attraction and aversion.
Nature, 459(7244), 218–223. https://doi.org/10.1038/nature07983
Seo, H. S., & Hummel, T. (2011). Auditory-olfactory integration: Congruent or pleasant sounds amplify odor pleasant-
ness. Chemical Senses, 36(3), 301–309. https://doi.org/10.1093/chemse/bjq129
Seo, H. S., Lohse, F., Luckett, C. R., & Hummel, T. (2014). Congruent sound can modulate odor pleasantness. Chemical
Senses, 39(3), 215–228. https://doi.org/10.1093/chemse/bjt070
Seo, H. S., Roidl, E., Müller, F., & Negoias, S. (2010). Odors enhance visual attention to congruent objects. Appetite,
54(3), 544–549. https://doi.org/10.1016/j.appet.2010.02.011
Serizawa, S., Miyamichi, K., & Sakano, H. (2004). One neuron-one receptor rule in the mouse olfactory system. Trends
in Genetics, 20(12), 648–653. https://doi.org/10.1016/j.tig.2004.09.006
Serizawa, S., Miyamichi, K., & Sakano, H. (2005). Negative feedback regulation ensures the one neuron-one receptor
rule in the mouse olfactory system. Chemical Senses, 30, Suppl. 1, i99–i100. https://doi.org/10.1093/chemse/bjh133
Sevelinges, Y., Desgranges, B., & Ferreira, G. (2009). The basolateral amygdala is necessary for the encoding and the
expression of odor memory. Learning and Memory, 16(4), 235–242. https://doi.org/10.1101/lm.1247609
Sevelinges, Y., Gervais, R., Messaoudi, B., Granjon, L., & Mouly, A. M. (2004). Olfactory fear conditioning induces
field potential potentiation in rat olfactory cortex and amygdala. Learning and Memory, 11(6), 761–769. https://doi.
org/10.1101/lm.83604
Shao, Z., Puche, A. C., Liu, S., & Shipley, M. T. (2012). Intraglomerular inhibition shapes the strength and temporal
structure of glomerular output. Journal of Neurophysiology, 108(3), 782–793. https://doi.org/10.1152/jn.00119.2012
Shepherd, G. M. (2004). The human sense of smell: Are we better than we think? PLOS Biology, 2(5), 572–575. https://
doi.org/ARTN e14610.1371/journal.pbio.0020146.
Small, D. M., Jones-Gotman, M., Zatorre, R. J., Petrides, M., & Evans, A. C. (1997). Flavor processing: More than the
sum of its parts. NeuroReport, 8(18), 3913–3917. https://doi.org/10.1097/00001756-199712220-00014
Small, D. M., Zatorre, R. J., Dagher, A., Evans, A. C., & Jones-Gotman, M. (2001). Changes in brain activity related to
eating chocolate: From pleasure to aversion. Brain, 124(9), 1720–1733. https://doi.org/10.1093/brain/124.9.1720
Smeets, M. A. M., Veldhuizen, M. G., Galle, S., Gouweloos, J., de Haan, A. J. A., Vernooij, J., Visscher, F., & Kroeze,
J. H. A. (2009). Sense of smell disorder and health-related quality of life. Rehabilitation Psychology, 54(4), 404–412.
https://doi.org/10.1037/a0017502
Snitz, K., Yablonka, A., Weiss, T., Frumin, I., Khan, R. M., & Sobel, N. (2013). Predicting odor perceptual similarity
from odor structure. PLOS Computational Biology, 9(9), e1003184. https://doi.org/10.1371/journal.pcbi.1003184
Sobel, N., Prabhakaran, V., Desmond, J. E., Glover, G. H., Goode, R. L., Sullivan, E. V.,  & Gabrieli, J. D. (1998).
Sniffing and smelling: Separate subsystems in the human olfactory cortex. Nature, 392(6673), 282–286. https://doi.
org/10.1038/32654
Sobel, N., Prabhakaran, V., Zhao, Z., Desmond, J. E., Glover, G. H., Sullivan, E. V., & Gabrieli, J. D. (2000). Time
course of odorant-induced activation in the human primary olfactory cortex. Journal of Neurophysiology, 83(1), 537–
551. https://doi.org/10.1152/jn.2000.83.1.537
Spence, C., Levitan, C. A., Shankar, M. U., & Zampini, M. (2010). Does food color influence taste and flavor perception
in humans? Chemosensory Perception, 3(1), 68–84. https://doi.org/10.1007/s12078-010-9067-z
Stettler, D. D., & Axel, R. (2009). Representations of odor in the piriform cortex. Neuron, 63(6), 854–864. https://doi.
org/10.1016/j.neuron.2009.09.005
Stevenson, R. J. (2010). An initial evaluation of the functions of human olfaction. Chemical Senses, 35(1), 3–20. https://
doi.org/10.1093/chemse/bjp083
Stowers, L., & Logan, D. W. (2010). Olfactory mechanisms of stereotyped behavior: On the scent of specialized circuits.
Current Opinion in Neurobiology, 20(3), 274–280. https://doi.org/10.1016/j.conb.2010.02.013
Stutz, R. S., Banks, P. B., Proschogo, N., & McArthur, C. (2016). Follow your nose: Leaf odour as an important foraging
cue for mammalian herbivores. Oecologia, 182(3), 643–651. https://doi.org/10.1007/s00442-016-3678-2
Suh, G. S., Wong, A. M., Hergarden, A. C., Wang, J. W., Simon, A. F., Benzer, S., Axel, R., & Anderson, D. J. (2004).
A  single population of olfactory sensory neurons mediates an innate avoidance behaviour in Drosophila. Nature,
431(7010), 854–859. https://doi.org/10.1038/nature02980

170
Odour aesthetics

Sündermann, D., Scheumann, M.,  & Zimmermann, E. (2008). Olfactory predator recognition in predator-
naive gray mouse lemurs (Microcebus murinus). Journal of Comparative Psychology, 122(2), 146–155. https://doi.
org/10.1037/0735-7036.122.2.146
Suyama, H., Egger, V., & Lukas, M. (2021). Top-down acetylcholine signaling via olfactory bulb vasopressin cells contrib-
utes to social discrimination in rats. Communications Biology, 4(1), 603. https://doi.org/10.1038/s42003-021-02129-7
Temmel, A. F., Quint, C., Schickinger-Fischer, B., Klimek, L., Stoller, E., & Hummel, T. (2002). Characteristics of
olfactory disorders in relation to major causes of olfactory loss. Archives of Otolaryngology–Head and Neck Surgery,
128(6), 635–641. https://doi.org/10.1001/archotol.128.6.635
Tholl, D., Sohrabi, R., Huh, J. H., & Lee, S. (2011). The biochemistry of homoterpenes—Common constituents of
floral and herbivore-induced plant volatile bouquets. Phytochemistry, 72(13), 1635–1646. https://doi.org/10.1016/j.
phytochem.2011.01.019
Uchida, N., Poo, C., & Haddad, R. (2014). Coding and transformations in the olfactory system. Annual Review of Neu-
roscience, 37, 363–385. https://doi.org/10.1146/annurev-neuro-071013-013941
Ulrich, D., Kecke, S., & Olbricht, K. (2018). What do we know about the chemistry of strawberry aroma? Journal of
Agricultural and Food Chemistry, 66(13), 3291–3301. https://doi.org/10.1021/acs.jafc.8b01115
Urban, N. N. (2002). Lateral inhibition in the olfactory bulb and in olfaction. Physiology and Behavior, 77(4–5), 607–612.
https://doi.org/10.1016/s0031-9384(02)00895-8
Uva, L., & de Curtis, M. (2005). Polysynaptic olfactory pathway to the ipsi- and contralateral entorhinal cortex mediated
via the hippocampus. Neuroscience, 130(1), 249–258. https://doi.org/10.1016/j.neuroscience.2004.08.042
Vincis, R., & Fontanini, A. (2016). Associative learning changes cross-modal representations in the gustatory cortex.
eLife, 5. https://doi.org/10.7554/eLife.16420
Vincis, R., Gschwend, O., Bhaukaurally, K., Beroud, J., & Carleton, A. (2012). Dense representation of natural odorants
in the mouse olfactory bulb. Nature Neuroscience, 15(4), 537–539. https://doi.org/10.1038/nn.3057
Yeshurun, Y., Lapid, H., Dudai, Y., & Sobel, N. (2009). The privileged brain representation of first olfactory associations.
Current Biology, 19(21), 1869–1874. https://doi.org/10.1016/j.cub.2009.09.066
Yeshurun, Y., & Sobel, N. (2010). An odor is not worth a thousand words: From multidimensional odors to unidi-
mensional odor objects. Annual Review of Psychology, 61, 219–241, C211–215. https://doi.org/10.1146/annurev.
psych.60.110707.163639
Zarzo, M. (2008). Psychologic dimensions in the perception of everyday odors: Pleasantness and edibility. Journal of Sen-
sory Studies, 23(3), 354–376, https://doi.org/10.1111/j.1745-459X.2008.00160.x
Zatorre, R. J., Jones-Gotman, M., Evans, A. C., & Meyer, E. (1992). Functional localization and lateralization of human
olfactory cortex. Nature, 360(6402), 339–340. https://doi.org/10.1038/360339a0
Zatorre, R. J., Jones-Gotman, M., & Rouby, C. (2000). Neural mechanisms involved in odor pleasantness and intensity
judgments. NeuroReport, 11(12), 2711–2716. https://doi.org/10.1097/00001756-200008210-00021
Zavitz, D., Youngstrom, I. A., Borisyuk, A.,  & Wachowiak, M. (2020). Effect of interglomerular inhibitory net-
works on olfactory bulb odor representations. Journal of Neuroscience, 40(31), 5954–5969. https://doi.org/10.1523/
JNEUROSCI.0233-20.2020
Zellner, D. A., Bartoli, A. M., & Eckard, R. (1991). Influence of color on odor identification and liking ratings. American
Journal of Psychology, 104(4), 547–561. https://doi.org/10.2307/1422940
Zhang, Y., Lu, H., & Bargmann, C. I. (2005). Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis
elegans. Nature, 438(7065), 179–184. https://doi.org/10.1038/nature04216
Zhang, Z., Liu, Q., Wen, P., Zhang, J., Rao, X., Zhou, Z., Zhang, H., He, X., Li, J., Zhou, Z., Xu, X., Zhang, X.,
Luo, R., Lv, G., Li, H., Cao, P., Wang, L., & Xu, F. (2017). Activation of the dopaminergic pathway from VTA to
the medial olfactory tubercle generates odor-preference and reward. eLife, 6. https://doi.org/10.7554/eLife.25423
Zhang, Z., Zhang, H., Wen, P., Zhu, X., Wang, L., Liu, Q., Wang, J., He, X., Wang, H., & Xu, F. (2017). Whole-brain
mapping of the inputs and outputs of the medial part of the olfactory tubercle. Frontiers in Neural Circuits, 11, 52.
https://doi.org/10.3389/fncir.2017.00052
Zhou, W., Jiang, Y., He, S., & Chen, D. (2010). Olfaction modulates visual perception in binocular rivalry. Current Biol-
ogy, 20(15), 1356–1358. https://doi.org/10.1016/j.cub.2010.05.059
Zou, L. Q., van Hartevelt, T. J., Kringelbach, M. L., Cheung, E. F. C., & Chan, R. C. K. (2016). The neural mechanism
of hedonic processing and judgment of pleasant odors: An activation likelihood estimation meta-analysis. Neuropsy-
chology, 30(8), 970–979. https://doi.org/10.1037/neu0000292
Zufall, F.,  & Leinders-Zufall, T. (2000). The cellular and molecular basis of odor adaptation. Chemical Senses, 25(4),
473–481. https://doi.org/10.1093/chemse/25.4.473

171
9
MOVEMENT APPRECIATION
Kohinoor M. Darda, Ionela Bara and Emily S. Cross

The speed and agility of a white-tailed deer; the dramatic falling waters of the Niagara; the powerful take-
off of the Concorde’s supersonic airliner; Anna Pavlova’s gentle yet fluid pirouette in The Dying Swan; and
the dynamicity of the Laocoön, the group statue at the Vatican Museum—movement, in all its splendour, has
captured people’s, and artists’, imagination for millennia. Many artists have sought new ways to give visual
expression to movement, while much more recently, scientists have begun to investigate how we perceive
and why we appreciate real and apparent motion.
Artists have discovered how to exploit the visual form in order to capture motion’s vitality in a wide range
of styles, including conceptual, abstract, and realistic artistic representations. The successful depiction of
motion in visual art represents a milestone in human creation and has shaped the way we understand, appre-
ciate, and disseminate artworks today. Yet the first instances of studying aesthetic appreciation of movement
from a psychological scientific perspective only date back to the early 20th century. With the advent of more
refined neuroimaging technologies over the past several decades, the field of neuroaesthetics has contributed
to our understanding of how the human brain perceives and appreciates movement in art.
In this chapter, we discuss movement appreciation from a neuroaesthetic perspective, with a special focus on
movement representations in visual art. We begin with a short history of movement representation in art images.
We continue to outline the antecedents and consequents of movement appreciation before moving on to discuss the
neurobiological processes, functions and mechanisms that underlie movement appreciation in different contexts.

History
In the field of neuroaesthetics, the study of movement appreciation has been primarily investigated using
dance as stimuli and/or dancers as participants, as dance is an art form whose very essence is the human
body in motion (Cross & Ticini, 2012). A number of studies have investigated cognitive and neural processes
engaged when observing and appreciating movement as a starting point to understand the aesthetic experi-
ence triggered by watching dance (Christensen & Calvo-Merino, 2013). The depiction of movement in
paintings, sculpture, and photography, however, brings its own set of challenges—how does one represent
movement and dynamism in a static medium? Photographers have long used suspended movement to give
the impression of motion to their viewers—the ability of the camera to freeze a literal split second and catch
details imperceptible to the human eye gives the viewer a strong sense of the (implied) motion that would
happen if the moment in the photograph were “un-frozen” (Scharf, 1962; see Figure 9.1B). In addition,

172 DOI: 10.4324/9781003008675-10


Movement appreciation

Figure 9.1 The representation of motion in photography, paintings, and sculpture: (A) Use of motion blur in photog-
raphy, (B) use of suspended movement in photography, (C) Edgar Degas’ Ballet Scene (1879) depicting bal-
lerinas in motion, (D) Wassily Kadinsky’s Yellow-Red-Blue (1925) depicting the use of action lines to imply
motion, (E) Bhimbetka Rock painting (c. 12,000–8000 BCE) showing a humanoid being attacked by a wild
boar (photo taken by Rhodia Colomes, published with permission from photographer), (F) an example of
extreme contrapposto—Myron’s Discobolus by Giovanni Battista Piranesi (18th century), and (G) an example
of Tribhanga (triple bend): statue of an Indian deity from the Halebidu temple in India (circa 12th century;
photo taken by Dr. Saurabh Kadekodi, published with permission from photographer).

173
Darda, Bara and Cross

motion blur, action sequencing, and visual flow, when incorporated deliberately in a photograph, can further
contribute to striking representations of visual dynamism (see Figure 9.1A).
Long before the invention of photography, however, artists were confronted with the challenge of rep-
resenting movement in static sculptures and paintings. That dynamism or motion in artwork is important
and appreciated is well evidenced—art theorists like Arnheim (1951) supported the idea that motion cues
in visual art might be salient and more engaging and therefore inextricably linked to aesthetic appreciation.
Artist and inventor Leonardo Da Vinci called rigid, static sculptures and paintings “doubly dead”—dead
once because artworks are after all an impression (of something or someone) and “doubly dead” because
they do not imply movement of the mind or of the body ( Justi, 1923). In addition, Gombrich (1964) and
Wölfflin  (1942/2012) have pointed out that  the use of motion cues in visual art shaped the standards in
art production and perception and contributed to a better recognition and classification of art styles. Both
Gombrich and Wölfflin discussed how Italian Renaissance paintings used closed lines to accentuate the
repose and stability of forms, while Baroque paintings emphasized dynamic effects by using strong diagonal
lines and open dynamic forms.
Some of the earliest examples of the depiction of movement or actions in static media can be found in
the Bhimbetka paintings of Madhya Pradesh, India, likely dating back to the Mesolithic period (12,000–
8000 BCE; see Figure 9.1E). Over time, as an alternative style to depicting rigid forms in sculptures and
paintings, Indian artists introduced the concept of tribhanga (“triple bend,” circa 2300 BCE)—a position
where the body bends in one direction at the knees and in the other direction at the hips and then at
the shoulders and the neck, an early example of how body asymmetry can be used to depict move-
ment (Chakraborty, 1986; see Figure 9.1G). The Greek equivalent is contrapposto (or counterpoise, circa
500 BCE), the positioning of the human figure (in a sculpture or a painting) with the shoulder and arms
turned in a different direction to the hips and legs—a position that suggests action and reaction in various
parts of the figure, lending it naturalness and dynamicity (Summers, 1972). An example of extreme con-
trapposto is Myron’s Discobolus, a sculpture depicting a Greek discus thrower whose body looks like a tense
spring set to uncoil in a furious burst of purposeful energy (Unger, 2014; see Figure 9.1F). In all these
artworks, the dynamism perceived, although not real, is a stylistic impression of motion. It is conveyed to
the viewer through several associations, regularities, and structural characteristics that the human brain is
able to rapidly and easily comprehend.

Form features of implied motion in art


Art theorists like Arnheim (1974) have suggested that perceived motion is derived not only from the
inherent characteristics of the subject that is depicted (for instance, a running human or a waterfall) but
also from the compositional features of the artwork as a result of the use of lines, shapes, and particular
directions. For example, in Edgar Degas’ Balletprobe (1873; see Figure 9.1C), we perceive the dancers to
be in a state of motion because we understand they cannot hold a position with one leg in the air for too
long. In contrast, Kandinsky’s Yellow-Red-Blue (1925; see Figure 9.1D) features strong directional, diago-
nal, and gestural lines; repetition; and object placement to imply dynamism/motion in the artwork. Apart
from action lines, artists use a variety of form features or cues to enhance or even create implied motion
in an artwork and to convey information about direction and speed (Pavan et al., 2011; Krekelberg et al.,
2005). These form cues include (but are not limited to) broken symmetry (i.e., an asymmetrical relationship
between different parts of the body), blur (i.e., a blurring of the subject or of the background), stroboscopic
effects (i.e., superimposing different moments of a moving object in the same painting), forward lean, and/
or dynamic balance (Cutting, 2002; see Figure 9.2 for an illustration of typical implied motion cues used in
visual art).

174
Movement appreciation

Figure 9.2 Typical motion cues used in visual art. (A) Blur—Rain, Steam, and Speed by William Turner (1844);
(B) forward lean—Liberty Leading the People by Eugene Delacroix (1830); (C) stroboscopic effects—Dynamism
of a Dog on a Leash by Giacomo Balla (1912); (D) The Wind by Hans Hofmann (1942). The use of form
features to convey implied motion in artworks reached its peak in the early 20th century—the representa-
tion of movement became the central issue of the poetics of the Futurist avant-garde. Futurists wanted to
capture the remarkable speed and dynamism that characterized their time: “the splendour of the world has
been enriched by a new beauty: the beauty of speed” (Futurist Manifesto; Marinetti, 1908; p. 286). This is
represented in Giacomo Balla’s painting Dynamism of a Dog on a Leash (1912; see Figure 9.2C), which uses
stroboscopic images as a means to communicate motion—one can almost feel the frantic energy of the dog
and its walker trying to keep up. Similarly, another paradigmatic example of using form features to convey
implied motion even in the absence of recognizable content or objects is the abstract action painting style
developed towards the middle of the 20th century, primarily associated with artists such as Jackson Pollock,
Franz Kline, and Hans Hofmann (see Figure 9.2D).

175
Darda, Bara and Cross

Movement appreciation—antecedents and consequences


As Figures 9.1 and 9.2 show, implied motion cues have long been used in all types of artwork, irrespective
of content. At the most fundamental level, content in visual art concerns the depiction of form: humans,
still lifes, landscapes, or natural scenes, as well as non-figurative or abstract representations (Harrison, 2009).
Evidence suggests that a majority of people tend to place higher aesthetic value on landscape paintings
compared to paintings with human content and that such preferences are stable over time (Vessel & Rubin,
2010; Augustin et al., 2012; Leder et al., 2016; Pugach et al., 2017; Vessel et al., 2018). This preference also
applies to artworks that depict implied motion. While both paintings portraying dynamic human forms and
dynamic landscapes are rated as more aesthetically pleasing than paintings depicting static landscapes and
human bodies, dynamic landscape paintings are rated as more aesthetically pleasing than dynamic paintings
with a human content (Cazzato et al., 2012; Nather et al., 2014; Orgs et al., 2013; Palmer & Langlois, 2017;
Massaro et al., 2012; Bara et al., 2021).
Non-figurative or abstract art also contains implied motion cues, even when the artwork contains no
representational or figurative content (see Figure 9.1D; Cutting et al., 2002). Overall, abstract art is preferred
less than representational art (Uusitalo et al., 2009; Vessel & Rubin, 2010; Brinkmann et al., 2014; Hayn-
Leichsenring et al., 2020). This preference is, however, modulated by art expertise, such that the preference
for abstract art is higher for art experts compared to art-naïve participants (Pihko et al., 2011; Van Paasschen
et al., 2015; Bimler et al., 2019). So far, the aesthetic appreciation of abstract artworks with implied move-
ment has only been modestly researched, but preliminary evidence suggests that dynamic abstract artworks
might be preferred to static abstract artworks (Cattaneo et al., 2017). In sum, behavioural evidence suggests
that across different types of content, dynamic artworks are rated as more beautiful and more aesthetically
pleasing than static artworks, suggesting a relationship between implied motion perception and aesthetic
appreciation (Bara et al., 2021; Massaro et al., 2012; Mastandrea & Umilta, 2016).
Emerging evidence also suggests that changing a painting’s title to a metaphorical title that is in congru-
ence with its meaning or content can enhance the perceived aesthetic value of the painting (Millis, 2001;
Leder et al., 2006; Mastandrea & Umiltà, 2016). In the context of this work, Mastandrea and Umiltà (2016)
found that when titles of Futurist artworks with movement-related words were manipulated such that some
titles were “increased” to enhance the sense of dynamism, and some titles were “decreased” to diminish the
sense of dynamism, paintings with the “increased” titles were rated as more dynamic. Paintings with more
movement (irrespective of title) were liked more than paintings with less movement. This finding suggests
that ratings of movement or dynamism are somewhat flexible and can be influenced by extraneous factors
and contexts.
Limited behavioural evidence also suggests that the level of dynamism or implied motion in a painting
can influence participants’ subjective estimation of time. Participants exposed for variable periods of time
to abstract and representational paintings varying in dynamism reported that they had been exposed to the
dynamic paintings for a longer duration than actually happened, suggesting that participants’ perception of
time is affected by the observation of implied motion cues (Bueno & Nather, 2012; Nather et al., 2012;
Nather et al., 2014).
In summary, the literature reviewed here suggests that implied motion cues affect the aesthetic apprecia-
tion of visual artworks. Taken together, the findings we have highlighted suggest that dynamic artworks, irre-
spective of whether they depict human bodies, landscapes, or abstract forms, are preferred to static artworks.

Neural mechanisms of movement appreciation


Visual neuroaesthetics has studied the brain-based foundations of aesthetic appreciation across a diverse set
of art forms, ranging from paintings to watching professional dance performances (see Chapter 6). Most of

176
Movement appreciation

this research has focussed on understanding the response to visual properties such as form, colour, symme-
try, complexity, luminance, or contrast (Bar & Neta, 2006; Jacobsen & Höfel, 2003; Jacobsen et al., 2006;
Palmer & Schloss, 2010; Nadal et al., 2010; Vartanian et al., 2013; Bona et al., 2015; Graham et al., 2016;
Van Geert & Wagemans, 2020; Hayn-Leichsenring et al., 2020; Iigaya et al., 2021).
Only recently has this research been expanded to also include studies that explore the way motion cues
influence neural activity associated with aesthetic appreciation (e.g., Kim & Blake, 2007; Di Dio et al., 2016;
Thakral et al., 2012; Bara et al., 2021). Neuroimaging studies that have examined this question have used
abstract and representational paintings as well as sculptures (Kim & Blake, 2007; Di Dio et al., 2011; Di Dio
et al., 2016), dynamic and static photographs of people (Proverbio et al., 2009; Kourtzi & Kanwisher, 2000;
David & Senior, 2000), glass patterns (Krekelberg et al., 2005), and line cartoons such as the Hokaisai manga
(Osaka et al., 2010) as stimuli. In the following sections, we first briefly summarize what is known about the
general neural mechanisms of aesthetic appreciation. We then review what we have learned about neurobio-
logical processes specifically involved in the appreciation of different types of movement stimuli.

Models of aesthetic appreciation


In psychology, aesthetic appreciation has been modelled as a system of information processing hubs, where
aesthetic responses to individual stimuli are considered an emergent function of both bottom-up and top-
down processing stages (e.g., Pelowski, Markey et al., 2016; Spee et al., 2018). These models often distinguish
between automatic and controlled processes. Thus, many existing models assume that stimulus properties
prompt a number of automatic processing stages involved in representing perceptual features such as motion,
shape, colour, symmetry, and complexity. The result of these processing stages then engenders a number of
controlled processing stages during which cognitive processes help extract the perceived aesthetic value and
meaning of the artwork (Leder & Nadal, 2014; Locher et al., 2010; Pelowski & Akiba, 2011; Locher et al.,
2010; Silvia, 2005; Graf, & Landwehr, 2015; Redies, 2015).
In addition to these psychologically based models, neuroimaging experiments have examined the neuro-
biological mechanism associated with aesthetic appreciation. Based on this work, Chatterjee and Vartanian
(2014) have proposed that aesthetic experiences emerge from an interplay between sensory–motor, emo-
tion–valuation, and meaning–knowledge processes (the so-called “aesthetic triad”; see also Chapter  10).
Consistent with the aesthetic triad model, neuroimaging studies have demonstrated that aesthetic responses
to art are not limited to the engagement of only single brain regions. Rather, they depend upon a distributed
network across both hemispheres (Boccia et al., 2016). For example, aesthetic appreciation of visual artworks
shows engagement of the occipital-temporal brain regions when appreciating visual features such as shape,
colour, and symmetry (Vartanian  & Skov, 2014), whereas evocative artworks engage the bilateral insula,
precuneus, and anterior cingulate cortex (Cupchik et al., 2009; Di Dio et al., 2007; Huang et al., 2011; Var-
tanian & Goel, 2004). Since the experience of art has been described as pleasurable and gratifying (Dutton,
2009), the engagement of the reward brain circuit when evaluating aesthetics of an artwork is reasonably
expected. Indeed, several studies have emphasized that aesthetic judgments are associated with the activation
of the mesolimbic reward circuit (for reviews, see Boccia et al., 2016; Brown et al., 2011; Di Dio & Gallese,
2009; Kirsch et al., 2016; see also Chapters 2, 3, 6, and 7 in this volume). Taken together, these findings sug-
gest that an extended network of brain regions known to be involved in sensory, perceptual, cognitive, and
reward processes all play a role in aesthetic appraisal.

Neural correlates of real and implied motion perception in general


Motion plays a critical role in our survival as a species, and it is thus unsurprising that across millennia,
humans have developed robust abilities in perceiving, interpreting, and responding to motion (Grossman

177
Darda, Bara and Cross

et  al., 2000; Thompson  & Parasuraman, 2012). The neural substrates of motion perception have been
investigated extensively in both animals and humans. Evidence from human neuroimaging and brain stimu-
lation studies suggests that the medial temporal complex (MT+/V5) plays a key role in motion processing
(Dupont et al., 1994; Goebel et al., 1998; Grossman et al., 2000; Watson et al., 1993; Zeki et al., 1991).
Neurons located within MT+/V5 are also involved in decoding the speed and direction of objects in motion
(Beckers & Zeki, 1995; Born & Bradley, 2005) and unify motion signals from primary visual cortex (V1) into
coherent global motion (Snowden et al., 1991). Interestingly, MT+/V5 is not only engaged when perceiv-
ing real motion but also illusory (Zeki et al., 1993; Kobayashi et al., 2002), imagined (Goebel et al., 1998;
Winawer et al., 2010), and implied motion (Kourtzi & Kanwisher, 2000; Proverbio et al., 2009; Osaka et al.,
2010; Senior et al., 2000).
Converging findings from a number of neuroimaging studies suggest that images that incorporate motion
cues and induce motion perception in observers also engage the same cortical regions (MT+/V5) respon-
sible for processing real motion (Kourtzi  & Kanwisher, 2000; Proverbio et  al., 2009; Osaka et  al., 2010;
Senior et al., 2000). Using functional magnetic resonance imaging (fMRI), these studies have reported a
higher blood-oxygen level dependent (BOLD) signal response in MT+/V5 to images that imply movement
(e.g., a ballet dancer executing pirouettes or ocean waves breaking) compared with static images without
implied movement. In addition, Senior et al. (2002) conducted a brain stimulation study to investigate the
causal role of medial temporal cortex in implied motion processing. They disrupted neural activity in MT+
by applying transcranial magnetic stimulation (TMS) pulses over this area of the brain, which led to a reduc-
tion in participants’ perception of implied motion. Such techniques enable researchers to better understand
the important functional contribution made by MT+ to implied movement processing.

Neural mechanisms of movement appreciation


One of the fundamental aims of neuroaesthetics is to examine the neurobiological bases of aesthetic experi-
ence, and most research paths are focused on investigating the aesthetic response to a varied set of art forms,
spanning visual artworks and dance performance to non-art forms, such as natural landscape or face aesthet-
ics (Pearce et al., 2016). Given that the appreciation of dance is covered in depth elsewhere in this volume
(Chapter 16), the rest of the chapter focuses on aesthetic appreciation of implied movement, primarily in
visual arts.
In one of the first neuroimaging studies investigating movement appreciation in visual art, Kim and
Blake (2007) investigated the extent to which the BOLD response within brain region MT+ is related to
the impression of movement from motion cues in art and also to its aesthetic appreciation. Higher aesthetic
evaluation scores were linked to increased motion perception, but only among experienced art viewers.
Overall, MT+ showed more engagement when participants viewed abstract artworks with implied motion
cues compared to more static artworks. However, from this study, it was not clear whether MT+ is more
engaged because it is involved in the aesthetic appreciation of the artwork, or simply because participants pay
more attention to dynamic artworks.
Using dynamic and static figurative artworks, Di Dio et al. (2016) reported a more robust BOLD response
in MT+ both when participants were making judgements about the dynamism of a painting as well as when
they were engaged in the task of assigning an aesthetic value to it. This finding consequently suggests that
visual motion areas may contribute to both aesthetic and non-aesthetic (motion) judgements. In contrast,
however, Thakral et al. (2012) measured BOLD responses in MT+ while participants viewed Van Gogh’s
dynamic paintings and found no correlation between MT+ responses and pleasantness ratings.
Further support for the involvement of the MT+ in aesthetic processing comes from a TMS study
conducted by Cattaneo et al. (2017). They applied TMS over either the MT+ or vertex while participants
viewed dynamic and static paintings and made judgements about implied motion and aesthetic appreciation.

178
Movement appreciation

The results showed that compared to vertex stimulation, TMS over the MT+ decreased implied motion
perception in dynamic artworks overall, as well as participants’ aesthetic preference for abstract artworks.
However, TMS over the MT+ did not have any influence on participants’ preference for figurative artworks.
These findings further elucidate the causal role of the MT+ in aesthetic appreciation and suggest that the
contribution of the MT+ to aesthetic appreciation may depend on the type or content of the artwork.

Neural correlates of movement appreciation modulated by content


Previous neurobiological evidence demonstrates that the aesthetic appreciation of movement begins in brain
regions associated with visual processing. Besides motion cues, this work suggests that other basic elements of
visual processing, such as colour, luminance, shape, and orientation, as well as higher-order object recogni-
tion (i.e., identifying faces, bodies, and scenes), are all decoded within visual brain regions before signals are
transferred to brain areas involved in extracting affective and semantic value (Chatterjee & Vartanian, 2014;
Kirsch et al., 2016). In order to understand the neural correlates of movement appreciation, it is thus vital
to consider an integrated approach, which recognizes the importance of content representations and how
distinct type of content representations engage different neural correlates.

Neural correlates of human movement appreciation


The cognitive and neural representations we hold for the human body (for ourselves and others) continu-
ously evolve and are updated following developmental, cultural, cognitive, and emotional changes. Human
body forms are rich vehicles mediating signals that relate to the self and others and promote social interac-
tions (Featherstone et al., 1991). The human body also fosters hedonic values associated with body aesthetics
(Aglioti et al., 2012; Gallese, 2003; Gallese & Di Dio, 2012). In most cultures, body attractiveness is associ-
ated with physical standards of body preference. The idea that a beautiful body can be objectively measured is
deeply rooted within European culture and can be traced back to ancient Greece (Clark, 1984; Livio, 2003).
Across multiple depictions, from visual art to dance, theatre, and film representations, the human body cap-
tures attention and is thought to convey a mixture of meaning-attractiveness-affective signals that ultimately
impact a perceiver’s aesthetic experience.
Body movements such as walking, running, and dancing provide various cues about a person’s health,
age, sex, emotions, and attractiveness (Kozlowski, & Cutting, 1977; Montepare, & Zebrowitz-McArthur,
1988; Loula et al., 2005; Dittrich et al., 1996; Fink et al., 2015). Our understanding of the brain mechanisms
involved in perceiving (and appreciating) the human body in motion is inherently linked to the neural corre-
lates of body and face processing. Previous neuroimaging findings have found the visual presentation of human
body forms to engage specific cortical regions, such as the extrastriate body area (EBA) and the fusiform body
area, two patches of cortical tissue located on the ventral surface of the temporal lobe (Peelen & Downing,
2007). In addition to engaging EBA, representations of the human body displaying (implied or real) dynamic
postures recruit additional brain regions, including the superior temporal sulcus (STS; Grossman & Blake,
2002; Peuskens et al., 2005; Peelen et al., 2006) and large portions of the fronto-parietal brain system (Urgesi
et al., 2006, 2007; Candidi et al., 2008; Rizzolatti & Craighero, 2004; Chong et al., 2008).
Several hypotheses explaining the relationship between body movement and aesthetic preference have
been put forward. One, taking its departure from Darwinian aesthetics (Grammer et al., 2003), proposes that
body movement aesthetics is linked to mate choice and sexual selection rituals (see also Chapter 11). It has
been suggested that the more extravagant the body postures and movements, the higher the reported ratings
on the subjective perception of fitness and consequently aesthetic preference.
Another hypothesis has linked the aesthetic appreciation of dynamic human body movements to the role
played by mirror neurons in the neural coupling of perception and action (Di Pellegrino et al., 1992; Gallese

179
Darda, Bara and Cross

et al., 1996). Mirror neurons, originally discovered within sensorimotor cortices of the non-human primate
brain, respond both when an action is performed and when the same action is observed being performed
by another actor. Since the initial discovery of mirror neurons in the monkey brain, considerable work has
examined the extent to which homologous regions of the human brain link action with perception. This
work has given rise to what is termed the action observation network (AON), comprising sensorimotor brain
regions, including premotor, parietal, and occipitotemporal cortices, that is believed to be engaged when
humans watch others engaged in action (Cross et al., 2009, 2011).
The discovery of mirror neurons and the subsequent characterization of the AON, including the role
played by physical embodiment in shaping this network (e.g., Cross et al., 2006), has informed an influential
aesthetics theory proposed by Freedberg and Gallese (2007). This theory places the body and bodily expe-
riences of the perceiver at the forefront of aesthetic experience. Known as the embodied simulation theory of
aesthetics, it posits that the engagement of the perceiver’s body in simulating real or implied emotions or
actions plays an important role in aesthetic processing. For example, Freedberg and Gallese (2007) hypoth-
esized that when viewing actions depicted on a canvas, either in the form of human body representations
or brushstrokes representing the artists’ movements, viewers covertly simulate the actions being observed,
with this simulation contributing to aesthetic appraisal. In support of this idea, Leder et al. (2012) found
that when participants performed hand movements congruent with the type of brushstroke used in a paint-
ing, their aesthetic appraisal of that painting increased compared to situations where they made incongruent
movements. Although the embodied theory of aesthetics was initially based on work involving responses to
paintings, it is also extremely relevant in the domain of performing arts (especially dance), as many research-
ers have since suggested and investigated (Calvo-Merino et al., 2008; Calvo-Merino et al., 2010; Cross et al.,
2011; Kirsch et al., 2015).
In the first-ever study to examine naïve observers’ aesthetic preferences for dance movements, Calvo-
Merino et al. (2008) used fMRI to examine brain responses while participants viewed dance clips relative to
a control task. Their findings showed stronger engagement of the right premotor and occipital cortices while
participants observed stimuli they later rated as beautiful. Following this work, Cross et al. (2011) investi-
gated the role of embodied simulation, in particular the involvement of sensorimotor brain areas in aesthetic
appraisal of dance performance, and how engagement of these brain regions related to dance-naïve observ-
ers’ actual physical abilities. The results demonstrated that higher activity in occipito-temporal and parietal
regions correlated with higher aesthetic appreciation ratings, especially for movements which were rated as
more difficult for observers to reproduce themselves. Taken together, these two studies suggest that aesthetic
appreciation of dynamic body forms involves sensorimotor portions of the brain associated with the AON,
even when the observed movements extend far beyond the observer’s own physical abilities.
Support for the role played by sensorimotor embodiment in the aesthetic appreciation of visual artworks
comes from studies showing engagement of motor and premotor cortices when participants engage with
these types of artworks. For example, motor simulation, as indexed by cortico-spinal excitability from a
wrist extension muscle, was facilitated when participants viewed a painting showing a hand extension move-
ment (Michelangelo’s Expulsion from Paradise) compared to other paintings with no hand movement being
displayed (although the researchers did not assess excitability during explicit aesthetic appreciation per se;
Battaglia et al., 2011). In a similar vein, an electroencephalography (EEG) experiment found suppression of
the mu rhythm (a measure of motor activation) during passive observation of Lucio Fontana’s slashed canvases
(where the actions of the artist can be inferred) but not during observation of modified versions of the same
canvases (where the actions of the artist cannot be inferred).
A related fMRI study conducted by Di Dio et al. (2011) examined the brain response when participants
observed classical sculptures compared with real photographs of athletes in two conditions: canonical versus
modified proportions and dynamic versus static postures. Although the authors did not explicitly com-
pare dynamic versus static body postures, they found an overall higher activation in the STS for dynamic

180
Movement appreciation

sculptures compared to baseline. In addition, by using dynamic and static paintings with human and land-
scape content, Di Dio and colleagues (2016) found activation of parietal and temporal areas mainly driven by
paintings depicting human content with implied motion compared to more static paintings which did not
show humans in action. Thus, converging evidence from human neuroscience techniques spanning fMRI,
TMS, and EEG suggests that motor simulation of implied motion in artworks critically shapes their aesthetic
appreciation.
With that being said, it is important to emphasize that engagement of sensorimotor brain regions alone
cannot explain the whole story of aesthetic appreciation. According to the neurobiological models already
mentioned, it is realistic to assume that visual areas, sensorimotor brain areas, and brain regions known to
be involved in emotion-hedonic valuation all inform aesthetic appreciation (Chatterjee & Vartanian, 2014;
Martin-Loeches et al., 2014). Accordingly, a recent brain imaging study examined the extent to which aes-
thetic appreciation for implied movement in art that features the human body engages brain regions that span
different circuits, including the EBA, MT+, and the reward brain circuit (Bara et al., 2021). During scanning,
participants made aesthetic and motion judgments of paintings representing human bodies in dynamic and
static postures. Using functional region-of-interest and Bayesian multilevel modelling approaches, the results
revealed no unique functional contribution within or between the main brain systems of interest—EBA,
MT+ and reward circuit—to aesthetic appreciation of paintings with implied motion. Instead, exploratory
whole-brain connectivity analyses showed functional coupling between neural systems associated with body
perception and dorsal parietal cortex for aesthetic appreciation of artworks. These results, although sugges-
tive, are consistent with hierarchical models of aesthetic processing which assume a continuous interaction
between perceptual and attentional neural systems (Iigaya et al., 2021).
Overall, the neuroimaging findings discussed in this section suggest that while theoretical accounts pro-
posing an interplay between perceptual and affective networks hold intuitive appeal, nascent empirical efforts
to characterize these relationships are far from straightforward and require further investigation.

Appreciation of dynamic landscapes


The pictorial tradition of depicting natural scenes has a rich history, with origins tracing back to 4th century
Chinese art and the European Renaissance in the 15th century (Clark, 1976). Behavioural evidence suggests
that landscape content images are appreciated more than images with human and abstract content (Vessel &
Rubin, 2010; Augustin et al., 2012; Leder et al., 2016; Pugach et al., 2017; Vessel et al., 2018). One possible
explanation for this finding is that, for adaptive reasons, people prefer savannah-like landscapes with unhin-
dered views and a presence of animal life and water bodies because such vistas have been associated with
safety, shelter, and access to food and water (Voland & Grammer, 2003; Heerwagen & Orians, 1993). An
alternative hypothesis has proposed that, compared to other types of content, landscape images are more eas-
ily perceived because of their visual symmetry and simplicity, leading them to be preferred over these other
types of content (Mayer & Landwehr, 2018). In addition to behavioural evidence, neuroimaging studies have
suggested that highly preferred and beautiful landscape images engage the parahippocampal place area (PPA)
as well as the reward circuit compared to less preferred and less beautiful landscape images or images featuring
other types of content (Kawabata & Zeki, 2004; Yue et al., 2007; Boccia et al., 2016).
Given people’s aesthetic preference for landscapes, it is surprising that the aesthetic appreciation of natu-
ral scenes as a distinct aesthetic category beyond arts was recognized only in the 18th century (Brown &
Dissanayake, 2017; Seeley, 2014). With the attention given to natural landscapes by poets, painters, and
photography (Kemal & Gaskell, 1993) and over time by cinematic genres, such as Westerns, as well as by
travel and adventure documentaries, a shift occurred in the aesthetic appreciation of landscapes from static
paintings to more dynamic natural scenes: increasingly, static viewpoints became replaced by more dynamic
motion displays (Sitney, 1993).

181
Darda, Bara and Cross

Although there is some behavioural evidence suggesting that paintings with dynamic landscapes are found
more aesthetically pleasing than those with static landscapes (Di Dio et  al., 2016; Massaro et  al., 2012),
dynamic representations of landscape scenes have only been modestly investigated from a neuroimaging per-
spective. An fMRI study conducted by Di Dio et al. (2016) sheds some light on the aesthetic processing of
dynamic landscapes. Comparing landscape scenes and human figures in dynamic and static artworks, Di Dio
and colleagues found that participants behaviourally preferred landscape paintings to human figure paintings
and also reported that dynamic landscapes were preferred over dynamic human paintings. Neuroimaging
results revealed more engagement of the inferior and middle temporal sulci as well as the posterior parietal
and intraparietal sulcus bilaterally for dynamic artworks compared to static artworks, irrespective of content.
In addition, static compared to dynamic landscapes engaged the central and posterior insula. The authors
suggest that the central insula may represent a locus for sensorimotor processes to interact with internal affec-
tive states. Based on this idea, they speculate that the perception of static paintings might involve additional
internally generated sensorimotor processing associated with the imaginary exploration of the depicted
scenery—an idea in line with Freedberg and Gallese’s (2007) embodied theory of aesthetic appreciation.
Evidence for this hypothesis remains to be seen, but recent TMS work (Finisguerra et al., 2021) has found
late muscle-specific activation to correlate with the observation of landscapes painted with a brushstroke
style rather than a pointillist-like style, suggesting that motor simulation of the painters’ movements might be
essential to subjective aesthetic preference.
Taken together, neuroimaging evidence suggests that aesthetic appreciation of dynamic landscapes likely
involves an extensive network that includes the visual occipital areas and the parahippocampal place area, as
well as the sensorimotor and reward brain circuits.

The role of abstract stimuli in movement appreciation


Previous studies investigating aesthetic appreciation of abstract art compared to figurative art found that
abstract art is less preferred to figurative art (such as landscapes and portraits; e.g., Uusitalo et  al., 2009;
Vessel & Rubin, 2010; Brinkmann et al., 2014; Hayn-Leichsenring et al., 2020). One potential explanation
for this observation is the so-called fluency theory, according to which stimuli that are easier to process are
more liked (Reber et al., 2004). In the case of abstract art, evidence shows that both perceptual fluency and
conceptual fluency (i.e., the ease with which one can extract meaning), is reduced. Possibly, the effortful
and laborious processing associated with abstract art affects the viewer’s interest and motivation, ultimately
diminishing the pleasure they feel (Graf & Landwehr, 2017; Mayer & Landwehr, 2018).
In one of the first neuroimaging studies using abstract art, Vartanian and Goel (2004) investigated brain
activity associated with aesthetic judgments of figurative and abstract artworks. These authors found greater
activation in the bilateral occipital gyrus, fusiform gyrus, and precuneus for figurative art, suggesting that
non-abstract art engages brain systems previously associated with object perception. Cattaneo et al. (2015)
found that applying TMS pulses over the lateral occipital cortex during an aesthetic appreciation task selec-
tively interfered with aesthetic preference for figurative art but not for abstract art. Overall, this evidence sup-
ports the idea that figurative art appreciation relies on neural systems involved in object recognition, whereas
the lateral occipital cortex does not play a causal role in abstract art appreciation. Additional evidence suggests
that aesthetic preferences for abstract paintings engage the posterior cingulate cortex, a brain region associ-
ated with the default mode network (Vessel et al., 2012). This finding implies that abstract art may relate
more to conceptual and internally directed processing of thought.
Contrasting neuroimaging evidence comes from investigations of implied movement in abstract artworks.
Artists have long used visual form cues to convey a sense of movement, such as stroboscopic effects or the
direction of speed cues (Cutting, 2002). While some early neuroimaging evidence suggests that implied
motion cues in abstract art activate the MT+ (Kim & Blake, 2007), the aesthetic appreciation of abstract

182
Movement appreciation

forms implying movement has not been extensively investigated. One example that may suggest a link
between aesthetic preference and kinetic abstract patterns comes from Zeki and Stutters’s (2012) study, which
reported that participants’ subjective preference for moving abstract patterns correlated with higher engage-
ment of visual areas, as well as with activity in the medial orbitofrontal cortex. Together, the authors inter-
preted this pattern of findings as support for the hypothesis that aesthetic appreciation for abstract movement
involves shared mechanisms with other forms or stimuli that imply motion. Furthermore, a recent study by
Cattaneo et al., 2017 examining the causal contribution of the MT+ to aesthetic appreciation of dynamic
figurative and abstract art found that applying TMS pulses over the MT+ reduced aesthetic preference for
abstract artworks but not for figurative artworks. The authors interpreted this finding as evidence that the
aesthetic appreciation of abstract art is mostly driven by sensory perceptual features, such as motion, shape,
and colour.
A recent study by Humphries et al. (2021) using “high-motion” Jackson Pollock and “low-motion” Piet
Mondrian abstract paintings showed altered art appreciation among patients diagnosed with Parkinson’s dis-
ease that was specifically linked to their altered ability to translate implied motion cues from abstract artworks
into movement representations. This suggests involvement of the motor system (which is disrupted in people
with Parkinson’s disease) in representing movement from motion cues in abstract art. The authors further
suggest that the motor system may integrate low-level visual features of the artwork to form abstract repre-
sentations of movements rather than simulating them as actions (as proposed by the embodiment simulation
account of aesthetics; Freedberg & Gallese, 2007).
In summary, the findings summarized in this section suggest that aesthetic evaluation of dynamic abstract
art engages more visual than sensorimotor areas in the brain. In contrast, the aesthetic evaluation of paintings
with dynamic representational content (especially human content) seems to involve the sensorimotor system,
supporting theories of embodied aesthetics that suggest motor responses are linked to aesthetic pleasure, even
in the context of visual art such as paintings and sculptures. However, this has been challenged by recent
evidence that shows no straightforward relationship between sensorimotor responses and aesthetic apprecia-
tion ratings for visual art (Bara et al., 2021; Humphries et al., 2021). Instead, recent findings suggest that a
broad fronto-parietal network is involved in aesthetic processing, and the aesthetic appreciation of artworks
relies on an interplay between regions of the visual ventral stream and regions in the prefrontal and parietal
areas, which are also part of the default mode network (Di Dio et al., 2016; Bara et al., 2021; Iigaya et al.,
2021; Vessel et al., 2019).

Summary of neural mechanisms underlying movement appreciation


Advances in human neuroimaging techniques during the past few decades have spurred the development of
a cognitive neuroscience of aesthetics, increasing our understanding of the underlying cognitive, affective,
and neural mechanisms that shape aesthetic experiences (Cela-Conde et al., 2011). Most models of aesthetic
appreciation recognize an interplay between perceptual, cognitive, and emotional or reward neural networks
and stages to aesthetic processing, even at the level of low-level feature or object processing (Cupchik et al.,
2009; Kirk et al., 2009; see also Brieber et al., 2014; Chatterjee, 2014; Leder et al., 2013). Movement percep-
tion and appreciation represents a meaningful instance of such interactions—the perception and appreciation
of movement involves the processing of both low-level features such as orientation and colour, as well as
high-level features such as content represented in the art. Neuroimaging evidence suggests that, in addition
to motion-sensitive visual areas such as the MT+, higher-level visuo-motor areas such as the motor and
premotor cortices, as well as the temporal and parietal cortices, are engaged in the evaluation and apprecia-
tion of movement (Thakral et al., 2012; Cattaneo et al., 2015; Proverbio et al., 2009; Battaglia et al., 2011;
Di Dio et al., 2011, 2016; Concerto et al., 2016). Most studies that have investigated human movement in
art have also found engagement of higher visual areas (EBA, STS) along with areas involved in motor-mirror

183
Darda, Bara and Cross

mechanisms (Di Dio et al., 2007; Proverbio et al., 2009; Calvo-Merino et al., 2010; Cross et al., 2011) and
their interaction (Concerto et al., 2016). Thus, the interplay between regions that encode the human body
and those implicated in motor processing is evident in the appreciation of human movement depicted in the
visual and performing arts.
However, the specifics of this relationship remain unclear at this stage and are especially unclear with
respect to artworks that lack a depiction of the human form (such as landscapes or non-figurative abstract
forms; Di Dio et al., 2016; Kim & Blake, 2007). Neuroimaging evidence continues to accrue that supports
the suggestion that aesthetic appreciation of visual art is brought about by an interplay of different networks
in the brain that span sensorimotor, visual perception, and reward networks (see Figure 9.3; Kurth et al.,
2010; Brown et al., 2011; Boccia et al., 2016; Vartanian & Skov, 2014; Kirsch et al., 2016). More recent evi-
dence, however, suggests that the aesthetic appreciation of artworks with implied movement in both abstract
art and art with representational human content relies less on sensorimotor systems and more on an inter-
play between regions of the visual ventral stream and regions in the prefrontal and parietal areas, which also
make up the default mode network (Bara et al., 2021; Humphries et al., 2021; Vessel et al., 2019). Overall,
therefore, this section highlights the fact that investigations on movement appreciation are raising more ques-
tions than they are answering, and future work will be essential in gleaning a more meaningful and nuanced
understanding of movement appreciation.

Figure 9.3 Brain networks implicated by previous research in movement appreciation include the perceptual/visual areas,
sensorimotor network, reward network, and regions of the default-mode network. Current evidence, how-
ever, does not allow us to determine whether particular brain areas perform specific functions (e.g., the role
of the vmPFC) or whether regions (such as the MT+) encode movement in a domain-general or domain-
specific manner. Abbreviations—vmPFC = ventral medial prefrontal cortex, mmPFC = medial medial pre-
frontal cortex, dmPFC = dorsal medial prefrontal cortex, ACC = anterior cingulate cortex, NAcc = nucleus
acumbens, AMG = amygdala, PCC = posterior cingulate cortex, M1 = primary motor cortex, S1 = pri-
mary sensory cortex, PMC  =  premotor cortex, PPA  =  parahippocampal place area, EVA  =  early visual
areas, FFA = fusiform face area, EBA = extrastriate body area, LTC = lateral temporal cortex, IPL = inferior
parietal lobule, pSTS = posterior superior temporal sulcus, aI = anterior insula, OFC = orbitofrontal cortex.

184
Movement appreciation

Major challenges, goals, and suggestions


In this section, we summarize the major challenges, goals, and suggestions for future research not just with
respect to the field of movement appreciation but also as they apply to the field of neuroaesthetics as a whole.

Functional specificity in movement appreciation and the development of a


cumulative science
The extant literature provides us with a foundation for the neural mechanisms underpinning movement
appreciation. A  more specific and functional characterization of neural networks underlying movement
appreciation, however, remains far from clear, at present. As an example, very few studies in the field have
defined the regions of interest functionally, making claims about functional specificity difficult. For instance,
the medial prefrontal cortex (mPFC) is known to play a role in internally driven thoughts as well as reward
processing more generally. However, engagement of the mPFC for aesthetic appreciation does not indicate
whether it is involved in reward processing or processing of internally driven thoughts (or both), and authors
often pick the interpretation that most closely fits their theory or hypothesis, without regard for what they
can actually conclude from their findings, given their study design (see Figure 9.3). Interpreting findings
in such a way is a classic example of reverse inference (Poldrack, 2006). Such approaches suffer from poor
functional specificity, given the macroscopic resolution and heterogeneous nature of swathes of neural tissue
that are typically under investigation when using techniques like fMRI. In a similar vein, claims regarding the
role of the frontoparietal cortex in “mirroring” processes, as well as a role for MT+ in sensitivity to motion
and aesthetic preferences (e.g . Di Dio et al., 2016), could also be explained by processing demands such as
attention.
While existing studies have provided us with a sound framework within which to work, the field of
movement appreciation, and neuroaesthetics as a whole, can benefit from moving toward more sophisticated
analysis pipelines that increase functional sensitivity such as by first identifying functional units and then test-
ing how they respond during an aesthetic judgement task ( Julian et al., 2012; Nieto-Castañón & Fedorenko,
2012). To add on to univariate analytical approaches, which can test relatively simple models of brain organi-
zation, a move toward more multivariate and connectivity analyses would enable researchers to test more
complex models of movement appreciation (e.g. Cela-Conde et al., 2013; Orgs et al., 2016; Bara et al., 2021;
Vessel et al., 2019; Iigaya et al., 2021). For instance, in the domain of movement perception, one important
question will be to clarify whether the MT+ or sensorimotor areas contain a domain-general representation
of movement appreciation across different types and forms of art or whether they represent aesthetic appeal
in a domain-specific fashion (see Figure 9.3).
In addition, given recent concerns over questionable research practices and low levels of reproducibility
in the fields of psychology and neuroscience, neuroaesthetics as a whole, and movement appreciation more
specifically, can also benefit from practices that encourage the development of a cumulative science of move-
ment appreciation (Button et al., 2013; Open Science Collaboration, 2015; Simmons et al., 2011). Before
moving on to construct elaborate theories, test complex models, and/or attempt to synthesize findings, it
may be sensible for the field to use more robust methodological approaches and embrace the credibility
revolution in order to develop a more credible foundation for future studies to build upon (Ramsey, 2020;
Vazire, 2018).

A more general and inclusive model


Overall, the aesthetic preference or appreciation of movement would gain more insight by investigating
stimuli across different domains, as well as by including populations from different cultures and ages and

185
Darda, Bara and Cross

with different neurological conditions. For instance, some evidence suggests that art and non-art images that
imply movement engage different neural regions (e.g. Di Dio et al., 2011). A comparison between art and
non-art stimuli would help assess different aesthetically pleasing motion cues and patterns that are common
across all stimuli, as well as identifying patterns (if any do indeed exist) that may be specific to art. In addi-
tion, movement appreciation may differ across cultures. For instance, the western world is thought to have
a strong preference for intense dynamic activity in the context of sports (Proverbio et al., 2009). Moreover,
an important aspect that might be addressed by future studies in movement appreciation is to examine how
movement appreciation changes during the life span from childhood through to advanced age. Some evi-
dence suggests that musical and visual art aesthetic preferences tend to be formed between 20 and 30 years of
age and that adults present higher stable aesthetic preferences than adolescents (Holbrook & Schindler, 1989;
Pugach et al., 2017). More research into the stability of aesthetic movement preference would deepen our
understanding into how the aesthetic preference is computed and the required heuristics needed to maintain
these preferences.
Although there have been some accounts taking neuropsychological perspectives toward art cognition
(Chatterjee, 2014; Lauring et al., 2019; Humphries et al., 2021), the field of neuro-aesthetics and especially
the aesthetic appreciation of movement would also benefit from research quantifying how art production and
art appreciation change as a function of neurological conditions. Such investigations can also serve as a useful
way of identifying the functional contributions of brain regions and networks in movement appreciation.
For instance, due to motor and spatial deficits, Parkinson’s disease patients might show different preferences
for more dynamic artworks compared to controls, further elucidating the role that motor processing systems
may play in the aesthetic appreciation of movement (e.g., Humphries et al., 2021).

Moving towards a neuroaesthetics of the performance space


The field of neuroaesthetics in general will benefit from encouraging more research to be undertaken within
real-world contexts, such as museums or theatres, especially as people tend to place greater aesthetic value
on artworks presented in art galleries rather than in laboratory environments (Brieber et al., 2014, 2015;
Locher et al., 2001; see also Chapters 22 and 24). A few early studies have investigated the neural correlates of
aesthetic appreciation in more naturalistic contexts by using functional near infrared spectroscopy (fNIRS),
a non-invasive and portable imaging technique, with higher temporal resolution and tolerance for motion
compared to fMRI (Cui et al., 2011; Pelowski, Oi et al., 2016; Kaimal et al., 2017). fNIRS thus offers an
exciting methodological opportunity in exploring the aesthetic experience in general, especially the aes-
thetic experience associated with observing and performing movement.

Conclusion
Given the ubiquitous presence and influence of movement in our lives, it is not surprising that researchers in
the domain of neuroaesthetics have investigated the neurocognitive underpinnings of movement apprecia-
tion. In this chapter, we have described considerable advances made in order to illuminate the complex cog-
nitive and neural mechanisms that subserve movement appreciation. With a focus on movement appreciation
in visual art, we first showed the importance of different implied motion cues that successfully create an
impression of dynamism and their link with aesthetic preference. Second, we discussed the neural correlates
of movement appreciation across different kinds of stimuli and content. Finally, we outlined future goals and
new avenues for the field, highlighting that research in the developing domain of movement appreciation
more specifically and neuroaesthetics more generally will benefit from more inclusive, naturalistic, and robust
methodological approaches in order to generate a more holistic and complete understanding of the aesthetics
of movement.

186
Movement appreciation

References
Aglioti, S. M., Minio-Paluello, I., & Candidi, M. (2012). The beauty of the body. Rendiconti Lincei, 23(3), 281–288.
https://doi.org/10.1007/s12210-012-0169-1
Arnheim, R. (1951). Perceptual and aesthetic aspects of the movement response. Journal of Personality, 19(3), 265–281. 
https://doi.org/10.1111/j.1467-6494.1951.tb01101.x
Arnheim, R. (1974). Art and Visual Perception, second edition. University of California Press.
Augustin, M. D., Wagemans, J., & Carbon, C. C. (2012). All is beautiful? Generality vs. specificity of word usage in visual
aesthetics. Acta Psychologica, 139(1), 187–201. https://doi.org/10.1016/j.actpsy.2011.10.004
Bar, M.,  & Neta, M. (2006). Humans prefer curved visual objects.  Psychological Science,  17(8), 645–648. https://doi.
org/10.1111/j.1467-9280.2006.01759.x
Bara, I., Darda, K. M., Kurz, A. S.,  & Ramsey, R. (2021). Functional specificity and neural integration in the aes-
thetic appreciation of artworks with implied motion. European Journal of Neuroscience, 54(9), 7231–7259. https://doi.
org/10.1111/ejn.15479
Battaglia, F., Lisanby, S. H., & Freedberg, D. (2011). Corticomotor excitability during observation and imagination of a
work of art. Frontiers in Human Neuroscience, 5, 79. https://doi.org/10.3389/fnhum.2011.00079
Beckers, G., & Zeki, S. (1995). The consequences of inactivating areas V1 and V5 on visual motion perception. Brain,
118(1), 49–60. https://doi.org/10.1093/brain/118.1.49
Bimler, D. L., Snellock, M., & Paramei, G. V. (2019). Art expertise in construing meaning of representational and abstract
artworks. Acta Psychologica, 192, 11–22. https://doi.org/10.1016/j.actpsy.2018.10.012
Boccia, M., Barbetti, S., Piccardi, L., Guariglia, C., Ferlazzo, F., Giannini, A. M., & Zaidel, D. W. (2016). Where does
brain neural activation in aesthetic responses to visual art occur? Meta-analytic evidence from neuroimaging stud-
ies. Neuroscience and Biobehavioral Reviews, 60, 65–71. https://doi.org/10.1016/j.neubiorev.2015.09.009
Bona, S., Cattaneo, Z.,  & Silvanto, J. (2015). The causal role of the occipital face area (OFA) and lateral occipi-
tal (LO) cortex in symmetry perception. Journal of Neuroscience, 35(2), 731–738. https://doi.org/10.1523/
JNEUROSCI.3733-14.2015
Born, R. T.,  & Bradley, D. C. (2005). Structure and function of visual area MT. Annual Review of Neuroscience, 28,
157–189. https://doi.org/10.1146/annurev.neuro.26.041002.131052
Brieber, D., Nadal, M., & Leder, H. (2015). In the white cube: Museum context enhances the valuation and memory of
art. Acta Psychologica, 154, 36–42. https://doi.org/10.1016/j.actpsy.2014.11.004
Brieber, D., Nadal, M., Leder, H., & Rosenberg, R. (2014). Art in time and space: Context modulates the relation
between art experience and viewing time. PLOS ONE, 9(6), e99019. https://doi.org/10.1371/journal.pone.0099019
Brinkmann, H., Commare, L., Leder, H., & Rosenberg, R. (2014). Abstract art as a universal language? Leonardo, 47(3),
256–257. https://doi.org/10.1162/LEON_a_00767
Brown, S., & Dissanayake, E. (2017). The arts are more than aesthetics: Neuroaesthetics as narrow aesthetics. In M. Skov &
O. Vartanian (Eds.), Neuroaesthetics. Routledge.
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. Neuroimage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Bueno, J. L. O., & Nather, F. C. (2012). Implied movement perception in different static artworks affects subjective time.
In Proceedings of the annual meeting of the international society for psychophysics (pp. 85–90). Ottawa, Canada.
Button, K. S., Ioannidis, J. P., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S., & Munafò, M. R. (2013). Power fail-
ure: Why small sample size undermines the reliability of neuroscience. Nature Reviews. Neuroscience, 14(5), 365–376.
https://doi.org/10.1038/nrn3475
Calvo-Merino, B., Jola, C., Glaser, D. E., & Haggard, P. (2008). Towards a sensorimotor aesthetics of performing art.
Consciousness and Cognition, 17(3), 911–922. https://doi.org/10.1016/j.concog.2007.11.003
Calvo-Merino, B., Urgesi, C., Orgs, G., Aglioti, S. M., & Haggard, P. (2010). Extrastriate body area underlies aesthetic
evaluation of body stimuli. Experimental Brain Research, 204(3), 447–456. https://doi.org/10.1007/s00221-010-2283-6
Candidi, M., Urgesi, C., Ionta, S.,  & Aglioti, S. M. (2008). Virtual lesion of ventral premotor cortex impairs visual
perception of biomechanically possible but not impossible actions. Social Neuroscience, 3(3–4), 388–400. https://doi.
org/10.1080/17470910701676269
Cattaneo, Z., Lega, C., Ferrari, C., Vecchi, T., Cela-Conde, C. J., Silvanto, J., & Nadal, M. (2015). The role of the lateral
occipital cortex in aesthetic appreciation of representational and abstract paintings: A TMS study. Brain and Cogni-
tion, 95, 44–53. https://doi.org/10.1016/j.bandc.2015.01.008
Cattaneo, Z., Schiavi, S., Silvanto, J., & Nadal, M. (2017). A TMS study on the contribution of visual area V5 to the
perception of implied motion in art and its appreciation. Cognitive Neuroscience, 8(1), 59–68. https://doi.org/10.1080/
17588928.2015.1083968

187
Darda, Bara and Cross

Cazzato, V., Siega, S., & Urgesi, C. (2012). ‘What women like’: Influence of motion and form on aesthetic body percep-
tion. Frontiers in Psychology, 3, 235. https://doi.org/10.3389/fpsyg.2012.00235
Cela-Conde, C. J., Agnati, L., Huston, J. P., Mora, F., & Nadal, M. (2011). The neural foundations of aesthetic apprecia-
tion. Progress in Neurobiology, 94(1), 39–48. https://doi.org/10.1016/j.pneurobio.2011.03.003
Cela-Conde, C. J., García-Prieto, J., Ramasco, J. J., Mirasso, C. R., Bajo, R., Munar, E., Flexas, A., del-Pozo, F., &
Maestú, F. (2013). Dynamics of brain networks in the aesthetic appreciation. Proceedings of the National Academy of Sci-
ences of the United States of America, 110(Suppl. 2), 10454–10461. https://doi.org/10.1073/pnas.1302855110
Chakraborty, S. (1986). Socioreligious and cultural study of the ancient Indian coins. BR. Publication Corp.
Chatterjee, A. (2014). Scientific aesthetics: Three steps forward. British Journal of Psychology, 105(4), 465–467. https://
doi.org/10.1111/bjop.12086
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics.  Trends in Cognitive Sciences,  18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Chong, T. T., Cunnington, R., Williams, M. A., Kanwisher, N., & Mattingley, J. B. (2008). fMRI adaptation reveals
mirror neurons in human inferior parietal cortex. Current Biology, 18(20), 1576–1580. https://doi.org/10.1016/j.
cub.2008.08.068
Christensen, J. F., & Calvo-Merino, B. (2013). Dance as a subject for empirical aesthetics. Psychology of Aesthetics, Creativ-
ity, and the Arts, 7(1), 76–88.  https://doi.org/10.1037/a0031827
Clark, K. (1976). Landscape into art. Harper & Row.
Clark, K. (1984). The nude: A study in ideal form. Princeton University Press.
Concerto, C., Infortuna, C., Mineo, L., Pereira, M., Freedberg, D., Chusid, E., Aguglia, E.,  & Battaglia, F. (2016).
Observation of implied motion in a work of art modulates cortical connectivity and plasticity. Journal of Exercise Reha-
bilitation, 12(5), 417–423. https://doi.org/10.12965/jer.1632656.328
Cross, E. S., & Ticini, L. F. (2012). Neuroaesthetics and beyond: New horizons in applying the science of the brain to
the art of dance. Phenomenology and the Cognitive Sciences, 11(1), 5–16. https://doi.org/10.1007/s11097-010-9190-y
Cross, E. S., Hamilton, A. F. D. C., & Grafton, S. T. (2006). Building a motor simulation de novo: Observation of dance
by dancers. Neuroimage, 31(3), 1257–1267. https://doi.org/10.1016/j.neuroimage.2006.01.033
Cross, E. S., Hamilton, A. F. D. C., Kraemer, D. J., Kelley, W. M., & Grafton, S. T. (2009). Dissociable substrates for body
motion and physical experience in the human action observation network. European Journal of Neuroscience, 30(7),
1383–1392. https://doi.org/10.1111/j.1460-9568.2009.06941.x
Cross, E. S., Kirsch, L., Ticini, L. F., & Schütz-Bosbach, S. (2011). The impact of aesthetic evaluation and physical ability
on dance perception. Frontiers in Human Neuroscience, 5, 102. https://doi.org/10.3389/fnhum.2011.00102
Cui, X., Bray, S., Bryant, D. M., Glover, G. H., & Reiss, A. L. (2011). A quantitative comparison of NIRS and fMRI
across multiple cognitive tasks. Neuroimage, 54(4), 2808–2821. https://doi.org/10.1016/j.neuroimage.2010.10.069
Cupchik, G. C., Vartanian, O., Crawley, A., & Mikulis, D. J. (2009). Viewing artworks: Contributions of cognitive con-
trol and perceptual facilitation to aesthetic experience. Brain and Cognition, 70(1), 84–91. https://doi.org/10.1016/j.
bandc.2009.01.003
Cutting, J. E. (2002). Representing motion in a static image: Constraints and parallels in art, science, and popular culture.
Perception, 31(10), 1165–1193. https://doi.org/10.1068/p3318
David, A. S., & Senior, C. (2000). Implicit motion and the brain. Trends in Cognitive Sciences, 4(8), 293–295. https://doi.
org/10.1016/S1364-6613(00)01511-4
Di Dio, C., Ardizzi, M., Massaro, D., Di Cesare, G., Gilli, G., Marchetti, A.,  & Gallese, V. (2016). Human, nature,
dynamism: The effects of content and movement perception on brain activations during the aesthetic judgment of
representational paintings. Frontiers in Human Neuroscience, 9, 705. https://doi.org/10.3389/fnhum.2015.00705
Di Dio, C., Canessa, N., Cappa, S. F., & Rizzolatti, G. (2011). Specificity of esthetic experience for artworks: An fMRI
study. Frontiers in Human Neuroscience, 5, 139. https://doi.org/10.3389/fnhum.2011.00139
Di Dio, C., & Gallese, V. (2009). Neuroaesthetics: A review. Current Opinion in Neurobiology, 19(6), 682–687. https://
doi.org/10.1016/j.conb.2009.09.001
Di Dio, C., Macaluso, E., & Rizzolatti, G. (2007). The golden beauty: Brain response to Classical and Renaissance sculp-
tures. PLOS ONE, 2(11), e1201. https://doi.org/10.1371/journal.pone.0001201
Di Pellegrino, G., Fadiga, L., Fogassi, L., Gallese, V., & Rizzolatti, G. (1992). Understanding motor events: A neuro-
physiological study. Experimental Brain Research, 91(1), 176–180. https://doi.org/10.1007/BF00230027
Dittrich, W. H., Troscianko, T., Lea, S. E., & Morgan, D. (1996). Perception of emotion from dynamic point-light dis-
plays represented in dance. Perception, 25(6), 727–738. https://doi.org/10.1068/p250727
Dupont, P., Orban, G. A., De Bruyn, B., Verbruggen, A., & Mortelmans, L. (1994). Many areas in the human brain
respond to visual motion. Journal of Neurophysiology, 72(3), 1420–1424. https://doi.org/10.1152/jn.1994.72.3.1420
Dutton, D. (2009). The art instinct: Beauty, pleasure, & human evolution. Oxford University Press.

188
Movement appreciation

Featherstone, M., Hepworth, M., & Turner, B. S. (Eds.). (1991). The body: Social process and cultural theory. Sage.
Finisguerra, A., Ticini, L. F., Kirsch, L. P., Cross, E. S., Kotz, S. A., & Urgesi, C. (2021). Dissociating embodiment
and emotional reactivity in motor responses to artworks.  Cognition,  212, 104663. https://doi.org/10.1016/j.
cognition.2021.104663
Fink, B., Weege, B., Neave, N., Pham, M. N., & Shackelford, T. K. (2015). Integrating body movement into attractive-
ness research. Frontiers in Psychology, 6, 220. https://doi.org/10.3389/fpsyg.2015.00220
Freedberg, D.,  & Gallese, V. (2007). Motion, emotion and empathy in esthetic experience.  Trends in Cognitive Sci-
ences, 11(5), 197–203. https://doi.org/10.1016/j.tics.2007.02.003
Gallese, V. (2003). The manifold nature of interpersonal relations: The quest for a common mechanism. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 358(1431), 517–528. https://doi.org/10.1098/
rstb.2002.1234
Gallese, V., & Di Dio, C. (2012). Neuroesthetics: The body in esthetic experience. In Encyclopedia of human behavior (2nd
ed., pp. 687–693). https://doi.org/10.1016/B978-0-12-375000-6.00251-2.
Gallese, V., Fadiga, L., Fogassi, L., & Rizzolatti, G. (1996). Action recognition in the premotor cortex. Brain, 119(2),
593–609. https://doi.org/10.1093/brain/119.2.593
Goebel, R., Khorram-Sefat, D., Muckli, L., Hacker, H., & Singer, W. (1998). The constructive nature of vision: Direct
evidence from functional magnetic resonance imaging studies of apparent motion and motion imagery.  European
Journal of Neuroscience, 10(5), 1563–1573. https://doi.org/10.1046/j.1460-9568.1998.00181.x
Gombrich, E. H. (1964). Moment and movement in art. In source. Journal of the Warburg and Courtauld Institutes, 27(1),
293–306. https://doi.org/10.2307/750521
Graf, L. K. M., & Landwehr, J. R. (2017). Aesthetic pleasure versus aesthetic interest: The two routes to aesthetic liking.
Frontiers in Psychology, 8, 15. https://doi.org/10.3389/fpsyg.2017.00015
Graf, L. K., & Landwehr, J. R. (2015). A dual-process perspective on fluency-based aesthetics: The pleasure-interest model
of aesthetic liking. Personality and Social Psychology Review, 19(4), 395–410. https://doi.org/10.1177/1088868315
574978
Graham, D., Schwarz, B., Chatterjee, A., & Leder, H. (2016). Preference for luminance histogram regularities in natural
scenes. Vision Research, 120, 11–21. https://doi.org/10.1016/j.visres.2015.03.018
Grammer, K., Fink, B., Møller, A. P.,  & Thornhill, R. (2003). Darwinian aesthetics: Sexual selection and the biol-
ogy of beauty.  Biological Reviews of the Cambridge Philosophical Society,  78(3), 385–407. https://doi.org/10.1017/
S1464793102006085
Grossman, E. D., & Blake, R. (2002). Brain areas active during visual perception of biological motion. Neuron, 35(6),
1167–1175. https://doi.org/10.1016/S0896-6273(02)00897-8
Grossman, E., Donnelly, M., Price, R., Pickens, D., Morgan, V., Neighbor, G.,  & Blake, R. (2000). Brain areas
involved in perception of biological motion.  Journal of Cognitive Neuroscience,  12(5), 711–720. https://doi.
org/10.1162/089892900562417
Harrison, C. (2009). An introduction to art. Yale University Press.
Hayn-Leichsenring, G. U., Kenett, Y. N., Schulz, K., & Chatterjee, A. (2020). Abstract art paintings, global image prop-
erties, and verbal descriptions: An empirical and computational investigation. Acta Psychologica, 202, 102936. https://
doi.org/10.1016/j.actpsy.2019.102936
Heerwagen, J. H., & Orians, G. H. (1993). Humans, habitats, and aesthetics. In S. R. Kellert & E. O. Wilson (Eds.), The
biophilia hypothesis (pp. 138–172). Island Press.
Holbrook, M. B., & Schindler, R. M. (1989). Some exploratory findings on the development of musical tastes. Journal of
Consumer Research, 16(1), 119–124. https://doi.org/10.1086/209200
Huang, M., Bridge, H., Kemp, M. J., & Parker, A. J. (2011). Human cortical activity evoked by the assignment of authen-
ticity when viewing works of art. Frontiers in Human Neuroscience, 5, 134. https://doi.org/10.3389/fnhum.2011.00134
Humphries, S., Workman, C. I., Hayn-Leichsenring, G., Hartnung, F., & Chatterjee, A. (2021). Movement in aesthetic
experiences: What we can learn from Parkinson disease. Journal of Cognitive Neuroscience, 33(7), 1329–1342. https://
doi.org/10.1162/jocn_a_01718
Iigaya, K., Yi, S., Wahle, I. A., Tanwisuth, K.,  & O’Doherty, J. P. (2021). Aesthetic preference for art can be pre-
dicted from a mixture of low- and high-level visual features. Nature Human Behaviour, 5(6), 743–755. https://doi.
org/10.1038/s41562-021-01124-6
Jacobsen, T.,  & Höfel, L. (2003). Descriptive and evaluative judgment processes: Behavioral and electrophysiological
indices of processing symmetry and aesthetics. Cognitive, Affective and Behavioral Neuroscience, 3(4), 289–299. https://
doi.org/10.3758/CABN.3.4.289
Jacobsen, T., Schubotz, R. I., Höfel, L., & Cramon, D. Y. V. (2006). Brain correlates of aesthetic judgment of beauty. Neu-
roimage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010

189
Darda, Bara and Cross

Julian, J. B., Fedorenko, E., Webster, J.,  & Kanwisher, N. (2012). An algorithmic method for functionally defin-
ing regions of interest in the ventral visual pathway.  Neuroimage,  60(4), 2357–2364. https://doi.org/10.1016/j.
neuroimage.2012.02.055
Justi, C. (1923). Winckelmann und seine Zeitgenossen, 3. FCW Vogel.
Kaimal, G., Ayaz, H., Herres, J., Dieterich-Hartwell, R., Makwana, B., Kaiser, D. H., & Nasser, J. A. (2017). Functional
near-infrared spectroscopy assessment of reward perception based on visual self-expression: Coloring, doodling, and
free drawing. Arts in Psychotherapy, 55, 85–92. https://doi.org/10.1016/j.aip.2017.05.004
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Kemal, S., & Gaskell, I. (1993). Landscape, natural beauty and the arts. Cambridge University Press.
Kim, C. Y., & Blake, R. (2007). Brain activity accompanying perception of implied motion in abstract paintings. Spatial
Vision, 20(6), 545–560. https://doi.org/10.1163/156856807782758395
Kirk, U., Skov, M., Christensen, M. S., & Nygaard, N. (2009). Brain correlates of aesthetic expertise: A parametric fMRI
study. Brain and Cognition, 69(2), 306–315. https://doi.org/10.1016/j.bandc.2008.08.004
Kirsch, L. P., Dawson, K., & Cross, E. S. (2015). Additive routes to action learning: How layering experience shapes action
observation network activity. Annals of the New York Academy of Sciences, 1337, 130–139. https://doi.org/10.1111/
nyas.12634
Kirsch, L. P., Urgesi, C., & Cross, E. S. (2016). Shaping and reshaping the aesthetic brain: Emerging perspectives on the
neurobiology of embodied aesthetics. Neuroscience and Biobehavioral Reviews, 62, 56–68. https://doi.org/10.1016/j.
neubiorev.2015.12.005
Kobayashi, Y., Yoshino, A., Ogasawara, T., & Nomura, S. (2002). Topography of evoked potentials associated with illu-
sory motion perception as a motion aftereffect. Brain Research. Cognitive Brain Research, 13(1), 75–84. https://doi.
org/10.1016/S0926-6410(01)00112-4
Kourtzi, Z., & Kanwisher, N. (2000). Activation in human MT/MST by static images with implied motion. Journal of
Cognitive Neuroscience, 12(1), 48–55. https://doi.org/10.1162/08989290051137594
Kozlowski, L. T., & Cutting, J. E. (1977). Recognizing the sex of a walker from a dynamic point-light display. Perception
and Psychophysics, 21(6), 575–580. https://doi.org/10.3758/BF03198740
Krekelberg, B., Vatakis, A., & Kourtzi, Z. (2005). Implied motion from form in the human visual cortex. Journal of Neu-
rophysiology, 94(6), 4373–4386. https://doi.org/10.1152/jn.00690.2005
Kurth, F., Zilles, K., Fox, P. T., Laird, A. R., & Eickhoff, S. B. (2010). A link between the systems: Functional differ-
entiation and integration within the human insula revealed by meta-analysis. Brain Structure and Function, 214(5–6),
519–534. https://doi.org/10.1007/s00429-010-0255-z
Lauring, J. O., Ishizu, T., Kutlikova, H. H., Dörflinger, F., Haugbøl, S., Leder, H., Kupers, R., & Pelowski, M. (2019).
Why would Parkinson’s disease lead to sudden changes in creativity, motivation, or style with visual art?: A review of
case evidence and new neurobiological, contextual, and genetic hypotheses. Neuroscience and Biobehavioral Reviews, 100,
129–165. https://doi.org/10.1016/j.neubiorev.2018.12.016
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Leder, H., Bär, S., & Topolinski, S. (2012). Covert painting simulations influence aesthetic appreciation of artworks. Psy-
chological Science, 23(12), 1479–1481. https://doi.org/10.1177/0956797612452866
Leder, H., Carbon, C. C., & Ripsas, A. L. (2006). Entitling art: Influence of title information on understanding and
appreciation of paintings. Acta Psychologica, 121(2), 176–198. https://doi.org/10.1016/j.actpsy.2005.08.005
Leder, H., Goller, J., Rigotti, T., & Forster, M. (2016). Private and shared taste in art and face appreciation. Frontiers in
Human Neuroscience, 10, 155. https://doi.org/10.3389/fnhum.2016.00155
Leder, H., Ring, A., & Dressler, S. G. (2013). See me, feel me! Aesthetic evaluations of art portraits. Psychology of Aesthet-
ics, Creativity, and the Arts, 7(4), 358–369.  https://doi.org/10.1037/a0033311
Livio, M. (2003). The golden ratio: The story of phi, the world’s most astonishing number. Broadway Books.
Locher, P. J., Smith, J. K.,  & Smith, L. F. (2001). The influence of presentation format and viewer training in the
visual arts on the perception of pictorial and aesthetic qualities of paintings. Perception, 30(4), 449–465. https://doi.
org/10.1068/p3008
Locher, P., Overbeeke, K.,  & Wensveen, S. (2010). Aesthetic interaction: A  framework. Design Issues, 26(2), 70–79.
https://doi.org/10.1162/DESI_a_00017
Loula, F., Prasad, S., Harber, K., & Shiffrar, M. (2005). Recognizing people from their movement. Journal of Experimental
Psychology. Human Perception and Performance, 31(1), 210–220.  https://doi.org/10.1037/0096-1523.31.1.210

190
Movement appreciation

Marinetti, F. T. (1908). The foundation and manifesto of Futurism. In H. B. Chipp (Ed.), Theories of modern art. A source
book by artists and critics (1968) (pp. 284–289). University of California Press.
Martín-Loeches, M., Hernández-Tamames, J. A., Martín, A., & Urrutia, M. (2014). Beauty and ugliness in the bod-
ies and faces of others: An fMRI study of person aesthetic judgement. Neuroscience, 277, 486–497. https://doi.
org/10.1016/j.neuroscience.2014.07.040
Massaro, D., Savazzi, F., Di Dio, C., Freedberg, D., Gallese, V., Gilli, G.,  & Marchetti, A. (2012). When art moves
the eyes: A  behavioral and eye-tracking study. PLOS ONE, 7(5), e37285. https://doi.org/10.1371/journal.pone.
0037285
Mastandrea, S., & Umiltà, M. A. (2016). Futurist art: Motion and aesthetics as a function of title. Frontiers in Human
Neuroscience, 10, 201–209. https://doi.org/10.3389/fnhum.2016.00201
Mayer, S., & Landwehr, J. R. (2018). Quantifying visual aesthetics based on processing fluency theory: Four algorithmic
measures for antecedents of aesthetic preferences. Psychology of Aesthetics, Creativity, and the Arts, 12(4), 399–431.
https://doi.org/10.1037/aca0000187
Millis, K. (2001). Making meaning brings pleasure: The influence of titles on aesthetic experiences. Emotion, 1(3),
320–329. https://doi.org/10.1037/1528-3542.1.3.320
Montepare, J. M.,  & Zebrowitz-McArthur, L. (1988). Impressions of people created by age-related qualities of their
gaits. Journal of Personality and Social Psychology, 55(4), 547–556.  https://doi.org/10.1037//0022-3514.55.4.547
Nadal, M., Munar, E., Marty, G., & Cela-Conde, C. J. (2010). Visual complexity and beauty appreciation. Explaining the
diverging of results. Empirical Studies of the Arts, 28(2), 173–191. https://doi.org/10.2190/EM.28.2.d
Nather, F. C., & Bueno, J. L. O. (2012). Exploration time of static images implying different body movements causes
time distortions. Perceptual and Motor Skills, 115(1), 105–110. https://doi.org/10.2466/27.07.24.PMS.115.4.105-110
Nather, F. C., Fernandes, P. A. M., & Bueno, J. L. O. (2014). Subjective time perception is affected by different durations
of exposure to abstract paintings that represent human movement. Psychology and Neuroscience, 7(3), 381–392. https://
doi.org/10.3922/j.psns.2014.046
Nieto-Castañón, A., & Fedorenko, E. (2012). Subject-specific functional localizers increase sensitivity and functional
resolution of multi-subject analyses.  Neuroimage,  63(3), 1646–1669. https://doi.org/10.1016/j.neuroimage.2012.
06.065
Open Science Collaboration. (2015). Estimating the reproducibility of psychological science. Science, 349(6251), aac4716.
https://doi.org/10.1126/science.aac4716
Orgs, G., Dovern, A., Hagura, N., Haggard, P., Fink, G. R., & Weiss, P. H. (2016). Constructing visual perception of
body movement with the motor cortex. Cerebral Cortex, 26(1), 440–449. https://doi.org/10.1093/cercor/bhv262
Orgs, G., Hagura, N., & Haggard, P. (2013). Learning to like it: Aesthetic perception of bodies, movements and choreo-
graphic structure. Consciousness and Cognition, 22(2), 603–612. https://doi.org/10.1016/j.concog.2013.03.010
Osaka, N., Matsuyoshi, D., Ikeda, T., & Osaka, M. (2010). Implied motion because of instability in Hokusai manga acti-
vates the human motion-sensitive extrastriate visual cortex: An fMRI study of the impact of visual art. NeuroReport,
21(4), 264–267. https://doi.org/10.1097/WNR.0b013e328335b371
Palmer, S. E., & Langlois, T. A. (2017). Effects of implied motion and facing direction on positional preferences in single-
object pictures. Perception, 46(7), 815–829. https://doi.org/10.1177/0301006617694189
Palmer, S. E., & Schloss, K. B. (2010). An ecological valence theory of human color preference. Proceedings of the National
Academy of Sciences of the United States of America, 107(19), 8877–8882. https://doi.org/10.1073/pnas.0906172107
Pavan, A., Cuturi, L. F., Maniglia, M., Casco, C.,  & Campana, G. (2011). Implied motion from static photographs
influences the perceived position of stationary objects. Vision Research, 51(1), 187–194. https://doi.org/10.1016/j.
visres.2010.11.004
Pearce, M. T., Zaidel, D. W., Vartanian, O., Skov, M., Leder, H., Chatterjee, A., & Nadal, M. (2016). Neuroaesthetics:
The cognitive neuroscience of aesthetic experience. Perspectives on Psychological Science, 11(2), 265–279. https://doi.
org/10.1177/1745691615621274
Peelen, M. V., & Downing, P. E. (2007). The neural basis of visual body perception. Nature Reviews. Neuroscience, 8(8),
636–648. https://doi.org/10.1038/nrn2195
Peelen, M. V., Wiggett, A. J., & Downing, P. E. (2006). Patterns of fMRI activity dissociate overlapping functional brain
areas that respond to biological motion. Neuron, 49(6), 815–822. https://doi.org/10.1016/j.neuron.2006.02.004
Pelowski, M., & Akiba, F. (2011). A model of art perception, evaluation and emotion in transformative aesthetic experi-
ence. New Ideas in Psychology, 29(2), 80–97. https://doi.org/10.1016/j.newideapsych.2010.04.001
Pelowski, M., Markey, P. S., Lauring, J. O., & Leder, H. (2016). Visualizing the impact of art: An update and com-
parison of current psychological models of art experience. Frontiers in Human Neuroscience, 10, 1–21. https://doi.
org/10.3389/fnhum.2016.00160

191
Darda, Bara and Cross

Pelowski, M., Oi, M., Liu, T., Meng, S., Saito, G., & Saito, H. (2016). Understand after like, viewer’s delight? A fNIRS
study of order-effect in combined hedonic and cognitive appraisal of art. Acta Psychologica, 170, 127–138. https://doi.
org/10.1016/j.actpsy.2016.06.005
Peuskens, H., Vanrie, J., Verfaillie, K., & Orban, G. A. (2005). Specificity of regions processing biological motion. Euro-
pean Journal of Neuroscience, 21(10), 2864–2875. https://doi.org/10.1111/j.1460-9568.2005.04106.x
Pihko, E., Virtanen, A., Saarinen, V. M., Pannasch, S., Hirvenkari, L., Tossavainen, T., Haapala, A., & Hari, R. (2011).
Experiencing art: The influence of expertise and painting abstraction level. Frontiers in Human Neuroscience, 5, 94.
https://doi.org/10.3389/fnhum.2011.00094
Poldrack, R. A. (2006). Can cognitive processes be inferred from neuroimaging data? Trends in Cognitive Sciences, 10(2),
59–63. https://doi.org/10.1016/j.tics.2005.12.004
Proverbio, A. M., Riva, F., & Zani, A. (2009). Observation of static pictures of dynamic actions enhances the activity of
movement-related brain areas. PLOS ONE, 4(5), e5389. https://doi.org/10.1371/journal.pone.0005389
Pugach, C., Leder, H., & Graham, D. J. (2017). How stable are human aesthetic preferences across the lifespan? Frontiers
in Human Neuroscience, 11, 289. https://doi.org/10.3389/fnhum.2017.00289
Ramsey, R. (2020). Advocating for the credibility revolution. Cognitive Psychology Bulletin, 5. https://doi.org/10.31234/
osf.io/3kwnu.
Reber, R., Schwarz, N.,  & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the per-
ceiver’s processing experience? Personality and Social Psychology Review, 8(4), 364–382. https://doi.org/10.1207/
s15327957pspr0804_3
Redies, C. (2015). Combining universal beauty and cultural context in a unifying model of visual aesthetic experience.
Frontiers in Human Neuroscience, 9, 218. https://doi.org/10.3389/fnhum.2015.00218
Rizzolatti, G., & Craighero, L. (2004). The mirror-neuron system. Annual Review of Neuroscience, 27, 169–192. https://
doi.org/10.1146/annurev.neuro.27.070203.144230
Scharf, A. (1962). Painting, photography, and the image of movement. The Burlington Magazine, 104(710), 186–195.
Seeley, W. P. (2014). Philosophy of art and empirical aesthetics: Resistance and rapprochement. In P. P. Tinio & J. K.
Smith (Eds.), The Cambridge handbook of the psychology of aesthetics and the arts. Cambridge University Press.
Senior, C., Barnes, J., Giampietro, V., Simmons, A., Bullmore, E. T., Brammer, M., & David, A. S. (2000). The func-
tional neuroanatomy of implicit-motion perception or “representational momentum”. Current Biology, 10(1), 16–22.
https://doi.org/10.1016/s0960-9822(99)00259-6
Senior, C., Ward, J., & David, A. S. (2002). Representational momentum and the brain: An investigation into the func-
tional necessity of V5/MT. Visual Cognition, 9(1–2), 81–92. https://doi.org/10.1080/13506280143000331
Silvia, P. J. (2005). Emotional responses to art: From collation and arousal to cognition and emotion. Review of General
Psychology, 9(4), 342–357. https://doi.org/10.1037/1089-2680.9.4.342
Simmons, J. P., Nelson, L. D., & Simonsohn, U. (2011). False-positive psychology: Undisclosed flexibility in data col-
lection and analysis allows presenting anything as significant.  Psychological Science,  22(11), 1359–1366. https://doi.
org/10.1177/0956797611417632
Sitney, A. P. (1993). Landscape in the cinema: The rhythms of the world and the camera. In S. Kemal & I. Gaskell (Eds.),
Landscape, natural beauty and the arts. Cambridge University Press.
Snowden, R. J., Treue, S., Erickson, R. G., & Andersen, R. A. (1991). The response of area MT and V1 neurons to transpar-
ent motion. The Journal of Neuroscience, 11(9), 2768–2785. https://doi.org/10.1523/JNEUROSCI.11-09-02768.1991
Spee, B., Ishizu, T., Leder, H., Mikuni, J., Kawabata, H., & Pelowski, M. (2018). Neuropsychopharmacological aesthet-
ics: A theoretical consideration of pharmacological approaches to causative brain study in aesthetics and art. In Progress
in Brain Research, 237, 343–372. https://doi.org/10.1016/bs.pbr.2018.03.021
Summers, D. (1972). Maniera and movement: The figura serpentine. Art Quarterly, 35, 269–230.
Thakral, P. P., Moo, L. R., & Slotnick, S. D. (2012). A neural mechanism for aesthetic experience. NeuroReport, 23(5),
310–313. https://doi.org/10.1097/WNR.0b013e328351759f
Thompson, J., & Parasuraman, R. (2012). Attention, biological motion, and action recognition. Neuroimage, 59(1), 4–13.
https://doi.org/10.1016/j.neuroimage.2011.05.044
Unger, M. J. (2014). Michelangelo: A life in six masterpieces. Simon & Schuster.
Urgesi, C., Candidi, M., Ionta, S., & Aglioti, S. M. (2007). Representation of body identity and body actions in extras-
triate body area and ventral premotor cortex. Nature Neuroscience, 10(1), 30–31. https://doi.org/10.1038/nn1815
Urgesi, C., Moro, V., Candidi, M., & Aglioti, S. M. (2006). Mapping implied body actions in the human motor sys-
tem. Journal of Neuroscience, 26(30), 7942–7949. https://doi.org/10.1523/JNEUROSCI.1289-06.2006
Uusitalo, L., Simola, J.,  & Kuisma, J. (2009). Perception of abstract and representative visual art. In  Proceedings of the
AIMAC, 10th conference of the international association of arts and cultural management. Dallas, TX.

192
Movement appreciation

Van Geert, E., & Wagemans, J. (2020). Order, complexity, and aesthetic appreciation. Psychology of Aesthetics, Creativity,
and the Arts, 14(2), 135–154. https://doi.org/10.1037/aca0000224
Van Paasschen, J., Bacci, F., & Melcher, D. P. (2015). The influence of art expertise and training on emotion and pref-
erence ratings for representational and abstract artworks. PLOS ONE, 10(8), e0134241. https://doi.org/10.1371/
journal.pone.0134241
Vartanian, O., & Goel, V. (2004). Neuroanatomical correlates of aesthetic preference for paintings. NeuroReport, 15(5),
893–897. https://doi.org/10.1097/00001756-200404090-00032
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov,
M. (2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceedings
of the National Academy of Sciences of the United States of America,  110(2), 10446–10453. https://doi.org/10.1073/
pnas.1301227110
Vartanian, O., & Skov, M. (2014). Neural correlates of viewing paintings: Evidence from a quantitative meta-analysis
of functional magnetic resonance imaging data.  Brain and Cognition,  87, 52–56. https://doi.org/10.1016/j.bandc.
2014.03.004
Vazire, S. (2018). Implications of the credibility revolution for productivity, creativity, and progress. Perspectives on Psycho-
logical Science, 13(4), 411–417. https://doi.org/10.1177/1745691617751884
Vessel, E. A., Isik, A. I., Belfi, A. M., Stahl, J. L., & Starr, G. G. (2019). The default-mode network represents aesthetic
appeal that generalizes across visual domains. Proceedings of the National Academy of Sciences of the United States of Amer-
ica, 116(38), 19155–19164. https://doi.org/10.1073/pnas.1902650116
Vessel, E. A., Maurer, N., Denker, A. H., & Starr, G. G. (2018). Stronger shared taste for natural aesthetic domains than
for artifacts of human culture. Cognition, 179, 121–131. https://doi.org/10.1016/j.cognition.2018.06.009
Vessel, E. A., & Rubin, N. (2010). Beauty and the beholder: Highly individual taste for abstract, but not real-world
images. Journal of Vision, 10(2), 1–14. https://doi.org/10.1167/10.2.18
Vessel, E. A., Starr, G. G., & Rubin, N. (2012). The brain on art: Intense aesthetic experience activates the default mode
network. Frontiers in Human Neuroscience, 6, 66. https://doi.org/10.3389/fnhum.2012.00066
Voland, E., & Grammer, K. (Eds.). (2003). Evolutionary aesthetics. Springer-Verlag.
Watson, J. D., Myers, R., Frackowiak, R. S., Hajnal, J. V., Woods, R. P., Mazziotta, J. C., Shipp, S., & Zeki, S. (1993).
Area V5 of the human brain: Evidence from a combined study using positron emission tomography and magnetic
resonance imaging. Cerebral Cortex, 3(2), 79–94. https://doi.org/10.1093/cercor/3.2.79
Winawer, J., Huk, A. C., & Boroditsky, L. (2010). A motion aftereffect from visual imagery of motion. Cognition, 114(2),
276–284. https://doi.org/10.1016/j.cognition.2009.09.010
Wölfflin, H. (1942/2012). Principles of art history. Dover Publications.
Yue, X., Vessel, E. A.,  & Biederman, I. (2007). The neural basis of scene preferences. NeuroReport, 18(6), 525–529.
https://doi.org/10.1097/WNR.0b013e328091c1f9
Zeki, S., & Stutters, J. (2012). A brain-derived metric for preferred kinetic stimuli. Open Biology, 2(2), 120001. https://
doi.org/10.1098/rsob.120001
Zeki, S., Watson, J. D., & Frackowiak, R. S. (1993). Going beyond the information given: The relation of illusory visual
motion to brain activity. Proceedings of the Royal Society of London. Series B: Biological Sciences, 252(1335), 215–222.
https://doi.org/10.1098/rspb.1993.0068
Zeki, S., Watson, J. D., Lueck, C. J., Friston, K. J., Kennard, C., & Frackowiak, R. S. (1991). A direct demonstration
of functional specialization in human visual cortex. Journal of Neuroscience, 11(3), 641–649. https://doi.org/10.1523/
JNEUROSCI.11-03-00641.1991

193
10
HOW ARCHITECTURAL DESIGN
INFLUENCES EMOTIONS,
PHYSIOLOGY, AND BEHAVIOR
Alex Coburn, Adam Weinberger and Anjan Chatterjee

Over the past decade, a new wave of research has emerged investigating the impact of architectural design on
human emotions, physiology, and behavior. This research, termed the neuroscience of architecture, leverages
evidence-based methods to explore how the human brain interacts with the built environment and to iden-
tify key features of buildings and neighborhoods that promote wellness and human flourishing. We begin this
chapter by outlining the historical context that motivated this field, including the importance of aesthetics in
vernacular architecture, the devaluation of humanistic design principles in mid-20th century construction,
and the birth of environmental psychology and evidence-based design movements in response to post-war
architectural mass production. We then review empirical findings that have emerged from research on the
neuroscience of architecture, including investigations of aesthetic responses to architectural design features;
the potential benefits of biophilic design for stress reduction, health, and wellness; and the impact of sensory
features on movement and navigation. Finally, we discuss how the probing of specific dimensions of emo-
tional experience in the built environment—including fascination, coherence, and hominess—can promote
more evidence-based design practices and foster more human-centered buildings.
People often spend most of their lives inside buildings (Evans, 2003). The design of the built environment
has a profound impact on people’s mental states and sense of well-being. Here, we review empirical research
that links these impacts to brains and behavior. Design can affect how comfortable (Baker  & Standeven,
1995; Brager et al., 2004) or focused (Mehta & Zhu, 2009) people feel and modulates excitement, fear, and
awe (Coburn et al., 2020); feelings of trust (Zhang et al., 2014); stress-related hormonal patterns (Fich et al.,
2014a; Küller & Lindsten, 1992; Tyrväinen et al., 2014; Valtchanov et al., 2010); and even complex social
behaviors such as criminality (Kotabe, 2016). Over time, repeated exposures to specific environments may
also affect health outcomes such as speed of recovery from surgery (Ulrich, 1984) and cardiac health (Kardan,
Gozdyra et al., 2015). These claims come from findings in several fields, including environmental psychol-
ogy, cognitive neuroscience, and neuroaesthetics (Coburn et al., 2017). Researchers try to define key aspects
of psychological experience that are affected by the built environment and identify important architectural
variables and design strategies that enhance wellness on a large scale.
A challenge in this approach is the difficulty of operationalizing relevant features of the built environment
(Cooper, 2014). Past studies have often assessed the psychological impact of easily quantifiable environmen-
tal variables, such as type of ventilation (i.e., mechanical vs. natural), number of windows in a room, and
intensity of ambient noise in a building (Hygge, 2003; Johnson, 2000; Lercher et al., 2003). These variables
can be operationalized and lend themselves easily to scientific inquiry. However, such isolated measures are

194 DOI: 10.4324/9781003008675-11


Architectural design and the brain

often too simple to capture complex real-world environments that people actually experience (Alexander,
2002). The limited external validity of these measures may be why evidence that emerges from this literature
is often contradictory (Cooper & Burton, 2014).
Aesthetic qualities of the built environment may predict psychological experience and overall wellness
better than any single design variable measured in isolation (Adams, 2014; Brown, 2014; Ellaway, 2014;
Kyttä et  al., 2011; Kyttä  & Broberg, 2014). Environments that are perceived as attractive and high quality,
for example, are consistently associated with positive mental health outcomes (Ellaway et al., 2005; Evans,
2003; Kyttä et al., 2011). Several studies report strong associations between the perceived attractiveness of
residential neighborhoods and self-reports of health, quality of life, and wellness after controlling for socio-
economic and demographic factors (Coles, 2014; Kyttä et al., 2011). In a large-scale study in Finland, per-
ceived environmental quality was strongly linked to perceived happiness, health, and quality of life, whereas
no significant associations were found between quantifiable environmental measures and wellness (Kyttä &
Broberg, 2014). This research underscores the importance of moving beyond simple quantitative measures
of architectural spaces and developing more nuanced ways of defining and operationalizing aesthetic qualities
of the built environment (Cooper & Burton, 2014).
To give a concrete example of an “aesthetic quality,” one promising category relates to naturalness. Natural
environments, as well as naturalistic features of the built environment, promote wellness across a wide range
of contexts and populations. The cognitive benefits of exposure to natural spaces include improving mood
(Barton & Pretty, 2010; Bowler et al., 2010; Valtchanov et al., 2010), reducing stress (Valtchanov et al., 2010;
Villani & Riva, 2011), heightening concentration and working memory (Berman et al., 2008, 2012; Berto,
2005; Bratman, Daily et al., 2015; Bratman et al., 2012; Bratman, Hamilton et al., 2015; Kaplan, 1995),
increasing self-esteem (Barton & Pretty, 2010; Pretty et al., 2005), enhancing vitality and energy (R. M.
Ryan & Huta, 2010), and increasing self-perceptions of health (Kardan, Gozdyra et al., 2015). Buildings and
rooms that offer views of nature are also linked to reduced criminality in residential neighborhoods (Kuo &
Sullivan, 2001), faster recovery from surgery in hospitals (Ulrich, 1984), more charitable giving behaviors,
and higher levels of trust (Zhang et al., 2014). Simply looking at images of nature and virtual representations
of natural landscapes may promote many of these benefits (Berman et al., 2008; Berto, 2005; Valtchanov
et al., 2010; Valtchanov & Ellard, 2015).
Although the natural environment is often framed as a distinct category of space from the built environ-
ment (Kaplan, 1995), many buildings also contain natural visual and sensory features. Examples of these fea-
tures include actual vegetation (i.e., plants, trees, water), symbolic references to nature, and abstracted natural
patterns such as fractal scaling (Capo, 2004; Goldberger, 1996; Hagerhall et al., 2004), color contrast, and
high density of curved edges in a scene (Coburn et al., 2019a; Ibarra et al., 2017; Kardan, Demiralp et al.,
2015).1 The incorporation of natural elements and patterns into the built environment is often referred to
as biophilic design ( Joye, 2007a; Kellert, 2003, 2015; Ryan et al., 2014). Biophilic design is associated with
improved mood and cognitive functioning and may confer similar psychological benefits as interacting with
natural landscapes (Coburn et al., 2019a; Kotabe, 2016; Kotabe et al., 2017, 2016; Lavdas & Schirpke, 2020a;
Salingaros, 2020a, 2020b). Specific aesthetic qualities of the built environment, including perceived beauty
and naturalness, can therefore enhance an individual’s psychological experience within an architectural space
and may contribute to better mental health and wellness.
The idea that aesthetic qualities of the built environment can impact wellness is not new. For millennia,
civilizations across the globe sought to understand how to optimize designs of buildings and landscapes in
order to improve social, functional, and spiritual aspects of the human experience. From ancient Rome to
Imperial China, cultures around the world developed sophisticated aesthetic rules to guide the construction
of buildings, neighborhoods, and cities, motivated by the belief that these aesthetic principles are as much
as form of science as they are a form of art (Coburn et al., 2017; Mak & Thomas Ng, 2005; Patra, 2009;
Vitruvius Pollio et al., 1914).

195
Coburn, Weinberger and Chatterjee

However, around the middle of the 20th century, a seismic shift occurred in Western culture that led to
the widespread rejection of these humanistic principles of construction in favor of a new set of design rules in
which measurable variables such as cost, speed, and efficiency were prioritized above less easily quantifiable
factors, including aesthetics and occupant experience. Although this shift achieved some progress in urban
development, it also brought about unintended social consequences ( Jacobs, 1992). The evidence-based
design movements of the 1970s arose in response to mid-century architectural mass production and paved
the way for research in urban sociology and environmental psychology. This early scholarship exploring
the effects of architectural design on social behavior and psychological experience laid the groundwork for
recent research on the neuroscience of architecture.
In the following section, we outline the historical context for the rise of research in the psychology and
neuroscience of architecture. We then summarize the key evidence that has emerged to date. Finally, we dis-
cuss the prospects and limitations of translating evidence-based research into environmental design strategies
and architectural practice and discuss the future directions of this emerging area of study.

Historical context

Aesthetics and vernacular architecture


For most of human history, architectural design and aesthetics developed hand in hand. From medieval
Iran, to ancient Japan, to indigenous North American cultures, the construction of vernacular buildings
was closely linked with advanced aesthetic principles of architectural design. Among the best known sets of
design principles are the Chinese feng shui and the Indian vaastu shastra (Mak & Thomas Ng, 2005; Patra,
2009). These design-oriented philosophies sought to create harmony and comfort in the built environ-
ment using wisdom derived from empirical observation. These complex systems of environmental aesthetics
emerged gradually, over the course of centuries, by integrating trial and error, spiritual reflection, and intel-
lectual scholarship. In many cases, the buildings generated by these design principles remain among the most
important cultural artifacts of the civilizations that built them.
Aesthetics was also integral to the development of Western European architecture. Vitruvius, the influen-
tial Roman architect and writer, declared that successful construction relies on three closely related princi-
ples: functionality (utilitas), structural integrity (firmitas), and beauty (venustas) (Vitruvius Pollio et al., 1914).
These three principles, known as the Vitruvian triad, not only reflected a design philosophy of the ancient
world but also served as the foundation for the architectural philosophy of Renaissance Europe. Indeed, the
Renaissance not only witnessed significant advances in structural engineering (firmitas) and efficient urban
planning (utilitas), it also brought about remarkable achievements in aesthetics (venustas) of the Vitruvian
triad. Humanistic principles of architectural aesthetics emerged through the prolific drawings and writings of
great Italian architects like Palladio, Michelangelo, and Alberti. At the same time, novel mathematical con-
cepts, such as the Golden Ratio, influenced the design of buildings large and small across Western Europe,
North Africa, and the Middle East. To the extent that the iconic buildings produced by these cultures were
a reflection of skilled engineering, they were just as much a product of sophisticated advances in the science
of architectural aesthetics.

Industrialization and the modernist aesthetic


In the mid-20th century, the dimension of aesthetics was forcefully rejected. Following the second Industrial
Revolution, advocates of the International Style, including Mies van der Rohe and Le Corbusier, spread
a new architectural gospel: buildings should be designed as “machines” for living and working (Corbusier,
1927). This new philosophy was based on the idea that “form follows function.” It declared that a building’s

196
Architectural design and the brain

design and appearance should no longer be shaped by humanistic principles of comfort, aesthetics, and
harmony but should instead conform to industrial values like cost, speed, density, and efficiency (Rosner,
2020). As part of this new way of thinking about buildings, the founders of Modernism rejected architectural
ornament and embraced new mottoes like “less is more” and “death to decoration” ( Johnson, 1947). In a
similar vein, these architects abandoned basic tenets of proportion and human scaling that had for centuries
guided human construction around the world (Alexander, 1979, 2002). In this way, the Vitruvian principle
of venustas was turned on its head, and architectural beauty was reformulated to signify the manifestation of
functionalist design.
This shift in thinking led to the widespread adoption of a new and highly specific aesthetic ideal. This
ideal embraced minimalist, reductive forms and celebrated materials that conformed easily to these shapes,
such as glass, steel, and concrete (Rosner, 2020). This movement spread internationally after the Second
World War, and its rapid proliferation was inseparable from the economic context out of which it was born.
This approach to architectural design depended on the availability of mass-produced, synthetic materials and
large-scale manufacturing (Murphy, 2012). Rapid architectural fabrication was promoted by industrial lead-
ers, who profited financially from its rise, and by political leaders, who embraced mass-produced buildings
as a large-scale and low-cost solution to post-war reconstruction.
Important social factors also contributed to the international proliferation of the neo-Modernist aesthetic.
Whereas design principles of vernacular architectural traditions developed gradually, over centuries and across
many different cultures, the aesthetic ideals of neo-Modernism were invented in the 1930s by an elite group
of white European men ( Johnson, 1947). In the following decades, these ideals were institutionalized by city
governments, urban planners, and real estate developers wherever low-cost property and perceived poverty
or social unrest could be found. In the United States, this paternalistic sensibility led to the demolition of
entire neighborhoods and privately owned homes inhabited largely by black and brown communities and the
construction of massive, monolithic apartment towers in their place (Austen, 2018).2 The application of neo-
Modernist design principles to these “urban renewal” projects generated substantial revenue for developers
and municipal governments by increasing the density of taxpayers and renters occupying each square foot of
real estate. The tower-block model also facilitated the geographic segregation of black and white communi-
ties into distinct buildings and neighborhoods as part of a larger set of systemically racist American housing
policies (Rainwater, 2006). The tower-block model also gained popularity in European cities as a means of
class-based, rather than race-based, segregation. In Copenhagen and Amsterdam, for instance, mass-produced
housing was used by municipal governments to move working-class people out of downtown neighborhoods
and into suburban high-rise apartments (Helleman & Wassenberg, 2004). Across the Western world, the neo-
Modernist aesthetic therefore served as an effective tool for both financial gain and social control.
The rapid spread of these aesthetic principles contributed to the international standardization of archi-
tectural form (Alexander, 2002). Supported by powerful socioeconomic forces, the neo-Modernist doctrine
pushed competing philosophies of design to the side and morphed into an international monopoly of archi-
tectural ideology (Salingaros, 2007). The spread of these principles was not driven by their grassroots popu-
larity or demonstrated social benefit. Instead, it was based on their ability to serve the financial and political
interests of a powerful minority of stakeholders. To be clear, this was not the intention of the Bauhaus archi-
tects who invented Modernism. They saw their movement as a force for equity and social progress. However,
these a priori theories were not tested in the real world, and the evidence supporting them amounted to little
more than idealistic conjecture.

Urban sociology and evidence-based design


Starting in the 1960s, the principles of Modernism began to face empirical scrutiny. In The Death and Life of
Great American Cities, Jane Jacobs (1992) documented how urban renewal projects were impoverishing the

197
Coburn, Weinberger and Chatterjee

lives of individuals and eroding the vibrancy of urban communities. This research challenged the founda-
tional tenants of neo-Modernist urbanist theory and led to grassroots political efforts opposing a number of
significant urban design projects. Among other projects, these efforts prevented the construction of a planned
expressway and real estate development in lower Manhattan that would have destroyed large swaths of Soho
and Little Italy (Flint, 2009).
Christopher Alexander, a professor of architecture at Berkeley, later embarked on a decade-long research
project investigating the social and psychological effects of different “pattern languages” of architectural
design. Using case studies of buildings and neighborhoods from around the world, he developed a system
of empirically derived design principles to improve human experience in the built environment (Alexander,
1979). Alexander employed these principles in several international projects, including the University of
Oregon campus design, residential houses in northern Mexico (Alexander et al., 1985), and Eishin School in
Japan (Alexander et al., 2012). These empirically driven research and design efforts served as the foundation
for evidence-based architectural practices like New Urbanism, which gained traction in the 1970s and 1980s
and challenged the previous aesthetic dogma of neo-Modernism (Mehaffy, 2017).3

Environmental psychology and neuroscience


In the early 1980s, research arose in the emerging field of environmental psychology, including the Kaplans’
groundbreaking work on evolutionary and environmental aesthetics (Kaplan, 1973; Kaplan & Kaplan, 1989;
Kaplan, 1987) and Ulrich’s seminal studies linking the aesthetics of healthcare spaces to health and wellness
(Ulrich, 1977, 1983, 1984; Ulrich et al., 1991). Harvard biologist E.O. Wilson published his foundational
biophilia hypothesis around this time (1984), which outlined novel evolutionary arguments illustrating the
psychological and spiritual connection between humans and natural life forms and landscapes. In collabora-
tion with Yale professor Stephen Kellert, Wilson later applied these broad ideas to architecture and urban
design (Kellert, 2003, 2005; Wilson & Kellert, 1995) and inspired biophilic architecture, a growing movement
of evidence-based design practices (Berto & Barbiero, 2017; Joye, 2007a; Kellert, 2015; Ryan et al., 2014;
Salingaros, 2020a).
In the early 2000s, this foundational literature on evolutionary and environmental psychology motivated
a new wave of research exploring links between architectural design and the human mind and brain (Eber-
hard, 2008, 2009, 2004; Eberhard & Patoine, 2004; Whitelaw, 2013). A new field of research thus emerged,
often referred to as the neuroscience of architecture. This field draws from several scientific disciplines, including
experimental psychology, empirical aesthetics, and cognitive neuroscience (Coburn et al., 2017). The pur-
pose of this field is to investigate the neurobiological underpinnings of architectural and aesthetic experience
and to measure the effects of architectural design on human experience using a combination of behavioral
studies, brain imaging, and other physiological biomarkers (e.g., tracking eye movements, measuring stress
hormones). Whereas previous research in environmental psychology was largely limited to self-report sur-
veys and observations of human behavior, the neuroscience of architecture has enabled researchers to move
beyond descriptive and observational measures in order to understand the neural structure and cognitive
mechanisms that mediate aesthetic experiences of the built environment. With this historical context in
mind, we offer a more detailed review of this nascent field of research.

Neuroscience of architecture
The neuroscience of architecture has begun to advance knowledge about how and why specific architectural
features affect people in specific ways. This section will begin by reviewing neuroscientific research that has
shown how properties of the built environment influence aesthetic experiences and instantiate psychological

198
Architectural design and the brain

and emotional responses. This research is placed within the theoretical framework of the aesthetic triad, a neu-
roscientific model originally developed for neuroaesthetics (Chatterjee & Vartanian, 2014), which was then
reformulated for the neuroscience of architecture (Coburn et al., 2017). As part of this discussion, we will
also consider how individual-level characteristics moderate responses to the built environment, recognizing
that some aspects of architectural experience vary from person to person, while others are more stable across
most people. The design of built environments can also have a meaningful impact on mood and health.
Consequently, this section concludes by highlighting important connections between architectural design
and wellness. Research on how built environments influence movement behavior is discussed in detail in
Chapter 17.
At the outset, it is important distinguish between descriptive and experimental neuroscientific claims.
The former uses observational knowledge about brain functioning to qualitatively map the biological or
cognitive nature of an aesthetic experience and to potentially develop testable theories and hypotheses. For
example, some neurons in primary visual cortex respond preferentially to edges and high visual contrast
(Brady & Field, 2000; Ramachandran & Hirstein, 1999). Thus, neuroscience has descriptively demonstrated
that the brain is sensitive to certain visual features. Experimental neuroscience, by contrast, directly tests
hypotheses, makes predictions, and yields quantitative data. For instance, Vartanian and colleagues (2013)
identified increased anterior cingulate activation when participants made judgments of interior spaces dur-
ing functional magnetic resonance imaging (fMRI), experimentally demonstrating the involvement of emo-
tion and reward pathways. Whereas much of the early research on the neuroscience of architecture took a
descriptive approach (see Barbara & Perliss, 2006; Eberhard 2004, 2008, 2009; Eberhard & Patoine, 2004),
research over the past decade has shifted towards using experimental methods (see Coburn et al., 2019, 2020;
Lavdas & Schirpke, 2020; Vartanian et al., 2013, 2015). Recognizing the value of both types of research, we
will return to this distinction at various points throughout this chapter.

The aesthetic triad


We developed a brain-based model of architectural experience (Coburn et al., 2017) by applying the aesthetic
triad (Chatterjee & Vartanian, 2014) to architecture. Using this model, we proposed that human–building
interactions are mediated by three large-scale networks in the brain: the sensory-motor, knowledge-
meaning, and emotion-valuation systems (Figure  10.1). The sensory-motor system addresses “bottom-up”
processing of the features of buildings, including visual (color, shape, size, materiality), as well as acoustic,
tactile, and even olfactory features of the built environment. The brain’s knowledge-meaning circuitry plays
an important role in mediating “top-down” processing of architectural environments. The brain’s baseline
response to a building’s sensory features may be either dampened or enhanced by an individual’s cultural
background, identity, and education, as well as their knowledge about a space. Finally, the emotion-valuation
system integrates information from the sensory-motor and knowledge-meaning systems, leading to aes-
thetic experiences. These experiences may range from profound feelings of joy or delight to interest or
even fear and disgust. We postulated that these three systems interact closely to create a holistic sense of
architecture.
In the following discussion, we review research on the psychology and neuroscience of architecture using
the aesthetic triad as an organizing framework. First, we outline key sensory features of architectural spaces
that influence psychological and neural responses across a wide range of building typologies, with a particular
emphasis on biophilic design. Next, we discuss emotional responses to architectural spaces and examine the
neural pathways that mediate different emotional experiences in the built environment. Finally, we discuss
key dimensions of the knowledge-meaning neural systems that contribute to individual differences in certain
aspects of architectural experience.

199
Coburn, Weinberger and Chatterjee

Figure 10.1 The Aesthetic Triad proposes that the human brain processes architectural environments via interac-
tions between sensory-motor, emotion-valuation, and knowledge-meaning neural systems. Chatterjee &
­Vartanian, 2014. Reprinted with permission from Trends in Cognitive Sciences.

Sensory perception of the built environment


We begin with a discussion of the sensory features of architectural design that are most closely linked to
positive aesthetic experiences, as well as the mechanisms of human perception that mediate these experi-
ences. A logical starting point is the perception of nature and naturalistic features of the built environment
(i.e., biophilic architecture), given the wealth of research related to this topic. The natural environment has
served as one the most important sources of architectural inspiration throughout history. The Hanging Gar-
dens of Babylon—one of the Seven Wonders of the Ancient World—was characterized by rich greenery.
Gardens were featured prominently in ancient Egyptian and Chinese societies. In the modern era, housing
costs fluctuate based on proximity to recreational natural spaces (Crompton, 2001), and vacationers spend
more money for rooms with ocean views (Fleischer, 2012). These trends reflect a long-standing desire to
maintain contact with nature while being in the built environment (Ulrich, 1993).
It comes as no surprise, then, that some of the most prominent neurobiological accounts of architecture
are concerned with natural qualities and aesthetics. For example, consistent with the biophilia hypothesis,
empirical evidence suggests a widespread human preference for natural environments compared to urban
spaces (Berman et  al., 2008; Kaplan, 1995), although recent work suggests individuals’ preferences may
gradually emerge over the course of development (Meidenbauer et al., 2019). People show a strong ten-
dency to interact with nature for recreation and relaxation, a preference observed globally and across cultures
(Chang et al., 2020). Interactions with nature are also salubrious, producing positive effects on physiological,
emotional, and cognitive functioning ( Joye, 2007b).
Kaplan’s attention restoration theory (ART; Kaplan, 1995) argues that the ameliorative effects of nature
stem from alterations to neurocognitive processing—specifically, an improved ability to focus. ART draws
from William James’ distinction between involuntary attention—which is automatically captured by sur-
rounding stimuli—and directed forms of attention that rely on cognitive control ( James, 1985), the latter
of which is effortful but replenished by interactions with nature. More precisely, ART indicates that natural
settings reduce the burden otherwise placed on effortful attention mechanisms (i.e., in an urban environ-
ment), thereby allowing directed-attention mechanisms to restore (because there is nothing demanding

200
Architectural design and the brain

them; Berman et al., 2008; Joye, 2007; Kaplan, 1995). Crucially, this restoration occurs independently of
emotional changes or aesthetic preferences (Kaplan, 1995; Meidenbauer et al., 2020). Consistent with ART,
experimental research indicates a modest improvement in attention after exposure to natural environments
(e.g., Berman et al., 2008; Bowler et al., 2010), although questions remain about individual differences and
proper control conditions.
Several evolutionarily based accounts argue that preferences for natural spaces are affectively and aestheti-
cally driven. Ulrich’s psychoevolutionary stress-reduction framework (1993) posits that, in our evolution-
ary history, humans were frequently confronted by threatening stimuli, leading to a rapid cortisol response,
which persisted until the setting became unthreatening. Unthreatening settings were typically open, calm,
and warm. Present-day environments rich in these attributes (among others) may reduce stress and improve
affect. The architect Don Ruggles emphasizes the importance of designing buildings that increase the base-
line tone of the parasympathetic autonomic nervous system as a target to reduce stress (Ruggles & Boak,
2020). Similarly, the “prospect-refuge” theory (Dosen & Ostwald, 2016) argues that environments which
are both open (i.e., prospect) and convey feelings of safety (i.e., refuge) were evolutionarily beneficial and,
therefore, remain aesthetically preferred. Although empirical evidence demonstrating one-to-one association
between specific natural features (e.g., “openness”) and stress reduction is lacking ( Joye & De Bloack, 2011),
a wealth of research reports that exposure to nature is associated with reduced stress and negative emotions,
such as anger and sadness (Bowler et al., 2010).
The neuroscience of architecture also draws from theories of aesthetic experience more broadly. In
particular, Rolf Reber’s theory of “processing fluency” proposes that aesthetic experience fluctuates based
on how efficiently an observer can processes the properties of an object. Objects—for example, built
environments—are experienced as pleasurable if they contain some complexity but can still be fluently
processed. In this way, processing fluency can “bridge the gap” between cognitive (e.g., ART) and affective
(e.g., stress-reduction) frameworks, as the theory describes both a reduction in cognitive resources (consist-
ent with attention restoration theory) and hedonic value associated with the condition (in line with stress-
reduction frameworks).
Descriptively, an extensive body of neuroscience research has demonstrated sensitivity to low-level stimu-
lus features. For example, brains are especially attuned to edges and high contrast (Brady  & Field, 2000;
Geisler, 2008), which likely evolved to support object identification. Other low-level features such as lumi-
nance, color, and motion are initially processed in primary sensory areas (Chatterjee, 2003), before being
processed in higher-level regions, such as the parahippocampal place area (PPA)—which responds prefer-
entially to scenes and buildings (Epstein & Kanwisher, 1998)—as well as the hippocampal and entorhinal
cortices that are crucial for spatial navigation (Spiers & Barry, 2015). Other research indicates sensitivity to
visual symmetry (Bertamini et al., 2019; Gartus & Leder, 2013; Rhodes et al., 1998).
These features are fluently processed and, therefore, plausibly contribute to aesthetic experience. Scenes
with repeating low-level features like those mentioned previously are judged as more fascinating and coher-
ent ( Joye & van den Berg, 2011), putatively because of easy recognition and error-free processing (Clark,
2013; Reber et al., 2004). These sensitivities could explain aesthetic appreciation for rhythmic architectural
designs such as alternating columns or color patterns in stained glass windows (Alexander, 2004; Coburn
et al., 2017). Natural environments, specifically, are characterized by a host of reoccurring low-level features,
including non-straight edges, low color saturation (Berman et al., 2014), and greater contrast (Coburn et al.,
2019). The prevalence of predictable low-level features may explain why humans are able to process natural
scenes more rapidly than human-made structures (Greene & Oliva, 2009; Rousselet et al., 2005).
Given this preference for nature—that is, biophilia—many designers have sought to bring a natural aes-
thetic into the built environment. Many global design initiatives encourage the use of natural elements in
architecture (Living-Future.Org, 2020; International WELL Building Institute, 2020). One way to do this is to
incorporate nature directly into the built environment. This use of natural elements could be as straightforward

201
Coburn, Weinberger and Chatterjee

as adding plants, water features, or small fires ( Joye, 2007b). Drawing on the ideas of prospect-refuge, archi-
tects can also incorporate large windows or balconies that provide extensive vistas of the outdoors. Even
more simply, people can arrange pictures or photographs of the outdoors around the home.
These approaches, however, may be overly simple. For one, inhabitants of such spaces may not judge
the water, plants, or pictures as being natural; they are aware that these elements were inserted into a built
space. Furthermore, merely placing an otherwise unnatural structure into a natural landscape is also unlikely
to meet biophilic ambitions; if there is nothing inherently natural about the structure itself, buildings may
integrate poorly with the surrounding environment and actually detract from the cognitive and affective
benefits of an otherwise natural setting.
An alternative approach is to incorporate low-level features or patterns that occur frequently in nature into
the design of human-made structures and spaces. Indeed, philosophers and artists have long posited that peo-
ple are innately drawn to human-made objects that echo organic, natural qualities. In one study, participants
who evaluated 240 interior and exterior architectural scenes based on perceived naturalness and preference
were found to exhibit strong preferences for buildings that contained high densities of naturalistic visual pat-
terns, such as edge density and contrast (measured quantitatively using image statistics). Notably, these two
nature-like patterns explained the most of the variance in preference ratings of both architectural facades
and interior scenes, even after controlling for the quantity of actual vegetation that was visible in each scene
(Coburn et al., 2019). These findings suggested that the degree of implicit naturalness perceived in an archi-
tectural environment might be just as important as (and possibly independent from) the amount of explicit
nature (i.e., water, plants, trees) conveyed in that scene.
This study builds on previous research showing that rooms with properties typical of the natural environ-
ment (e.g., non-straight edges) are preferred to spaces with unnatural features (e.g., straight edges; Vartanian
et al., 2013). The salubrious effects of interacting with nature may stem largely from these preferences; in
one recent study, when participants were presented with equally preferred urban and natural images, no
differences in affective state were observed (Meidenbauer et al., 2020). That is, people experienced positive
emotions in natural settings because of the prevalence of preferred visual inputs rather than because of unique
qualities of nature itself. Thus, low-level features characteristic of natural environments may evoke positive
affect and aesthetic appreciation if they are incorporated into the built environment, even without the pres-
ence of explicitly natural elements.
Crucially, low-level stimuli can elicit neural responses that characterize the “whole.” For instance, brains
contain a powerful face detection mechanism—instantiated in the fusiform face area (FFA)—that responds
strongly to faces. Yet the region not only responds to actual human faces but also to basic, low-level stimulus
features that make up a face (Yue et al., 2011). The FFA is activated by sparse schematic representations of
a face (i.e., a smiley face) or even the front of a car (Kühn et al., 2014; Windhager et al., 2010). Moreover,
these low-level features can induce emotional responses similar to these evoked by real human faces (Aiken,
1998; Joye, 2007b).
Analogous results have been experimentally demonstrated for scene perception. In one study (Kravitz
et al., 2011), participants viewed a series of built and natural spaces during fMRI. Results indicated that the
primary factor that influenced PPA activation was not whether the space was natural or human-made but
rather the extent of openness conveyed in each image. That is, the PPA was sensitive to a particular feature
(i.e., openness of the space) regardless of image categorization (i.e., natural vs. built). Viewing open spaces is
also associated with activation of temporal lobe structures sensitive to visual motion (Vartanian et al., 2015),
suggesting a connection between openness and a desire to move in space (Coburn et al., 2017). Consistent
with this interpretation, open interior spaces are rated as more natural (Coburn et al., 2019) and beautiful
(Vartanian et al., 2015) and are preferred over enclosed environments (Dosen & Ostwald, 2016).
Other work indicates that binocular eye movements may be specifically attuned to statistical regularities in
the natural environment (Gibaldi & Banks, 2019), further suggesting that humans are highly sensitive to the

202
Architectural design and the brain

occurrence of certain low-level features prevalent in nature. Similarly, aesthetic preferences for scenes with
a high density of curved edges and color contrast have been demonstrated in several studies and across mul-
tiple environmental contexts, including natural landscapes (Berman et al., 2014; Ibarra et al., 2017; Kardan,
Demiralp et al., 2015), architectural facades (Coburn et al., 2019a; Ibarra et al., 2017; Kardan, Demiralp
et al., 2015; Lavdas & Schirpke, 2020b), and building interiors (Coburn et al., 2019a, 2020; Vartanian et al.,
2013). Taken together, these findings point to the powerful influence of low-level features on aesthetic
experience and further indicate that built environments rich with such features found in nature may trigger
biophilic responses.
A key low-level feature that influences aesthetic experience is fractal scaling (i.e., “fractals”). Fractals,
a hierarchy of self-similar patterns that repeat at increasingly fine sizes, provide a sense of “organized
complexity.” Fractals are common in nature; clouds, trees, plants, waves, fire, lightning, and mountains are
all composed of repeating patterned elements.4 When incorporated into the built environment, fractals
evoke feelings of naturalness and are consistently preferred compared to non-fractal design ( Joye, 2007b;
Lavdas & Schirpke, 2020b; Taylor, 2021). Historically, fractal design has been prevalent across many civi-
lizations and architectural practices. As detailed by Richard Taylor (2021), fractals are seen in traditional
African settlements, 8th-century temples, 13th-century castles of the Holy Roman Empire, Gothic-period
cathedrals, Buddhist temples, Islamic minarets, Gaudi’s Sagrada Familia, and the organic houses of Frank
Lloyd Wright. More recent initiatives have incorporated fractals into the design of floors, carpets, walls,
solar panels, and window shades. Several contemporary architecture firms have also incorporated frac-
tal scaling into their design practices (e.g., Light Earth Designs, Center for Environmental Structure),
although non-fractal design practices are more prominent in contemporary Western architecture ( Joye,
2006; Salingaros, 2007).
Neuroscience research—both descriptive and experimental—provides insight into the historical appeal of
fractals. First, it is clear that the visual system is proficient at grouping together repeating elements (Biederman,
1987; Reber et al., 2004) and process fractals automatically and fluently (Spehar et al., 2015). This is true
for humans as well as primates (Finn et al., 2019). Moreover, neurons in the primary visual cortex appear to
show a preference for fractals, suggesting that fractals may play a crucial role in adapting the visual system to
the natural environment (Yu et al., 2005). Consistent with these findings, patients with neurological damage
to higher-order visual processing regions do not show any deficits in fractal gaze dynamics (Marlow et al.,
2015). Fractals may also partially explain associations between nature (which is often rich in fractal patterns)
and stress reduction ( Joye, 2007b); results using EEG report heightened attentional states when people view
naturally patterned fractals (Hagerhall et al., 2008). All of these findings are broadly consistent with attention
restoration theory, stress reduction, and processing fluency perspectives.
It is also evident that higher-level visual and semantic features of an environment influence aesthetic expe-
rience. In a recent fMRI study, researchers identified decodable neural representations of architectural styles
and buildings in high-level visual regions but not cortical regions devoted to low-level features such as the
primary visual cortex (Choo et al., 2017). Ibarra et al. (2017) also demonstrated that high-level visual features
play an important role in aesthetic evaluations of urban and natural landscapes. They found that high-level
scene features—such as the shape and undulation of the skyline, the presence of water in the scene, and the
distribution of buildings—mediated the relationship between low-level scene features and aesthetic prefer-
ence ratings by explaining over half of their shared statistical variance. Thus, aesthetic judgments of built
environments may involve complex interactions between low-level and high-level stimulus features.
Aesthetic experiences may also be influenced by design features related to spatial navigation. The grid cells
of the hippocampus create a “cognitive map” of one’s environment to facilitate navigation (McNaughton
et al., 2006; O’Keefe & Nadel, 1978) and may be retrieved during subsequent re-exposure to a given loca-
tion (Astur et al., 2002; Maguire et al., 2000). Grid cell encoding is influenced by the shape of the environ-
ment, with distinct patterns for different geometrical spaces (Krupic et al., 2015). Notably, these differences

203
Coburn, Weinberger and Chatterjee

have relevance to the ease with which one is are able to navigate their environment, with orienting abilities
significantly worse in circular rooms with rotational symmetry (Kelly et al., 2008).
An additional class of neurons (head direction cells) signal an animal’s directional position and are espe-
cially critical for navigation (Knierim et al., 1995; Taube, 1998). Environmental information can also be
found in medial temporal lobe, where representations have been found to be modulated by the spatial
location of other individuals (Stangl et al., 2021). Homogenous spaces (e.g., uniform coloring, no obvious
landmarks) detract from the functioning of these neural processes and complicate the extent to which one is
able to form a complex mental representation of their spatial location (Evans & McCoy, 1998). Feeling lost
in one’s surrounding environment is a negatively valenced experience. Other research has found that beauty
judgments of buildings vary with neural activity in the global pallidus (Vartanian et al., 2013), a brain struc-
ture responsible for regulating voluntary movement.
Collectively, these findings suggest that the ease with which one can move through a space may influence
aesthetic experiences of built environments. Analogous to perceptual fluency, motor fluency might influ-
ence aesthetic experiences. Architects should consider how specific design features can facilitate more fluid
spatial navigation. In a study of nursing home design, monotony of architectural composition and absence
of reference points and signage were found to impede wayfinding and induce anxiety among patients with
advanced dementia (Passini et al., 2000). In another study of corridor design and navigation among healthy
adults, hallways with warmer colors were found to more memorable and attractive than hallways with
cooler colors (Hidayetoglu et al., 2012). These design factors are also likely to be crucial for large-scale built
environments—such as train stations or hospitals—in which people typically move about rapidly.
Neuroscience research has also begun to examine how non-visual environmental properties influence
aesthetic experience. For instance, specific environmental odors may trigger memories and past experiences,
putatively because of the anatomical connections between the olfactory and limbic systems (Ward et  al.,
2015). Historically, olfactory considerations played an important role in architectural design and provided an
important sensory and emotional link between people and places. Contemporary efforts to eliminate odors
and sterilize architectural interiors may contribute to feelings of sensory isolation and detachment in institu-
tional settings such as hospitals, schools and apartment buildings (Barbara & Perliss, 2006). Acoustic features
are also relevant to aesthetic experience. Loud environments increase blood pressure (Payne et  al., 2014)
and disrupt neural development (Gilbert & Galea, 2014). Several studies have investigated how the acoustic
design of classrooms influences childhood learning. Acute exposure to classroom noise can impair speech
recognition ( Johnson, 2000), decrease children’s performance on complex listening tasks, and interfere with
memory encoding processes (Hygge, 2003). Designs that attenuate noise might therefore be desirable, par-
ticularly in urban settings. Other features of a building, such as temperature and ventilation strategies, can
influence occupants’ perceptions of comfort and even contribute to perceptions of spatial beauty (Nicol &
Humphreys, 2002; Thorsson et al., 2007).

Emotional responses to architecture


Humans experience deeply emotional responses to beautiful objects, including architecture (Chatterjee &
Vartanian, 2014). Stress-reduction frameworks argue that certain visual properties convey feelings of calm-
ness or warmth, leading to improved affect (Tyrväinen et al., 2014; Ulrich, 1983; Ulrich et al., 1991). The
prospect-refuge theory expands on the evolutionary bases for this association. Processing-fluency accounts
indicate that efficiently processed stimuli are hedonically marked. These theoretical perspectives are sup-
ported by empirical findings linking architectural experiences with neural pathways related to reward and
stress response.
A meta-analysis of 93 neuroimaging studies pointed to a crucial role of the right anterior insula in
positively valenced aesthetic appraisals (Brown et al., 2011). The anterior insula has been highlighted as an

204
Architectural design and the brain

important brain area for processing negative emotions (e.g., disgust, sadness, pain), as well as the regulation
of the autonomic nervous system (Cechetto, 2014; Nagai et al., 2010), suggesting a link between aesthetic
experience and stress reduction. Further, anterior portions of the insula are also part of the gustatory cortex
and show activation when individuals are presented with pictures of food. Thus, aesthetic “taste” and gusta-
tory taste may share neural substrates. While somewhat speculative, it is plausible that a neural system initially
evolved to appraise food sources may have been co-opted for the appraisal of aesthetics. Aesthetic appraisals
also engaged the orbitofrontal cortex (OFC), a core region of the brain’s emotional and reward circuitry
(Bechara et al., 2000; see also Chapter 3).
Research on architectural aesthetics has extended these findings. In one study examining approach-
avoidance responses, the anterior midcingulate cortex (aMCC) was engaged when people viewed
enclosed interior spaces that elicited exit decisions (Vartanian et al., 2015). Because the aMCC receives
projections from the amygdala—indicating a potential role in fear processing—negative emotions may
be involved in processing architectural spaces, particularly ones that people wish to leave. Other work
report heightened fear, stress, and cortisol levels when people are immersed in a virtual simulation of
an enclosed room (Fich et al., 2014b). Together, this work highlights an underlying negative emotional
component (i.e., fear) that may drive aesthetic experiences of the built environment. These findings are
broadly consistent with stress-reduction frameworks that emphasize affective responses to the environ-
ment based on automatic, evolutionarily evolved processing of the environment (Ulrich, 1993). It is
important to note, however, that emotional responses are not automatic; the involvement of prefrontal
and hippocampal brain regions in beauty judgments of architecture suggest that conscious reasoning and
memory retrieval exert top-down influences on initial, automatic emotional reactions (Coburn et al.,
2017; Vartanian et al., 2013)
Following up on this work, Coburn et al. (2020) conducted a study in which participants evaluated 200
interior architectural scenes across a variety of aesthetic rating scales. Principal component analysis was con-
ducted to search for statistical patterns of overlap among thousands of ratings. Nearly 90% of the variance
in responses was explained by just three underlying psychological dimensions: coherence, fascination, and
hominess. Coherence describes the degree to which a space feels organized to the viewer. Fascination refers to
the visual richness and complexity of a space and is closely linked to a viewer’s sense of excitement and desire
to explore it. Hominess represents the extent to which a space feels comfortable, personal, and “home-like”
to the viewer. As a reliability check, the same experiment and analysis was repeated with a separate group of
600 participants. Again, the vast majority of variation in aesthetic responses was explained by the same three
dimensions. Coherence, fascination, and hominess have also been identified for images of building exteriors
(Weinberger et al., 2021), further suggesting that these three dimensions may broadly applicable to the built
environment.
Taking these analyses a step further, the authors also examined whether these psychological dimensions
were associated with specific neural signatures (Coburn et al., 2020). This hypothesis was tested by integrat-
ing the PCA scores of the architectural scenes with fMRI data from Vartanian and colleagues (2013), who
had previously evaluated the same images in the scanner via approach-avoidance and beauty judgment tasks.
The degree of fascination covaried with neural activity in the right lingual gyrus for both tasks. Coherence
was associated with neural activity in the left inferior occipital gyrus only when participants judged beauty,
and hominess covaried with activation of the left cuneus exclusively for the approach-avoidance task. Criti-
cally, these neural data were collected years before the three psychological dimensions had been identified
and in a separate group of participants. The authors concluded that the three dimensions of psychological
experience may be hardwired into the visual cortex, with each dimension carrying its own distinct neural
imprint. If these insights are shown to extend beyond the specific stimuli used in these studies, and to gen-
eralize to other architectural spaces, they could critically inform how buildings and urban environments are
designed and evaluated.

205
Coburn, Weinberger and Chatterjee

Knowledge, meaning, and individual variability


The neuroscience of architecture is also broadly interested in how differences at the individual level relate to
aesthetic experience. Within the framework of the aesthetic triad (Chatterjee & Vartanian, 2014), individual
differences are likely to involve knowledge-meaning systems (Coburn et al., 2017). Some general factors that
moderate individual-level differences in architectural experience include familiarity, expertise, context, and
cultural upbringing.
Here, again, we begin by considering findings from neuroscience and neuroaesthetics broadly. The mere
exposure effect, for instance, is a domain-general phenomenon in which people show a preference for stim-
uli that they see repeatedly (Bornstein & D’agostino, 1992; Montoya et al., 2017). But there is also a serial
effect; people prefer art that is preceded by an attractive painting compared to when it is preceded by a less
attractive piece (Kim et al., 2019). Aesthetic experience can also be modulated by context and culture, such
that people provide more favorable artistic judgments if presented with information about a piece’s cultural
value (Kirk et al., 2009a). Differences in preference are evident in greater activity of the medial orbitofrontal
and ventro-medial prefrontal cortex, pointing towards the role of memory and expectation. More recent
work indicates that hedonic responses to specific architectural styles are associated with differences in anterior
prefrontal cortex grey matter volume (Skov et al., 2021), further pointing towards the role of executive con-
trol on aesthetic experience. Knowledge about a building’s function may also influence aesthetic experience,
with buildings devoted to more positively valenced purposes (e.g., an art museum or temple) likely to receive
more positive judgments than negatively valenced ones (e.g., a prison or funeral home; Coburn et al., 2017).
Other research suggests that expertise within a specific artistic or stylistic domain biases the ways in which a
viewer attends to a stimulus (Seeley, 2013). Taken together, these findings speak directly to the knowledge-
meaning component of the aesthetic triad.
The upshot is that these factors vary widely across individuals. Person-to-person variability in familiarity,
knowledge, memory, and expectation are all likely to moderate the ways in which they engage with—and
respond emotionally to—the built environment. That is, the history of an individual’s exposure to a wide
array of environments could result in profound differences how they view a particular architectural space.
To illustrate this point more concretely, consider again the previously mentioned findings involving spa-
tial navigation. Notably, grid cells encode more than visual features alone; they are also used to represent
memories in a given environment (Eberhard, 2009). Thus, when people return to a familiar locale, they are
likely to recall their emotional states during prior visits. Frustrating past encounters (e.g., trouble navigating
a space) could plausibly lead one to develop negative feelings towards that space, or even other environ-
ments with shared physical features. In this way, past exposure mediates responses to a given environment
(Eberhard, 2009). This perspective has been broadly supported by experimental neuroscience research; par-
ticipation in a virtual spatial learning task was associated with increased resting-state functional connectivity
between left posterior hippocampus and dorsal caudate, brain regions with well-established roles in spatial
navigation (Woolley et al., 2015). Because the observed effects occurred at rest—that is, when participants
were not in the virtual space—this study demonstrated that repeated engagement of brain regions for specific
purposes may alter “baseline” functional connections between them (Wig et al., 2011). Perhaps experiences
in a built environment have lasting ramifications on neural organization and future interactions in that space,
as well as other similar environments.
Other experimental work has demonstrated this concept more directly. In one study, architecture stu-
dents showed lower “neural cost” when viewing buildings than students from other disciplines. That is, they
recruited fewer brain regions upon repeated presentations of buildings (Wiesmann & Ishai, 2011). This is
largely consistent with notions of fluent processing; given their expertise, the architecture students were able
to process the architecture images more efficiently. In another study comparing architects and controls, archi-
tects exhibited greater engagement of the reward pathway when making aesthetic judgments about buildings

206
Architectural design and the brain

(Kirk et al., 2009b). They also showed greater activation of memory structures, consistent with the theory
that these differences are experience and/or memory based. The extent to which these effects are driven by
top-down or bottom-up processes remains an open question, although results from other domains of empiri-
cal aesthetics suggests both are likely to play a role (Leder, 2013).

Health and wellness


Last, a growing body of work has begun to indicate how specific design features of the built environment can
influence psychological functioning and wellness. Exposure to nature appears to yield wide-ranging benefits
( Joye, 2007b), such as greater happiness (MacKerron & Mourato, 2013), improved attention (Berman et al.,
2008; Bowler et al., 2010), enhanced creativity (Mehta & Zhu, 2009), and better working memory (Bratman
et al., 2019). The use of brief exposure to natural environments may also improve outcomes in clinical popu-
lations (e.g., Beute & de Kort, 2018; Roe & Aspinall, 2011). Biophilic design—either by inserting greenery,
allowing views of the outdoors, or incorporating low-level features prevalent in nature—may be one way
to access these salubrious effects in human-made spaces. Additionally, some research suggests that lighting
conditions influence mood and academic performance, although the results of other studies run contrary
to these claims. Buildings that provide access to daylight may improve circadian rhythms and sleep quality
(Dutton, 2014) and benefit student learning. However, excess natural light can also cause unwanted effects
like glare if improperly designed.
Research on spatial navigation is also relevant to wellness. For instance, circular rooms with rotational
symmetry impede orientation efforts and can lead to higher levels of stress (Kelly et al., 2008). This adverse
effect may be heightened when visual reference points and/or exterior views are absent (Passini et al., 2000).
It is also clear that open and closed spaces differentially engage fear and emotional neural circuitry, with
closed spaces associated with a host of negative emotions and increased stress (Brown et al., 2011; Fich et al.,
2014b; Vartanian et al., 2013). Architects may be able to improve the mental health of inhabitants by design-
ing buildings that are open and facilitate fluid navigation. Such designs may be especially useful in clinical
populations such as people with Alzheimer’s disease who might be inclined to wander and be frustrated by
barriers to movement (Passini et al., 2000).
It is important to note, however, no one-size-fits-all framework will apply to how architecture can
improve well-being. For one, we have highlighted ways in which individuals vary in response to architec-
tural design. What works for one person may not work for all. Furthermore, responses to the environment
can even vary within an individual. Such variability could stem from differences in the knowledge-meaning
or emotion-valuation systems, but responses may also vary based on the function of the space itself. For
instance, Graham and colleagues (2015) identified characteristics that people generally valued in home ambi-
ances, such as invitingness, organization, and relaxation. However, they also observed a strong effect of
room type. That is, participants endorsed different desired aesthetic qualities based on which room they
were considering. Different parts of a building—or even different spaces within a single room—will be
experienced differently. Assuming that pairing a room with its desired ambiance or aesthetic is beneficial for
the inhabitants’ well-being, this research clearly indicates the need for adaptable and fluid design features.
Consistent with this perspective, there is also evidence demonstrating that humans prefer settings with varied
stimulus properties, especially those over which they are able to exert some amount of control. For instance,
although neutral luminance levels are widely viewed as comfortable, stimulating environments characterized
by varying levels of lighting are more favorable for productivity (Gou et al., 2014). Similarly, the ability to
control the temperature of one’s environment is associated with greater feelings of comfort, and occupants
with greater freedom to open and close windows in naturally ventilated buildings feel comfortable at a wider
range of temperatures (Brager et al., 2004). Personal choice has been found to activate the extended reward
network, with diminished activation for people showing depressive symptoms (Romaniuk et  al., 2019),

207
Coburn, Weinberger and Chatterjee

consistent work showing increased stress and feelings of helplessness when exposed to uncontrollable envi-
ronments (Evans & McCoy, 1998). Thus, designs that provide diverse kinds of stimulation and/or the ability
for occupants to regulate sensory qualities may reduce feelings of stress.

Challenges and future directions


Insights gained from the neuroscience of architecture can be used to improve models of how and why spe-
cific architectural features affect people in specific ways. However, as we have previously detailed (Coburn
et al., 2017), a number of challenges must be addressed to more clearly establish how buildings can be con-
structed to improve human experience and well-being. Here, we briefly touch on these outstanding issues
and highlight potential strategies to resolve them.
Architectural spaces encompass a wide range of functions and settings. Thus, features that relate to aes-
thetic experience of the built environment may not be universally shared. There are also practical limitations
based on physical setting; an architect cannot merely insert a large window that overlooks water or a forest
into every building. Physical and financial constraints further limit potential design choices. The function of
a space also cannot be overlooked, both in terms of the design features and the experiences of the inhabit-
ants. For instance, the experiences of a patient in a hospital are distinct from those of a student in a school
or someone inside their own home. Design elements that foster wellness are unlikely to be consistent across
these different settings, which complicates the ability to apply findings from one set of stimuli to another or
to make broad generalizations.
Challenges also relate to the measurement of aesthetic experience. To date, the most studies of neu-
roaesthetics have used 2D images (oftentimes while participants lie horizontally in an MRI scanner). This
approach, while practical, overlooks features such as scale and texture and introduces additional confounding
factors associated with presenting works of art on a computer screen and in a loud machine. Experimental
stimuli are also carefully selected and/or modified to control for potentially confounding variables like light-
ing and pixilation. While this approach makes it easier to identify the source of an observed effect (e.g., dif-
ferences in beauty judgments are not related to the “crispness” of an image), generalizability becomes more
difficult. These problems are further complicated for the study of architecture. The reasons for this are fairly
obvious; buildings are three dimensional, immersive, interactive, and multisensory. There are also a wide
range of contextual (e.g., outside noise, surrounding environment) and functional (e.g., hospital, museum,
school) factors that cannot be adequately conveyed by images. Advances in virtual reality have the potential
to mitigate some of these problems but are still unlikely to fully capture the multidimensional nature of
architecture.
Another challenge concerns temporal dynamics of aesthetic experiences. Repeated viewings of the same
stimulus are associated with more positive appraisals (Bornstein & D’agostino, 1992); thus, spending more
time in a building is likely to be associated with fluctuations in aesthetic judgments. Most research studies
present participants with an image for only a few seconds, even though it is fairly well established that aes-
thetic experiences vary over longer durations (Chatterjee & Vartanian, 2014; Coburn et al., 2017). This raises
issues both methodological (e.g., How long should participants view an image?) and theoretical (e.g., When
can aesthetic experience be most accurately measured? Do we need to obtain multiple aesthetic judgments
at different timepoints?).
Perhaps the best way to address the previous issues is to move experiments out of the laboratory and into
the “wild.” Rather than presenting people with images, data can be collected at actual buildings or structures.
This approach has been successfully used to measure feelings of nostalgia at heritage sites (Prayag  & Del
Chiappa, 2021). Experience sampling methods could be paired with this approach to obtain aesthetic judg-
ments on different spatial and temporal scales. For instance, participants could respond to a series of prompts
on their cell phones at specific times or locations. Further, thanks to recent advances in mobile EEG and

208
Architectural design and the brain

fNIRS, researchers can pair these behavioral ratings with neural data across a wide range of settings. Mobile
imaging techniques have provided novel insights about neural processing in art museums (Kontson et al.,
2015) and collaboration in the classroom (Dikker et al., 2017) and are now being applied to architecture
(Djebbara et al., 2021; see Chapter 17). While these data will be inherently “messier” and complicate the
separation of neural signal from noise, results gained in the field may be more ecologically relevant than those
collected in the laboratory.

Conclusion
The neuroscience of architecture is a very young field that has witnessed significant progress over the past
decade. In this chapter, we discussed the historical and economic context that led to the mass-standardization
of architectural aesthetics in mid–20th-century Western cultures. We then explored how evidence-based
design and research movements emerged as a reaction to these trends in building design and urban develop-
ment. The neuroscience of architecture represents an important arm of evidence-based architectural research
that focuses on understanding sensory, emotional, and psychological dimensions of human experiences in
response to architectural design. Many subtle aspects of architectural and aesthetic experience were previ-
ously difficult to measure and validate using research methods of behavioral science and environmental
psychology; the research tools of neuroimaging and cognitive science, by contrast, have enabled researchers
to increase the magnification of their lens, as it were, and observe more closely the mechanisms of the mind
and brain that mediate human-architectural interactions. These tools have also enabled researchers to move
beyond descriptive and theoretical approaches in order to test specific hypotheses about how people perceive
and respond to buildings.
Among the most promising areas of research discussed in this chapter include explicit investigation of the
aesthetic experience of built spaces; the potential benefits of biophilic design for stress reduction, wellness,
and health; the importance of sensory features of spaces that facilitate wayfinding and navigation; and the
probing of specific dimensions of emotional experience in the built environment—such as hominess, fasci-
nation, and coherence. We also explored factors of the individual that contribute to person-level variation
in architectural experiences, including culture, expertise, and familiarity. These individual differences under-
score the importance of considering the specific needs of prospective building occupants when designing
a space rather than taking a one-size-fits-all approach. Yet a significant body of evidence also suggests that
certain types of aesthetic experiences, such as sensory perceptions of biophilic design features, may be shared
by many types of building occupants across different spaces. This research highlights the potential value of
integrating evidence-based design principles into architectural education, training, and practice. In this way,
the neuroscience of architecture stands at the frontier of architectural innovation and is poised to contribute
to the design and construction of human-centered buildings and urban spaces.

Notes
1 See the section “Neuroscience of Architecture” for further discussion of fractal scaling and other natural patterns in
architecture.
2 One such community, DeSoto-Carr in St. Louis, was a black neighborhood bulldozed in the early 1950s to make way
for the infamous Pruitt-Igoe housing project. Designed in the classic Le Corbusian aesthetic, these three dozen mono-
lithic high-rise towers soon degenerated into some of the most derelict and poverty-stricken buildings in America.
All 33 buildings were demolished within 20 years of their construction, meeting the same eventual fate as hundreds of
similar towers in Chicago, Philadelphia, and Baltimore, among other American cities.
3 Alexander’s projects were also an early model for the design-build method of architectural construction, which has
gained popularity since the turn of the millennium. A design-build firm takes responsibility for all facets of planning,
construction, and financing of a project, instead of dividing the work among subsets of specialists (i.e., architects,
contractors, and sub-contractors).

209
Coburn, Weinberger and Chatterjee

4 An oak tree is an intuitive example of fractal scaling found in nature. Self-similar shapes and colors can be found at
many scales in the tree, including the large trunk, the medium-sized branches, and the smaller twigs. Even the tiny
veins of an oak leaf echo the curvature and shape of the larger trunk and branches, creating a sense of self-similarity
and unity across all scales of the structure.

References
Adams, M. (2014). Quality of urban spaces and wellbeing. In Wellbeing and the environment (Vol. 2, pp. 249–270). John
Wiley & Sons, Inc.
Aiken, N. E. (1998). Human cardiovascular response to the eye spot threat stimulus. Evolution and Cognition, 4(1),
51–62.
Alexander, C. (1979). The timeless way of building. Oxford University Press.
Alexander, C. (2002). The phenomenon of life: An essay on the art of building and the nature of the universe (Vol. 1). Center for
Environmental Structure.
Alexander, C. (2004). The phenomenon of life: Book one: The nature of order: An essay on the art of building and the nature of
the universe. Taylor & Francis.
Alexander, C., Alexander, P. in the D. of A. C., & Davis, H. (1985). The production of houses. Oxford University Press.
Alexander, C., Neis, H. J., & Alexander, M. M. (2012). The battle for the life and beauty of the earth: A struggle between two
world-systems. Oxford University Press.
Astur, R. S., Taylor, L. B., Mamelak, A. N., Philpott, L., & Sutherland, R. J. (2002). Humans with hippocampus damage
display severe spatial memory impairments in a virtual Morris water task. Behavioural Brain Research, 132(1), 77–84.
https://doi.org/10.1016/S0166-4328(01)00399-0
Austen, B. (2018). High-risers: Cabrini-Green and the fate of American public housing. HarperCollins.
Baker, N., & Standeven, M. (1995). A behavioural approach to thermal comfort assessment in naturally ventilated build-
ings. In Proceedings of the CIBSE national conference. The Chartered Institution of Building Services Engineers Confer-
ence, Eastbourne.
Barbara, A., & Perliss, A. (2006). Invisible architecture: Experiencing places through the sense of smell. Skira.
Barton, J., & Pretty, J. (2010). What is the best dose of nature and green exercise for improving mental health? A multi-
study analysis. Environmental Science and Technology, 44(10), 3947–3955. https://doi.org/10.1021/es903183r
Bechara, A., Damasio, H., & Damasio, A. R. (2000). Emotion, decision making and the orbitofrontal cortex. Cerebral
Cortex, 10(3), 295–307. https://doi.org/10.1093/cercor/10.3.295
Berman, M. G., Hout, M. C., Kardan, O., Hunter, M. R., Yourganov, G., Henderson, J. M., Hanayik, T., Karimi, H., &
Jonides, J. (2014). The perception of naturalness correlates with low-level visual features of environmental scenes.
PLOS ONE, 9(12), e114572. https://doi.org/10.1371/journal.pone.0114572
Berman, M. G., Jonides, J., & Kaplan, S. (2008). The cognitive benefits of interacting with nature. Psychological Science,
19(12), 1207–1212. https://doi.org/10.1111/j.1467-9280.2008.02225.x
Berman, M. G., Kross, E., Krpan, K. M., Askren, M. K., Burson, A., Deldin, P. J., Kaplan, S., Sherdell, L., Gotlib, I.
H., & Jonides, J. (2012). Interacting with nature improves cognition and affect for individuals with depression. Journal
of Affective Disorders, 140(3), 300–305. https://doi.org/10.1016/j.jad.2012.03.012
Bertamini, M., Rampone, G., Makin, A. D. J., & Jessop, A. (2019). Symmetry preference in shapes, faces, flowers and
landscapes. PeerJ, 7, e7078. https://doi.org/10.7717/peerj.7078
Berto, R. (2005). Exposure to restorative environments helps restore attentional capacity. Journal of Environmental Psychol-
ogy, 25(3), 249–259. https://doi.org/10.1016/j.jenvp.2005.07.001
Berto, R., & Barbiero, G. (2017). The biophilic quality index. A tool to improve a building from “green” to restorative.
Visions for sustainability, 8.
Beute, F., & de Kort, Y. A. W. (2018). The natural context of wellbeing: Ecological momentary assessment of the influ-
ence of nature and daylight on affect and stress for individuals with depression levels varying from none to clinical.
Health and Place, 49, 7–18. https://doi.org/10.1016/j.healthplace.2017.11.005
Biederman, I. (1987). Recognition-by-components: A theory of human image understanding. Psychological Review, 33.
https://doi.org/10.1037/0033-295X.94.2.115
Bornstein, R. F., & D’agostino, P. R. (1992). Stimulus recognition and the mere exposure effect. Journal of Personality and
Social Psychology, 63(4), 545–552. https://doi.org/10.1037//0022-3514.63.4.545
Bowler, D. E., Buyung-Ali, L. M., Knight, T. M.,  & Pullin, A. S. (2010). A  systematic review of evidence for the
added benefits to health of exposure to natural environments. BMC Public Health, 10(456), 1–10. https://doi.
org/10.1186/1471-2458-10-456

210
Architectural design and the brain

Brady, N., & Field, D. J. (2000). Local contrast in natural images: Normalisation and coding efficiency. Perception, 29(9),
1041–1055. https://doi.org/10.1068/p2996
Brager, G., Paliaga, G., & De Dear, R. (2004). Operable windows, personal control and occupant comfort. ASHRAE
Transactions, 110(2), 17–35.
Bratman, G. N., Anderson, C. B., Berman, M. G., Cochran, B., Vries, S. de, Flanders, J., Folke, C., Frumkin, H., Gross,
J. J., Hartig, T., Kahn, P. H., Kuo, M., Lawler, J. J., Levin, P. S., Lindahl, T., Meyer-Lindenberg, A., Mitchell, R.,
Ouyang, Z., Roe, J., . . . Daily, G. C. (2019). Nature and mental health: An ecosystem service perspective. Science
Advances, 5(7), eaax0903. https://doi.org/10.1126/sciadv.aax0903
Bratman, G. N., Daily, G. C., Levy, B. J., & Gross, J. J. (2015). The benefits of nature experience: Improved affect and
cognition. Landscape and Urban Planning, 138, 41–50. https://doi.org/10.1016/j.landurbplan.2015.02.005
Bratman, G. N., Hamilton, J. P.,  & Daily, G. C. (2012). The impacts of nature experience on human cogni-
tive function and mental health. Annals of the New York Academy of Sciences, 1249(1), 118–136. https://doi.
org/10.1111/j.1749-6632.2011.06400.x
Bratman, G. N., Hamilton, J. P., Hahn, K. S., Daily, G. C., & Gross, J. J. (2015). Nature experience reduces rumination
and subgenual prefrontal cortex activation. Proceedings of the National Academy of Sciences of the United States of America,
112(28), 8567–8572. https://doi.org/10.1073/pnas.1510459112
Brown, S. C. (2014). Neighborhoods and social interaction. In Wellbeing and the environment (Vol. 2). John Wiley  &
Sons, Inc.
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. NeuroImage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Capo, D. (2004). The fractal nature of the architectural orders. Nexus Network Journal, 6(1), 30–40. https://doi.
org/10.1007/s00004-004-0004-9
Cechetto, D. F. (2014). Cortical control of the autonomic nervous system. Experimental Physiology, 99(2), 326–331.
https://doi.org/10.1113/expphysiol.2013.075192
Chang, C. C., Cheng, G. J. Y., Nghiem, T. P. L., Song, X. P., Oh, R. R. Y., Richards, D. R., & Carrasco, L. R. (2020).
Social media, nature, and life satisfaction: Global evidence of the biophilia hypothesis. Scientific Reports, 10(1), 4125.
https://doi.org/10.1038/s41598-020-60902-w
Chatterjee, A. (2003). Prospects for a cognitive neuroscience of visual aesthetics. Bulletin of Psychology and the Arts, 4,
55–60.
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Choo, H., Nasar, J. L., Nikrahei, B., & Walther, D. B. (2017). Neural codes of seeing architectural styles. Scientific Reports,
7(1), 40201. https://doi.org/10.1038/srep40201
Clark, A. (2013). Whatever next? Predictive brains, situated agents, and the future of cognitive science. Behavioral and
Brain Sciences, 36(3), 181–204. https://doi.org/10.1017/S0140525X12000477
Coburn, A., Kardan, O., Kotabe, H., Steinberg, J., Hout, M. C., Robbins, A., MacDonald, J., Hayn-Leichsenring, G., &
Berman, M. G. (2019). Psychological responses to natural patterns in architecture. Journal of Environmental Psychology,
62, 133–145. https://doi.org/10.1016/j.jenvp.2019.02.007
Coburn, A., Vartanian, O., & Chatterjee, A. (2017). Buildings, beauty, and the brain: A neuroscience of architectural
experience. Journal of Cognitive Neuroscience, 29(9), 1521–1531. https://doi.org/10.1162/jocn_a_01146
Coburn, A., Vartanian, O., Kenett, Y. N., Nadal, M., Hartung, F., Hayn-Leichsenring, G., Navarrete, G., González-
Mora, J. L., & Chatterjee, A. (2020). Psychological and neural responses to architectural interiors. Cortex; a Journal
Devoted to the Study of the Nervous System and Behavior, 126, 217–241. https://doi.org/10.1016/j.cortex.2020.01.009
Coles, R. (2014). Environmental interaction and engagement. In Wellbeing and the environment (Vol. 2). John Wiley &
Sons, Inc.
Cooper, R. (2014). Wellbeing and the environment: An overview. In Wellbeing and the environment (Vol. 2, pp. 653–668).
John Wiley & Sons, Inc.
Cooper, R., & Burton, E. (2014). Wellbeing and the environmental implications for design. In Wellbeing and the environ-
ment (Vol. 2, pp. 653–668). John Wiley & Sons, Inc.
Corbusier, L. (1927). Vers une architecture. Towards a new architecture . . . Translated from the Thirteenth (French ed) with
an Introduction by Frederick Etchells. John Rodker.
Crompton, J. L. (2001). The impact of parks on property values: A review of the empirical evidence. Journal of Leisure
Research, 33(1), 1–31. https://doi.org/10.1080/00222216.2001.11949928
Dikker, S., Wan, L., Davidesco, I., Kaggen, L., Oostrik, M., McClintock, J., Rowland, J., Michalareas, G., Van Bavel,
J. J., Ding, M., & Poeppel, D. (2017). Brain-to-brain synchrony tracks real-world dynamic group interactions in the
classroom. Current Biology, 27(9), 1375–1380. https://doi.org/10.1016/j.cub.2017.04.002

211
Coburn, Weinberger and Chatterjee

Djebbara, Z., Fich, L. B., & Gramann, K. (2021). The brain dynamics of architectural affordances during transition.
Scientific Reports, 11(1), 2796. https://doi.org/10.1038/s41598-021-82504-w
Dosen, A. S., & Ostwald, M. J. (2016). Evidence for prospect-refuge theory: A meta-analysis of the findings of envi-
ronmental preference research. City, Territory and Architecture, 3(1), 4. https://doi.org/10.1186/s40410-016-0033-1
Dutton, R. (2014). The built housing environment, wellbeing, and older people. In Wellbeing: A complete reference guide
(pp. 1–38). Wiley.
Eberhard, J. P. (2004). Neuroscience and architecture: The new frontier. Design with spirit. Proceedings of the 35th Annual
Conference of the Environmental Design Research Association (pp. 2–6).
Eberhard, J. P. (2008). Brain landscape: The coexistence of neuroscience and architecture. Oxford University Press.
Eberhard, J. P. (2009). Applying neuroscience to architecture. Neuron, 62(6), 753–756. https://doi.org/10.1016/j.
neuron.2009.06.001
Eberhard, J. P., & Patoine, B. (2004). Architecture with the brain in mind. Cerebrum, 6(2), 71–84.
Ellaway, A. (2014). The impact of the local social and physical local environment on wellbeing. In Wellbeing and the envi-
ronment (Vol. 2). John Wiley & Sons, Inc.
Ellaway, A., Macintyre, S., & Bonnefoy, X. (2005). Graffiti, greenery, and obesity in adults: Secondary analysis of Euro-
pean cross sectional survey. BMJ, 331(7517), 611–612. https://doi.org/10.1136/bmj.38575.664549.F7
Epstein, R.,  & Kanwisher, N. (1998). A  cortical representation of the local visual environment. Nature, 392(6676),
598–601. https://doi.org/10.1038/33402
Evans, G. W. (2003). The built environment and mental health. Journal of Urban Health: Bulletin of the New York Academy
of Medicine, 80(4), 536–555. https://doi.org/10.1093/jurban/jtg063
Evans, G. W., & McCoy, J. M. (1998). When buildings don’t work: The role of architecture in human health. Journal of
Environmental Psychology, 18(1), 85–94. https://doi.org/10.1006/jevp.1998.0089
Fich, L. B., Jönsson, P., Kirkegaard, P. H., Wallergård, M., Garde, A. H., & Hansen, Å. (2014a). Can architectural design
alter the physiological reaction to psychosocial stress? A  virtual TSST experiment. Physiology and Behavior, 135,
91–97. https://doi.org/10.1016/j.physbeh.2014.05.034
Fich, L. B., Jönsson, P., Kirkegaard, P. H., Wallergård, M., Garde, A. H., & Hansen, Å. (2014b). Can architectural design
alter the physiological reaction to psychosocial stress? A  virtual TSST experiment. Physiology and Behavior, 135,
91–97. https://doi.org/10.1016/j.physbeh.2014.05.034
Finn, K. R., Crutchfield, J. P., & Bliss-Moreau, E. (2019). Macaques preferentially attend to visual patterns with higher
fractal dimension contours. Scientific Reports, 9(1), 1–11.
Fleischer, A. (2012). A  room with a view—A valuation of the Mediterranean sea view. Tourism Management, 33(3),
598–602. https://doi.org/10.1016/j.tourman.2011.06.016
Flint, A. (2009). Wrestling with Moses: How Jane Jacobs Took on New York’s master builder and transformed the American city.
Random House Publishing Group.
Gartus, A., & Leder, H. (2013). The small step toward asymmetry: Aesthetic judgment of broken symmetries. i-Perception,
4(5), 361–364. https://doi.org/10.1068/i0588sas
Geisler, W. S. (2008). Visual perception and the statistical properties of natural scenes. Annual Review of Psychology, 59(1),
167–192. https://doi.org/10.1146/annurev.psych.58.110405.085632
Gibaldi, A., & Banks, M. S. (2019). Binocular eye movements are adapted to the natural environment. The Journal of
Neuroscience, 39(15), 2877–2888. https://doi.org/10.1523/JNEUROSCI.2591-18.2018
Gilbert, E., & Galea, S. (2014). Urban neighborhoods and mental health across the life course. N Wellbeing: A complete
reference guide (pp. 1–28). Wiley.
Goldberger, A. L. (1996). Fractals and the birth of Gothic: Reflections on the biologic basis of creativity. Molecular Psy-
chiatry, 1(2), 99–104.
Gou, Z., Lau, S. S.-Y., & Ye, H. (2014). Visual alliesthesia: The gap between comfortable and stimulating illuminance
settings. Building and Environment, 82, 42–49. https://doi.org/10.1016/j.buildenv.2014.08.001
Graham, L. T., Gosling, S. D., & Travis, C. K. (2015). The psychology of home environments: A call for research on
residential space. Perspectives on Psychological Science, 10, 346–356.
Greene, M. R., & Oliva, A. (2009). The briefest of glances: The time course of natural scene understanding. Psychological
Science, 20(4), 464–472. https://doi.org/10.1111/j.1467-9280.2009.02316.x
Hagerhall, C. M., Laike, T., Taylor, R. P., Küller, M., Küller, R., & Martin, T. P. (2008). Investigations of human EEG
response to viewing fractal patterns. Perception, 37(10), 1488–1494. https://doi.org/10.1068/p5918
Hagerhall, C. M., Purcell, T.,  & Taylor, R. (2004). Fractal dimension of landscape silhouette outlines as a predictor
of landscape preference. Journal of Environmental Psychology, 24(2), 247–255. https://doi.org/10.1016/j.jenvp.2003.
12.004

212
Architectural design and the brain

Helleman, G., & Wassenberg, F. (2004). The renewal of what was tomorrow’s idealistic city. Amsterdam’s Bijlmermeer
high-rise. Cities, 21(1), 3–17. https://doi.org/10.1016/j.cities.2003.10.011
Hidayetoglu, M. L., Yildirim, K., & Akalin, A. (2012). The effects of color and light on indoor wayfinding and the
evaluation of the perceived environment. Journal of Environmental Psychology, 32(1), 50–58. https://doi.org/10.1016/j.
jenvp.2011.09.001
Hygge, S. (2003). Classroom experiments on the effects of different noise sources and sound levels on long-term recall
and recognition in children. Applied Cognitive Psychology, 17(8), 895–914.
Ibarra, F. F., Kardan, O., Hunter, M. R., Kotabe, H. P., Meyer, F. A. C., & Berman, M. G. (2017). Image feature types
and their predictions of aesthetic preference and naturalness. Frontiers in Psychology, 8, 632. https://doi.org/10.3389/
fpsyg.2017.00632
Jacobs, J. (1992). The death and life of great American cities (Reissue ed.). Vintage Book Company.
James, W. (1985). Psychology: The briefer course (1892) (G. Allport, Ed.). University of Notre Dame Press.
Johnson, C. E. (2000). Childrens’ phoneme identification in reverberation and noise. Journal of Speech, Language, and
Hearing Research, 43(1), 144–157. https://doi.org/10.1044/jslhr.4301.144
Johnson, P. C. (1947). Mies van der Rohe [New York, 1947]. Museum of Modern Art.
Joye, Y. (2006). An interdisciplinary argument for natural morphologies in architectural design. Environment and Planning
B: Planning and Design, 33(2), 239–252. https://doi.org/10.1068/b31194
Joye, Y. (2007). Architectural lessons from environmental psychology: The case of biophilic architecture. Review of General
Psychology, 11(4), 305–328. https://doi.org/10.1037/1089-2680.11.4.305
Joye, Y., & De Block, A. (2011). ‘Nature and I are two’: A critical examination of the biophilia hypothesis. Environmental
Values, 20(2), 189–215. https://doi.org/10.3197/096327111X12997574391724
Joye, Y., & van den Berg, A. (2011). Is love for green in our genes? A critical analysis of evolutionary assumptions in
restorative environments research. Urban Forestry and Urban Greening, 10(4), 261–268. https://doi.org/10.1016/j.
ufug.2011.07.004
Kaplan, R. (1973). Predictors of environmental preference: Designers and “clients.” In Environmental design research.
Dowden, Hutchinson & Ross.
Kaplan, R., & Kaplan, S. (1989). The experience of nature—A psychological perspective. Cambridge University Press.
Kaplan, S. (1987). Aesthetics, affect, and cognition: Environmental preference from an evolutionary perspective. Environ-
ment and Behavior, 19(1), 3–32. https://doi.org/10.1177/0013916587191001
Kaplan, S. (1995). The restorative benefits of nature: Toward an integrative framework. Journal of Environmental Psychology,
15(3), 169–182. https://doi.org/10.1016/0272-4944(95)90001-2
Kardan, O., Demiralp, E., Hout, M. C., Hunter, M. R., Karimi, H., Hanayik, T., Yourganov, G., Jonides, J., & Berman,
M. G. (2015). Is the preference of natural versus man-made scenes driven by bottom—Up processing of the visual
features of nature? Frontiers in Psychology, 6, 471. https://doi.org/10.3389/fpsyg.2015.00471
Kardan, O., Gozdyra, P., Misic, B., Moola, F., Palmer, L. J., Paus, T., & Berman, M. G. (2015). Neighborhood greenspace
and health in a large urban center. Scientific Reports, 5, 11610. https://doi.org/10.1038/srep11610
Kellert, S. R. (2003). Kinship to mastery: Biophilia in human evolution and development. Island Press.
Kellert, S. R. (2005). Building for life: Designing and understanding the human-nature connection. Island Press.
Kellert, S. R. (2015, October  26). What is and is not biophilic design? Metropolis. http://www.metropolismag.com/
architecture/what-is-and-is-not-biophilic-design/
Kelly, J. W., McNamara, T. P., Bodenheimer, B., Carr, T. H., & Rieser, J. J. (2008). The shape of human navigation:
How environmental geometry is used in maintenance of spatial orientation. Cognition, 109(2), 281–286. https://doi.
org/10.1016/j.cognition.2008.09.001
Kim, S., Burr, D., & Alais, D. (2019). Attraction to the recent past in aesthetic judgments: A positive serial dependence
for rating artwork. Journal of Vision, 19(12), 19–19. https://doi.org/10.1167/19.12.19
Kirk, U., Skov, M., Christensen, M. S., & Nygaard, N. (2009b). Brain correlates of aesthetic expertise: A parametric
fMRI study. Brain and Cognition, 69(2), 306–315. https://doi.org/10.1016/j.bandc.2008.08.004
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009a). Modulation of aesthetic value by semantic con-
text: An fMRI study. NeuroImage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009
Knierim, J. J., Kudrimoti, H. S., & McNaughton, B. L. (1995). Place cells, head direction cells, and the learning of land-
mark stability. The Journal of Neuroscience, 15(3 Pt 1), 1648–1659. https://doi.org/10.1523/JNEUROSCI.15-03-01
648.1995
Kontson, K. L., Megjhani, M., Brantley, J. A., Cruz-Garza, J. G., Nakagome, S., Robleto, D., White, M., Civillico, E., &
Contreras-Vidal, J. L. (2015). ‘Your Brain on Art’: Emergent cortical dynamics during esthetic experiences. Frontiers
in Human Neuroscience, 9, 626. https://doi.org/10.3389/fnhum.2015.00626

213
Coburn, Weinberger and Chatterjee

Kotabe, H. P. (2016). Disorder, naturalness, and their influence on aesthetics and behavior. University of Chicago. https://
knowledge.uchicago.edu/handle/11417/214
Kotabe, H. P., Kardan, O., & Berman, M. G. (2016). Can the high-level semantics of a scene be preserved in the low-
level visual features of that scene? A study of disorder and naturalness. Proceedings of the 38th Annual Meeting of the
Cognitive Science Society, 38, 1721–1726.
Kotabe, H. P., Kardan, O., & Berman, M. G. (2017). The nature-disorder paradox: A perceptual study on how nature
is disorderly yet aesthetically preferred. Journal of Experimental Psychology: General, 146(8), 1126–1142. https://doi.
org/10.1037/xge0000321
Kravitz, D. J., Peng, C. S., & Baker, C. I. (2011). Real-world scene representations in high-level visual cortex: It’s the
spaces more than the places. Journal of Neuroscience, 31(20), 7322–7333. https://doi.org/10.1523/JNEUROSCI.4588-
10.2011
Krupic, J., Bauza, M., Burton, S., Barry, C.,  & O’Keefe, J. (2015). Grid cell symmetry is shaped by environmental
geometry. Nature, 518(7538), 232–235. https://doi.org/10.1038/nature14153
Kühn, S., Brick, T. R., Müller, B. C., & Gallinat, J. (2014). Is this car looking at you? How anthropomorphism pre-
dicts fusiform face area activation when seeing cars. PLOS ONE, 9(12), e113885. https://doi.org/10.1371/journal.
pone.0113885
Küller, R., & Lindsten, C. (1992). Health and behavior of children in classrooms with and without windows. Journal of
Environmental Psychology, 12(4), 305–317. https://doi.org/10.1016/S0272-4944(05)80079-9
Kuo, F. E., & Sullivan, W. C. (2001). Environment and crime in the inner city: Does vegetation reduce crime? Environ-
ment and Behavior, 33(3), 343–367. https://doi.org/10.1177/0013916501333002
Kyttä, M., & Broberg, A. (2014). The multiple pathways between environment and health. In Wellbeing and the environ-
ment (Vol. 2). John Wiley & Sons, Inc.
Kyttä, M., Kahila, M., & Broberg, A. (2011). Perceived environmental quality as an input to urban infill policy-making.
Urban Design International, 16(1), 19–35. http://doi.org/10.1057/udi.2010.19
Lavdas, A. A.,  & Schirpke, U. (2020). Aesthetic preference is related to organized complexity. PLOS ONE, 15(6),
e0235257. https://doi.org/10.1371/journal.pone.0235257
Leder, H. (2013). Next steps in neuroaesthetics: Which processes and processing stages to study? Psychology of Aesthetics,
Creativity, and the Arts, 7(1), 27–37. https://doi.org/10.1037/a0031585
Lercher, P., Evans, G. W., & Meis, M. (2003). Ambient noise and cognitive processes among primary schoolchildren.
Environment and Behavior, 35(6), 725–735. https://doi.org/10.1177/0013916503256260
Living building challenge | Living-Future.org. (2020, February 19). https://living-future.org/lbc/
Mackerron, G., & Mourato, S. (2013). Happiness is greater in natural environments. Global Environmental Change, 23(5),
992–1000. https://doi.org/10.1016/j.gloenvcha.2013.03.010
Maguire, E. A., Mummery, C. J., & Büchel, C. (2000). Patterns of hippocampal-cortical interaction dissociate tempo-
ral lobe memory subsystems. Hippocampus, 10(4), 475–482. https://doi.org/10.1002/1098-1063(2000)10:4<475::
AID-HIPO14>3.0.CO;2-X
Mak, M. Y., & Thomas Ng, S. (2005). The art and science of feng shui—A study on architects’ perception. Building and
Environment, 40(3), 427–434. https://doi.org/10.1016/j.buildenv.2004.07.016
Marlow, C. A., Viskontas, I. V., Matlin, A., Boydston, C., Boxer, A.,  & Taylor, R. P. (2015). Temporal structure of
human gaze dynamics is invariant during free viewing. PLOS ONE, 10(9), e0139379. https://doi.org/10.1371/
journal.pone.0139379
McNaughton, B. L., Battaglia, F. P., Jensen, O., Moser, E. I., & Moser, M. B. (2006). Path integration and the neural basis
of the “cognitive map.” Nature Reviews. Neuroscience, 7(8), 663–678. https://doi.org/10.1038/nrn1932
Mehaffy, M. W. (2017). Cities alive. Off the common books. Sustasis Press.
Mehta, R., & Zhu, R. J. (2009). Blue or red? Exploring the effect of color on cognitive task performances. Science,
323(5918), 1226–1229. https://doi.org/10.1126/science.1169144
Meidenbauer, K. L., Stenfors, C. U. D., Bratman, G. N., Gross, J. J., Schertz, K. E., Choe, K. W., & Berman, M. G.
(2020). The affective benefits of nature exposure: What’s nature got to do with it? Journal of Environmental Psychology,
72, 101498. https://doi.org/10.1016/j.jenvp.2020.101498
Meidenbauer, K. L., Stenfors, C. U. D., Young, J., Layden, E. A., Schertz, K. E., Kardan, O., Decety, J., & Berman, M.
G. (2019). The gradual development of the preference for natural environments. Journal of Environmental Psychology,
65, 101328. https://doi.org/10.1016/j.jenvp.2019.101328
Montoya, R. M., Horton, R. S., Vevea, J. L., Citkowicz, M., & Lauber, E. A. (2017). A re-examination of the mere
exposure effect: The influence of repeated exposure on recognition, familiarity, and liking. Psychological Bulletin,
143(5), 459–498. https://doi.org/10.1037/bul0000085
Murphy, D. (2012). The architecture of failure. John Hunt Publishing.

214
Architectural design and the brain

Nagai, M., Hoshide, S., & Kario, K. (2010). The insular cortex and cardiovascular system: A new insight into the brain-
heart axis. Journal of the American Society of Hypertension, 4(4), 174–182. https://doi.org/10.1016/j.jash.2010.05.001
Nicol, J. F., & Humphreys, M. A. (2002). Adaptive thermal comfort and sustainable thermal standards for buildings.
Energy and Buildings, 34(6), 563–572. https://doi.org/10.1016/S0378-7788(02)00006-3
O’Keefe, J., & Nadel, L. (1978). The hippocampus as a cognitive map. Clarendon Press.
Passini, R., Pigot, H., Rainville, C., & Tétreault, M.-H. (2000). Wayfinding in a nursing home for advanced dementia
of the Alzheimer’s type. Environment and Behavior, 32(5), 684–710. https://doi.org/10.1177/00139160021972748
Patra, R. (2009). Vaastu shastra: Towards sustainable development. Sustainable Development, 17(4), 244–256. https://doi.
org/10.1002/sd.388
Payne, S., Potter, R., & Cain, R. (2014). Linking the physical design of health-care environments to wellbeing indicators.
In Wellbeing: A complete reference guide (pp. 1–28). Wiley.
Prayag, G., & Del Chiappa, G. (2021). Nostalgic feelings: Motivation, positive and negative emotions, and authenticity
at heritage sites. Journal of Heritage Tourism, 1–16. https://doi.org/10.1080/1743873X.2021.1874000
Pretty, J., Peacock, J., Sellens, M., & Griffin, M. (2005). The mental and physical health outcomes of green exercise.
International Journal of Environmental Health Research, 15(5), 319–337. https://doi.org/10.1080/09603120500155963
Rainwater, L. (2006). Behind ghetto walls: Black families in a federal slum. Routledge.
Ramachandran, V. S., & Hirstein, W. (1999). The science of art: A neurological theory of aesthetic experience. Journal
of Consciousness Studies, 6(6–7), 15–51.
Reber, R., Schwarz, N.,  & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the per-
ceiver’s processing experience? Personality and Social Psychology Review, 8(4), 364–382. https://doi.org/10.1207/
s15327957pspr0804_3
Rhodes, G., Proffitt, F., Grady, J. M., & Sumich, A. (1998). Facial symmetry and the perception of beauty. Psychonomic
Bulletin and Review, 5(4), 659–669. https://doi.org/10.3758/BF03208842
Roe, J., & Aspinall, P. (2011). The restorative benefits of walking in urban and rural settings in adults with good and poor
mental health. Health and Place, 17(1), 103–113. https://doi.org/10.1016/j.healthplace.2010.09.003
Romaniuk, L., Sandu, A. L., Waiter, G. D., McNeil, C. J., Xueyi, S., Harris, M. A., Macfarlane, J. A., Lawrie, S. M.,
Deary, I. J., Murray, A. D., Delgado, M. R., Steele, J. D., McIntosh, A. M., & Whalley, H. C. (2019). The neurobiol-
ogy of personal control during reward learning and its relationship to mood. Biological Psychiatry. Cognitive Neuroscience
and Neuroimaging, 4(2), 190–199. https://doi.org/10.1016/j.bpsc.2018.09.015
Rosner, V. (2020). Machines for living: Modernism and domestic life. Oxford University Press.
Rousselet, G., Joubert, O., & Fabre-Thorpe, M. (2005). How long to get to the “gist” of real-world natural scenes? Visual
Cognition, 12(6), 852–877. https://doi.org/10.1080/13506280444000553
Ruggles, D. H., & Boak, J. (2020). Bonding with beauty: The connection between facial patterns, design and our well-
being. In Urban experience and design (pp. 40–57). Routledge.
Ryan, C. O., Browning, W. D., Clancy, J. O., Andrews, S. L., & Kallianpurkar, N. B. (2014). Biophilic design patterns:
Emerging nature-based parameters for health and well-being in the built environment. International Journal of Architec-
tural Research: Archnet-IJAR, 8(2), 62–76. https://doi.org/10.26687/archnet-ijar.v8i2.436
Ryan, R. M.,  & Huta, V. (2010). Pursuing pleasure or virtue: The differential and overlapping well-being ben-
efits of hedonic and eudaimonic motives. Journal of Happiness Studies, 11(6), 735–762. https://doi.org/10.1007/
s10902-009-9171-4
Salingaros, N. A. (2007). A theory of architecture. Umbau-Verlag.
Salingaros, N. A. (2020a). Proposing a biophilic healing index of design and architecture. Urban experience and design: Contempo-
rary perspectives on improving the public realm.
Salingaros, N. A. (2020b). The biophilic healing index predicts effects of the built environment on our wellbeing. Journal
of Biourbanism, 8, 13–34.
Seeley, W. P. (2013). Art, meaning, and perception: A question of methods for a cognitive neuroscience of art. British
Journal of Aesthetics, 53(4), 443–460. https://doi.org/10.1093/aesthj/ayt022
Skov, M., Vartanian, O., Navarrete, G., Modroño, C., Chatterjee, A., Leder, H., Gonzalez-Mora, J. L., & Nadal, M.
(2021). Differences in regional gray matter volume predict the extent to which openness influences judgments of
beauty and pleasantness of interior architectural spaces. Annals of the New York Academy of Sciences, 1507(1), 133–145.
https://doi.org/10.1111/nyas.14684
Spehar, B., Wong, S., van de Klundert, S., Lui, J., Clifford, C. W. G., & Taylor, R. P. (2015). Beauty and the beholder:
The role of visual sensitivity in visual preference. Frontiers in Human Neuroscience, 9, 514. https://doi.org/10.3389/
fnhum.2015.00514
Spiers, H. J., & Barry, C. (2015). Neural systems supporting navigation. Current Opinion in Behavioral Sciences, 1, 47–55.
https://doi.org/10.1016/j.cobeha.2014.08.005

215
Coburn, Weinberger and Chatterjee

Stangl, M., Topalovic, U., Inman, C. S., Hiller, S., Villaroman, D., Aghajan, Z. M., Christov-Moore, L., Hasulak, N. R.,
Rao, V. R., Halpern, C. H., Eliashiv, D., Fried, I., & Suthana, N. (2021). Boundary-anchored neural mechanisms of
location-encoding for self and others. Nature, 589(7842), 420–425. https://doi.org/10.1038/s41586-020-03073-y
Taube, J. S. (1998). Head direction cells and the neurophysiological basis for a sense of direction. Progress in Neurobiology,
55(3), 225–256. https://doi.org/10.1016/S0301-0082(98)00004-5
Taylor, R. (2021). The potential of biophilic fractal designs to promote health and performance: A review of experiments
and applications. Sustainability, 13(2), 823. https://doi.org/10.3390/su13020823
Thorsson, S., Honjo, T., Lindberg, F., Eliasson, I., & Lim, E.-M. (2007). Thermal comfort and outdoor activity in Japa-
nese urban public places. Environment and Behavior, 39(5), 660–684. https://doi.org/10.1177/0013916506294937
Tyrväinen, L., Ojala, A., Korpela, K., Lanki, T., Tsunetsugu, Y., & Kagawa, T. (2014). The influence of urban green
environments on stress relief measures: A field experiment. Journal of Environmental Psychology, 38, 1–9. https://doi.
org/10.1016/j.jenvp.2013.12.005
Ulrich, R. S. (1977). Visual landscape preference: A model and application. Man-environment Systems, 7(5), 279–293.
Ulrich, R. S. (1983). Aesthetic and affective response to natural environment. In Behavior and the natural environment
(pp. 85–125). Springer. http://link.springer.com/chapter/10.1007/978-1-4613-3539-9_4.
Ulrich, R. S. (1984). View through a window may influence recovery from surgery. Science, 224(4647), 224–225.
Ulrich, R. S. (1993). Biophilia, biophobia, and natural landscapes. Biophilia Hypothesis, 7, 73–137.
Ulrich, R. S., Simons, R. F., Losito, B. D., Fiorito, E., Miles, M. A., & Zelson, M. (1991). Stress recovery during expo-
sure to natural and urban environments. Journal of Environmental Psychology, 11(3), 201–230. https://doi.org/10.1016/
S0272-4944(05)80184-7
Valtchanov, D., & Ellard, C. G. (2015). Cognitive and affective responses to natural scenes: Effects of low level visual
properties on preference, cognitive load and eye-movements. Journal of Environmental Psychology, 43, 184–195. https://
doi.org/10.1016/j.jenvp.2015.07.001
Valtchanov, D., Barton, K. R., & Ellard, C. (2010). Restorative effects of virtual nature settings. Cyberpsychology, Behavior
and Social Networking, 13(5), 503–512. https://doi.org/10.1089/cyber.2009.0308
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Gonzalez-Mora, J. L., Leder, H., Modroño, C., Nadal, M.,
Rostrup, N., & Skov, M. (2015). Architectural design and the brain: Effects of ceiling height and perceived enclosure
on beauty judgments and approach-avoidance decisions. Journal of Environmental Psychology, 41, 10–18. https://doi.
org/10.1016/j.jenvp.2014.11.006
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov, M.
(2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceedings of the
National Academy of Sciences of the United States of America, 110(Suppl. 2), 10446–10453. https://doi.org/10.1073/
pnas.1301227110
Villani, D., & Riva, G. (2011). Does interactive media enhance the management of stress? Suggestions from a controlled
study. Cyberpsychology, Behavior and Social Networking, 15(1), 24–30. https://doi.org/10.1089/cyber.2011.0141
Vitruvius Pollio, M., Morgan, M. H., & Warren, H. L. (1914). Vitruvius: The ten books on architecture. Harvard University
Press. http://library.wur.nl/WebQuery/clc/128812
Ward, A., Hong, W., Favaloro, V., & Luo, L. (2015). Toll receptors instruct axon and dendrite targeting and participate
in synaptic partner matching in a drosophila olfactory circuit. Neuron, 85(5), 1013–1028. https://doi.org/10.1016/j.
neuron.2015.02.003
Weinberger, A. B., Christensen, A. P., Coburn, A., & Chatterjee, A. (2021). Psychological responses to buildings and
natural landscapes. Journal of Environmental Psychology, 77, 101676. https://doi.org/10.1016/j.jenvp.2021.101676
WELL Certified | International WELL Building Institute | IWBI. (n.d.). Retrieved March 10, 2021, from http://www.
wellcertified.com/
Whitelaw, A. (2013). Introducing ANFA, the academy of neuroscience for architecture. Intelligent Buildings International,
5(Suppl. 1), 1–3. https://doi.org/10.1080/17508975.2013.818764
Wiesmann, M., & Ishai, A. (2011). Expertise reduces neural cost but does not modulate repetition suppression. Cognitive
Neuroscience, 2(1), 57–65. https://doi.org/10.1080/17588928.2010.525628
Wig, G. S., Schlaggar, B. L., & Petersen, S. E. (2011). Concepts and principles in the analysis of brain networks. Annals
of the New York Academy of Sciences, 1224(1), 126–146. https://doi.org/10.1111/j.1749-6632.2010.05947.x
Wilson, E. O. (1984). Biophilia. Harvard university press.
Wilson, E. O., & Kellert, S. R. (1995). The biophilia hypothesis. Island Press.
Windhager, S., Hutzler, F., Carbon, C. C., Oberzaucher, E., Schaefer, K., Thorstensen, T., Leder, H., & Grammer, K.
(2010). Laying eyes on headlights: Eye movements suggest facial features in cars. Collegium Antropologicum, 34(3),
1075–1080.

216
Architectural design and the brain

Woolley, D. G., Mantini, D., Coxon, J. P., D’Hooge, R., Swinnen, S. P., & Wenderoth, N. (2015). Virtual water maze
learning in human increases functional connectivity between posterior hippocampus and dorsal caudate: Spatial
learning and functional connectivity. Human Brain Mapping, 36(4), 1265–1277. https://doi.org/10.1002/hbm.22700
Yu, Y., Romero, R., & Lee, T. S. (2005). Preference of sensory neural coding for 1/f signals. Physical Review Letters,
94(10), 108103. https://doi.org/10.1103/PhysRevLett.94.108103
Yue, X., Cassidy, B. S., Devaney, K. J., Holt, D. J., & Tootell, R. B. H. (2011). Lower-level stimulus features strongly
influence responses in the fusiform face area. Cerebral Cortex, 21(1), 35–47. https://doi.org/10.1093/cercor/bhq050
Zhang, J. W., Piff, P. K., Iyer, R., Koleva, S., & Keltner, D. (2014). An occasion for unselfing: Beautiful nature leads to
prosociality. Journal of Environmental Psychology, 37, 61–72. https://doi.org/10.1016/j.jenvp.2013.11.008

217
11
SEXUAL SELECTION, AESTHETIC
APPRECIATION AND MATE
CHOICE
Michael J. Ryan

Darwin proposed a theory for the evolution of sexual beauty in animals—sexual selection (1859, 1871).
Darwin’s emphasis was on one particular mechanism of sexual selection, mate choice modulated by what he
called “a taste for the beautiful” (1871, p. 63). Darwin was also prescient about the role of underlying neural
biases promoting the evolution of sexual beauty:

When male animals utter sounds in order to please the females, they would naturally employ those
which are sweet to the ears of the species; and it appears that the same sounds are often pleasing to
widely different animals, owing to the similarity of their nervous systems, as we ourselves perceive
in the singing of birds and even in the chirping of certain tree-frogs giving us pleasure.
(1872, p. 91)

It has taken more than a century and a half, but recently neuroaesthetics has become an important, although
sometimes implicit rather than explicit, approach to understanding where all this beauty comes from. To
explore these ideas, we first have to have an understanding of the fundamental tenets of Darwin’s theory of
sexual selection.

Sexual selection
Beauty abounds in nature (Figure 11.1). One type of natural beauty that for some of us, including Darwin,
is the most enchanting includes the melodious songs of birds, the striking colours of coral reef fishes, and the
nocturnal flashes of fireflies. These and many other exemplars of animal beauty share something in common:
they are all recruited in the service of sex. The discussion of beauty in this chapter refers to animal sexual
beauty.
The peacock is something of a mascot for sexual beauty, and it is all because of his tail. He has 200
feathers up to 4 feet long. These feathers are adorned with eye-like spots and have an iridescent sheen that
causes them to sparkle brilliantly in the sunlight. While courting, he develops an erection, and once those
feathers are erect, he shakes, rattles, and rolls them, causing the tail to hum like an engine and his eye-spots
to vibrate hypnotically. A true thing of beauty to most of us, including female peahens, but surprisingly not
to Darwin. In 1860, he wrote to Asa Gray, a famous North American botanist, “The sight of a feather in a
peacock’s tail, whenever I gaze at it, makes me sick!” How is it that Charles Darwin, the co-founder of the

218 DOI: 10.4324/9781003008675-12


Sexual selection and mate choice

Figure 11.1 Examples of sexually selected traits. (from top left to right): A calling male túngara frog at a typical breeding
site in central Panama. The male’s large vocal sac is as distinctive as his complex call (photo by Ryan Taylor).
A male peacock erecting its tail feathers or train while courting a female. Although the tail is enchanting
to females, it is this beautiful structure which, Darwin declared, made him sick every time he saw it (photo
by Jyshah Jysha). The golden-headed lion tamarin is an endangered species found in lowland tropical forests
in the state of Bahia, Brazil. They live in social groups in which both males and females care for the young
and for juveniles. Little else is known about their mating system. It is considered by some the world’s most
beautiful primate (photo by Steve Wilson). The red bird of paradise is a native of Indonesia. The males, one of
which is shown here, are characterized by a pair of long tail wires. During courtship, these tail wires seem
to outline the male in the middle of a heart (photo by Tim Laman).
(from middle left to right): The quetzal is the national bird of Guatemala, and its image adorns the country’s
coat of arms. Some consider the male resplendent quetzal the world’s most beautiful bird. My binoculars
began to shake in my hands the first time I saw one (photo by Dominic Sherony). A male swordtail characin,
right, a native of Trinidad, Tobago, Venezuela, and Colombia, extends its pectoral fin-ray with a piece of
flesh that resembles a food item to the female, on the left. The female is attracted to this faux food item, at
which time the male initiates courtship (photo by Nicolas Kolm). Male and female fireflies engage in spec-
tacular nocturnal visual displays. As with many other courtship displays, the patterns of flashes are distinc-
tive for each species. This image is a time-lapse photograph of synchronous fireflies from the Great Smoky
Mountains National Park, near Elkmont, Tennessee (photo by Radim Schreiber). A  male hairy caterpillar
extruding its hair pencils. The tubes, or coremata, are inflated by blood pressure, causing sex pheromones
to be secreted through the hairs (photo by Rodney and Smudge Foster Rentz). A bee orchid pseudo-copulating
with an orchid. Although this behaviour appears maladaptive, it makes perfect sense in the context of the
bee’s strategy for finding females. As females are few and far between, it behoves the male bee to copulate
with anything resembling a female (photo by Nicolas J. Vereecken).
(from bottom left to right): Sexual selection often results in extreme differences between males and females.
In many species, the male is more adorned than the female, as seen here in the collared lizard in which the
more colourful male, top, is contrasted with the less colourful female, bottom (photo by A.K. Lappin). The
peacock spider is a type of jumping spider; the male’s colourful display is reminiscent of a peacock. His
beautifully adorned abdomen is only raised when the male courts the female, at which time he waves it back
and forth in an invitation to mate (photo by Jurgen Otto). Guppies are known not only for their spectacular
colours but also for incredible variation in those colours, especially among streams in Trinidad. Only a small
sample of the striking variation is shown here (photo by Cara Gibson and Anne Houde). (From Ryan, 2018).

219
Michael J. Ryan

theory of evolution by natural selection, could be turned off rather than turned on by such a magnificent
sight?
Charles Darwin and Alfred Wallace proposed natural selection as a mechanism that caused the evolution
of traits that enhance the survivorship of their bearers. Three steps were necessary for this to happen: there
had to be variation in the physical attributes of individuals, that variation had to result in some individuals
having a higher survivorship than others, and there needed to be a genetic basis to this physical variation. In
short, evolution by natural selection requires variation, selection, and heritability.
So why such consternation over the peacock’s tail? Yes, a male peacock displaying his wares to a female is
a magnificent sight to us and to his females, but when watching that same male trying to flee from a predator,
he seems pathetic. The tail slows down the male whether he is running or flying, and it is clearly a detriment
to survival. This is what concerned Darwin. If natural selection is a powerful force in nature, how to explain
the evolution of traits, such as the peacock’s tail, that are clearly detrimental to survival?
To reconcile the seeming contradiction, Darwin proposed another theory to supplement natural
selection—sexual selection. The two most important components of fitness are survival and reproduction.
You can’t reproduce if you don’t survive. But if you survive and don’t reproduce, you might as well be dead—
at least from a Darwinian perspective.
These types of sexually selected traits, the beautiful ornaments rather than the powerful armaments,
are common throughout most sexually reproducing animals yet quite diverse in their form: songs of birds,
crickets, and frogs; brilliant colours of butterflies, fish, and some primates; and striking odours of moths,
deer, and human perfumes. These traits all seem to share some similarities: they are clearly maladaptive for
survival, they are involved in sexual behaviour, and they are usually more extremely developed in males than
in females, although there are quite a few exceptions (Rosenthal, 2017). Like natural selection, three steps
are necessary for evolution by sexual selection: variation in traits, differential mating success due to these
traits, and genetic variation underlying these traits. Thus some traits enhance survivorship, while other traits
enhance mating success, and these two types of selection can impose conflicting forces on the evolution of
a trait. The male peacock can evolve tails that become more and more attractive until the benefit of this
elaborate trait is confounded by the costs that compromise survivorship.

Why males are often the sex with armaments and ornaments
Darwin noted that it was usually males that have become elaborated under sexual selection. The underlying
theory for this was not well developed until Trivers (1972) suggested his idea of parental investment, an idea
that arose from experiments on fruit flies conducted by Bateman in the 1940s (1948). Bateman showed that
male reproductive success increased with his number of matings, while this metric had no important effect
on the number of females’ offspring. Despite a number of exceptions (Arnqvist & Rowe, 2005), a recent
meta-analysis has shown strong support for this generalization from Bateman ( Janicke et al., 2016).
But why this big difference in the effect of mating success on reproductive success? The answer all boils
down to gamete size (Trivers, 1972). In many animals, the largest cell in the species is the female’s egg, and
the smallest cell in the same species is the male’s sperm. The result is that males produce many more gametes
than do females; in humans the difference is billions of sperm versus hundreds of eggs. The result is that the
reproductive success of sperm is limited by the relative scarcity of eggs. Thus males are under selection to
compete for access to females.
This competition can take two forms: combat between members of the same sex or mate choice exerted
by one sex on the other sex. Males have evolved a number of sexual traits that are used for combat with
other males to gain access to females; these include claws, canines, antlers, and a myriad of other armaments
(Emlen, 2014). Alternatively, males can evolve sexual ornaments. These are the traits of sexual beauty which
Darwin and legions of naturalists find so enchanting.

220
Sexual selection and mate choice

Why do females show preferences for sexual ornaments?


Darwin’s suggestion that males evolved weapons for sexual combat was readily accepted by his contempo-
raries. But the idea that males evolved ornaments to enchant females was rejected by many, including his
most ardent supporter, Alfred Wallace (Richards, 2017). There are several reasons this might have happened.
Certainly, empowering females to make mating decisions ran contrary to then-current Victorian social
mores (Cronin, 1991; Richards, 2017). But there were also some scientific objections, the main one being
no cogent argument as to why females would evolve preferences for elaborate ornaments that would hasten
the demise of the males that bore them. Of course, as noted previously, Darwin did provide an explanation,
that of female aesthetics. But this was viewed as merely kicking the can down the road and begs the question
of why females possess these sexual aesthetics.
There are now a few theories, not necessarily mutually exclusive, that can explain the evolution of these
female preferences (reviewed in Rosenthal, 2017; Ryan, 2018; Rosenthal & Ryan, 2022). In some cases,
a male’s ornaments might indicate direct benefits to females, an increase in the number of offspring birthed.
For example, male birds with brighter colours might have greater physical vigour and thus defend a larger
territory with more food for the females. Alternatively, the same brighter colours might also indicate those
males might have good genes that promote both greater vigour and survivorship, which the female would pass
on to her offspring. This is classified as an indirect benefit since her reproductive success does not vary with
the attractiveness of the male, but the offsprings’ later survivorship does. A tricky hypothesis, called runaway
sexual selection, assumes that both sexes possess genes for male traits and female preferences for those traits but
only express the gene appropriate for their sex. In this scenario, the preference genes and the colour genes
become linked in the genome. The female preference causes the evolution of brighter male colours, and
then the preference itself increases in frequency as it hitchhikes through the generations with the male colour
genes. This is also sometimes called the sexy son hypothesis since the indirect benefit gained by the female
derives from the production of more attractive male offspring.
Darwin’s hypothesis was far less utilitarian. He suggested that females had aesthetic preferences, much like
our own, and as the previous quote illustrates, males would evolve traits that the females found attractive.
This is Darwin’s hypothesis of neuroaesthetics, which has been resurrected more than 100 years later under
the rubric of sensory drive, sensory bias, and sensory exploitation.

Sensory biases and sexual beauty


All of us species of animal reside in our own sensory world. Standing in the middle of a tropical forest, our
senses are bombarded by a diversity of stimuli: the scents of numerous flowers, the flashing of fireflies, and
the chirping of katydids. But we are anosmic, blind, and deaf to the dense cloud of moth pheromones, the
ultraviolet signage on flowers that advertise to pollinators, and the echolocation calls of the bats swirling all
around us. In the late 1800s, Jacob von Uexküll described these sensory worlds as the animal’s Umvelt. This
is now one of the fundamental guiding principles of sensory ecology, and in the last decade, it has finally
infiltrated how we think about sexual beauty (Ryan, 2011; Caves et al., 2019).
There is no beauty without a brain, just as there is no sound when a tree falls in the forest if there is no ear
to hear it (Ryan, 2018). Courtship traits are only salient if they can be detected and perceived by choosers.
We must have some understanding of the species’ perceptual and cognitive biology to understand why it is that
some courtship traits are more attractive than others. In a sense, we need to view these traits through the eyes,
ears, and nares, or better yet the brains, of the beholders. Nevertheless, there are a few generalities of what
sexual traits are found more attractive. In hundreds of cases, choosers preferred sexual traits of courters that
were greater in magnitude: song and call amplitude and complexity, tail length, colour contrast, and quantity
of odours are only a few of the examples (Ryan & Keddy-Hector, 1992; Andersson, 1994; Rosenthal, 2017).

221
Michael J. Ryan

In the 1990s, studies of sexual beauty started to show a shift in emphasis from the potential information
sexual traits might transmit to the choosers to understanding how those traits interacted with the chooser’s
brain. Several similar hypotheses emerged almost simultaneously: sensory traps (West-Eberhard, 1979), sen-
sory biases (Endler, 1992), pre-existing biases (Basolo, 1990b), and sensory exploitation (Ryan, 1990). The
similarities and differences among these hypotheses have been described elsewhere (Endler, 1998; Ryan &
Cummings, 2013; Cummings & Endler, 2018); they all combine to point out the importance of the environ-
ment and the brain on the evolution of sexual traits and preferences for these traits. Interestingly, many of the
sensory, perceptual, and cognitive biases that lead to trait preferences are domain general and not specific to
the task of mate choice. In many cases, courters evolve traits that exploit pre-existing or hidden preferences
in choosers. Although olfactory, electrical, and tactile stimulation are all quite important in sexual behaviour
(Rosenthal, 2017), I will restrict this discussion to the two modalities in which we know the most about
sexual attraction: sights and sounds.

Visual beauty
The sensory end-organs, eyes, ears, olfactory, electrical, and taste receptors, are the portals by which envi-
ronmental stimuli enter the animal’s sensory, perceptual and cognitive systems. Responses of all sensory
end-organs to external stimuli are nonlinear. All sense organs respond more to some stimuli than they do to
others. Traits that elicit substantial stimulation of end-organs are not necessarily considered sexually attrac-
tive, especially if they have nothing to do with sex (Rosenthal, 2018). Nevertheless, there is often a tight fit
between the tuning of sensory organs and the properties of sexually attractive traits.
Many studies of sensory biases focus on the sensory receptors, and many of these studies are conducted
with fish. Cummings and Endler (2018) suggest this is probably because fish reside in environments in which
the ambient light conditions can vary substantially among habitats. This will result in strong selection on
photoreceptors to be able to detect food items in a given light background that, in turn, will result in males
evolving courtship traits that match the tuning of the photoreceptors. Cummings (2007) has shown that this
is exactly what happened in surf perch in the kelp forests off the coast of California.
Variation in background light might also play a role in the spectacular burst of speciation in cichlids in
the great African Lakes. Seehausen and his colleagues (2008—see also Maan et al., 2006) have shown that
eutrophication and water depth contract the visible light spectrum, which also influences photoreceptor
tuning, which in turn influences female mating preferences based on male colour. This interaction of ambi-
ent light environment, photoreceptor tuning, and mate choice preferences conspired to influence the rate at
which speciation takes place in these animals.
Guppies, Poecilia reticulata, are small brightly coloured fish that are well known in the pet trade, but their
natural environment is in streams and small rivers on the island of Trinidad. They are one of the most vari-
ably coloured of all vertebrates, and this variation seems to result from an interaction of predation pressure
and female preferences for more colourful males, especially males exhibiting more orange (Houde, 1997).
Not only does the amount of male orange colouration vary amongst populations, but so does the female’s
strength of preference for orange (Endler & Houde, 1995). Where in the sensory system does this preference
come from?
Oranges and yellows that adorn the skin of many animals are usually due to carotenoid-based pigments,
and most animals cannot synthesize carotenoids but must obtain them from the environment, usually by
feeding on small invertebrates. The classic explanation of the female guppy’s orange-based preference was
that orange was an “indicator” trait; it indicated to females the foraging ability of males. By being attracted to
males with more orange, the argument went, females might be mating with males with genes that promoted
foraging abilities (Endler, 1983).

222
Sexual selection and mate choice

Archer et al. (1982) used microspectrophotometer studies to show that the spectral sensitivity of long-
wavelength cones varied substantially among individual females, and they suggested that this variation in
spectral sensitivity contributes to the variation among females in their preference for orange colouration.
More recently, Sakai and colleagues (2018) showed that differences in longwave-sensitive opsin genotypes
interact with variation in the light environment to result in variation in opsin gene expression. They also
showed that the expression of multiple opsin genes correlates with the strength of female preference for
orange males. Thus the aesthetic preference for orange males, like many preferences for beautiful traits, can
be a bit idiosyncratic, and in these fish, this can be linked to variation in gene expression of the photopig-
ments that determine visual sensitivity.
Just because a preference for orange is exhibited in the domain of mate choice, it does not necessarily
mean that this preference is domain specific. In fact, Rodd and her colleagues (1999) presented compelling
evidence that this preference originated in the foraging domain and then had pleiotropic effects on colour
preference for mates. These authors noted that guppies often feed on small orange fruits that float on the
water’s surface. They presented guppies from different populations a series of chips of different colours and
measured the amount of time that both males and females spent inspecting the chips. They showed that
both sexes exhibited a preference for orange, and the strength of the preference for orange varied with
the strength of the female preference for orange males in that population. If the preference for orange had
evolved to promote mate choice, there is no reason to think that males would show the same preference
for inanimate orange objects that females express. Bourne and Watson (2009) showed similar results with
another closely related fish, Poecilia parae, in the same genus as guppies. Furthermore, Cole and Endler (2015)
conducted artificial selection experiments in guppies and showed a response to selection for sensitivity to red
wavelengths related to food after only five generations. The heritabilities (h2; the proportion of phenotypic
variation explained by additive genetic variation) in the two lines tested were 0.25 and 0.30. As with the
previous studies by Rodd et al., there is no difference between the sexes in their response to selection; thus,
the evolution of red sensitivity probably has nothing to do with mate preference, but it is assumed it would
influence mate preference as a pleiotropic effect.
So we know that female guppies prefer males with more orange, there is variation in the strength of this
female preference, this preference variation seems to be related to variation in photopigment sensitivity and
gene expression, sensitivity to red has a heritable genetic component, and that this chain of related phe-
nomena probably evolved to enhance foraging for orange fruit with the incidental consequence that orange
males prove more attractive. One final question is when did this preference for orange evolve? Clearly it did
not evolve only in guppies, since one of their close relatives expresses the same preference. How far back in
time and how deep in the phylogenetic history of these fishes was there this general preference for orange?
Foisy (2017) collected male colouration data for 232 species of poeciliids, the family in which guppies
resides, as well as some closely related species. He used some sophisticated phylogenetic analyses that pre-
dicted orange tended to evolve in males when there is evidence of a pre-existing bias for the colour orange.
Using this phylogeny, he identified 14 species lacking orange but for whom it was predicted there should
be a hidden preference for orange. Ten of the 14 species exhibited this hidden preference. The pre-existing
preference might be even more widespread, as Spence and Smith (2008) showed that zebrafish have a prefer-
ence for red fish even though this species lacks red. Therefore, not only do guppies prefer orange because
they evolve the preference for orange fruit, but this general preference for orange and also for red seems
to be true of many fish, including those in which males lack orange or red courtship colours. These pre-
existing preferences set the stage for the rapid evolution of these courtship colours. A male in an otherwise
bland species that evolves orange or red colouration will have a ready appreciator in the audience of females.
Some studies also suggest that there can be pre-existing preferences for the existence and size of orna-
ments in fish. Basolo (1990a) studied two types of fish in the genus Xiphophorus, swordtails and platyfish.

223
Michael J. Ryan

Although it has been difficult to resolve in detail the precise phylogenetic relationships within this genus (Cui
et al., 2013), in general it is thought that swordtails are one monophyletic group and the platyfish another
monophyletic group. The swordtails have extended caudal appendages, the swords, and females show prefer-
ences for swords and sometimes for longer swords (Rosenthal et al., 2001). The platyfish lacks swords. This
suggests that swords evolved at the base of the swordtail clade after this clade diverged from the platyfish
clade. By attaching plastic swords to male platyfish, Basolo showed that females had a pre-existing preference
for swords even though their males lacked swords. This pre-existing preference even extends outside of this
genus into some closely related genera (Basolo, 1995).
Gould (1999) came to a similar conclusion about the ubiquity of hidden preferences with a different set
of experiments. Mosquitofish, Gambusia, lack ornamented males, who mostly rely on forcing copulation
with females. The researchers presented females with a model of a typical male compared to tens of other
models that had exaggerated fins, the addition of swords, and various kinds of patterning. In many of these
cases, females preferred models with these novel traits to the model of the typical male. Thus the lack of
ornamented males in this species is not due to lack of interest by females but lack of evolutionary innovation
by males. The females have a pre-existing bias for a whole range of ornaments.

Sounds of beauty
Compared to eyes, ears have not been as rich in the insights they provide into sexual beauty, despite the fact
that acoustic sexual displays are widespread throughout the animal kingdom. An exception is the mating calls
of frogs and the inner ears that respond to them.
There are about 7000 species of frogs, most of them produce loud and conspicuous mating calls, and each
species has its own unique call. There is strong selection on female frogs to mate with males of their own spe-
cies: mating with a male of another species rarely produces viable offspring. This species recognition function
is facilitated by a pair of matched filters in the frog’s inner ear (Gerhardt & Huber, 2002).
Unlike birds and mammals, frogs have two inner ear organs that are sensitive to airborne sounds, the
amphibian papilla (AP) and the basilar papilla (BP). There are a number of differences between these two
end-organs, the most critical one being that the AP is most sensitive to lower frequencies, usually below 1500
Hz, and that the BP is tuned to higher frequencies, usually above 1500 Hz. The tuning of one or both of
these end-organs matches the distribution of spectral energy in that frog’s mating call. The calls of some frogs
have only lower emphasized frequencies, which match the AP; others have only higher emphasized frequen-
cies, which match the tuning of the BP; and some frogs have a pair of emphasized frequencies, one of which
matches the AP and one of which matches the BP (Gerhardt & Schwartz, 2001; Figure 11.2).
The túngara frog has a mating call that is about 300 ms in duration, sweeps from a high to low frequency,
and has a dominant frequency of about 700 Hz—this is the whine. The whine is both necessary and sufficient
to elicit a sexual response from the female, but males can make their calls more attractive by adding chucks, a
sound that is about 35 ms in duration, with a large number of harmonics, and a dominant frequency around
2200 Hz. Males can add up to seven chucks to their whine (Ryan et al., 2019; Figure 11.2). The males have
an unusual larynx that permits these two call components that are acoustically so different from one another.
The whine results from vibration of the vocal cords, as most frog calls are produced, but the chuck results
from vibration of a relatively large fibrous mass that hangs from the vocal cords (Griddi-Papp et al., 2006;
Kime et al., 2019; Figure 11.2). There is congruence between the whine’s dominant frequency and AP tun-
ing and the chuck’s dominant frequency and BP tuning (Ryan et al., 1990). Immediate early gene studies
have shown that the simultaneous stimulation of both inner ear organs results in greater neural activation, as
estimated from gene expression levels, in the frog’s main auditory nucleus, sensorimotor areas, and in some
brain nuclei that are part of the mesolimbic reward system (summarized in Wilczynski & Ryan, 2010). The
whine-chuck call is five times more attractive than that same whine by itself. Unfortunately for the calling

224
Sexual selection and mate choice

Figure 11.2 (A) The túngara frog has an unusual larynx characterized by a large fibrous mass (FM) that protrudes from
the vocal cords (VC). (B) Comparative studies (Ryan, 1990), biomechanical models (Kime et al., 2019),
and ablation of the fibrous mass (Griddi-Papp et al., 2006) all show that the vocal cord vibration is primar-
ily responsible for the whine (blue circle) and the vibration of the fibrous mass for the production of the
chuck (red circle). (C) The dominant frequency of the whine, about 750 Hz, matches the average most
sensitive frequency of the amphibian papilla (AP), while the dominant frequency of the chuck, about 2500
Hz, is a close match to the average most sensitive frequency of the basilar papilla (BP), the location of these
two sensory end-organs in the inner ear are indicated by the blue and red arrows, respectively (Ryan et al.,
1990). (D) Information from the two inner ear-organs enter the brain via the VIIIth cranial nerve. As the
information ascends through the brain it is processed in the sensory, sensorimotor, and motor areas of the
brain and results in movement of the female to the call, i.e., phonotaxis (see details in Wilczynski & Ryan,
2010). (E) Stimulation of both inner ear organs by the whine-chuck results in a fivefold increase in the
attractiveness of the call compared to the whine only (Ryan et al., 2019). Females prefer males making
complex calls; they choose these males as mates; and together with the male, they construct a foam nest
with approximately 250 fertilized eggs (Ryan, 1985) (from Ryan, 2021).

males, chucks also increase the male’s attractiveness to frog-eating bats by about the same proportion (Tut-
tle & Ryan, 1981; Ryan et al., 1982, 2019).
All frogs have both an AP and a BP, even if one of those inner year organs is not used in communication.
There is only one relative of the túngara frog that facultatively adds a chuck-like syllable to its whine-like
introductory call (Boul et al., 2007). The other species all have only whines, and the dominant frequencies
of their whines match their AP tuning. None of these frogs have call frequencies that would substantially
excite the BP. Interestingly, almost all of these frogs have BPs of the same tuning (Wilczynski et al., 2001).
Combining the neurophysiological data on the auditory system with the phylogeny of these frogs and the
calls they produce shows that túngara frogs evolved chucks to match a pre-existing bias in the female’s audi-
tory system; that is, the tuning of the BP (Figure 11.2).

225
Michael J. Ryan

The matched filter hypothesis of anuran acoustic processing (Capranica, 1965; Capranica & Moffat, 1983)
predicts a tight fit between the response properties of the auditory end-organs and the acoustic properties
of the species-specific advertisement call. A similarly tight fit between a neural template somewhere in the
central nervous system and acoustic properties of songs is made about song recognition in birds (Marler,
1997). But that fit might not be so tight.
The permissive aesthetics that Gould et al. uncovered in female mosquitofish has also been demonstrated
in túngara frogs. Recognition of and attraction to the whine appears not to be permissive. Substantial
change to the whine will disrupt its saliency (Rand et al., 1992). The female aesthetics for the chuck, how-
ever, is quite different. Ryan et al. (2007) showed that dozens of sounds attached to the whine in place of
the chuck are as attractive as the chuck. Interestingly, these studies did not reveal any supernormal stimuli,
that is, syllables that are more attractive than the chuck (Ryan et al., 2007). Thus the sexual potency of
the chuck—remember that it increases male attractiveness fivefold—is not unique to this particular sound.
A  variety of other sounds would have been just as effective in increasing the male’s sexual beauty; it is
just the chuck that happened to evolve first. In game theory jargon (Maynard Smith, 1982), the chuck is
an evolutionarily stable strategy in that once established in the population, it cannot be invaded by other
types of calls since they are as attractive but no more attractive than the chuck. Crickets also exhibit hidden
preferences for some call characteristics that have yet to evolve. Gray et al. (2016) showed that two species
of field crickets that produce slow pulse rates have hidden preferences for faster pulse rates than exist in
their species.
The songs of songbirds are much more complex than the calls of frogs or the chirping of crickets.
Although there are some exceptions (Feng et al., 2002), the acoustic repertoires of frogs and crickets usually
consist of only a handful of different components or syllables (Gerhardt & Huber, 2002). Song repertoires in
birds, however, can be quite extensive, and many studies have shown that female songbirds find larger song
repertoires more sexually attractive (Catchpole & Slater, 2003).
Hartshorne (1956, 1973) suggested the monotony threshold hypothesis to explain the evolution of song
repertoire. The hypothesis speaks to territorial interactions between males but, as we will see, is also applica-
ble to female preference for song repertoire. He suggested that if the male produces a simple song defending
his territory, then eventually his neighbours will habituate to that song and aggressive interactions will occur.
Habituation is less likely when a male produces more and different syllables in his songs. There is support
for the monotony threshold hypothesis both in the context of male–male interactions (Kroodsma, 1978) and
female song preferences in songbirds (Searcy, 1992; Catchpole & Slater, 2003; Lyu et al., 2016). Moreover,
there is some understanding as to how increased diversity within a song repertoire influences the underlying
neural and genetic bases for song preferences.
Studies of zebra finches and canaries, both of which habituate to repetition of identical song syllables,
show that both electrophysiological responses and gene expression also habituate to the same repeated song
stimuli (Stripling et al., 1997; Dong & Clayton, 2009). In addition, the molecular changes that are involved
when birds go from experiencing silence to novel calls to habituation are extreme. Exposure to novel songs
results in changes in expression of thousands of RNAs, while habituation leads to a very different gene
expression profile. We have now begun to understand the behavioural, neurophysiological, and gene expres-
sion changes that occur when birds become bored with a male’s song (Dong et al., 2009).

Music
It is tempting to draw comparisons between birdsong and music (e.g., Taylor, 2017). A mascot for this rela-
tionship is the story of Mozart’s starling, which was not only a pet but also a muse for the composer (Haupt,
2017). Both song in birds and music in humans are known to modulate emotions. Centuries ago, Christian

226
Sexual selection and mate choice

Schubart described the emotional correlates of different musical keys in his Ideen zu einer Ästhetik der Tonkunst
(translation by Steblin, 2002):

D Major, The key of triumph, of Hallelujahs, of war-cries, of victory-rejoicing. Thus, the invit-
ing symphonies, the marches, holiday songs and heaven-rejoicing choruses are set in this key;
D Minor, Melancholy womanliness, the spleen and humors brood; F# Minor, A gloomy key:
it tugs at passion as a dog biting a dress. Resentment and discontent are its language, A♭ Major,
Key of the grave. Death, grave, putrefaction, judgment, eternity lie in its radius. And finally,
getting closer to our own concerns, A Major, This key includes declarations of innocent love,
satisfaction with one’s state of affairs; hope of seeing one’s beloved again when parting; youthful
cheerfulness and trust in God; B♭ Major, Cheerful love, clear conscience, hope, aspiration for a
better world.
(from Ryan, 2018, pp. 102–103)

Barlow and his colleagues (Mitchell et al., 1998) took a more experimental approach to demonstrating how
music can influence physiological aspects of sexual motivation. The researchers exposed men to what was
described as happy or sad music prior to watching a pornographic video. The subjects then self-reported
their sexual arousal, and their penis tumescence was measured. Both were more elevated in men who were
exposed to the happy music compared to the men who were exposed to the sad music, even though both
groups of men were viewing the same video.
The effect of music can reach far beyond the sex organs deep into the recesses of the brain. We know that
music elicits responses from various nuclei in the mesolimbic reward system, as does the male song in zebra
finches (Maney, 2013; Spool & Riters, 2017, see the following). Blood and Zatorre (2001) showed that PET
scans detected increased blood flow to various regions of the reward system when the subject reported the
sensation of “shivers down the spine” or “chills” when listening to certain musical pieces. These studies were
confirmed and extended by Menon and Levitin (2005) with fMRIs, which provide more resolution than
PET scans (see also Chapters 7 and 15).
Investigating the neural underpinnings of hedonic pleasures of music is now a major venture in cognitive
neuroscience that is revealing a detailed understanding as to how various parts of the brain respond to music.
For example, Kim et al. (2019), also using fMRI studies, documented the interaction between spectral and
temporal properties of music throughout neural circuits in the frontal cortex. It seems clear that there will
be even more fascinating details of the neural architecture underlying music appreciation as neuroscientists
continue to dig even deeper.

Cognitive biases and sexual beauty


It is somewhat arbitrary to draw a distinction between an animal’s sensory system and its cognitive system.
In cognitive ecology and cognitive ethology, the two fields that apply notions of cognition to animals, the
definition of cognition is usually the acquisition and analysis of information to inform decision-making
(Shettleworth, 1998). Sensory ecology in general, and sensory investigations into the basis of sexual attraction
specifically, have usually grown out of an interaction between neuroethology and behavioural/evolutionary
ecology. But much of these fields also draw inspiration from studies of human cognition (Vasconcelos et al.,
2015). For example, many animal studies are interested in nonlinearities and are modelled on studies of
human psychophysics (e.g., Shepard, 1987). Since mate choice often prefers sexual traits that are greater in
magnitude, the relationship between magnitude and preference has important implications for how evolu-
tion might proceed.

227
Michael J. Ryan

Comparisons in assessments of sexual beauty


Just as sensory systems show nonlinear responses to changes in stimulus magnitude, cognitive evaluations of
difference in magnitudes between stimuli can also be nonlinear. An example of this is Weber’s law. We know
in humans that the magnitude of difference between stimuli to allow a just noticeable difference ( JND) fol-
lows a power law such that the greater the magnitude of the stimuli, the greater the difference between them
is needed for a JND. Animal studies have adopted the notion of a JND into a more naturalistic concept of
the JMD, the just meaningful difference (Nelson & Marler, 1990). We know that in a variety of tasks, such
as foraging decisions, the JMD follows a Weber function (Akre & Johnsen, 2014). If the same were true for
traits involved in sexual attraction, this would have important consequences for the tempo of evolution of
these traits.
As noted previously, túngara frogs can add chucks to the basic whine component of their mating call to
increase their attractiveness. Although a single chuck increases the attraction of the call fivefold, additional
chucks further increase call attractiveness. Akre et al. (2011) asked if the sexual attraction of these complex
calls increased linearly with additional chucks or if, alternatively, they followed a Weber function. The results
were quite clear. Additional chucks did increase attractiveness, but the relationship between attractiveness and
chuck number followed a Weber function. The difference in attractiveness of calls with two versus one chuck
was much greater than the difference in attractiveness between calls with five versus four chucks, for example.
Also, as noted previously, frog-eating bats are attracted to the túngara frog call and are more attracted to
calls with chucks than without chucks. Interestingly, this animal with a very different auditory system and a
much more complicated brain than the túngara frog followed a statistically identical function in their attrac-
tion to calls with more chucks. Thus whether it is frogs searching for a mate or bats searching for a meal, the
attraction of more complex calls followed similar power functions.
We know that in sexual selection, choosers can exert strong directional selection on traits to become more
attractive; usually the traits respond to sexual selection by evolving an increase in some component of mag-
nitude, conspicuousness, or complexity. But sexual selection is opposed by natural selection since these traits
also have higher costs, whether they be metabolic, developmental, or due to predation risk. The study of the
túngara frog shows that given the power function underlying female attraction to these calls, there might also
be a cognitive brake on the evolution of more attractive traits. As the peacock tail gets longer, it requires an
even greater increment in size to become more attractive to its competitors.
Studies of animal decision-making also draw inspiration from what economists sometimes call irrational
behaviours and have been especially influenced by prospect theory as developed by Tversky and Kahneman
(e.g., Tversky & Kahneman, 1989, 1992; Kahneman & Tversky, 2013). This latter cross-fertilization comes
about because both behavioural economics and evolutionary biology predict that individuals should behave
in a manner that maximizes some utility, monetary or happiness gains in humans, and Darwinian fitness gains
in animals. Irrational behaviours are therefore interesting either as test of the fundamental assumptions in
both of these fields or, more commonly, examples of the exceptions that might prove the rules (Vasconcelos
et al., 2015).
One of the major assumptions of economic rationality is that utilities assigned to different objects are not
influenced by irrelevant alternatives. For instance, an individual is given the following choice: a trip to Paris
versus a trip to Rome, both with a free breakfast included. She may not show a strong preference between
the two. If a third option is added, a trip to Rome without a free breakfast, then the trip to Rome with a
free breakfast quickly becomes the preferred option (Ariely, 2009). This is considered a violation of eco-
nomic rationality because the relative preference between two options should not change in the presence of
an irrelevant third option.
Numerous studies have shown animals are susceptible to the influence of irrelevant alternatives in the
context of foraging, but only recently have these ideas been applied to female decisions based on a male’s

228
Sexual selection and mate choice

sexual attraction (Bateson & Healy, 2005). The túngara frog again was a useful model for addressing socially
induced variation in perceived attractiveness. Lea and Ryan (2015) identified three natural mating calls of
males that differed in static attractiveness, A, B, and C. They experimentally manipulated the call rate of
each of these calls (females are usually more attracted to calls produced at a more rapid rate). Thus, the three
calls varied over two dimensions of attractiveness: static attractiveness and call rate. Preliminary tests showed
that call C was the least attractive. In one set of phonotaxis experiments, there was no significant preference
between calls A and B in a binary choice test, but when C was added in a trinary choice tests, there was a
significant preference for call A. In another set of experiments, there was a preference for call A versus B, but
when call C was added in a trinary choice test, the preference for A vanished. It is an interesting paradox that
what one finds attractive can be influenced by what one finds unattractive.
Weber’s law and departures from regularity add an interesting twist to understanding what goes on in an
animal’s head when it is making decisions about sexual beauty. It seems the further we get from the sense
organs, the more complex the interpretations.

Perceptual fluency and processing biases


Revisiting von Uexküll, we must remember not only that different animals can have different perceptions of
their environment but their environments can also have important effects on how they perceive and extract
meaning from the world around them. Inevitably, some objects will be perceptually processed more easily
than others. Recent studies suggest that ease of perceptual processing may deliver hedonic rewards, poten-
tially revealing the biochemical basis of appreciation of some types of beauty (for further discussions of this
literature, see also Chapters 6 and 26).
Reber at al. (2004) argued that in humans, ease of processing, or perceptual fluency, can result in aesthetic
pleasures. They make several points. First, they argue that objects will vary in the fluency with which they
are processed. Interestingly, features that are highlighted by objectivist theories of beauty, such as symmetry
and contrasts with background, also enhance perceptual fluency. The second point the authors make is that
processing fluency is, in their words, “hedonically marked”; that is, interacting with such object results in
aesthetic pleasure, as has been illustrated in psychophysiological studies (Reber et al., 2004). Renoult and
Mendelson (2019) extend this idea of perceptual fluency to mate attraction. They suggest that preference
for certain sexual traits can be emotionally rewarding if those traits exploit a processing bias due to efficient
information processing. This idea of a processing bias is an important extension to theories of sensory and
perceptual biases in mate choice (e.g., Ryan & Cummings, 2013).
The first test of this general idea that perceptual fluency is related to attractiveness was addressed in a
study of human facial attractiveness by Renoult et al. (2016). The type of perceptual fluency studied by these
authors was efficient neural coding, specifically sparse coding. Sparse coding is the representation of an object
by the activation of a relatively small number of neurons—the more sparse the coding, the more efficient
it is. Renoult et al. modelled the activity of simple cells in V1, the primary visual cortex, of humans, as it
processed images of women’s faces. They found that the sparseness of coding in the model was significantly
and positively correlated with men’s evaluations of the attractiveness of these same faces. The sparseness of
the code predicted 17% of the variation in the attractiveness scores given by men. The authors also point out
that sparse coding in the primary visual cortex is not an adaptation for spatial facial recognition but probably
evolved to solve the more general problem of extracting patterns from complex natural scenes.
If sparse coding in visual systems evolved to analyse natural scenes, then the correlation between sparse
coding and facial attraction suggests that some aspects of facial phenotypes might have evolved to exploit
sparse coding that was already in place for its more general domain function. This hypothesis fits well
with studies of the composition of natural scenes and the patterns of letters utilized in different languages.

229
Michael J. Ryan

Changizi and his colleagues (2006; Changizi, 2010) posit that the shapes most commonly used for letters
across languages should be drawn from patterns that are most abundant in natural scenes. They analysed 93
speechwriting systems and found that the average number of strokes per letter was three. This was a fairly
close match to the average in natural visual scenes. They also showed, quite convincingly, that there is an
almost perfect correlation between the use of 19 visual patterns in these alphabets and in natural visual
scenes. This study highlights an interesting interaction between the environment and culture. The environ-
ment selected for processing biases that could extract patterns from natural scenes; later in human evolution,
culture resulted in different groups using these patterns in their writing systems. This scenario has striking
similarities to sexual selection by sensory exploitation (Ryan, 1990; Verpooten & Nelissen, 2012; Renoult &
Mendelson, 2019).
Returning to beauty, there seems to be an interesting relationship between patterns in natural scenes and
artists’ portraits of human faces. Natural scenes are sometimes characterized as scale invariant. Redies et al.
(2007) showed that artists often enhance the attractiveness of human faces, which are not scale invariant,
with patterns that approximate the scale-invariant properties of natural scenes. Again, this might be another
example of a cultural expression being shaped by a domain general processing bias.
Mendelson and her colleagues (Hulse et al., 2020) apply a similar rationale to the evolution of patterns
of male courtship traits in a group of fish called darters. As with other studies, they use Fourier analysis to
decompose spatial patterns in natural scenes where the fish reside to the spatial patterns across the bodies of
both males and females. The slope of the log-log plot (power as a function of frequency) of the Fourier spec-
trum was used to characterize this pattern (Hulse et al., 2020). They showed that there was strong concord-
ance between the habitat statistics and the male patterning statistics but not the female patterning statistics.
They argued quite sensibly that camouflage was an unlikely explanation for these results; if that were the case,
one would expect both males and females to be equally advantaged by camouflage. Instead, they argue that
by matching the habitat statistics these male patterns were more attractive to females because of the hedonic
pleasure of perceptual fluency.
The previous studies suggest that our visual processing systems have evolved to extract information from
natural visual scenes and that letters in various alphabets, portrait paintings, and patterns of courtship traits in
some fish can all be explained to some degree by their match to natural scenes. Can the same argument be
applied to natural acoustic scenes?
Such an approach was adopted by Purves and his colleagues (Schwartz et al., 2003; Bowling & Purves,
2015). Much as Changizi showed how natural visual scenes influenced the details of a cultural adaptation,
letters in alphabets, Purves et al. asked how natural acoustic scenes influenced universals in another cultural
trait, music (Mehr et al., 2019). These authors pointed out that the dominant acoustic scene of humans was
human speech.
The frequency spectrum of human speech sounds results from an interaction of the spectrum at the source,
which generate triangle waves which are then filtered by the resonance of the vocal tract. The researchers
normalized the amplitude of all the frequencies to that frequency with the greatest amplitude. The resulting
Fourier transform illustrates frequencies as ratios relative to this loudest frequency (Figure 11.3). The result-
ing spectra were almost identical for speech in the four languages that they examined.
Sound frequency is a continuous variable, but in musical scales, these frequencies are categorized into dis-
crete tones. In Western music terminology, the chromatic scale has 12 tones within an octave (Figure 11.3).
Each of these tones can be assigned a frequency ratio such that the unison tone by definition is 1.0, while the
fifth is 1.5 and the octave is 2.0. When one superimposes the ratios of these notes onto the power spectrum
of human speech, there is remarkable concordance (Figure 11.3). Thus the tones used in the chromatic scale,
as well as other scales, were chosen to match the power output of human speech sounds.
The acoustic structure of speech sounds not only predicts musical tones but also our preference for such
tones. Studies of the consonance of two-tone notes have been conducted over more than a century and show

230
Sexual selection and mate choice
231

Figure 11.3 (A) A power spectrum of human speech normalized to the frequency with the greatest amplitude (B) A portion of a piano keyboard indicating the
chromatic scale tones over one octave, their names, and their frequency ratios with respect to the tonic in one of the major tuning systems, “Just
Intonation,” used in Western music. (C) The majority of the musical intervals of the chromatic scale (arrows) correspond to the mean amplitude peaks
in the normalized spectrum of human speech sounds, shown here over a single octave. (D) Consonance rankings of musical dyads as a function of the
normalized spectrum of speech sounds (from Schwartz et al., 2003).
Michael J. Ryan

repeatable results. For example, the octave, the fifth, and the fourth are quite pleasing to our ears, while the
major second, the major seventh, and the minor second are less so. When one compares the normalized
amplitude of the higher of the two tones in the musical dyad to its consonance ranking, there is also a strong
correlation (Figure 11.3). It is not clear how this finding applies to non-Western music cultures.
This concept of perceptual fluency is a new one in addressing animal sexual beauty. It loans further cre-
dence to Darwin’s idea of animal aesthetics by suggesting that choosers might be attracted to the traits of
some courters for the mere pleasure it delivers to them, perhaps similar to why we might be attracted to
certain pieces of art or music.

Incentive salience in mate choice decision-making


Rosenthal (2018) correctly pointed out that sensory, perceptual, and cognitive biases can be important com-
ponents of the chooser’s preference for attractive sexual traits, but these are not sufficient to account for an
animal’s sexual aesthetics. A bright red cap on a courting male bird might be sexually attractive to a female,
but a bright red band on a coral snake would elicit aversion from the same bird even if the stimulation of
long wavelength cones were identical in both contexts. He points out that there are a variety of evaluat-
ing mechanisms that can result in positive, neutral, or negative hedonic value to a stimulus. Further, these
evaluative mechanisms can vary among and within individuals at any point in time, and these differences
can be modified by genetic, environmental, or internal physiological mechanisms. One of these evaluative
mechanisms that is quite important in hedonic evaluation is the mesolimbic reward system (see Chapter 2).
Lynch and Ryan (2020) recently discussed the potential role of incentive salience in sexual selection by
mate choice per se, not merely its more general involvement in reproductive behaviour. This section borrows
heavily from that review.
In the context of an animal’s sexual aesthetics, we are interested in the role of dopamine in modulating the
animal’s “wanting” rather than “liking” or “learning” of a stimulus. This distinction was illustrated by studies
of rats in which Berridge and colleagues showed that dopamine antagonists did not influence the animal’s
liking of a food reward but did influence the degree that the animal would work to obtain such a reward, that
is, the degree to which the animal wanted the food (reviewed in Kringelbach & Berridge, 2012). Berridge
(2007, 2019; see also Chapter 3) reviews evidence for the roles of dopamine in liking, learning, and wanting
and suggests that the evidence strongly supports a role in modulating wanting. Dopamine assigns incentive
salience to an object or action. If we apply this role of dopamine to the context of mate choice, it can result
in reward-seeking behaviour toward sexual signals of some males over others. Dopamine alone, however, is
likely not sufficient to stimulate reward seeking toward sexual signals, as a sexually motivational state must
first be initiated by reproductive steroids such as oestrogen. In this case, once oestrogen (and other reproduc-
tive steroids) heightens a female’s interest in sexual stimuli, dopamine may act to “mark” salience of signals
and stimulate pursuit of those signals. It is doubtful that dopamine alone, without elevated reproductive ster-
oids, would play this role in sexual behaviour. For example, in the few studies that examine dopamine’s role
in incentive salience during sexual decision-making, the animals tested were either treated with exogenous
steroid hormones or peptides that increase steroid hormones, or the animals were in a natural breeding con-
dition, indicating they had elevated reproductive hormones.
Lynch and Ryan (2020) reviewed three types of studies that speak to the potential role of dopamine in
mate choice: (1) dopamine agonist or antagonist administered to females during mate choice, (2) measures
of neural activity in dopaminergic neural circuits during mate choice, and (3) social regulation of dopamine
in females when exposed to mate signals.
One avenue for exploring the role of dopamine in mate choice is to treat females with dopamine agonists
or antagonists immediately prior to mate choice. Riters and her colleagues (Pawlisch & Riters, 2010) took

232
Sexual selection and mate choice

this approach to investigate responses, courtship solicitation displays, to male courtship song in reproductively
primed female European starlings (Sternus vulgaris). They stimulated dopamine receptors with peripheral
injections of a non-selective dopamine reuptake inhibitor (GBR-12909, an indirect dopamine receptor ago-
nist). Dopamine eliminated selective female responses to male song so much that it induced female responses
to heterospecific as well as conspecific songs.
These results suggest that dopamine transforms a previously unattractive stimulus into one that initiates an
appetitive sexual response from the female. In a recent preliminary study of female tùngara frogs dopamine
also similarly expanded the female’s preference landscape. These researchers administered a non-selective
dopamine agonist (apomorphine) to reproductively active females just prior to exposure to a synthetic non-
conspecific mating call known to be unattractive from a previous study (Ryan & Rand, 2003). None of the
control females responded to this typically unattractive call, which was also true in the earlier study, while
20% of the females treated with apomorphine showed a positive phonotaxis response to those same calls. This
is another case in which dopamine plays a role in expanding the females’ preference landscape (Ryan et al.,
unpublished data), but it is not known whether this occurs via a learning, liking, or wanting mechanism.
Nonetheless, this result raises the interesting possibility that dopamine might play a creative role in the evolu-
tion of the amazing diversity of courtship traits that we see in nature.
Another approach to documenting the role of dopamine in sexual aesthetics is to measure proxies of neu-
ral activity in dopaminergic neural circuits when individuals are presented with members of the opposite sex.
In humans, the presentation of a picture of one’s sexual partner increased activation of dopaminergic
circuitry in an fMRI study (Fisher et al., 2005). In a related study, homosexual and heterosexual men and
women rated the attractiveness of faces, and fMRI results showed similar amounts of activation in areas of the
brain involved in face recognition among the four groups. But when reward areas of the brain were meas-
ured, there was variation among subjects and sexual orientation; heterosexual women and homosexual men
both showed enhanced activation of the reward circuit when they viewed men’s faces, while homosexual
women and heterosexual men showed enhanced activation of the same circuit when they viewed women’s
faces (Kranz & Ishai, 2006). These results indicate the reward system exhibits sexual preference–based rather
than gender-based responses. It also highlights the difference between liking and wanting. In a sense, men
and women, heterosexuals and homosexuals, exhibited similar liking of faces, but wanting of faces was a
function of sexual preference.
In nonhuman animals, studies that have probed reward circuits often utilize activity-dependent genes such
as immediate early genes (IEG) as a proxy for neural excitation. Such studies have shown that female white-
throated sparrows (Zonotrichia albicollis) and túngara frogs exhibit a correlation between neural activity in the
mesolimbic reward system after exposure to conspecific mate signals (Maney, 2013; Chakraborty et al., 2010;
Hoke et al., 2010). While we do not know if these IEG responses are specific to dopaminergic cells within
the mesolimbic reward system, these studies do implicate the reward system in assessment of mating signals.
These studies, and some others, are noteworthy because they specifically examine how these reward circuits
respond to different sexual signals (Lynch & Ryan, 2020).
Not only does dopamine seem to regulate the female’s response to the male’s sexual signals, there is
a reciprocal interaction of the male’s sexual signals socially regulating the female’s dopamine levels. Alger
et al. (2011) showed in female zebra finches that tyrosine hydroxylase (TH) labelling density, indicative of
dopamine-containing cells, in two dopamine-dense nuclei of the reward circuit, the ventromedial nucleus
of the hypothalamus and the rostral ventral tegmental area (VTA), were positively related to the amount of
courtship received from their partners. While elevated TH within reward circuitry of female zebra finches
does not respond with increased IEG expression to simple playback of songs (Svec et al., 2009; Lynch et al.,
2012), other regions containing TH do respond during simple song playback experiments, particularly the
locus coeruleus, a region that is predominately noradrenergic in nature (Lynch et al., 2012).

233
Michael J. Ryan

Finally, in a study of plainfin midshipmen, Sisneros and co-workers (Forlano et al., 2017) showed that
female responses to advertisement calls are correlated with the activation of a specific subset of catecholamin-
ergic (CA) and social decision-making network (SDM) nuclei underlying auditory-driven sexual motivation.
The dopaminergic reward system has been intensively studied and is known to stamp in incentive salience
in a number of domains. Moreover, there has been intensive study of the role of dopamine in consummatory
sexual behaviour. There has been surprisingly little consideration, however, of how dopamine, particularly
within the mesolimbic reward system, can assign incentive salience to courtship signals. In other words, our
understanding of how dopamine acts during the decision-making or appetitive phases of sex is lagging. The
data that are available show that it plays a similar role in assessing sexual signals as it does in some other systems
such as human appreciation for loved ones and music. Some studies also suggest that dopamine might play
a creative role in the evolution of sexual displays by expanding the stimuli that females find desirable. This
attraction of previously unencountered stimuli might have some analogies to the involvement of dopamine
in obsessive and addictive behaviour of humans such as drug use (Volkow et al., 2011), gambling (Comings
et al., 1996), overeating (Stice et al., 2010), and compulsive viewing of pornography (Hilton, 2013).

Major challenges, goals, and suggestions


Darwin presented his theory of sexual selection, with its emphasis on sexual aesthetics, 150 years ago. But
it was not until 50 years ago that there was a renewed interest in Darwin’s theory in general, specifically the
potency of mate choice in influencing the evolution of sexual beauty. Initially, studies concentrated on what
type of information sexual signals contain about their bearer. More recently, there is a new interest in the
mechanisms that form the bases of sexual aesthetics. Detailed and controlled manipulations of sexual traits
started to give some insights as to why these traits are sexually attractive. These interests led to studies of the
brain using both electrophysiological and gene expression techniques to give us some understanding of how
sexual traits are processed and why they are so appreciated.
More recently, however, this research area has been strongly influenced by studies of humans in fields such as
psychophysics (e.g., Weber’s law), behavioural economics (e.g., competitive decoys), and neuroaesthetics (e.g.,
perceptual fluency). Many of the questions are similar; for example, why is it animals find that certain visual
and acoustic patterns enhance the sexual attractiveness of a mate, and to what extent do these traits themselves
induce hedonic pleasure? Researchers in sexual selection, most of whom have a firm grounding in naturalistic
studies of animals in the wild, can benefit greatly by immersing themselves in these parallel disciplines.
The elephant in the room, or at least the elephant wandering through this review, is whether non-human
animals and humans share the same aesthetics. There is no doubt that sexual beauty and appreciation for
sexual beauty are similar to the manifestations of liking and wanting in humans who are exposed to potential
sexual partners, music, and art. Are these systems homologous, shared through a common ancestor, or do
they represent convergent similarities? Certainly the colours of bird plumage and fish scales have no shared
history with one another, let alone with the paints that artists put to canvas. But all of these animals have
some similarities in how they sense the world around them and how their brains begin to process informa-
tion. Importantly, we should keep in mind that all vertebrates, including us, have homologous reward and
decision-making systems (O’Connell & Hofmann, 2012). Is sexual aesthetics a phenomenon that exhibits
continuity between other animals and us, or are these common solutions to similar problems? This question
is still open for debate (Skov, unpublished; White, 2020).

References
Akre, K. L., Farris, H. E., Lea, A. M., Page, R. A., & Ryan, M. J. (2011). Signal perception in frogs and bats and the
evolution of mating signals. Science, 333(6043), 751–752. https://doi.org/10.1126/science.1205623

234
Sexual selection and mate choice

Akre, K. L., & Johnsen, S. (2014). Psychophysics and the evolution of behaviour. Trends in Ecology and Evolution, 29(5),
291–300. https://doi.org/10.1016/j.tree.2014.03.007
Alger, S. J., Juang, C., & Riters, L. V. (2011). Social affiliation relates to tyrosine hydroxylase immunolabeling in male and
female zebra finches (Taeniopygia guttata). Journal of Chemical Neuroanatomy, 42(1), 45–55. https://doi.org/10.1016/j.
jchemneu.2011.05.005
Andersson, M. (1994). Sexual selection. Princeton University Press.
Archer, S. N., Endler, J. A., Lythgoe, J. N., & Partridge, J. C. (1982). Visual pigment polymorphism in the guppy Poecilia
reticulata. Vision Research, 8, 1243–1252.
Ariely, D. (2009). Predictably irrational (rev. and expanded ed). HarperCollins Publishers.
Arnqvist, G., & Rowe, L. (2005). Sexual conflict. Princeton University Press.
Basolo, A. L. (1990a). Female preference for male sword length in the green swordtail, Xiphophorus helleri (Pisces: Poecili-
idae). Animal Behaviour, 40(2), 332–338. https://doi.org/10.1016/S0003-3472(05)80928-5
Basolo, A. L. (1990b). Female preference predates the evolution of the sword in swordtail fish. Science, 250(4982),
808–810. https://doi.org/10.1126/science.250.4982.808
Basolo, A. L. (1995). Phylogenetic evidence for the role of a pre-existing bias in sexual selection. Proceedings of the Royal
Society of London. Series B: Biological Sciences, 259(1356), 307–311. https://doi.org/10.1098/rspb.1995.0045
Bateman, A. J. (1948). Intrasexual selection in Drosophila. Heredity, 2(3), 349–368. https://doi.org/10.1038/hdy.
1948.21
Bateson, M., & Healy, S. D. (2005). Comparative evaluation and its implications for mate choice. Trends in Ecology and
Evolution, 20(12), 659–664. https://doi.org/10.1016/j.tree.2005.08.013
Berridge, K. C. (2007). The debate over dopamine’s role in reward: The case for incentive salience. Psychopharmacology,
191(3), 391–431. https://doi.org/10.1007/s00213-006-0578-x
Berridge, K. C. (2019). Affective valence in the brain: Modules or modes? Nature Reviews. Neuroscience, 20(4), 225–234.
https://doi.org/10.1038/s41583-019-0122-8
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Boul, K. E., Funk, W. C., Darst, C. R., Cannatella, D. C., & Ryan, M. J. (2007). Sexual selection drives speciation
in an Amazonian frog. Proceedings of the Royal Society Biological Sciences Series B, 274(1608), 399–406. https://doi.
org/10.1098/rspb.2006.3736
Bourne, G. R., & Watson, L. C. (2009). Receiver-bias implicated in the nonsexual origin of female mate choice in the
pentamorphic fish Poecilia parae Eigenmann, 1894. Aquaculture, Aquarium, Conservation & Legislation, 2(3), 299–317.
Bowling, D. L., & Purves, D. (2015). A biological rationale for musical consonance. Proceedings of the National Academy of
Sciences, 112(36), 11155–11160. https://doi.org/10.1073/pnas.1505768112
Capranica, R. R. (1965). The evoked vocal response of the bullfrog—A study of communication by sound. MIT Press.
Capranica, R. R., & Moffat, A. J. (1983). Neurobehavioral correlates of sound communication in anurans. In Advances
in vertebrate neuroethology (pp. 701–730). Springer.
Catchpole, C. K., & Slater, P. J. (2003). Bird song: Biological themes and variations. Cambridge University Press.
Caves, E. M., Nowicki, S., & Johnsen, S. (2019). Von Uexküll revisited: Addressing human biases in the study of animal
perception. Integrative and Comparative Biology, 59(6), 1451–1462. https://doi.org/10.1093/icb/icz073
Chakraborty, M., Mangiamele, L. A., & Burmeister, S. S. (2010). Neural activity patterns in response to interspecific and
intraspecific variation in mating calls in the túngara frog. PLoS One, 5(9), e12898. https://doi.org/10.1371/journal.
pone.0012898
Changizi, M. A. (2010). The vision revolution: How the latest research overturns everything we thought we knew about human
vision. BenBella Books.
Changizi, M. A., Zhang, Q., Ye, H., & Shimojo, S. (2006). The structures of letters and symbols throughout human
history are selected to match those found in objects in natural scenes. The American Naturalist, 167(5), E117-–E139.
https://doi.org/10.1086/502806.
Cole, G. L., & Endler, J. A. (2015). Artificial selection for food colour preferences. Proceedings of the Royal Society B:
Biological Sciences, 282(1804), 20143108. https://doi.org/10.1098/rspb.2014.3108
Comings, D. E., Rosenthal, R. J., Lesieur, H. R., Rugle, L. J., Muhleman, D., Chiu, C., Dietz, G., & Gade, R. (1996).
A study of the dopamine D2 receptor gene in pathological gambling. Pharmacogenetics, 6(3), 223–234. https://doi.
org/10.1097/00008571-199606000-00004
Cronin, H. (1991). The ant and the peacock: Altruism and sexual selection from Darwin to today. Cambridge University Press.
Cui, R., Schumer, M., Kruesi, K., Walter, R., Andolfatto, P., & Rosenthal, G. G. (2013). Phylogenomics reveals exten-
sive reticulate evolution in Xiphophorus fishes. Evolution, 67(8), 2166–2179. https://doi.org/10.1111/evo.12099

235
Michael J. Ryan

Cummings, M. E. (2007). Sensory trade-offs predict signal divergence in surfperch. Evolution, 61(3), 530–545. https://
doi.org/10.1111/j.1558-5646.2007.00047.x
Cummings, M. E., & Endler, J. A. (2018). 25 years of sensory drive: The evidence and its watery bias. Current Zoology,
64(4), 471–484. https://doi.org/10.1093/cz/zoy043
Darwin, C. (1859). On the origin of species. Murray.
Darwin, C. (1860). Letter 2743 of Darwin correspondence project. Cambridge University Press. http://www.darwinproject.
ac.uk/entry2743.
Darwin, C. (1871). The descent of man and selection in relation to sex. J. Murray.
Darwin, C. (1872). The expression of the emotions in man and animals. J. Murray.
Dong, S.,  & Clayton, D. F. (2009). Habituation in songbirds. Neurobiology of Learning and Memory, 92(2), 183–188.
https://doi.org/10.1016/j.nlm.2008.09.009
Dong, S., Replogle, K. L., Hasadsri, L., Imai, B. S., Yau, P. M., Rodriguez-Zas, S., Southey, B. R., Sweedler, J. V., &
Clayton, D. F. (2009). Discrete molecular states in the brain accompany changing responses to a vocal signal. Proceed-
ings of the National Academy of Sciences, 106(27), 11364–11369. https://doi.org/10.1073/pnas.0812998106
Emlen, D. J. (2014). Animal weapons: The evolution of battle. Henry Holt, and Company.
Endler, J. A. (1983). Natural and sexual selection on colour patterns in poeciliid fishes. Environmental Biology of Fishes,
9(2), 173–190. https://doi.org/10.1007/BF00690861
Endler, J. A. (1992). Signals, signal conditions and the direction of evolution. American Naturalist, 139, S125–S153.
https://doi.org/10.1086/285308
Endler, J. A., & Basolo, A. L. (1998). Sensory ecology, receiver biases and sexual selection. Trends in Ecology and Evolution,
13(10), 415–420. https://doi.org/10.1016/s0169-5347(98)01471-2
Endler, J. A., & Houde, A. E. (1995). Geographic variation in female preferences for mail traits in Poecilia reticulata. Evolu-
tion, 49(3), 456–468. https://doi.org/10.1111/j.1558-5646.1995.tb02278.x
Feng, A. S., Narins, P. M., & Xu, C. H. (2002). Vocal acrobatics in a Chinese frog, Amolops tormotus. Naturwissenschaften,
89(8), 352–356. https://doi.org/10.1007/s00114-002-0335-x
Fisher, H., Aron, A., & Brown, L. L. (2005). Romantic love: An fMRI study of a neural mechanism for mate choice.
Journal of Comparative Neurology, 493(1), 58–62. https://doi.org/10.1002/cne.20772
Foisy, M. (2017). Phylogenetic and experimental evidence for an evolutionary precursor to male colouration in poeciliid fishes and their
relatives [MS Thesis, University of Toronto, Toronto].
Forlano, P. M., Licorish, R. R., Ghahramani, Z. N., Timothy, M., Ferrari, M., Palmer, W. C., & Sisneros, J. A. (2017).
Attention and motivated response to simulated male advertisement call activates forebrain dopaminergic and social
decision-making network nuclei in female midshipman fish. Integrative and Comparative Biology, 57(4), 820–834.
https://doi.org/10.1093/icb/icx053
Gerhardt, H. C., & Huber, F. (2002). Acoustic communication in insects and anurans. University of Chicago Press.
Gerhardt, H. C., & Schwartz, J. J. (2001). Auditory tunings and frequency preferences in anurans. In M. J. Ryan (Ed.),
Anuran communication (pp. 73–85). Smithsonian Institution Press.
Gould, J. L., Elliott, S. L., Masters, C. M., & Mukerji, J. (1999). Female preferences in a fish genus without female mate
choice. Current Biology, 9(9), 497–500. https://doi.org/10.1016/s0960-9822(99)80217-6
Gray, D. A., Gabel, E., Blankers, T., & Hennig, R. M. (2016). Multivariate female preference tests reveal latent per-
ceptual biases. Proceedings of the Royal Society B: Biological Sciences, 283(1842), 20161972. https://doi.org/10.1098/
rspb.2016.1972
Gridi-Papp, M., Rand, A. S., & Ryan, M. J. (2006). Complex call production in túngara frogs. Nature, 441(7089), 38.
https://doi.org/10.1038/441038a
Hartshorne, C. (1956). The monotony-threshold in singing birds. Auk, 73(2), 176–192. https://doi.org/10.2307/4081470
Hartshorne, C. (1973). Born to sing. Indiana University Press.
Haupt, L. L. (2017). Mozart’s starling. Hachette.
Hilton, D. L. (2013). Pornography addiction—A supranormal stimulus considered in the context of neuroplasticity.
Socioaffective Neuroscience and Psychology, 3(1), 20767. https://doi.org/10.3402/snp.v3i0.20767
Hoke, K. L., Ryan, M. J., & Wilczynski, W. (2010). Sexually dimorphic sensory gating drives behavioural differences in
túngara frogs. Journal of Experimental Biology, 213(20), 3463–3472. https://doi.org/10.1242/jeb.043992
Houde, A. (1997). Sex, colour, and mate choice in guppies. Princeton University Press.
Hulse, S. V., Renoult, J. P., & Mendelson, T. C. (2020). Sexual signaling pattern correlates with habitat pattern in visually
ornamented fishes. Nature Communications, 11(1), 2561. https://doi.org/10.1038/s41467-020-16389-0
Janicke, T., Häderer, I. K., Lajeunesse, M. J., & Anthes, N. (2016). Darwinian sex roles confirmed across the animal
kingdom. Science Advances, 2(2), e1500983. https://doi.org/10.1126/sciadv.1500983

236
Sexual selection and mate choice

Kahneman, D.,  & Tversky, A. (2013). Prospect theory: An analysis of decision under risk. In Handbook of the fun-
damentals of financial decision making: Part I (pp.  99–127). World Scientific. https://doi.org/10.1142/9789814417
358_0006
Kim, S. G., Mueller, K., Lepsien, J., Mildner, T., & Fritz, T. H. (2019). Brain networks underlying aesthetic appreciation
as modulated by interaction of the spectral and temporal organisations of music. Scientific Reports, 9(1), 19446. https://
doi.org/10.1038/s41598-019-55781-9
Kime, N. M., Ryan, M. J.,  & Wilson, P. S. (2019). Modelling the production of complex calls in the túngara frog
(Physalaemus pustulosus). Bioacoustics, 28(4), 345–363. https://doi.org/10.1080/09524622.2018.1458249
Kranz, F., & Ishai, A. (2006). Face perception is modulated by sexual preference. Current Biology, 16(1), 63–68. https://
doi.org/10.1016/j.cub.2005.10.070
Kringelbach, M. L.,  & Berridge, K. C. (2012). The joyful mind. Scientific American, 307(2), 40–45. https://doi.
org/10.1038/scientificamerican0812-40
Kroodsma, D. E. (1978). Continuity and versatility in bird song: Support for the monotony–threshold hypothesis. Nature,
274(5672), 681–683. https://doi.org/10.1038/274681a0
Lea, A. M., & Ryan, M. J. (2015). Irrationality in mate choice revealed by túngara frogs. Science, 349(6251), 964–966.
https://doi.org/10.1126/science.aab2012
Lynch, K. S., Diekamp, B., & Ball, G. F. (2012). Colocalization of immediate early genes in catecholamine cells after
song exposure in female zebra finches (Taeniopygia guttata). Brain, Behavior and Evolution, 79(4), 252–260. https://doi.
org/10.1159/000337533
Lynch, K. S., & Ryan, M. J. (2020). Understanding the role of incentive salience in mate choice decision-making. Integra-
tive and Comparative Biology. https://doi.org/10.1093/icb/icaa094
Lyu, N., Li, J., & Sun, Y.-H. (2016). Can simple songs express useful signals for mate choice? Avian Research, 7(1), 10.
https://doi.org/10.1186/s40657-016-0045-2
Maan, M. E., Hofker, K. D., van Alphen, J. J., & Seehausen, O. (2006). Sensory drive in cichlid speciation. American
Naturalist, 167(6), 947–954. https://doi.org/10.1086/503532
Maney, D. L. (2013). The incentive salience of courtship vocalizations: Hormone-mediated “wanting” in the auditory
system. Hearing Research, 305, 19–30. https://doi.org/10.1016/j.heares.2013.04.011
Marler, P. (1997). Three models of song learning: Evidence from behaviour. Journal of Neurobiology, 33(5), 501–516.
https://doi.org/10.1002/(SICI)1097-4695(19971105)33:5<501::AID-NEU2>3.0.CO;2-8
Maynard Smith, J. (1982). Evolution and the theory of games. Cambridge University Press.
Mehr, S. A., Singh, M., Knox, D., Ketter, D. M., Pickens-Jones, D., Atwood, S., Lucas, C., Jacoby, N., Egner, A. A., &
Hopkins, E. J. (2019). Universality and diversity in human song. Science, 366(6468). https://doi.org/10.1126/science.
aax0868
Menon, V., & Levitin, D. J. (2005). The rewards of music listening: Response and physiological connectivity of the mes-
olimbic system. Neuroimage, 28(1), 175–184. https://doi.org/10.1016/j.neuroimage.2005.05.053
Mitchell, W. B., DiBartolo, P. M., Brown, T. A.,  & Barlow, D. H. (1998). Effects of positive and negative mood
on sexual arousal in sexually functional males. Archives of Sexual Behavior, 27(2), 197–207. https://doi.
org/10.1023/a:1018686631428
Nelson, D. A., & Marler, P. (1990). The perception of birdsong and an ecological concept of signal space. In W. C. Steb-
bins & M. A. Berkeley (Eds.), Comparative perception: Complex signals. John Wiley & Sons.
O’Connell, L. A.,  & Hofmann, H. A. (2012). Evolution of a vertebrate social decision-making network. Science,
336(6085), 1154–1157. https://doi.org/10.1126/science.1218889
Pawlisch, B. A.,  & Riters, L. V. (2010). Selective behavioural responses to male song are affected by the dopamine
agonist GBR-12909 in female European starlings (Sturnus vulgaris). Brain Research, 1353, 113–124. https://doi.
org/10.1016/j.brainres.2010.07.003
Rand, A. S., Ryan, M. J., & Wilczynski, W. (1992). Signal redundancy and receiver permissiveness in acoustic mate
recognition by the túngara frog Physalaemus pustulosus. American Zoologist, 32(1), 81–90. https://doi.org/10.1093/
icb/32.1.81
Reber, R., Schwarz, N.,  & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the per-
ceiver’s processing experience? Personality and Social Psychology Review, 8(4), 364–382. https://doi.org/10.1207/
s15327957pspr0804_3
Redies, C., Hänisch, J., Blickhan, M., & Denzler, J. (2007). Artists portray human faces with the Fourier statistics of
complex natural scenes. Network, 18(3), 235–248. https://doi.org/10.1080/09548980701574496
Renoult, J. P., Bovet, J., & Raymond, M. (2016). Beauty is in the efficient coding of the beholder. Royal Society Open
Science, 3(3), 160027. https://doi.org/10.1098/rsos.160027

237
Michael J. Ryan

Renoult, J. P.,  & Mendelson, T. C. (2019). Processing bias: Extending sensory drive to include efficacy and effi-
ciency in information processing. Proceedings of the Royal Society B, 286(1900), 20190165. https://doi.org/10.1098/
rspb.2019.0165
Richards, E. (2017). Darwin and the making of sexual selection. University of Chicago Press.
Rodd, F. H., Hughes, K. A., & Pitcher, T. E. (1999). Sex, colour and mate choice in guppies. Reviews in Fish Biology and
Fisheries, 9(2), 203–204. https://doi.org/10.1023/A:1008950613746
Rosenthal, G. G. (2017). Mate choice: The evolution of sexual decision making from microbes to humans. Princeton University
Press.
Rosenthal, G. G. (2018). Evaluation and hedonic value in mate choice. Current Zoology, 64(4), 485–492. https://doi.
org/10.1093/cz/zoy054
Rosenthal, G. G., Flores Martinez, T. Y., García de León, F. J., & Ryan, M. J. (2001). Shared preferences by predators
and females for males ornaments in swordtails. American Naturalist, 158(2), 146–154. https://doi.org/10.1086/321309
Rosenthal, G. G., & Ryan, M. J. (2022). Sexual selection and the ascent of women: Mate choice research since Darwin.
Science, 375(6578), eabi6308. https://doi.org/10.1126/science.abi6308
Ryan, M. J. (1985). The túngara frog, A study in sexual selection and communication. University of Chicago Press.
Ryan, M. J. (1990). Sensory systems, sexual selection, and sensory exploitation. Oxford Surveys in Evolutionary Biology,
7, 157–195.
Ryan, M. J. (2011). The brain as a source of selection on the social niche: Examples from the psychophysics of mate
choice in túngara frogs. Integrative and Comparative Biology, 51(5), 756–770. https://doi.org/10.1093/icb/icr065
Ryan, M. J. (2018). A taste for the beautiful: The evolution of attraction. Princeton University Press.
Ryan, M. J. (2021). Darwin, sexual selection, and the brain. Proceedings of the National Academy of Sciences, 118(8). https://
doi.org/10.1073/pnas.2008194118
Ryan, M. J., Akre, K. L., Baugh, A. T., Bernal, X. E., Lea, A. M., Leslie, C., Still, M. B., Wylie, D. C., & Rand, A. S.
(2019). Nineteen years of consistently positive and strong female mate preferences despite individual variation. Ameri-
can Naturalist, 194(2), 125–134. https://doi.org/10.1086/704103
Ryan, M. J., Bernal, X. E., & Rand, A. S. (2007). Patterns of mating call preferences in túngara frogs, Physalaemus pustu-
losus. Journal of Evolutionary Biology, 20(6), 2235–2247. https://doi.org/10.1111/j.1420-9101.2007.01420.x
Ryan, M. J., & Cummings, M. E. (2013). Perceptual biases and mate choice. Annual Review of Ecology, Evolution, and
Systematics, 44(1), 437–459. https://doi.org/10.1146/annurev-ecolsys-110512-135901
Ryan, M. J., Fox, J. H., Wilczynski, W., & Rand, A. S. (1990). Sexual selection for sensory exploitation in the frog
Physalaemus pustulosus. Nature, 343(6253), 66–67. https://doi.org/10.1038/343066a0
Ryan, M. J., & Keddy-Hector, A. (1992). Directional patterns of female mate choice and the role of sensory biases.
American Naturalist, 139, S4–S35. https://doi.org/10.1086/285303
Ryan, M. J., & Rand, A. S. (2003). Sexual selection and female preference space: How female túngara frogs perceive
and respond to complex population variation in acoustic mating signals. Evolution, 57(11), 2608–2618. https://doi.
org/10.1111/j.0014-3820.2003.tb01503.x
Ryan, M. J., Tuttle, M. D., & Rand, A. S. (1982). Sexual advertisement and bat predation in a Neotropical frog. American
Naturalist, 119(1), 136–139. https://doi.org/10.1086/283899
Sakai, Y., Kawamura, S., & Kawata, M. (2018). Genetic and plastic variation in opsin gene expression, light sensitivity,
and female response to visual signals in the guppy. Proceedings of the National Academy of Sciences, 115(48), 12247–
12252. https://doi.org/10.1073/pnas.1706730115
Schwartz, D. A., Howe, C. Q., & Purves, D. (2003). The statistical structure of human speech sounds predicts musi-
cal universals. Journal of Neuroscience, 23(18), 7160–7168. https://doi.org/10.1523/JNEUROSCI.23-18-07160.2003
Searcy, W. A. (1992). Song repertoire and mate choice in birds. American Zoologist, 32(1), 71–80. https://doi.org/10.1093/
icb/32.1.71
Seehausen, O., Terai, Y., Magalhaes, I. S., Carleton, K. L., Mrosso, H. D., Miyagi, R., van der Sluijs, I., Schneider, M.
V., Maan, M. E., Tachida, H., Imai, H., & Okada, N. (2008). Speciation through sensory drive in cichlid fish. Nature,
455(7213), 620–626. https://doi.org/10.1038/nature07285
Shepard, R. N. (1987). Toward a universal law of generalization for psychological science. Science, 237(4820), 1317–
1323. https://doi.org/10.1126/science.3629243
Shettleworth, S. (1998). Cognition, evolution, and behaviour. Oxford University Press.
Skov, M. (unpublished). Animal preferences: Implications of sexual selection research for empirical aesthetics. https://doi.org/
10.31234/osf.io/8ycdt
Spence, R., & Smith, C. (2008). Innate and learned colour preference in the zebrafish, Danio rerio. Ethology, 114(6),
582–588. https://doi.org/10.1111/j.1439-0310.2008.01515.x

238
Sexual selection and mate choice

Spool, J. A., & Riters, L. V. (2017). Associations between environmental resources and the “wanting” and “liking” of
male song in female songbirds. Integrative and Comparative Biology, 57(4), 835–845. https://doi.org/10.1093/icb/
icx117
Steblin, R. K. (2002). Key characteristics in the 18th and early 19th centuries: A historical approach. University of Rochester
Press.
Stice, E., Yokum, S., Zald, D., & Dagher, A. (2010). Dopamine-based reward circuitry responsivity, genetics, and over-
eating. In Behavioural neurobiology of eating disorders (pp. 81–93). Springer. https://doi.org/10.1007/7854_2010_89
Stripling, R., Volman, S. F.,  & Clayton, D. F. (1997). Response modulation in the zebra finch neostriatum: Rela-
tionship to nuclear gene regulation. Journal of Neuroscience, 17(10), 3883–3893. https://doi.org/10.1523/
JNEUROSCI.17-10-03883.1997
Svec, L. A., Lookingland, K. J., & Wade, J. (2009). Estradiol and song affect female zebra finch behaviour independent of
dopamine in the striatum. Physiology and Behavior, 98(4), 386–392. https://doi.org/10.1016/j.physbeh.2009.07.003
Taylor, H. (2017). Is birdsong music?: Outback encounters with an Australian songbird. Indiana University Press.
Trivers, R. L. (1972). Parental investment and sexual selection. In B. Campbell (Ed.), Sexual selection and descent of man
(pp. 136–179). Aldine Publishing, Co.
Tuttle, M. D., & Ryan, M. J. (1981). Bat predation and the evolution of frog vocalizations in the Neotropics. Science,
214(4521), 677–678. https://doi.org/10.1126/science.214.4521.677
Tversky, A., & Kahneman, D. (1989). Rational choice and the framing of decisions. In Multiple criteria decision making and
risk analysis using microcomputers (pp. 81–126). Springer. https://doi.org/10.1086/296365
Tversky, A., & Kahneman, D. (1992). Advances in prospect theory: Cumulative representation of uncertainty. Journal of
Risk and Uncertainty, 5(4), 297–323. https://doi.org/10.1007/BF00122574
Vasconcelos, M., Monteiro, T., & Kacelnik, A. (2015). Irrational choice and the value of information. Scientific Reports,
5(1), 13874. https://doi.org/10.1038/srep13874
Verpooten, J., & Nelissen, M. (2012). Sensory exploitation: Underestimated in the evolution of art as once in sexual selection
theory? In Philosophy of behavioural biology (pp. 189–216). Springer. https://doi.org/10.1007/978-94-007-1951-4_9
Volkow, N. D., Wang, G. J., Fowler, J. S., Tomasi, D., & Telang, F. (2011). Addiction: Beyond dopamine reward circuitry.
Proceedings of the National Academy of Sciences, 108(37), 15037–15042. https://doi.org/10.1073/pnas.1010654108
West-Eberhard, M. J. (1979). Sexual selection, social competition, and evolution. Proceedings of the American Philosophical
Society, 123, 222–234.
White, A. (2020). Sex and beauty for birds, frogs, and humans. https://sites.williams.edu/awhite/files/2020/03/White-
2020-Sex-and-Beauty-for-Birds-Frogs-and-Humans.pdf
Wilczynski, W., Rand, A. S., & Ryan, M. J. (2001). Evolution of calls and auditory tuning in the Physalaemus pustulosus
species group. Brain, Behavior and Evolution, 58(3), 137–151. https://doi.org/10.1159/000047268
Wilczynski, W., & Ryan, M. J. (2010). The behavioural neuroscience of anuran social signal processing. Current Opinion
in Neurobiology, 20(6), 754–763. https://doi.org/10.1016/j.conb.2010.08.021

239
12
AESTHETIC SENSITIVITY
Origin and development

Ana Clemente

To provide general explanations for the influence of object features on people’s liking, preference, or appre-
ciation is one of the central goals of empirical aesthetics and neuroaesthetics (Berlyne, 1971; Fechner, 1876;
Martindale, 2001; Skov, 2019, 2020). Such explanations often invoke general perceptual, cognitive, and
affective processes to account for regular and predictable ways in which people value object features, such as
symmetry and complexity in the visual domain (Gartus & Leder, 2017) or uncertainty and surprise in the
musical domain (Cheung et al., 2019). For instance, people generally prefer complex and symmetric visual
designs (Westphal-Fitch et al., 2013; Nadal et al., 2010; Friedenberg, 2018; Jacobsen & Höfel, 2001) and
high surprise and low uncertainty or low surprise and high uncertainty in music (Cheung et al., 2019; Gold
et al., 2019). The reason for broadly shared preferences for sensory features is that sensory valuation relies on
common perceptual, cognitive, and affective processes.
General preference patterns, however, do not necessarily imply uniformity of preferences. In fact, there
is substantial evidence for differences in how people assess the value of visual and musical features (Corradi
et al., 2019, 2020; Clemente, Pearce, & Nadal, 2021; Clemente, Pearce, Skov et al., 2021). Such differ-
ences have been attributed to the effects of personality (Chamorro-Premuzic et al., 2009; Mastandrea et al.,
2009; McManus  & Furnham, 2006), intelligence (Chamorro-Premuzic  & Furnham, 2004; Furnham  &
Chamorro-Premuzic, 2004), expertise (Belke et al., 2010; Pang et al., 2013; Silvia & Barona, 2009), and
other personal traits.
The study of these individual differences in the appreciation of art and aesthetics has a long history, begin-
ning in the 1920s with the search for efficient measures of artistic and aesthetic abilities in school children
that could be used to test achievement and for vocational guidance (Burt, 1927, 1933; Meier, 1927, 1928;
Thorndike, 1916, 1917). Among such measures, aesthetic sensitivity stood out for its prognostic value, suit-
ability for laboratory research, and crucial role in several influential paradigms. Although the concept of aes-
thetic sensitivity had been used occasionally in the late 19th century, it developed during the central decades
of the 20th century, reaching its heyday between 1960 and 1980 (Figure 12.1). Aesthetic sensitivity became
a central concept in the thinking of Norman C. Meier (1927), Cyril L. Burt (1933), Hans J. Eysenck (1940),
and Irving L. Child (1962) and a conceptual tool for defining and explaining variance in aesthetic apprecia-
tion that was inextricably linked to different views on aesthetics, beauty, and human cognition.
In this chapter, I discuss three different approaches to aesthetic sensitivity: The first views sensitivity as
subordinate to a normative standard of aesthetic taste, according to which a person can be more or less sensi-
tive to an object’s true aesthetic value and people are born with unalterable degrees of good or bad taste. The

240 DOI: 10.4324/9781003008675-13


Aesthetic sensitivity
241

Figure 12.1 Frequency of books mentioning aesthetic sensitivity (blue), esthetic sensitivity (red), aesthetic sensitiveness (g reen), and esthetic sensitiveness (orange) from 1800
to 2019. Source: Google Ng ram Viewer (Michel et al., 2011).
Ana Clemente

second also assumes an ideal of good taste that people deviate from or approach but considers aesthetic sen-
sitivity dependent on culture and environment and improvable through learning. Finally, the third disputes
the existence of norms of objectively good taste, regarding aesthetic sensitivity as differences in the extent to
which stimulus features are factored in aesthetic valuation.

Historical background
The earliest conceptions of aesthetic sensitivity emerged from attempts to identify efficient ways of measur-
ing artistic talent in the context of psychological testing in schools for the purpose of measuring aptitude and
offering vocational guidance (Burt, 1933; Meier, 1926, 1927, 1928; Thorndike, 1916, 1917). Thorndike, for
instance, developed a measure of the merit of children’s drawings that was based on psychophysical scaling
rules (1913, 1924) and a measure of aesthetic merit and appreciation ability that was based on agreement and
disagreement with majority norms (1916, 1917). He also compared the “good taste” of different communi-
ties (Powel et al., 1942; Thorndike & Woodyard, 1943), concluding that it seemed “positively associated
with differences in the intelligence, morality, and competence of their residents” (Thorndike & Woodyard,
1943, p. 59). Likewise, McAdory (1929) developed a measure of artistic taste that was based on rankings
derived from academic grades (Siceloff & Woodyard, 1933), just as Binet and Simon’s (1916) scale of intel-
ligence included pairs of prettier/uglier drawings of faces.
Meier (1927, 1928, 1939) argued that aesthetic sensitivity, understood as “the ability to recognize composi-
tional excellence in representative art-situations, or the ability to ‘sense’ quality (beauty?) in an aesthetic organ-
ization” (Meier, 1928, p. 185; also 1939), was the most efficient and predictive measure of artistic talent. To
measure aesthetic sensitivity, Meier and Seashore (1929; also Meier, 1942) developed the Art Judgment Test.
Later, Meier (1940) published the Meier Art Tests: I. Art Judgment, premised upon the belief that the aes-
thetic character of art resides in the organization of parts according to universal principles of goodness and that
aesthetic judgment involves detecting such principles. The Meier Art Tests: II. Aesthetic Perception (1963)
was designed to assess the perceptual-facility factor of artistic talent, that is, the ability to detect subtle aspects of
aesthetic significance. For Meier, aesthetic sensitivity was a measure of agreement with norms of artistic value
determined by the original artworks versus their distortions—following Abbott and Trabue’s method (1921),
or the so-called controlled-alteration process. Crucially, Meier (1934, 1939) regarded aesthetic sensitivity and
perceptual facility largely acquired components of artistic aptitude and thus fruit of experience and training.
These studies led to two approaches to individual differences in aesthetic appreciation and their underly-
ing factors. Both considered objective factors to form a common foundation for aesthetic appreciation and
developed into the determinist and educational normative views of aesthetic sensitivity, understood as degrees
of convergence with, or deviation from, some external reference. Alternative approaches treated objective
and subjective factors as constituents of aesthetic appreciation, intertwined and with different relevance for
different people. This ultimately crystallized as the subjective responsiveness view of aesthetic sensitivity.

The determinist view of aesthetic sensitivity


Cyril L. Burt was the leading figure and pioneer of the determinist view of aesthetic sensitivity. He was
strongly drawn to Galton’s ideas on statistics, individual differences, mental tests, and eugenics. In 1913,
Burt was appointed part-time chief psychologist at the London County Council (Arnold, 2013), where he
worked on developing measures of aptitude and achievement. He defended the view that intelligence and
mental abilities were genetically based and the main determinants of social position and that these abilities
could be objectively and accurately measured using mental tests (Norton, 1981). Burt’s goal was to introduce
such tests to the educational system and use measures of children’s performance to determine the vocational

242
Aesthetic sensitivity

path that best fitted the individual child’s natural aptitude. Amongst his test battery were tasks targeted at
measuring children’s aptitude for literary, musical, and visual appreciative and creative abilities.
Burt found two factors driving aesthetic judgment. What he called the general factor of artistic ability was
thought to account for good taste: the ability to appreciate relations among elements in art, music, and poetry
(Burt, 1933, 1949, 1960). For him, this general factor of artistic ability was unitary, inherited, unalterable,
and measurable through simple tests (Bulley & Burt, 1933; Burt, 1960). Burt (1939) believed that aesthetic
appreciation should be measured against expert judgments, which established the true rank of aesthetic values.
A second bipolar factor distinguished between objective or classical and subjective or romantic types (Burt, 1915,
1933)—resembling Bullough’s (1908, 1910) types. This second factor was more pronounced when control-
ling for the first factor and for younger or less artistically sophisticated people, for which “irrelevant factors
become more obvious” (Burt, 1915; Dewar, 1938; Stephenson, 1936). This twofold factor was deemed
analogous to those in Binet’s (1903) intelligence tests and Burt’s (1912) temperamental differences and close
to Jungian extravert/introvert types (Dewar, 1938). Burt’s theories exerted a broad, profound, and long-lasting
influence on education, society, and politics.
Burt’s doctoral student Hans J. Eysenck continued his psychometric approach to aesthetics. Eysenck’s
(1940) factor analysis uncovered what he deemed a general objective factor of aesthetic appreciation. Eysenck
asserted that this factor explained performance on virtually any aesthetic appreciation test and claimed that
his factor was universal, biologically determined, and innate (Eysenck, 1941b, 1941c, 1942, 1981). He
equated this factor, t (for taste), with an ability to appreciate objective beauty (Eysenck, 1941c, 1942, 1981)
and described it as distinct—because “[this ability], independently of intelligence and personality, determines
the degree of good or bad taste” (Eysenck, 1983, p. 213)—general—for “it covers a large number of, prob-
ably all, pictorial tests” (Eysenck, 1940, p. 100)—stable—as “[it] presumably [has] a genetic foundation in the
structure of the nervous system” (Götz et al., 1979, p. 801)—and insensitive to experience —“[it] is independ-
ent of teaching, tradition, and other irrelevant associations” (Eysenck, 1940, p.  102)—and culture—given
the “comparative absence of cultural factors determining aesthetic judgments” (Eysenck & Iwawaki, 1971,
p. 817; Eysenck & Iwawaki, 1975; Soueif & Eysenck, 1971). He later identified a second bipolar factor, k
(Eysenck, 1941a; Frois & Eysenck, 1995), characterized by “brightness or intensity as opposed to darkness
or lack of intensity” (Eysenck, 1983, p. 91).
According to Eysenck (1941a, 1942), aesthetic sensitivity scaled with the degree to which liking approxi-
mated true aesthetic value, as determined by group agreement or expert opinion (Eysenck, 1972a, 1981;
Eysenck & Iwawaki, 1971). In this way, someone’s aesthetic sensitivity could easily be calculated by compar-
ing their average liking ratings with group averages or with expert judgments. Eysenck first examined ratings
of artworks (Eysenck, 1940). Later, he used simple geometric designs (Eysenck, 1972b; Eysenck & Castle,
1971), borrowed from Birkhoff (1933), and the Figure Preference Test (Barron & Welsh, 1952; Welsh &
Barron, 1949). Finally, he developed the Visual Aesthetic Sensitivity Test (VAST; Chan et al., 1980; Götz
et al., 1979; Iwawaki et al., 1979).
However, the VAST exhibited low internal consistency and structural validity, and its scores were
explained by intelligence, figural creativity, and personality traits (Chamorro-Premuzic & Furnham, 2004;
Furnham & Chamorro-Premuzic, 2004; Myszkowski et al., 2018; Myszkowski et al., 2014; Payne, 1967).
Thus, contrary to Eysenck’s (1941a, 1942) claims, aesthetic sensitivity appeared not to be a distinct ability but
to draw upon general cognitive processes, learning, and experience. To overcome this issue, Myszkowski and
colleagues have recently suggested two mutually compatible amends: First, Myszkowski and Zenasni (2016)
proposed a composite measure of aesthetic aptitude that brings together measures of aesthetic exploration, art
expertise, sensitivity to complexity, aesthetic empathy, and aesthetic sensitivity as aesthetic balance recognition.
Second, Myszkowski and Storme (2017) introduced a revised version of the VAST with improved internal
consistency and structural validity.

243
Ana Clemente

Regardless of the instrument used to measure it, normative conceptions of aesthetic sensitivity assume
that aesthetic appreciation adheres to an objective standard of taste: that aesthetic sensitivity, or good taste,
can be defined as an ability that approximates a true order of taste (Eysenck, 1972a) or an absolute beauty
(Eysenck, 1972a; Eysenck & Iwawaki, 1971) with some objects objectively superior to others. Furthermore,
this deterministic view maintains that aesthetic taste is innate and impervious to experience, despite the
scarce and contradictory (e.g., Eysenck, 1972a, 1983) empirical support for these claims—particularly if
considering expert judgments norms, given that expertise is acquired by definition.

The educative view of aesthetic sensitivity


The educative view of aesthetic sensitivity agrees with the determinist view that objective aesthetic values and
standards of taste exist. However, it does not agree that aesthetic taste is innate and immutable. Instead, the
educative view suggests that aesthetic tastes are acquired through learning and, therefore, culturally and
socially moulded. Especially dominant in the 20th century, the educative view of aesthetic sensitivity has
been espoused by a wide range of academics, including social scientists (e.g., Dai & Shader, 2001), political
scientists (Suga, 2003), philosophers (Mitchells, 1966), and artists (Fehl, 1953; Smets & Knops, 1976), as
well as psychologists and educators (Adler, 1929; Anderson, 1972, 1975; Bullock, 1971; Day, 1976; Gernet,
1940; Hahn, 1954; Hevner, 1930; Kwalwasser & Dykema, 1930; Kyme, 1967; Reimer, 1965, 1968a, 1968b;
Taunton, 1982; Trabue, 1923; Vernon, 1930; Webster, 1988a, 1988b).
The educative view of aesthetic sensitivity emphasizes contextual, situational, and personal factors.
Accordingly, Hevner (1937; Hevner & Mueller, 1939) showed that information about an object was able to
modulate its aesthetic appreciation. Similarly, Voss (1936) observed notable improvements in aesthetic analy-
sis and judgment in children aware of criteria for aesthetic merit, and Clair (1939) overtly refuted the condi-
tion of disinterestedness because “critical and appraising analysis of works of art . . . intensifies . . . [their]
appreciation” (p. 67). Remarkably, Carroll (1932) showed that the relationship between the ability to appre-
ciate art, literature, and music was very slight. This finding not only discredited the determinist claim of gen-
erality but pointed to a modality-specific basis of aesthetic appreciation, even if it still accorded to standards.
Among other determinants of creativity, Barron and Welsh investigated aesthetic judgments defined as
the ability “to discriminate the good from the poor (as judged by experts)” (Barron, 1952, p. 387, 1963,
1969; Barron  & Welsh, 1952). In Welsh’s (1949) study, a group of artists—ostensiably representatives of
good taste—exhibited a preference for complexity and asymmetry, dissident personality, and unconventional
political views (Barron & Welsh, 1952). Barron (1952) confirmed that artists preferred complex and asym-
metric figures and were rebellious against authority and tradition (1953). The paintings artists liked at that
time were “ ‘modern’ art movements as Primitivism, Expressionism, Impressionism, and Cubism” (Barron,
1952, p. 391), known for rebelling against traditional art traditions—thus evincing the mutable character of
expert opinion and, consequently, of aesthetic sensitivity conceived as agreement with expert judgment.
Irving L. Child was the central proponent of this educative conception of aesthetic sensitivity in empiri-
cal aesthetics. Child (1962) was sceptical about Eysenck’s assumption that average rankings represented true
aesthetic value and that the extent to which individual preferences agreed with the average constituted a valid
measure of aesthetic sensitivity. He therefore tested preference versus aesthetic value as determined by experts
(Child, 1962). He developed a measure of aesthetic value as the average rating of expert judges weighted by
agreement with the other judges and compared it to non-experts’ averaged preferences. He found that “the
degree to which preferences are related to aesthetic value is a very stable characteristic of the individual . . .
[and that the] degree of agreement with an aesthetic standard is an even more consistent characteristic than
[the] degree of agreement with group preferences” (Child, 1962, p. 504). Finally, he found a negative or
absent correlation between individual measures of preferences defined as the extent to which they resemble
one kind of standard or the other. He concluded that “the degree of agreement between one’s preferences

244
Aesthetic sensitivity

and aesthetic value is an index of aesthetic sensitivity . . . [whereas the] degree of agreement with group
preferences does not correspond to an external criterion of aesthetic sensitivity” (p. 506).
Child (1964, 1965) understood aesthetic sensitivity as “the extent to which, when a person judges the
esthetic value of works of art, his judgments agree with an appropriate external standard of their esthetic
value . . . provided by the judgment of experts” (p. 476). Child believed that “aesthetic sensitivity is expressed
in a tendency to prefer the aesthetically good” (p. 508) and that “[i]f one set out to measure aesthetic sensitiv-
ity, he would ordinarily not ask people to express personal preferences, he would generally do better to ask
them to make aesthetic judgments, as in the Bulley Test” (p. 510).
Thus, like Eysenck, Child thought of aesthetic sensitivity as a measure of the ability to make aesthetic
judgments according to standards of value. However, in contrast to Eysenck, Child (1962, 1965) believed that
aesthetic sensitivity was cultivated with practice and resulted not from a specific ability but from a general
cognitive style or personality (Child, 1964, 1965; Iwao & Child, 1966). For him, high aesthetic sensitivity
was the manifestation of an “actively inquiring mind, seeking out experience that may be challenging because
of complexity or novelty, even alert to the potential experience offered by stimuli not already in the focus of
attention” (Child, 1965, p. 508). Thus, a highly sensitive person would be “interested in understanding each
experience thoroughly and for its own sake rather than contemplating it superficially and promptly filing it
away in a category, and able to do all this with respect to the world inside himself as well as the world outside”
(p. 508)—emphasizing the relevance of motivation, theoretical interest, and personal and contextual factors.
In the 1980s, this educative view of aesthetic sensitivity was revised and updated. Bamossy and colleagues
(1985) criticized previous work for inappropriately distinguishing between aesthetic responses, preferences, and
judgments. Only the latter are about the object, and thus only those “can be relevant or irrelevant,” meaning
susceptible to value judgments; whereas responses “cannot validly be categorized as ‘appropriate or inap-
propriate’, or ‘relevant or irrelevant’,” and one “cannot be better or worse at having preferences” (Bamossy
et al., 1985, pp. 64–65). As noted by Bamossy and colleagues (1985), instruments such as Welsh’s (1959;
Barron & Welsh, 1952; Welsh & Barron, 1949), Graves’s (1939, 1948), and Thorndike’s (1916) “have not
observed this distinction and consequently seem to lack validity” (p. 65). They also observed that manipulat-
ing masterpieces was a standard technique in artistic movements like Pop Art, so they questioned the utility
of tests based on comparisons between originals and alterations and emphasized the mutable character of
expert judgments. They criticized the direct use of expert judgment in Child’s (1962), Graves’s (1939, 1948),
Meier’s (1940), Welsh’s (1959; also Barron  & Welsh, 1952; Welsh  & Barron, 1949), Thorndike’s (1916),
Bottorf ’s (1946), and Williams and Hattwick’s (1932) tests, as it is “easier to get agreement among experts
on reasons for judgments than on judgments themselves” (Bamossy et al., 1985, p. 67). For these reasons,
Bamossy and colleagues (1985) rejected the prevailing notion of aesthetic sensitivity while advocating for
the existence of objective value and the utility of assessing aesthetic ability. Their Aesthetic Judgment Ability
Test (Bamossy et al., 1985), based on the theory of cognitive development of aesthetic judgment (Parsons &
Durham, 1979; Parsons et al., 1978), measured judgments’ sophistication, with expert criteria representing
higher-stage reasons for aesthetic judgments.
Another aspect of aesthetic sensitivity that was revised in this decade was the idea that it represented a sin-
gle factor or ability. Winner et al. (1986) studied the development of aesthetic sensitivity in children, finding
that it was art form specific—that is, not a single factor—and property specific—that is, multiple. In this line,
Elliot (1995) asserted that “it is highly doubtful that there is any such general capacity as aesthetic sensitivity.
Multiple intelligence theories and contemporary studies of creativity argue against such possibility” (p. 249).

The responsiveness view of aesthetic sensitivity


The two normative conceptions of aesthetic sensitivity reviewed in the preceding two sections were domi-
nant during the 20th century. However, an alternative notion emerged during the early decades of the 20th

245
Ana Clemente

century. As part of her research on factors modulating aesthetic experience, Washburn introduced the con-
cept of affective sensitiveness to distinguish between people with a strong tendency to like and dislike materials
of different sorts—for example, tones, colours, and speech sounds—from people who are relatively indif-
ferent to them (Babbitt et al., 1915; Clark et al., 1913). Affective sensitiveness was calculated as “the ratio
of the sum of the number of judgments of extreme pleasantness and extreme unpleasantness to the number
of judgments of indifference” (Washburn et al., 1923, p. 105; Clark et al., 1913). Washburn’s experiments
demonstrated that affective sensitiveness depended to certain degree on different circumstances and condi-
tions. They showed, for instance, that fatigue reduced aesthetic sensitivity (Robbins et al., 1915), experience
and expertise in art and aesthetics increased aesthetic sensitivity (Washburn et al., 1923), and art interest led
people to approximate expert’s rankings (Cattell et al., 1918).
A century later, Corradi and colleagues (2019, 2020) developed another responsiveness approach to aes-
thetic sensitivity. They defined aesthetic sensitivity as the extent to which someone’s liking is determined by a
specific stimulus feature, such as symmetry or complexity. In this sense, someone is aesthetically sensitive to
complexity when their liking is influenced by complexity: for example, they prefer complex to simple ones
or simple to complex designs. Conversely, someone is aesthetically insensitive to complexity when their lik-
ing is uninfluenced by complexity: that is, their preference has little to do with the complexity or simplicity
of designs.
The idea resembles Washburn’s but with one fundamental difference: whereas Washburn’s notion of
affective sensitiveness reflected the magnitude of individual responsiveness to visual and auditory stimuli, it
did not reflect responsiveness to variations in specific stimulus features. In contrast, Corradi and colleagues’
(2019, 2020) notion is closer to Fechner’s (1876) psychophysical conception of empirical aesthetics in that it
captures the extent to which variations in stimulation (e.g., complexity or symmetry) translate into variations
in valuation (e.g., liking or preference).
Corradi and colleagues (2019, 2020) showed that people differ in the extent to which their aesthetic valu-
ation takes into account the balance, contour, symmetry, and complexity of visual designs. Clemente and
colleagues (Clemente, Pearce, & Nadal, 2021; Clemente, Pearce, Skov et al., 2021) reported similar differ-
ences in the extent to which people’s liking for melodies depends on their balance, contour, symmetry, and
complexity. These studies found little co-variation among sensitivities to different features. People can be
highly sensitive to some features but not to others: someone’s liking for visual designs might depend more on
their symmetry than their complexity, while someone else’s liking might depend more on their complexity
than their symmetry (Corradi et al., 2020). Likewise, some people are more aesthetically sensitive to musical
balance than contour, while others are more sensitive to musical contour than balance (Clemente, Pearce, &
Nadal, 2021).
This approach to aesthetic sensitivity has allowed asking new questions, such as whether sensitivity to one
feature entails sensitivity to an analogous feature in a different sensory modality. Is someone highly sensitive
to visual complexity also highly sensitive to musical complexity; that is to say, does complexity influence
this person’s liking regardless of sensory modality? Clemente, Pearce, Skov, and colleagues (2021) showed
that this is not the case, but aesthetic sensitivity is modality specific: for example, symmetry might influence
someone’s liking for visual designs but not for musical motifs, and complexity might influence someone’s
liking for musical motifs but not for visual designs. The exception seems to be contour: people who disliked
jagged visual contours also tended to dislike jagged melodies (Clemente, Pearce, Skov et al., 2021).
Although these patterns of aesthetic sensitivity for visual and musical features vary between people, they
seem to be relatively stable in time within people, as suggested by the similarity of liking ratings for the same
stimuli weeks apart (Corradi et al., 2020; Clemente, Pearce, & Nadal, 2021). However, aesthetic sensitivities
appear to be unrelated to other individual traits such as personality, intelligence, and interest in art (Clemente,
Pearce, Skov et al., 2021; Corradi et al., 2020).

246
Aesthetic sensitivity

Conclusions
Empirical aesthetics and neuroaesthetics strive to explain how cognitive, affective, and neural systems allow us
to enjoy perceiving objects around us, to be pleased by music and art. Like other domains of the behavioural
and brain sciences, they aim to provide general principles across cases or circumstances while accounting
for individual differences. The need to find a different explanation for each person or object of apprecia-
tion would entail no accounting for taste, preference, or pleasure. One of the main goals of the foundational
text of empirical aesthetics (Fechner, 1876) was to elaborate on several principles of aesthetic pleasure. The
principle of union of diverse elements, for instance, states that objects that produce aesthetic pleasure are
often composite and manifold. The quest for general principles of aesthetic appreciation continues today. For
instance, fluency theory holds that ease of processing leads to a pleasant feeling of fluency, which can translate
into assessments of beauty, likeability, and attractiveness: “The more fluently perceivers can process an object,
the more positive their aesthetic response” (Reber et al., 2004, p. 364).
There is extensive evidence supporting these general trends. For example, people usually find objects
composed of arranged elements aesthetically pleasing, and ease of processing often enhances this experience.
Nevertheless, extensive evidence suggests that aesthetic appreciation is not the result of only these sorts of
aesthetic principles. If this were the case, we would all share the same aesthetic preferences. The obvious fact
that people are aesthetically pleased by different painting and musical styles, literary and cinema genres, or
clothing fashions begs an explanation. If certain principles influence aesthetic appreciation, why do people
differ so much in what they find aesthetically pleasing? This question has fascinated philosophers for centu-
ries and psychologists and neuroscientists for decades. It is also the question that work on aesthetic sensitivity
seeks to answer.
Aesthetic sensitivity has a central place in empirical aesthetics. It was initially conceived as an ability to
recognize objective beauty in arrangements and compositions, easy to measure and predictive of artistic abili-
ties. For almost a century, it has continued to be associated with individual differences in aesthetic valuation.
Such differences, however, have been understood in substantially different ways, reflecting fundamental ideas
or views about art, aesthetics, and human cognition.
From the perspective that I have referred to as the determinist view, there is such a thing as objective beauty
or true aesthetic value. Aesthetics and art are consubstantial, in the sense that the greatness of art depends on
the extent to which it embodies aesthetic values such as beauty or sublimity in unique ways. The human
mind apprehends aesthetic value through an aesthetic sense, a sense of beauty, or taste. As with other mental
capacities, some people have better taste than others, and these differences are conditioned mainly by bio-
logical factors, relatively insensitive to learning and experience. From this perspective, aesthetic sensitivity is
a specific fixed ability to value—with greater or lesser accuracy—the beauty of objects, most notably art, as
determined by group average or expert opinion.
From the perspective that I have referred to as the educative view, cultures and societies arrive at a con-
sensus about what is considered beautiful or aesthetically valuable. Art is a material embodiment of cultural
and aesthetic values that vary from culture to culture and must be learned and cultivated. Experts are able
to recognize the beautiful and the good art because of their extensive experience and knowledge about
that particular culture or society’s values. People vary in the extent to which their appreciation agrees with
experts because they differ in the experiences, personality traits, and interests that foster attention to aesthet-
ics and art. From this perspective, aesthetic sensitivity is a general and learned capacity to recognize—with
greater or lesser accuracy—the sort of qualities experts and connoisseurs rely on to recognize in an artwork
a great embodiment of aesthetic and cultural values.
From the perspective that I have referred to as the responsiveness view, beauty is not a quality of objects
but a quality of our experience of objects. Beauty and other feelings result from general sensory valuation

247
Ana Clemente

processes that involve assigning hedonic value (pleasure–displeasure) to sensory information in the context
of judgments, decisions, and choices. Art is one of the many classes of objects susceptible to valuation. Peo-
ple vary in the extent to which their valuation relies on different object properties or features. From this
perspective, aesthetic sensitivity to a given object feature is the extent to which people take that feature into
account when assessing the hedonic value of that object. From this perspective, aesthetic sensitivity is not
measured against any external standard or norm but is a measure of pleasurable responsiveness to variations
in a particular kind of sensory information.
There is little hope for reconciliation between these notions of aesthetic sensitivity: they emerge from
diverging ideas about beauty, art, and human cognition. Nevertheless, these disagreements continue to
stimulate important advances in the study of aesthetic sensitivity in recent years. They have led to the clarifi-
cation and refinement of arguments and concepts and new experiments yielding decisive data. Perhaps most
critically, they emphasize the importance of continuing to study the reasons people differ in what they like
or find beautiful, because any comprehensive explanation of aesthetic appreciation needs to account for such
differences.

References
Abbott, A., & Trabue, M. R. (1921). A measure of ability to judge poetry. Teachers College Record, 22(2), 101–126.
Adler, M. J. (1929). Music appreciation: An experimental approach to its measurement. Archives of Psychology, 17(110),
102.
Anderson, L. E. (1972). The effects of divergent musical literature used in high school performance organizations in developing aes-
thetic sensitivity to music [Doctoral Dissertation, University of California, Berkeley].
Anderson, L. E. (1975). The effects of music literature in developing aesthetic sensitivity to music. Journal of Research in
Music Education, 23(1), 78–84. https://doi.org/10.2307/3345205
Arnold, C. (2013). The rise of education. In C. Arnold & H. Hardy (Eds.), British educational psychology: The first hundred
years (pp. 19–20). British Psychological Society.
Babbitt, M., Woods, M.,  & Washburn, M. F. (1915). Affective sensitiveness to colours, tone intervals, and articulate
sounds. American Journal of Psychology, 26(2), 289–291. https://doi.org/10.2307/1413259
Bamossy, G., Johnston, M., & Parsons, M. (1985). The assessment of aesthetic judgment ability. Empirical Studies of the
Arts, 3(1), 63–79. https://doi.org/10.2190/1U13-U2DB-0BE3-A4FN
Barron, F. X. (1952). Personality style and perceptual choice.  Journal of Personality, 20(4),  385–401.  https://doi.
org/10.1111/j.1467-6494.1952.tb01116.x
Barron, F. X. (1953). Complexity-simplicity as a personality dimension. Journal of Abnormal Psychology, 48(2), 163–172.
https://doi.org/10.1037/h0054907
Barron, F. X. (1963). Creativity and psychological health. Van Nostrand.
Barron, F. X. (1969). Creative person, creative process. Holt, Rinehart & Winston.
Barron, F. X., & Welsh, G. S. (1952). Artistic perception as a possible factor in personality style: Its measurement by a
figure preference test. Journal of Psychology, 33(2), 199–203. https://doi.org/10.1080/00223980.1952.9712830
Belke, B., Leder, H., Strobach, T., & Carbon, C. C. (2010). Cognitive fluency: High-level processing dynamics in art
appreciation. Psychology of Aesthetics, Creativity, and the Arts, 4(4), 214–222. https://doi.org/10.1037/a0019648
Berlyne, D. E. (1971). Aesthetics and psychobiology. Appleton-Century-Crofts.
Binet, A. (1903). L’étude expérimentale de l’intelligence. Schleicher frères & cie.
Binet, A., & Simon, T. (1916). The development of intelligence in children (the Binet-Simon scale) (E. S. Kite, Trans.). Wil-
liams & Wilkins Co.  https://doi.org/10.1037/11069-000
Birkhoff, G. D. (1933). Aesthetic measure. Harvard University Press.
Bottorf, E. A. (1946). A study comparing art abilities and general intelligence of college students. Journal of Educational
Psychology, 37(7), 398–426.  https://doi.org/10.1037/h0058445
Bulley, M. H., & Burt, C. L. (1933). Have you good taste?: A guide to the appreciation of the lesser arts. Methuen, and Co., Ltd.
Bullock, W. J. (1971). Construction and evaluation of a test of musico-aesthetic attitude [Doctoral Dissertation, Florida State
University, Tallahassee].
Bullough, E. (1908). The perceptive problem in the aesthetic appreciation of single colours. British Journal of Psychol-
ogy, 2(4), 406–463. https://doi.org/10.1111/j.2044-8295.1908.tb00189.x

248
Aesthetic sensitivity

Bullough, E. (1910). The perceptive problem in the aesthetic appreciation of simple colour-combinations. British Journal
of Psychology, 3(1), 406.
Burt, C. (1912). The meaning of social science. Eugenics Review, 4(1), 99.
Burt, C. (1915). General and specific factors underlying the primary emotions. British Associations Annual Reports, 84,
694–696.
Burt, C. (1927). The measurement of mental capacities: A review of the psychology of individual differences [Doctoral Disserta-
tion]. Oliver & Boyd.
Burt, C. (1933). How the mind works. Allen & Unwin.
Burt, C. (1939). A judgment test for measuring intelligence. Mental Welfare, 20(2), 45–48.
Burt, C. (1949). The structure of the mind: A review of the results of factor analysis. British Journal of Educational Psychol-
ogy, 19(3), 176–199.  https://doi.org/10.1111/j.2044-8279.1949.tb01621.x
Burt, C. (1960). The general aesthetic factor. III.  British Journal of Statistical Psychology,  13(1), 90–92. https://doi.
org/10.1111/j.2044-8317.1960.tb00044.x
Carroll, H. A. (1932). A preliminary report on a study of the interrelationships of certain appreciations. Journal of Educa-
tional Psychology, 23(7), 505–510.  https://doi.org/10.1037/h0073183
Cattell, J., Glascock, J., & Washburn, M. F. (1918). Experiments on a possible test of aesthetic judgment of pictures.
American Journal of Psychology, 29(3), 333–336. https://doi.org/10.2307/1414125
Chamorro-Premuzic, T., & Furnham, A. (2004). Art judgment: A measure related to both personality and intelligence?
Imagination, Cognition and Personality, 24(1), 3–24. https://doi.org/10.2190/U4LW-TH9X-80M3-NJ54
Chamorro-Premuzic, T., Reimers, S., Hsu, A., & Ahmetoglu, G. (2009). Who art thou? Personality predictors of artistic
preferences in a large UK sample: The importance of openness. British Journal of Psychology, 100(3), 501–516. https://
doi-org.sire.ub.edu/10.1348/000712608X366867
Chan, J., Eysenck, H. J., & Götz, K. O. (1980). A new visual aesthetic sensitivity test: III Cross-cultural comparison
between Hong Kong children and adults, and English and Japanese samples. Perceptual and Motor Skills, 50(3_suppl),
1325–1326. https://doi.org/10.2466/pms.1980.50.3c.1325
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology,
29(23), 4084–4092.e4. https://doi.org/10.1016/J.CUB.2019.09.067
Child, I. L. (1962). Personal preferences as an expression of aesthetic sensitivity. Journal of Personality, 30(3), 496–512. 
https://doi.org/10.1111/j.1467-6494.1962.tb02319.x
Child, I. L. (1964). Observations on the meaning of some measures of esthetic sensitivity. The Journal of Psychology, 57,
49–64. https://doi.org/10.1080/00223980.1964.9916671
Child, I. L. (1965). Personality correlates of esthetic judgment in college students. Journal of Personality, 33(3), 476–511. 
https://doi.org/10.1111/j.1467-6494.1965.tb01399.x
Clair, M. B. (1939). Variation in the perception of aesthetic qualities in paintings. Psychological Monographs, 51(5), 52–67. 
https://doi.org/10.1037/h0093476
Clark, H., Quackenbush, N., & Washburn, M. F. (1913). A suggested coefficient of affective sensitiveness. American Jour-
nal of Psychology, 24(4), 583–585. https://doi.org/10.2307/1413458
Clemente, A., Pearce, M. T., & Nadal, M. (2021). Musical aesthetic sensitivity. Psychology of Aesthetics, Creativity, and the
Arts. Advance online publication. https://doi.org/10.1037/aca0000381
Clemente, A., Pearce, M. T., Skov, M., & Nadal, M. (2021). Evaluative judgment across domains: Liking balance, cur-
vature, symmetry, and complexity in musical motifs and visual designs. Brain and Cognition, 151, 105729. https://doi.
org/10.1016/j.bandc.2021.105729
Corradi, G., Belman, M., Currò, T., Chuquichambi, E. G., Rey, C., & Nadal, M. (2019). Aesthetic sensitivity to curva-
ture in real objects and abstract designs. Acta Psychologica, 197, 124–130. https://doi.org/10.1016/j.actpsy.2019.05.012
Corradi, G., Chuquichambi, E. G., Barrada, J. R., Clemente, A., & Nadal, M. (2020). A new conception of visual aes-
thetic sensitivity. British Journal of Psychology, 111(4), 630–658. https://doi.org/10.1111/bjop.12427
Dai, D. Y., & Schader, R. (2001). Parents’ reasons and motivations for supporting their child’s music training. Roeper
Review, 24(1), 23–26. https://doi.org/10.1080/02783190109554121
Day, M. D. (1976). Effects of instruction on high school students’ art preferences and art judgments. Studies in Art Educa-
tion, 18(1), 25–39. https://doi.org/10.2307/1319516
Dewar, H. (1938). A comparison of tests of artistic appreciation. British Journal of Educational Psychology, 8(1), 29–49. 
https://doi.org/10.1111/j.2044-8279.1938.tb03181.x
Elliot, D. J. (1995). Music matters. Oxford University Press.
Eysenck, H. J. (1940). The “general factor” in aesthetic judgements. British Journal of Psychology, 31, 94–102.

249
Ana Clemente

Eysenck, H. J. (1941a). ‘Type’-factors in esthetic judgments. British Journal of Psychology, 31, 262–270.
Eysenck, H. J. (1941b). A critical and experimental study of colour preferences. American Journal of Psychology, 54(3),
385–394. https://doi.org/10.2307/1417683
Eysenck, H. J. (1941c). The empirical determination of an aesthetic formula. Psychological Review, 48(1), 83–92.  https://
doi.org/10.1037/h0062483
Eysenck, H. J. (1942). The experimental study of the “good Gestalt”: A  new approach.  Psychological Review, 49(4),
344–364.  https://doi.org/10.1037/h0057013
Eysenck, H. J. (1972a). Personal preferences, aesthetic sensitivity and personality in trained and untrained subjects. Journal
of Personality, 40(4), 544–557.  https://doi.org/10.1111/j.1467-6494.1972.tb00079.x
Eysenck, H. J. (1972b). Preference judgments for polygons, designs and drawings. Perceptual and Motor Skills, 34(2),
396–398. https://doi.org/10.2466/pms.1972.34.2.396
Eysenck, H. J. (1981). Aesthetic preferences and individual differences. In D. O’Hare (Ed.), Psychology and the arts
(pp. 76–101). Harvester Press.
Eysenck, H. J. (1983). A  new measure of “good taste” in visual art. Leonardo, 16(3), 229–231. https://doi.
org/10.2307/1574921
Eysenck, H. J., & Castle, M. (1971). Comparative study of artists and nonartists on the Maitland Graves design judgment
test. Journal of Applied Psychology, 55(4), 389–392.  https://doi.org/10.1037/h0031469
Eysenck, H. J., & Iwawaki, S. (1971). Cultural relativity in aesthetic judgments: An empirical study. Perceptual and Motor
Skills, 32(3), 817–818. https://doi.org/10.2466/pms.1971.32.3.817
Eysenck, H. J., & Iwawaki, S. (1975). The determination of aesthetic judgment by race and sex. Journal of Social Psychol-
ogy, 96(1), 11–20. https://doi.org/10.1080/00224545.1975.9923256
Fechner, G. T. (1876). Vorschule der Ästhetik. Breitkopf und Härtel.
Fehl, P. P. (1953). Tests of taste. College Art Journal, 12(3), 232–248. https://doi.org/10.2307/773820
Friedenberg, J. (2018). Geometric regularity, symmetry and the perceived beauty of simple shapes. Empirical Studies of the
Arts, 36(1), 71–89. https://doi.org/10.1177/0276237417695454
Frois, J. P., & Eysenck, H. J. (1995). The visual aesthetic sensitivity test applied to Portuguese children and fine art stu-
dents. Creativity Research Journal, 8(3), 277–284. https://doi.org/10.1207/s15326934crj0803_6
Furnham, A., & Chamorro-Premuzic, T. (2004). Personality, intelligence, and art. Personality and Individual Differences,
36(3), 705–715. https://doi.org/10.1016/S0191-8869(03)00128-4
Gartus, A., & Leder, H. (2017). Predicting perceived visual complexity of abstract patterns using computational measures:
The influence of mirror symmetry on complexity perception. PloS One, 12(11), e0185276. https://doi.org/10.1371/
journal.pone.0185276
Gernet, S. K. (1940). Musical discrimination at various age and grade levels. College Press.
Gold, B. P., Pearce, M. T., Mas-Herrero, E., Dagher, A.,  & Zatorre, R. J. (2019). Predictability and uncertainty in
the pleasure of music: A  reward for learning?  Journal of Neuroscience,  39(47), 9397–9409. https://doi-org.sire.
ub.edu/10.1523/JNEUROSCI.0428-19.2019
Götz, K. O., Borisy, A. R., Lynn, R., & Eysenck, H. J. (1979). A new visual aesthetic sensitivity test: I Construction and
psychometric properties. Perceptual and Motor Skills, 49(3), 795–802. https://doi.org/10.2466/pms.1979.49.3.795
Graves, M. (1939). What is your I. Q. in design? Art lnstructor, 4(1), 11–14.
Graves, M. (1948). Design judgment test. Psychological Corporation.
Hahn, M. E. (1954).  A proposed technique for investigating the relationship between musical preferences and personality struc-
ture [Doctoral dissertation, University of Kansas].
Hevner, K. (1930). Tests for esthetic appreciation in the field of music. Journal of Applied Psychology, 14(5), 470–477. 
https://doi.org/10.1037/h0071346
Hevner, K. (1937). An experimental study of the affective value of sounds in poetry. American Journal of Psychology, 49(3),
419–434. https://doi.org/10.2307/1415776
Hevner, K., & Mueller, J. H. (1939). The effectiveness of various types of art appreciation aids. Journal of Abnormal and
Social Psychology, 34(1), 63–72. https://doi.org/10.1037/h0059214
Iwao, S., & Child, I. L. (1966). Comparison of esthetic judgments by American experts and by Japanese potters. Journal
of Social Psychology, 68(1), 27–33. https://doi.org/10.1080/00224545.1966.9919662
Iwawaki, S., Eysenck, H. J.,  & Götz, K. O. (1979). A  new visual aesthetic sensitivity test (VAST): II. Cross-cultural
comparison between England and Japan. Perceptual and Motor Skills, 49(3), 859–862. https://doi.org/10.2466/
pms.1979.49.3.859
Jacobsen, T., & Höfel, L. (2001). Aesthetics electrified: An analysis of descriptive symmetry and evaluative aesthetic judg-
ment processes using event-related brain potentials. Empirical Studies of the Arts, 19(2), 177–190. https://doi-org.sire.
ub.edu/10.2190/P7W1-5F1F-NJK9-X05B

250
Aesthetic sensitivity

Kwalwasser, D. J., & Dykema, P. W. (1930). Manual of directions, K-D tests. Fischer.


Kyme, G. H. (1967). A study of the development of musicality in the junior high school and the contributions of musi-
cal composition to this development. Bulletin of the Council for Research in Music Education, 15–24. http://www.jstor.
org/stable/40316931
Martindale, C. (2001). How does the brain compute aesthetic preference? General Psychologist, 36, 25–35.
Mastandrea, S., Bartoli, G., & Bove, G. (2009). Preferences for ancient and modern art museums: Visitor experiences
and personality characteristics. Psychology of Aesthetics, Creativity, and the Arts, 3(3), 164–173. https://doi.org/10.1037/
a0013142
McAdory, M. (1929). McAdory art test. Teach Coll Bur Publisher.
McManus, I. C., & Furnham, A. (2006). Aesthetic activities and aesthetic attitudes: Influences of education, background
and personality on interest and involvement in the arts. British Journal of Psychology, 97(4), 555–587. https://doi.org/
10.1348/000712606X101088
Meier, N. C. (1926). Aesthetic judgment as a measure of art talent. Series of Aims & Progress of Research (Vol. 1, pp. 19,
30). University of Iowa Studies.
Meier, N. C. (1927). Can art talent be discovered by test devices? Western Arts Association Bulletin, 11, 74–79.
Meier, N. C. (1928). A  measure of art talent.  Psychological Monographs, 39(2), 184–199.  https://doi.org/10.1037/
h0093346
Meier, N. C. (1934). What we now know about talent in children (pp. 143–163). Western Arts Association Annual Report.
Meier, N. C. (1939). Factors in artistic aptitude: Final summary of a ten-year study of a special ability. Psychological Mono-
graphs, 51(5), 140–158. https://doi.org/10.1037/h0093484
Meier, N. C. (1940). Meier Art tests. I. Art judgment. State University of Iowa, Bureau of Educational Research and
Service.
Meier, N. C. (1942). The Meier art judgment test. Bureau of Educational Research and Service, University of Iowa.
Meier, N. C. (1963). Meier art tests. II: Aesthetic perception. Bureau of Educational Research and Service, University of
Iowa.
Meier, N. C., & Seashore, C. E. (1929). The Meier-Seashore art judgment test. Bureau of Educational Research, University
of Iowa.
Michel, J. B., Shen, Y. K., Aiden, A. P., Veres, A., Gray, M. K., Google Books Team, Pickett, J. P., Hoiberg, D., Clancy,
D., Norvig, P., Orwant, J., Pinker, S., Nowak, M. A., & Aiden, E. L. (2011). Quantitative analysis of culture using
millions of digitized books. Science, 331(6014), 176–182. https://doi.org/10.1126/science.1199644
Mitchells, K. (1966). Aesthetic perception and aesthetic qualities. In  Proceedings of the Aristotelian society  (Vol. 67,
pp. 53–72). Aristotelian Society, Wiley. www.jstor.org/stable/4544736
Myszkowski, N., Çelik, P., & Storme, M. (2018). A meta-analysis of the relationship between intelligence and visual
“taste” measures. Psychology of Aesthetics, Creativity, and the Arts, 12(1), 24–33.  https://doi.org/10.1037/aca0000099
Myszkowski, N., & Storme, M. (2017). Measuring “good taste” with the visual aesthetic sensitivity test-revised (VAST-
R). Personality and Individual Differences, 117, 91–100. https://doi.org/10.1016/j.paid.2017.05.041
Myszkowski, N., Storme, M., Zenasni, F., & Lubart, T. (2014). Is visual aesthetic sensitivity independent from intel-
ligence, personality and creativity? Personality and Individual Differences, 59, 16–20. https://doi.org/10.1016/j.
paid.2013.10.021
Myszkowski, N.,  & Zenasni, F. (2016). Individual differences in aesthetic ability: The case for an aesthetic quotient.
Frontiers in Psychology, 7, 750. https://doi.org/10.3389/fpsyg.2016.00750
Nadal, M., Munar, E., Marty, G., & Cela-Conde, C. J. (2010). Visual complexity and beauty appreciation: Explaining the
divergence of results. Empirical Studies of the Arts, 28(2), 173–191. https://doi-org.sire.ub.edu/10.2190/EM.28.2.d
Norton, B. (1981). Psychologists and class. In C. Webster (Ed.), Biology, medicine and society 1840–1940 (pp. 289–314).
Cambridge University Press.
Pang, C. Y., Nadal, M., Müller-Paul, J. S., Rosenberg, R., & Klein, C. (2013). Electrophysiological correlates of looking
at paintings and its association with art expertise. Biological Psychology, 93(1), 246–254. https://doi.org/10.1016/j.
biopsycho.2012.10.013
Parsons, M., & Durham, M. (1979). A cognitive-developmental approach to aesthetic experience. In R. L. Mosher (Ed.),
Adolescents’ development and education (pp. 290–335). University of California Press.
Parsons, M., Johnston, M., & Durham, R. (1978). Developmental stages in children’s aesthetic responses. Journal of Aes-
thetic Education, 12(1), 83–104. https://doi.org/10.2307/3331850
Payne, E. (1967). Musical taste and personality.  British Journal of Psychology,  58(1), 133–138. https://doi.org/
10.1111/j.2044-8295.1967.tb01066.x
Powel, L., Thorndike, E. L., & Woodyard, E. (1942). The aesthetic life of communities. Journal of Aesthetics and Art Criti-
cism, 2(7), 51–58. https://doi.org/10.1111/1540_6245.jaac2.7.0051

251
Ana Clemente

Reber, R., Schwarz, N., & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the perceiver’s
processing experience? Personality and Social Psychology Review, 8(4), 364–382.
Reimer, B. (1965). The development of aesthetic sensitivity.  Music Educators Journal,  51(3), 33–36. https://doi.
org/10.2307/3390339
Reimer, B. (1968a). Developing aesthetic sensitivity in the junior high school general music class. Journal of Aesthetic
Education, 2(2), 97–107. https://doi.org/10.2307/3331264
Reimer, B. (1968b). Performance and aesthetic sensitivity.  Music Educators Journal,  54(7), 27–114. https://doi.
org/10.2307/3391241
Robbins, H., Smith, D., & Washburn, M. F. (1915). The influence of fatigue on affective sensitiveness to colours. Ameri-
can Journal of Psychology, 26(2), 291–292. https://doi.org/10.2307/1413260
Siceloff, M. M., Woodyard, E, & staff of the division of psychology. (1933). Validity and standardization of the McAdory art
test. Teacher’s College Press.
Silvia, P. J., & Barona, C. M. (2009). Do people prefer curved objects? Angularity, expertise, and aesthetic preference.
Empirical Studies of the Arts, 27(1), 25–42. https://doi.org/10.2190/EM.27.1.b
Skov, M. (2019). Aesthetic appreciation: The view from neuroimaging. Empirical Studies of the Arts, 37(2), 220–248.
https://doi.org/10.1177/0276237419839257
Skov, M. (2020). The neurobiology of sensory valuation. In The Oxford handbook of empirical aesthetics. Oxford University
Press. https://doi.org/10.1093/oxfordhb/9780198824350.013.7.
Smets, G., & Knops, L. (1976). Measuring visual esthetic sensitivity: An alternative procedure. Perceptual and Motor Skills,
42(3), 867–874. https://doi.org/10.2466/pms.1976.42.3.867
Soueif, M. I., & Eysenck, H. J. (1971). Cultural differences in aesthetic preferences 1. International Journal of Psychol-
ogy, 6(4), 293–298. https://doi.org/10.1080/00207597108246695
Stephenson, W. (1936). The inverted factor technique.  British Journal of Psychology,  26(4), 344–361. https://doi.
org/10.1111/j.2044-8295.1936.tb00803.x
Suga, Y. (2003). “Purgatory of taste” or projector of industrial Britain? The British Institute of Industrial Art. Journal of
Design History, 16(2), 167–185. https://doi.org/10.1093/jdh/16.2.167
Taunton, M. (1982). Aesthetic responses of young children to the visual arts: A review of the literature. Journal of Aesthetic
Education, 16(3), 93–109. https://doi.org/10.2307/3332196
Thorndike, E. L. (1913). The measurement of achievement in drawing. Teachers College Record, 14, 345–382.
Thorndike, E. L. (1916). Tests of esthetic appreciation.  Journal of Educational Psychology, 7(9), 509–522.  https://doi.
org/10.1037/h0073375
Thorndike, E. L. (1917). Individual differences in judgments of the beauty of simple forms. Psychological Review, 24(2),
147–153.  https://doi.org/10.1037/h0073175
Thorndike, E. L. (1924). A scale for general merit of children’s drawings. Teacher’s College Press, Columbia University.
Thorndike, E. L., & Woodyard, E. (1943). The relation between the aesthetic status of a community and its status in
other respects. American Journal of Sociology, 49(1), 59–59. https://doi.org/10.1086/219310
Trabue, M. R. (1923). Scales for measuring judgment of orchestral music. Journal of Educational Psychology, 14(9), 545–
561.  https://doi.org/10.1037/h0074161
Vernon, P. E. (1930). A method for measuring musical taste. Journal of Applied Psychology, 14(4), 355–362.  https://doi.
org/10.1037/h0071071
Voss, M. D. (1936). A study of conditions affecting the functioning of the art appreciation process at the child-level. Psy-
chological Monographs, 48(1), 1–39.  https://doi.org/10.1037/h0093364
Washburn, M. F., Hatt, E., & Holt, E. B. (1923). Affective sensitiveness in poets and in scientific students. American Journal
of Psychology, 34(1), 105–106. https://doi.org/10.2307/1413933
Webster, P. R. (1988a). Creative thinking and music education. Design for Arts in Education, 89(5), 33–37. https://doi.org
/10.1080/07320973.1988.9935522
Webster, P. R. (1988b). New perspectives on music aptitude and achievement. Psychomusicology, 7(2), 177–194.  https://
doi.org/10.1037/h0094169
Welsh, G. S. (1949). A projective figure-preference test for diagnosis of psychopathology: 1. A preliminary investigation [Doctoral
Dissertation, University of Minnesota].
Welsh, G. S. (1959). Preliminary manual. Welsh Figure Preference Test.
Welsh, G. S., & Barron, F. X. (1949). Barron-Welsh art scale. Consulting Psychology Press.

252
Aesthetic sensitivity

Westphal-Fitch, G., Oh, J., & Fitch, W. T. (2013). Studying aesthetics with the method of production: Effects of context
and local symmetry. Psychology of Aesthetics, Creativity, and the Arts, 7(1), 13–26. https://doi.org/10.1037/a0031795
Williams, H. M., & Hattwick, M. S. (1932). The measurement of musical development (Vol. 11, No. 2). University of Iowa.
Winner, E., Rosenblatt, E., Windmueller, G., Davidson, L., & Gardner, H. (1986). Children’s perception of “aesthetic”
properties of the arts: Domain-specific or pan-artistic?  British Journal of Developmental Psychology,  4(2), 149–160. 
https://doi.org/10.1111/j.2044-835X.1986.tb01006.x

253
13
THE EVOLUTION OF SENSORY
VALUATION SYSTEMS
Esther Ureña and Marcos Nadal

Ernst Mayr (1961), one of the architects of the modern evolutionary synthesis, distinguished two sorts of
questions that can be asked about biological systems. Functional questions, on the one hand, ask about the
way structural elements operate and interact to make the system work. Evolutionary questions, on the other
hand, ask about the historical events that brought those elements and their features into existence. The focus
of most chapters in this volume is on functional questions about aesthetics: how do different brain regions
work together to attach hedonic value to objects and events (Chapter 2)? How do they generate the experi-
ences of liking and disliking (Chapters 3 and 4)? How do feelings of pleasure and appreciation arise in differ-
ent sensory domains and for different classes of objects (Chapters 6 to 10)? The focus of this chapter, unlike
the preceding ones, is on evolutionary questions about the biological systems underlying aesthetic valuation.
Its aim is to summarize what is known about the evolution of sensory valuation systems and their features.
The possibility of answering evolutionary questions about aesthetics has roused a good deal of scepticism,
even among evolutionary biologists. This scepticism is grounded on the view that aesthetic valuation is a
uniquely human trait, unparalleled in other animals:

Aesthetics and art are usually regarded as exclusively human possessions. Sensitivity to beauty and
making or doing things that are perceived as ‘beautiful’ are among the traits that elevate man above
the brutes. This renders the problem of the origin and biological meaning of art and aesthetics in
human evolution particularly challenging.
(Dobzhansky, 1962, p. 214)

If aesthetic valuation is a uniquely human trait, it follows that it must have appeared along the human
evolutionary lineage, after it diverged from the lineage of chimpanzees: “the aesthetic sense evolved in
response to important new selection pressures long after the human lineage had separated from that of the
apes” (Washburn, 1970, p. 824). To complicate matters further, there seems to be little hope of finding mate-
rial evidence for the evolution of this peculiarly human trait: “It does not seem possible to determine the
phylogenetic origin of the human capacity for appreciating beauty. Neither fossil nor archaeological records
provide convincing evidence to ascertain the appearance of this competence” (Ayala, 2017, p. 613).
The view that aesthetic liking and pleasure are somehow different to other forms of liking and pleasure is
not uncommon (Nadal & Skov, 2018; Skov & Nadal, 2018, 2020). There is a long history of efforts directed
toward identifying what makes aesthetic liking and pleasure special, from the earliest days of psychology to

254 DOI: 10.4324/9781003008675-14


Evolution of sensory valuation systems

the present (Bain, 1883; Christensen, 2017; Fingerhut  & Prinz, 2020; James, 1884, 1890; Menninghaus
et al., 2017, 2019; Sully, 1892). The historical reasons behind this view have been explained in Chapter 1 of
this volume and have to do with assumptions about aesthetics and the mind that filtered from 18th-century
philosophy to 19th-century psychology and to 20th-century neuroscience.

Aesthetic valuation is rooted in common sensory valuation systems


As presented in detail in the preceding chapters of this volume, neuroaesthetics has produced abundant evi-
dence showing that aesthetic valuation is rooted in general sensory valuation systems. Sensory valuation is a
process by which sensory information is conveyed to the reward circuit, where it is imbued with a hedonic
value that depends on its relevance to the internal state, needs and goals, and the external situation and con-
text. The reward circuit is a distributed system of brain regions that include the nucleus accumbens, caudate
nucleus, pallidum, amygdala, orbitofrontal cortex, anterior cingulate cortex, and insula (Bartra et al., 2013;
Brown et  al., 2011; Haber  & Knutson, 2010; Sescousse et  al., 2013). Reward signals computed in these
regions attach a hedonic value to information relayed from sensory cortices about the perceptual attributes of
objects (Becker et al., 2019; Berridge & Kringelbach, 2015; Skov, 2019). Let us illustrate this with two stud-
ies. Salimpoor and colleagues (2013) measured blood oxygenation level–dependent activity while partici-
pants listened to excerpts of unfamiliar music and placed economic bids to listen to them again. Their results
showed that activity in the nucleus accumbens was the best predictor of the amount participants were willing
to bid and that functional connectivity between the nucleus accumbens and the primary and surrounding
auditory cortices increased significantly when participants listened to the excerpts they found most desirable.
In another study, Cheung and colleagues (2019) showed that musical pleasure arises from combinations of
the uncertainty of perceivers’ musical expectations and their surprise when musical events deviate from those
expectations: musical pleasure is greatest when events are highly surprising in a low-uncertainty context
or when events are not very surprising in a high uncertainty context. Moreover, the interaction between
uncertainty and surprise was related to brain activity in the amygdala, hippocampus, and auditory cortex.
Sensory valuation, therefore, involves integrating information about perceptual attributes, such as tonal
patterns in the auditory domain or contour and symmetry in the visual domain, with information about
hedonic attributes, such as reward prediction or reward value. The integration of sensory and hedonic pro-
cesses is crucial for two reasons. First, sensory information that is not relayed to the reward circuit does not
acquire hedonic value. This is the case with specific musical anhedonia (SMA), the inability to experience
pleasure from music. People with SMA have reduced white matter connectivity between auditory brain
regions and the ventral striatum, a key region of the brain’s reward circuit (Sachs et al., 2016). Even in people
without SMA, individual differences in experienced musical pleasure correlates with differences in connec-
tivity between the auditory cortex and the reward circuit (Loui et al., 2017; Martínez-Molina et al., 2016).
Second, the integration of sensory and hedonic information marks the distinction between different sorts of
hedonic values. The same reward circuit is involved in the pleasurable experiences we get from many sources,
including music, painting, food, sex, and drugs (Levy & Glimcher, 2012; Mallik et al., 2017; Nadal & Skov,
2018). What distinguishes those pleasures from each other is the sort of sensory information that is relayed to
the reward circuit and the path it is relayed along (Mas-Herrero et al., 2021; Sescousse et al., 2013).
There are good reasons, therefore, to doubt that anything like a distinct aesthetic faculty appeared at some
point in human evolution after our lineage split from the chimpanzee lineage, giving us a taste for beauty and
art. The neuroscientific evidence shows that aesthetic liking and aesthetic pleasure rely on brain systems we
share with other animals and that evolved to signal relevant objects and situations by making them pleasur-
able (Nadal & Skov, 2018; Skov & Nadal, 2018). But not only is it the case that we share common reward
systems with other animals; there is abundant evidence that preference for sensory features and configurations
is widespread in the animal kingdom.

255
Esther Ureña and Marcos Nadal

Preferences are ubiquitous among animals


Empirical aesthetics and neuroaesthetics have devoted much effort to understanding the psychological and
neurobiological mechanisms underlying people’s liking and preference for object attributes (Berlyne, 1972;
Meier, 1942; Nadal & Vartanian, 2022; Valentine, 1962). Some of these attributes, including colour hue, col-
our brightness, regularity and symmetry, proportion, and curvature, have received much attention, because
they are often regarded as the basic formal or design elements of plastic artworks (Feldman, 1971; Lopes,
2005). We might wonder why humans evolved a preference for symmetry or curvature or what the adaptive
advantage conferred to early humans by such preferences was, as if these were aspects of a uniquely human
aesthetic sense. As it happens, however, humans are not unique in having preferences for colours, shapes,
and patterns. Many classes of animals, such as birds, rodents, and primates, also have preferences for colours,
shapes, arrangements, and movements.

Colour preferences
Studies on human preferences for colours go back to the earliest days of empirical aesthetics (Nadal & Ureña,
2021). Some of the first studies suggested there was no accounting for colour preferences, but these were
fraught with methodological problems (Eysenck, 1941). Later studies improved experimental control of
materials and procedures, and showed that people generally agree in their colour preferences: People gener-
ally prefer cooler colour hues, like green and blue, to warmer colour hues, like red, orange, and yellow, and
they prefer bright to dim colours (Palmer et al., 2013; Vartanian, 2022).
Humans are far from the only animals that have colour preferences. Newly hatched chicks show colour
preferences during imprinting. These young birds prefer to follow blue objects rather than red objects, and
any of these rather than green objects. The object colours they least prefer to follow are orange, grey, yellow,
and white (Schaefer & Hess, 1959). Young chicks, herring gulls, and pigeons also exhibit colour preferences
when pecking on objects: they preferentially peck on objects in the blue and orange regions of the spectrum
(Delius & Thompson, 1970; Hess, 1956; Sahgal & Iversen, 1975). These preferences seem to be innate, as
they are observed even when the chicks hatched in total darkness (Fischer et al., 1975; Salzen et al., 1971),
as they are not modified by pre-hatch stimulation with different coloured lights (Fischer et al., 1975), and as
chicks learn faster in colour discrimination tasks when they are reinforced for pecking on preferred colours
than when they are reinforced for pecking on nonpreferred colours (Kovach & Hickox, 1971). However,
early exposure to colours interacts with these innate preference tendencies (Kovach, 1971; Salzen et  al.,
1971): Chicks reared for three days with red or blue stimuli show strong preferences for the colours they
were reared with, but chicks reared for 3 days with green stimuli prefer red and green equally (Salzen et al.,
1971). Chicks also show brightness preferences. They prefer bright-coloured stimuli to dim-coloured stimuli
(Zolman & Lattin, 1972), and this preference is not eliminated by incubating, hatching, or rearing for 3 days
in darkness (Zolman & Lattin, 1972), though it becomes stronger with exposure to light during the first 3
days of life (Zolman & Lattin, 1972). Finally, colour preferences have also been observed in nonhuman pri-
mates. Several studies coincide in finding that the preferred colour of macaques and squirrel monkeys is blue,
followed by green, yellow, orange, and red (Green et al., 1966; Humphrey, 1971, 1972; Sahgal et al., 1975).
Thus, well-controlled studies show that animal and human colour preferences are quite similar (McManus
et al., 1981). What role might preference for colour hue and brightness play in chicks and monkeys? It has
been suggested that they are the result of adaptive perceptual templates that draw attention to, help discrimi-
nate, and elicit approach toward informative features of the environment during development (Kovach &
Wilson, 1981). At certain stages, these preferences seem to be quite fixed, while at other stages learning expe-
riences through interaction with the environment can lead to modifications of some of these preferences.

256
Evolution of sensory valuation systems

Preference for curvature


People generally prefer curvature to angularity. They prefer objects (Bar & Neta, 2006, 2007), designs (West-
erman et al., 2012), geometric figures (Bertamini et al., 2016; Palumbo & Bertamini, 2016; Palumbo et al.,
2015; Silvia & Barona, 2009), rooms (Vartanian et al., 2013), and artworks (Ruta et al., 2021) with curved
contours to those with sharp-angled contours. Preference for curvature has been observed using several dif-
ferent experimental paradigms (Gómez-Puerto et al., 2016; Palumbo & Bertamini, 2016) and even in young
infants ( Jadva et al., 2010) and newborns (Fantz & Miranda, 1975). Preference for curvature might be com-
mon, but it by no means fixed. Studies have shown that is modulated by the semantic content of the objects,
presentation time, experience, expertise, and familiarity (Chuquichambi et al., 2021; Corradi et al., 2019;
Silvia & Barona, 2009; Vartanian et al., 2018).
Preference for curvature has also been observed in birds, rodents, and nonhuman primates. Very young
chicks approach, and peck on, curved forms, such as circles and ellipsoids, more than angular forms, such
as triangles and stars (Fantz, 1957; Goodwin & Hess, 1969). Preference for curvature seems to be innate,
as it has been observed in the absence of any prior visual experience, it persists for weeks even when unre-
warded, and it takes chicks longer to learn not to respond to curved forms than to learn not to respond to
angular forms (Fantz, 1957; Zolman, 1969; Zolman et al., 1975). Rodents have also shown preference for
curvature: Rats (Rattus norvergicus) prefer to explore rounded objects over cylindrical ones (Winne et al.,
2015). Nonhuman primates, including orangutans, gorillas, and chimpanzees, also prefer to look at or
choose objects with curved contours over objects with angular contours (Ebel et al., 2020; Munar et al.,
2015).

Preference for order, regularity, and symmetry


Humans generally prefer regular, ordered, and symmetrical arrangements to irregular, disordered, and asym-
metrical ones (Bode et al., 2017; Nadal et al., 2010; Van Geert & Wagemans, 2019; Westphal-Fitch et al.,
2012). This, of course, does not mean that everyone prefers order and regular designs on every occasion.
These preferences seem to be modulated by a multitude of factors. Even the general preference for symmetry
can be modulated by affect, personality, familiarization, and expertise (Gollwitzer & Clark, 2019; Gollwitzer
et al., 2020; Leder et al., 2019; Pecchinenda et al., 2014; Tinio & Leder, 2009; Weichselbaum et al., 2018)
Again, far from being a uniquely human attribute, preference for regularity is widespread among different
classes of animals. Birds (Coloeus monedula, Corvus corone) show a preferential selection for regular and sym-
metrical geometrical patterns (Rensch, 1958; Tigges, 1963). Rensch (1957) showed that a capuchin monkey
(Cebus apella) and a vervet monkey (Chlorocebus aethiops) preferred to interact with geometric patterns that
contained lines with good continuation, radial and bilateral symmetry, regularity, and repetition of patterns
than with similar patterns that lacked these features and were thus irregular. Anderson and colleagues (2005)
replicated and extended Rensch’s (1957) original observations: When presented with pairs of geometric pat-
terns differing in regularity and symmetry, primates—capuchin monkeys (Cebus apella) and squirrel monkeys
(Saimiri sciureus)—systematically approached the regular and symmetrical ones first.

Preference for biological motion


When given a choice, people usually prefer to look at patterns of biological motion than other motion
patterns. This is true even for human newborns in the first days of life, suggesting that we are born with a
predisposition to latch onto biological motion patterns (Sifre et al., 2018; Simion et al., 2008), regardless of
whether the movement corresponds to another person or an animal (Bardi et al., 2011).

257
Esther Ureña and Marcos Nadal

Humans share with other animals this preference for biological motion. Other animals seem to discrimi-
nate and prefer autonomously moving objects to objects that move because they have been pushed: Newly
hatched, visually naïve domestic chicks prefer moving objects that are self-propelled to moving objects
caused by physical contact (Mascalzoni et al., 2010). Moreover, newly hatched chicks that were reared and
hatched in darkness, and thus have no visual experience whatsoever, prefer to approach biological motion
patterns over other kinds of motion patterns (Sifre et al., 2018). This preference is not specific to the motion
of hens but also potential predators, such as cats (Vallortigara & Regolin, 2006; Vallortigara et al., 2005). It
has been suggested that preference for biological motion in chicks is crucial for imprinting during the early
stages of development (Miura & Matsushima, 2016). Preference for biological motion has also been observed
in fish and seems to serve the recognition of animal species, sex, and group members and be crucial in social
engagement and in promoting shoaling behaviour (Larsch  & Baier, 2018; Nakayasu  & Watanabe, 2014;
Nunes et al., 2020).

Looking at sensory valuation through the lens of evolution


The evidence presented previously demonstrates that humans are not alone in having preferences for percep-
tual attributes. We have seen that, like humans, many kinds of animals, from fish to primates, have prefer-
ences for colours, curved contours, order and regularity, and biological motion. Although we are far from
understanding exactly what their function is, how they evolved, and how the neural systems underlying
them differ from one class of animals to another, these are widespread and robust preferences, which suggest
that they play a crucial role in development and survival (Bardi et al., 2011; Sifre et al., 2018). They seem
to be the result of basic perceptual mechanisms that bias animals’ attention towards individually and socially
relevant environmental cues. Colours, shapes, and movements, together with a plethora of information from
other sensory domains we have not examined here, serve as signals to attend to and engage with objects
and conspecifics during key developmental stages, triggering the relevant consummatory (e.g., pecking) and
affiliative behaviours (e.g., following and shoaling). Because some of these preferences require no previous
stimulation, they seem to be innate predispositions. But that does not mean they are fixed and unchang-
ing. As noted previously, some of these innate predispositions are malleable to a certain extent by learning
experiences.
It would seem, therefore, that human preferences for colour, curvature, order and regularity, and biologi-
cal motion—often referred to as aesthetic preferences—did not appear anew along the human lineage after
it split from the chimpanzee lineage. They evolved upon the foundations of animal preferences that served
to direct attention to relevant aspects of the environment, like conspecifics and potential food, and to create
opportunities for learning about the environment and others. A thorough explanation of aesthetic liking
requires understanding the evolution of the underlying neural systems: What were the circumstances that
led to their appearance and subsequent modifications? Why did they take the form they did? What was their
function? What was their adaptive value? Many of the details of the evolution of sensory valuation systems
remain obscure and might remain so forever. But there is much to be learnt about the features of sensory
valuation systems if we look at them through the lens of evolution. And the first thing we learn is that they
originated in some of the basic functions of deceptively simple single-celled organisms.

From single cells to the first nervous systems


No living cell is isolated from its environment: All cells respond to different sorts of stimuli and process the
signals they carry (Lichtneckert  & Reichert, 2009). Complex networks of biochemical reactions enable
single celled protozoans to organize their behaviour based on inner and outer conditions (Bray, 1995; Fitch,
2021). Single celled organisms, such as paramecia, amoebae, bacteria, and ciliates, respond to variations in

258
Evolution of sensory valuation systems

their environments, compute and store changes in chemical gradients, locate prey and potential mates, avoid
predators and aversive stimuli, learn through habituation and repeated exposure to the same conditions, and
anticipate periodic stimulation (Coyle et al., 2019; Dexter et al., 2019; Dussutour, 2021; Lan & Tu, 2016;
Tang & Marshall, 2018). These elemental forms of behaviour are the product of basic mechanisms of infor-
mation processing and storage carried out by cellular ion channels, receptors, vesicular transporters, signal-
ling molecules, and genetic regulatory circuits (Fitch, 2021; Göhde et al., 2021).
Thus, the molecular substrates of the computations underlying sensory valuation predate the appearance
of metazoans (Figure 13.1):

most of the basic elements of cognition were already present and functional before the nervous sys-
tem evolved. The ability to selectively perceive specific stimuli, the discrimination between favour-
able and unfavourable, the assessment of the overall valence of a situation, the retention of memory,
and the integration of information for decision-making—all of this was in place in one form or
another in unicellular organisms and early metazoans that did not (yet) possess a nervous system.
(Arendt, 2021, p. 1)

Figure 13.1 Simplified representation of the evolutionary relations among the major groups of organisms mentioned
in the main text. The groups highlighted in bold correspond to the major transitions in the evolution of
sensory valuation systems.

259
Esther Ureña and Marcos Nadal

Metazoans are multicellular eukaryotic organisms that evolved from colonies of protists about 900 mil-
lion years ago. Except for sponges and placozoans, all metazoans possess a network of excitable cells with
projections that connect to each other and to other cells that conform different sorts of nervous system
(Burkhardt  & Jékely, 2021). Neurons evolved over 700  million years ago from epithelial excitable secre-
tory cells that became able to transduce external mechanical, electromagnetic, or chemical stimulation and
transmit these signals to adjoining cells (Cisek, 2022; Göhde et al., 2021; Varoqueaux & Fasshauer, 2017)
(Figure 13.2). Neurons show many of the excitatory, regulatory, and secretory features of single-celled organ-
isms, such as choanoflagellates: they receive information from the interior and exterior of the organisms,
process and store it, and use it to organize behaviour and are also capable of generating activity endogenously
(Göhde et al., 2021). What sets neurons apart is their unique morphological and functional specialization for
electrical and chemical transmission of signals, secretion of biologically active substances, and endogenous
generation of pace-making activity through rhythmically oscillating potentials (Lichtneckert  & Reichert,
2009; Liebeskind et al., 2017; Schneider, 2014).
Networks of neurons endowed early multicellular organisms with the capacity to coordinate whole-body
movements during feeding and locomotion and the capacity to integrate different sorts of information from
the environment and the body’s own movement (Arendt, 2021; Cisek, 2022). These networks produced a
system of intercellular communication that allowed integration, that was faster and more targeted, and that
included a new kind of memory in the form of synapses (Ginsburg & Jablonka, 2021). The Ctenophora and
Cnidaria are the oldest extant animal lineages with nervous systems (Burkhardt & Jékely, 2021; Schneider,
2014). Ctenophora, or comb jellies, are planktonic gelatinous animals that generally have two tentacles for
capturing prey and eight rows of swimming ciliary combs (Figure 13.3). Their nervous system is a diffuse
network of neurons that from tracts below the combs and around the mouth and pharynx. These organ-
isms have sensory nerve cells interspersed among the epithelial cells and concentrated at the apical organ,
consisting of nerve cells and a statocyst, the balance sensory receptor. The apical organ directs the comb jel-
lies’ movement through pace-making activity transmitted through the comb plate rows, and excitatory and
inhibitory paths coordiante the activity of comb plate cells and tentacle movements that enable capturing and
ingesting prey. Cnidarians have evolved a broad array of nervous systems, including diffuse two-dimensional
nets formed by different types of neurons and with different concentrations; intermediate neurons between
sensory neurons and motor neurons; ring-shaped nerve tracts that serve as integrative centres where periph-
eral paths converge and behaviour is generated; and specialized sensory organs around the bell margin, such
as statocysts and ocelli and even lens eyes (Lichtneckert & Reichert, 2009; Schneider, 2014).

The origin of the central nervous system in early bilaterians


The nervous system of bilaterians evolved, about 540–560  million years ago (Burkhardt  & Jékely, 2021;
Marshall, 2006), from the ectodermal nerve net in a cnidarian ancestor (Figure 13.1). Several similarities in
the expression of regional patterning genes in cnidarians and bilaterians, as well as in genes that confer neural
identity, suggest that the blastopore of a cnidarian-like ancestor of bilaterians became elongated and formed
a slit. This gastric slit subsequently closed partially, creating a separate mouth and anus (Nielsen et al., 2018).
The mouth evolved at one end and the anus at the other, and a nerve cord appeared from the fused tissue in
between, with an anterior concentration of neurons (Holland, 2017). The appearance of an anteroposterior
axis also brought about the differentiation of body segments aligned along this axis and the development of
specialized body parts. Similarities in the expression of genes that mediate anteroposterior and mediolateral
patterning of the central nervous system (CNS) in lampreys, amphibians, mammals, annelids, and arthropods
suggest that early bilaterians had an anteriorly placed brain with a sensory-neurosecretory apical system capa-
ble of orienting the organism in relation to light, gravity, and chemical gradients, followed by several sections
and a single or a pair of nerve cords, forming a mechanosensory griddle around the ventral midline. This

260
Figure 13.2 Proposal for the evolution of neurons and synapses. Choanoflagellates might have had a polarized vesicle
transport system. Initially, there was no chemical signal transduction at soma or filopodial plasma membrane
contact sites (1). In a hypothetical colonial ancestor of metazoans, the apical–basal directed vesicle transport
translocated to soma and/or filopodial plasma membrane contact sites. This resulted in one cell becom-
ing a signal donor (2) and another a signal receiver (3). This relationship was stabilized in an epithelialized
multicellular animal ancestor. From this condition, more stable presynaptic (4) and postsynaptic (5) relation-
ship evolved in different groups of early branching animals. Reproduced with permission from Göhde and
colleagues (2021), originally published by the Royal Society under the terms of the Creative Commons
Attribution License.
Esther Ureña and Marcos Nadal

Figure 13.3 
Mertensia ovum, or Arctic comb jelly, is a species of ctenophore. Visible in this photograph is the apical
organ, at the top right, the rows of swimming ciliary combs running down the body, and the two tentacles.
Photograph by Alexander Smenov, reproduced with permission.

would later evolve into the preoral and postoral chain of ganglia in invertebrates and the brain and spinal cord
in vertebrates (Arendt et al., 2021; Cisek, 2022; Formery et al., 2019; Fritzsch & Glover, 2009; Tosches &
Arendt, 2013). The evolution of this basic CNS in early bilaterians involved determining the disposition and
size of the CNS within the ectoderm during embryonic development and the specification of morphologi-
cally and functionally diverse types of neurons. Both are interdependent processes, in that the developmental
patterning produces migratory pathways that lead specific neuron types to their correct places; dendritic and
axonal growth patterns that connect neurons to each other, creating tracts and nerves; and the regulation of
neural proliferation and survival (Fritzsch & Glover, 2009).
The radical change in the bilaterian body plan and the appearance of the CNS was associated not only
with specific genes that regulate axis specification and the induction of mesoderm and endoderm. It was also
associated with the appearance of genes that orchestrate the development of the CNS and regulate its func-
tion. Among these bilaterian-specific genes are those that code neurotrophin receptors, crucial elements of
the signalling system that regulates aspects of neural development and plasticity, including neuron survival,
synapse formation, and axon guidance; those that code monoamine neurotransmitter receptors for serotonin,
adrenaline, and dopamine; and those that code hormone receptors involved in the regulation of feeding,
homeostasis, and stress (Heger et al., 2020). Bilaterians thus represent not only a new body plan with differ-
ent and complex parts that result from regional patterning and specification (Peterson & Davidson, 2000).
They represent a new way of engaging with the environment, one which entailed a centralized nervous
system linked to physiological regulation of the organism through diverse signalling molecules, such as dopa-
mine or serotonin (Barron et al., 2010; Moroz et al., 2021), that enabled associative learning (Ginsburg &
Jablonka, 2021) and modulated approach and avoidance movements in response to appetitive or threatening
stimuli, exploratory movements, reward seeking, learning, memory, and reproductive and activity cycles,
among others.

262
Evolution of sensory valuation systems

Evolution of the vertebrate brain


The early worm-like bilaterians eventually gave rise to deuterostomes and protostomes (Figure 13.1). At
some point during the evolution of deuterostomes, there was a dorsal–ventral inversion, with the ventral
nerve cord becoming placed dorsally (Butler & Hodos, 2005). The nervous system of the cephalochordate
amphioxus is considered a good approximation of that of the last common ancestor of cephalochordates and
vertebrates (Figure 13.1). Similarities in the patterns of neural connectivity in adults and embryos, of neuro-
chemical circuits, and of gene expression (the genes mediating anteroposterior and dorsoventral patterning
are expressed similarly) of amphioxus and lampreys ( jawless fish) indicate that chordates evolved from an
ancestor with a CNS with differentiated diencephalic forebrain, midbrain, hindbrain, and spinal cord (Fig-
ure 13.4) (Cisek, 2022; Fritzsch & Glover, 2009; Holland, 2017; Pombal & Megías, 2017).
The early vertebrates evolved from an ancestral chordate. They became larger and powerful swimmers.
They had pharyngeal muscles, which allowed them to increase the rate of food ingestion, and vascularized
gills, which facilitated the exchange of gases. These changes came with innovations in the sense organs,
including a pair of image-forming eyes, a complex vestibular system, mechanosensory and electrosensory
lateral lines, taste receptors, and an olfactory system (Striedter & Northcutt, 2020). The CNS of vertebrates
evolved through the progressive elaboration of the chordate basic scheme, with the addition and refine-
ment of systems for sensory processing, motor control, motivational and homeostatic control, memory,
anticipation, and planning (Butler, 2000; Butler & Hodos, 2005; Lacalli, 2022; Schneider, 2014). Forward
locomotion led to innovations in sensory processing and the motor control of orienting towards or fleeing
from objects. The sensory inputs were conveyed from the diencephalon to the rhombencephalon, where
chemosensory processing evolved into a gustatory system that processed not only food selection and rejec-
tion, but also the motivation to search based on stored information about past encounters with suitable or
unsuitable foods, where vestibular processing evolved into a system that kept track of the orientation and
position of body and head, where many reflexes and fixed action patterns, such as fleeing, became organized
and where the control of vital functions, such as breathing, wakefulness, and arousal, became centralized.
These systems linked to centres for sensorimotor integration in the mesencephalon that received inputs from
multiple sensory modalities, coordinated orienting, feeding, escape, aggression, mating, parenting, and other
action patterns triggered by sensory input and exerted motivation control over these behaviours (Cisek,
2022; Schneider, 2014).
This elaboration of rhombencephalon and mesencephalon neural circuits occurred concurrently with the
elaboration of embryological proliferative zones that develop into prosencephalon circuits. The ventral aspect
of the prosencephalon evolved into the striatum and amygdala, the dorsomedial aspect evolved into the
medial pallium and eventually into the hippocampus, the dorsal aspect evolved into the dorsal pallium and

Figure 13.4 Schematic representation of the central nervous system of the vertebrate ancestor with differentiated
diencephalon, mesencephalon, rhombencephalon, and nerve cord. Adapted from Butler (2000; Butler &
Hodos, 2005).

263
Esther Ureña and Marcos Nadal

eventually into the cortex, and the lateral aspect remained the olfactory cortex. This prosencephalic expan-
sion seems to have initially been driven by the adaptive value of elaborating two links between the olfactory
system and systems for motor control, object identification, and spatial memory. One of these links was the
precursor of the ventral striatum. By connecting systems that processed olfactory stimuli, systems for motion
control, and motivational systems in the diencephalon and mesencephalon, the ventral striatum acquired a
fundamental role in disinhibiting approach and avoidance behaviours (Cisek, 2022). As a link between the
olfactory bulbs and the hypothalamus, the ventral striatum modulated endocrine and autonomic nervous
control. As a link to the midbrain, the ventral striatum became involved in the control of orienting move-
ments and approach and avoidance behaviours. A fundamental feature of this early ventral striatum link was
its capacity to establish associations between sensory information and actions and their outcomes by relying
on phasic dopamine releases as reinforcement signals. This extension of the role of dopamine from a tonic
signal of average food intake to a phasic signal of momentary increases or decreases in intake boosted learn-
ing and reinforcing state-action associations (Cisek, 2022). The capacity to learn responses and create habits
conferred by this plasticity probably led to the subsequent projection of other sensory modalities from the
thalamus and subthalamic regions to the striatum. Initially, these new projections would have overlapped
with olfactory projections, but eventually they segregated, leading to the differentiation of the dorsal and
ventral striatum (Schneider, 2014). The second link was the incipient medial pallium. With its projections
to the ventral striatum, hypothalamus, and centres for long-range locomotion, it became specialized in stor-
ing associations between gradients of cues in the environment and the positive or negative consequences of
being in, or approaching, those locations, and enabled places to be remembered as safe and favourable or as
threatening or unfavourable (Cisek, 2022; Manns & Eichenbaum, 2009; Schneider, 2014).
These innovations must have occurred very early on in the history of the vertebrate lineage, as the CNS
of lampreys, the oldest group of living vertebrates that diverged from the lineage that leads to mammals close
to 560 million years ago (Figure 13.1), shares many features with the CNS of jawed vertebrates (Butler, 2000;
Grillner, 2021; Grillner & Robertson, 2017; Suryanarayana et al., 2021). They share, for instance, the same
rostrocaudal organization: telencephalon, diencephalon, mesencephalon, rhombencephalon, and spinal cord.
Lampreys share with other vertebrate lineages the general pattern of connectivity, chemical markers, devel-
opment, and expression of homeobox genes in the basal ganglia (Smeets et al., 2000): in mammals and lam-
preys, the striatum functions as the input structure to the basal ganglia and receives excitatory signals from the
thalamus and pallium/cortex and modulatory signals from the dopamine, serotonin, and histamine systems.
Lampreys and mammals also share general projection patterns from pallium to downstream structures medi-
ated by glutamatergic projection neurons. They also share the topography, connectivity, and histochemistry
of the hippocampus (Bingman et al., 2017). They also have neural nuclei that resemble in development, con-
nectivity, and function, the homologue of the amygdala. The vertebrate amygdala emerged as a neural centre
that emotionally tagged odours by associating them with appetitive or aversive stimuli detected through the
olfactory or vomeronasal systems. Given the volatile nature of odours, this would have conferred early ver-
tebrates a significant advantage in detecting the appetitive or aversive significance of potentially beneficial or
pernicious objects and guiding behaviour accordingly (Martínez-García et al., 2009; Medina et al., 2017).
These similarities in overall brain partitioning, connectivity, topographical specialization, and neural sig-
nalling suggest that some of the fundamental features of the CNS of vertebrates appeared very early on (Grill-
ner, 2021; Grillner & Robertson, 2017; Schneider, 2014; Smith, 2021; Striedter, 2005). The basal ganglia
are a good example of this structural and functional conservation, especially among tetrapods (Marín et al.,
1998; Stephenson-Jones et al., 2011). All tetrapods (Figure 13.1) have a striatum that receives dopaminergic
projections from the posterior diencephalon and the midbrain tegmentum, a nucleus accumbens, a pallidum,
similar pathways connecting these structures, and a direct output pathway related to the release of specific
behavioural patterns and an indirect output pathway related to the inhibition of competing behavioural pat-
terns (Grillner & Robertson, 2016; Reiner, 2009; Smeets et al., 2000; Stephenson-Jones et al., 2011). This is

264
Evolution of sensory valuation systems

not to say that there was little evolution of the brain along the vertebrate lineage. This basic vertebrate CNS
plan was elaborated in different ways and to different extents along the vertebrate lineage, which is especially
clear in the cortex, the amygdala, and the basal ganglia.

Elaboration of the vertebrate dorsal pallium/cortex


The telencephalon of amphibians, which diverged from tetrapods around 350  million years ago (Iyer  &
Briggman, 2021), is relatively simple (Figure  13.5, top). It possesses a medial pallium, which is probably
homologous to the mammalian hippocampus; a lateral pallium, which receives projections from the olfactory
bulb and is probably homologous to the mammalian olfactory cortex; and a dorsal pallium, which receives
projections from the dorsal thalamus and is probably homologous to the mammalian neocortex (Manns &
Eichenbaum, 2009; Striedter, 2005). Nevertheless, the amphibian dorsal pallium differs from the mammalian
cortex in that it has only two layers, its neurons lack basal dendrites, it is not the main target of thalamic
projections (the amphibious thalamus projects mainly to the medial and lateral pallia), its thalamic inputs are
mainly multimodal, it has dominant bidirectional connections with the olfactory bulb, and it does not pro-
ject out of the telencephalon (Butler & Hodos, 2005; Striedter, 2005). The amphibian dorsal pallium is so
unclearly defined that it has been considered more of a transition area between the medial and lateral pallia,
implying that the dorsal pallium appeared with the first amniotes (Striedter & Northcutt, 2020).
Together with birds, reptiles are the living representatives of sauropsids, which appeared about 340 mil-
lion years ago. The telencephalon of reptiles evolved to be larger and more complex than that of amphibian
(Figure 13.5, middle). Their dorsal cortex receives stronger inputs from the dorsal thalamus, lacks recipro-
cal connections with the olfactory bulb, contains neurons that are more like mammalian pyramidal neu-
rons, is organized forming three better-defined layers (one layer of cells between two layers of axons and
dendrites that extend radially and tangentially), and projects to the medial cortex—the homologue of the
hippocampus—and various targets outside of the telencephalon. Unlike mammals, however, the dorsal cor-
tex of reptiles does not project to the spinal cord, it lacks reciprocal connections with the contralateral dorsal
cortex, and it does not receive strong projections from the sensory nuclei of the thalamus (which project
mostly to the ventral pallium in sauropsids) (Striedter, 2005; Striedter & Northcutt, 2020).

Elaboration of the vertebrate amygdala


Like sauropsids and mammals, amphibians possess an amygdaloid formation, with the main pallial and sub-
pallial components involved in orchestrating and regulating visceral and behavioural responses to stimuli and
locations. However, the internal organization of the central and medial extended amygdala of amphibians
is much simpler, which would indicate a less sophisticated and flexible organization of fear and anxiety
responses to threatening stimuli and of responses to sexual and social stimuli. Also, the amphibian amygdala
lacks differentiated cortical and basolateral nuclei, and it is mainly dedicated to processing olfactory and
vomeronasal information. This is related to the fact that the cerebral hemispheres of amphibians lack pallial
visual, auditory, and somatosensory centres: thalamic sensory inputs reach only the striatum (Martínez-
García et al., 2009; Medina et al., 2017). The reptilian and avian amygdala homologues are also divided into
pallial and subpallial components. However, in contrast to amphibians, reptiles and birds have pallial regions
that receive sensory information from the thalamus, and project to the basolateral pallial amygdala. Thus, the
appearance of sensory pallial regions in amniotes—vertebrates that evolved membranes that covered their
eggs, allowing them to survive on land (Figure 13.1)—occurred in tandem with a differentiation within the
pallial amygdala of superficial cortical nuclei that continued to receive direct olfactory inputs and a basolateral
amygdala connected with the pallial areas that processed other sensory modalities (Martínez-García et al.,
2009; Medina et al., 2017). Moreover, the general connectivity pattern of the reptilian and avian amygdala is

265
Esther Ureña and Marcos Nadal

Figure 13.5 Representation of the telencephalon of an amphibian (salamander, top row), reptilian (turtle, middle row),
and mammal (hedgehog, bottom row). On the left is a representation of regional homologies across the
three vertebrate groups. On the right, low-magnification views showing pyramidal or pyramidal-like neu-
rons in black and ascending thalamic inputs in red. Modified from Striedter (2005).

266
Evolution of sensory valuation systems

very similar to that of mammals. This suggests that the ancestral amniote already possessed a pallial amygdala
with lateropallial and centropallial components and a subpallial extended amygdala with medial and central
components that were involved integrating multiple signals, learning, and organizing responses to affectively
and socially relevant stimuli in complex and changing ecological niches and that these aspects have been
conserved throughout evolution (Martínez-García et al., 2009; Medina et al., 2017; Pessoa et al., 2022).

Elaboration of the vertebrate basal ganglia


The elaboration of the different pathways through the basal ganglia varies considerably among groups of
vertebrates (Smeets et  al., 2000). In amphibians, the basal ganglia are cell sparse and receive little teg-
mental dopaminergic input; the striatum receives its main sensory input from sensory nuclei in the dorsal
thalamus, and projects mainly to the pallidum, which projects directly and indirectly to the optic tectum
(Reiner et  al., 1998). In comparison to amphibians, the basal ganglia of amniotes are larger, have more
interneurons (Reiner, 2009) and are enriched with tegmental dopaminergic projections and cortical input,
which gives basal ganglia neurons their reward-dependent plasticity (Hikosaka et al., 2014). The connectiv-
ity scheme of basal ganglia in reptiles resembles that of amphibians, except that the sensory inputs arrive to
the striatum from the dorsal ventricular ridge, suggesting that the striatum receives more highly processed
sensory information in reptiles than in amphibians. In mammals, basal ganglia interneurons are even more
abundant than in reptiles, and the connectivity pattern has also changed: most of the sensory inputs to the
striatum arrive from the neocortex, from where almost all areas project to the striatum, and most of the
outputs project to the dorsal thalamus and from there to the neocortex, suggesting a further elaboration of
the sensory information arriving to the basal ganglia (Reiner et al., 1984, 1998). Thus, the appearance of
amniotes and, subsequently, of mammals was accompanied by increases in the number and complexity of
connections within the basal ganglia and of pallio-/cortico-striatal projections, at each step increasing the
involvement of the cortex, or cortical homologues, in the processing of sensory information conveyed to
the basal ganglia and the detail of the functional cortical map at the striatal level (Reiner, 2009; Schneider,
2014; Smeets et al., 2000). This gradual elaboration of the basal ganglia structure and connectivity enabled
an increase in the flexibility of reward-based learning, the creation of habits based on repeated choices, and
the capacity to direct gaze in anticipation of rewards. Together, these changes increased the possibility of
using different kinds of sensory information to detect stimuli in the environment that motivated approach
and avoidance behaviours and of executing a larger and more sophisticated repertoire of value-based and
experience-dependent behavioural patterns, which were crucial for the adaptation to a fully terrestrial way of
life (Cisek, 2022; Grillner & Robertson, 2016; Hikosaka et al., 2014; MacIver & Finlay, 2022; Reiner et al.,
1998; Stephenson-Jones et al., 2011).

Evolution of the mammalian brain


About 315 million years ago, early amniotes diverged into the synapsid and sauropsid lineages (Figure 13.1).
Sauropsids evolved into the lineages that led to current-day reptiles and birds, and close to 200 or 250 mil-
lion years ago, synapsids led to the early mammals. These had a small body; they were endothermal and
nocturnal, fed on small invertebrates and vertebrates, and cared for their offspring. Being nocturnal, they had
well-developed olfactory and auditory systems but poor vision. Their brains were small, with a differentiated
but modest neocortex (Kaas, 2017; Rowe, 2017). These primitive mammals evolved into a lineage that led
to present-day monotremes (platypuses and echidnas) and the lineage of therian mammals. About 150 mil-
lion years ago, therian mammals branched into the marsupial and placental mammals (Murphy et al., 2004).
The early evolution of the CNS in mammals was deeply influenced by changes in mastication, head move-
ment, ventilation, and sensory systems related to olfaction, dentition, and integument. These changes led to

267
Esther Ureña and Marcos Nadal

some of the particular features of mammals’ CNS, physiology, and behaviour (Kaas, 2017; Schneider, 2014).
Mammals retained the primitive tetrapod orthonasal smell, in which sniffing draws airborne molecules into
the nose, where they activate the olfactory epithelium. Early mammals evolved a flexible cranio-vertebral
articulation, which enabled unique modes of head movement; specialized thoracic and lumbar vertebrae and
ribs, which enabled diaphragmatic ventilation; and a huge olfactory receptor genome. Together, these innova-
tions conferred unique features on mammals’ orthonasal smell, such as the ability to track scents. In addition,
they evolved a secondary palate, which changed the way they chewed and swallowed, with the tongue becom-
ing the major guide of food around the mouth and toward the oesophagus (Rowe & Shepherd, 2016). These
transformations in olfaction, mastication, head movement and ventilation paved the way for retronasal smell,
in which odour molecules produced by chewing are carried by exhaled air from the mouth across the olfac-
tory epithelium. In retronasal smell, olfaction combines with taste, vision, and hearing to produce flavour (see
also Chapter 8 of this volume). Information from the orthonasal and retronasal stimulation of the olfactory
epithelium converges with somatosensory information from the lips, tongue, cheeks, and teeth in the orbito-
frontal cortex. This system evolved as a multimodal map that allowed the integration of different kinds and
combinations of information in new and unique ways (Kaas, 2017; Rowe, 2017; Rowe & Shepherd, 2016).
Improvements in mammalian audition were also related to changes in mastication that led to the struc-
tural and functional uncoupling of chewing and hearing: the jaw bones became specialized in chewing, and
the bones that had previously joined the lower jaw to the skull turned into the chain of middle ear bones
specialized in conveying vibrations. Together with this, the appearance of eardrums, long coiled cochleas,
and hair cells that amplify intracochlear vibrations greatly improved mammals’ ability to detect sounds and
expanded the range of (especially high) frequencies that could be heard. This new capability was exploited
by the appearance of the specifically mammalian auditory cortex (Kaas, 2017; Striedter, 2005), organized
as tonotopic maps, regions that respond to specific temporal patterns, and regions that respond to specific
natural sounds, such as conspecific vocalizations. Processing in the auditory cortex enabled mammals an
unprecedented capacity to identify acoustic patterns informative not only of the location in space of the
source, but also of its identity (Schneider, 2014).
One of the most crucial innovations in mammals was the evolution of a six-layered cortex that received
thalamic input radially (Figure 13.5, bottom) and could process, integrate, and store sensory information
with great detail and organize movements with great precision (Kaas, 2017; Striedter, 2005). Mammalian
cortical neurons are packed together, forming functional units called minicolumns, densely interconnected
neurons aligned in vertical rows. This microcircuitry of the mammalian neocortex is more intricate and
elaborate than reptiles’ and is richer in intracortical connections than in thalamic input (Barsotti et al., 2021;
Kaas, 2017). The cortex is subdivided across its surface, forming specialized areas. The neocortex of early
mammals probably had around 20 cortical areas and included at least a primary visual area and two secondary
visual areas, one auditory area, cingulate, retrosplenial, prefrontal, and perirhinal areas, and a primary soma-
tosensory area, and up to five secondary somatosensory areas (Figure 13.6). The primary motor and premo-
tor areas appeared, by partially taking over the motor functions of the somatosensory areas, only in placental
mammals (Kaas, 2017; Krubitzer, 2007). The addition of cortical primary and secondary sensory and motor
areas enabled a more accurate and detailed perception of the world, a greater precision of movement control,
and a greater ability to anticipate stimuli and plan actions in advance (Nudo & Frost, 2009; Schneider, 2014).
As the neocortex of mammals evolved, it established connections with all levels of the CNS. For instance,
unlike amphibians or reptiles, and unlike the output of other telencephalic components, the neocortex of
mammals sends projections that bypass the midbrain and target the hindbrain and spinal cord (Cisek, 2022;
Nudo  & Frost, 2009; Schneider, 2014). As the neocortex enlarged and specialized into novel functional
areas, its outputs began reaching the striatum, thalamus, hypothalamus, midbrain, and cerebellum, exerting
increasing influence on behaviour and its regulation, enabling, among other things, the transition from the
control of whole-body locomotion to the control of forelimbs for grasping and manipulating objects. The

268
Evolution of sensory valuation systems

Figure 13.6 Phylogenetic tree showing the relations among major groups of mammals and illustrating the extension
of primary and secondary visual, auditory, and somatosensory cortical areas. Reproduced with permission
from Krubitzer (2007).

evolution of the neocortex was coupled with a substantial increase in the connectivity of the circuits that
linked the cortex and basal ganglia, as noted previously, with the cortex becoming increasingly involved
in the processing of the thalamic sensory information projected to the basal ganglia and with the basal
ganglia informing cortical sensory and motor circuits, creating integrative corticostriatal loops that play a
fundamental role in action selection, associative learning, and goal-directed behaviours (Cisek, 2022; Kaas,
2017; Schneider, 2014; Smeets et al., 2000; Striedter & Northcutt, 2020). The neocortex also established
interconnections with the amygdala, and these became critical for the timely and contextually appropriate
engagement of behaviours driven by motivationally and socially relevant goals, such as approach-avoidance

269
Esther Ureña and Marcos Nadal

learning, foraging, defence against predators, and social signalling (Dixon & Dweck, 2021; Murray & Fel-
lows, 2021; Pessoa et al., 2019).
In mammals, the amygdala reached its greatest complexity and structural differentiation. It became a
true multisensory hub that contextualizes sensory information and uses it, together with learnt associations
between stimuli and between stimuli and outcomes, to determine the emotional, motivational, and social
relevance of sensory information and to initiate the neuroendocrine, autonomic and behavioural aspects of
approach and avoidance responses (Martínez-García et al., 2009; Medina et al., 2017; Nieuwenhuys et al.,
2008; Pessoa et al., 2022). It conserves the same two main subdivisions that appeared early in the vertebrate
lineage, which differ in the principal neurotransmitter they rely on, their primary connections, and function.
The cortex-like pallial amygdala of mammals includes the cortical (which processes mostly olfactory infor-
mation) and basolateral (which processes information from different modalities) divisions. The striatum-like
subpallial amygdala includes the central and medial nuclei, which are continuous with the bed nucleus of
the stria terminalis and together form the centromedial extended amygdala. All mammals, including mono-
tremes, marsupials, and placentals, share the same basic subdivisions of the amygdala, although the relative
sizes of these vary among the groups (Medina et  al., 2017). The mammalian amygdala receives sensory
information from the olfactory, vomeronasal, gustatory, viscerosensory, and nocioceptive systems; from the
thalamus; and from different areas of the neocortex. It also receives numerous projections that modulate its
functioning: cholinergic projections from the nucleus basalis-substantia innominate complex, dopaminergic
projections from the ventral tegmental area and substantia nigra, serotoninergic projections from the raphe
complex and locus coeruleus, and projections from several nuclei in the hypothalamus. The amygdala has
three main output pathways that modulate behaviour and physiology. The pathway descending from the
central extended amygdala to centres in the hypothalamus and brainstem is involved in the orchestration
of the motor, vegetative, and endocrine aspects of fear and anxiety. Projections from the medial extended
amygdala and portions of the pallial amygdala to centres in the hypothalamus mediate the behavioural and
neuroendocrine responses to conspecific chemical signals related to reproductive and agonistic behaviours.
Projections from the basolateral amygdala to the ventral striatum are related to the processing of reward, the
generation of positive emotions, appetitive behaviours, reward expectation, and learning through stimulus-
reward associations (Martínez-García et al., 2009).

Evolution of the primate brain


Primates diverged from other orders of placental mammals around 80–90 million years ago. Early primates
were small-bodied and small-brained nocturnal predators that foraged for insects and small vertebrates in the
fine branches of trees and shrubs (Fleagle & Seiffert, 2017; Ho et al., 2021; Kaas et al., 2022; Preuss, 2009;
Ross & Martin, 2009; Sussman, 1991; Wise, 2017). As primates evolved, their brains became larger in rela-
tion to body mass than the brains of other mammals and acquired a number of specializations, the most con-
spicuous of which is their enlarged neocortex (Preuss, 2009; Striedter, 2005). Among the earliest impetus for
this increase in brain and neocortex size were somatosensory and motor innovations related to new ways of
feeding and moving in the terminal branches of angiosperm trees and to an expanded and elaborated visual
system (Cartmill, 1992; DeCasien & Higham, 2019; Ho et al., 2021; Ross & Martin, 2009).
Grasping extremities appeared very early on in primate evolution, linked to the early primate form of
feeding and locomotion, consisting of leaping with the hindlimbs and grasping with the forelimbs. The
development of primate grasping for feeding and locomotion was accompanied by the elaboration of cer-
ebral motor systems. Whereas most mammal species have few premotor areas, primates have nine or more.
They are special in having a ventral premotor area specialized in hand and mouth movements that projects
directly to the spinal cord, conferring more dexterity in hand and arm movements involved in reaching,
grasping, and manipulating food items and bringing them to the mouth. In addition, the evolution of

270
Evolution of sensory valuation systems

touch-sensitive finger and toe tips was accompanied by an enlargement of the somatosensory cortex, increas-
ing the amount of somatosensory information available, especially related to hand haptics and the sensory
control of grasping (Preuss, 2009; Striedter, 2005).
These early primates evolved forward-facing eyes, which increased the scope of stereoscopic vision,
enabling better depth perception (Preuss, 2009; Striedter, 2005; Wise, 2017). The frontal-facing eyes were
later complemented by modifications to the retinal receptors, neural circuitry underlying colour vision, and
changes in retinal projections to the mesencephalon: in most other orders, the projection is almost com-
pletely crossed, but in primates, each retina projects roughly in equal proportions to both superior colliculi.
The lateral geniculate and the superior colliculus became larger and more differentiated, as did the cortical
regions devoted to visual processing (Kaas et al., 2022; Striedter, 2005). In most primates, the visual system
encompasses almost half of the cortex. It includes an enlarged cortical primary visual area V1 and several
extrastraite retinotopically organized visual areas that are not present in other mammals. The projections
from V1 and related visual areas to the parietal and temporal lobes, which had depended more on informa-
tion relayed on information from the superior colliculus in the primate ancestors, became the primary source
for the cortical processing of information. The dominance of cortical visual processing areas led to two
functionally specialized paths: a ventral stream, toward the inferior temporal cortex and related mainly to the
visual identification of objects, and the dorsal stream, toward the posterior parietal cortex, a multimodal area
with major visual inputs devoted mainly to organizing specific eye and hand movements towards objects in
nearby space, such as reaching and grasping, taking hand to mouth, or protecting face or body (Goodale &
Milner, 1992; Goodale & Westwood, 2004; Kaas et al., 2022; Kaas & Stepniewska, 2016; Preuss, 2009).
Some of the distinctive features of primate brains are found in the prefrontal cortex. Primates are unique
among mammals in that they have granular areas in the prefrontal cortex (Preuss, 2009; Preuss & Wise, 2021;
Rudebeck & Izquierdo, 2021; Wise, 2017). The distinction between agranular and granular cortex has to
do with the number and density of cell bodies in layer 4, the internal granular layer: Granular cortical areas
have a conspicuous layer 4, whereas agranular cortical areas have fewer cell bodies located in layer 4. The
appearance of primates was associated with the development of granular areas in the orbitofrontal and caudal
prefrontal cortex, including the frontal eye fields. Together with the new premotor and dorsal visual stream
areas noted previously, these new prefrontal areas improved early primates’ capacity to search and attend to
valuable objects and to compare and update the values of those objects based on their visual features, current
biological needs, and previous choices (Cisek, 2022; Passingham & Wise, 2012; Wise, 2017).
Early primates eventually gave rise to anthropoids, which became larger, diurnal, arboreal quadrupeds that
foraged over larger territories, relying less on olfaction and more on visual cues. In anthropoids, the dorsal
and ventral processing streams were elaborated further, with the posterior parietal cortex representing spatial
references and several different metrics, including relative numerosity, duration, and distance, and the tem-
poral cortex representing the identity of objects based on visual and auditory cues. These innovations were
coupled with the evolution of new granular areas in the dorsolateral, dorsomedial, ventral, and polar prefron-
tal cortex, forming part of a larger system of association areas. These new areas enabled anthropoid primates
to extend their foraging range; to anticipate seasonal changes and to reduce their foraging errors and the risk
of predation by representing abstract goals, learning and using abstract rules, representing events extended
in time, keeping goals and rules active in memory until they are needed, planning ahead, and anticipating
the appetitive or aversive outcomes of their actions (Cisek, 2022; Passingham & Wise, 2012; Preuss & Wise,
2021; Rudebeck & Izquierdo, 2021; Wise, 2017).

Evolution of the human brain


The most striking feature of the human brain is its disproportionately large size in relation to the human
body. As illustrated in Figure 13.7, humans are between three and four times as encephalized as African great

271
Esther Ureña and Marcos Nadal

Figure 13.7 Illustration of the evolutionary relations among hominoid species, with macaque monkeys as outgroup,
with a comparison of adult brain sizes. Reproduced with permission from Sousa and colleagues (2017).

apes (Barton, 2006; Preuss, 2017b; Rilling & Insel, 1999). This increase in brain size is clearly recognizable
in the fossil record with the origin of the genus Homo, about 2.5 million years ago, and especially in Homo
erectus, which appeared about 1.8 million years ago (Brunet et al., 2002; Falk, 2015; Holloway, 2015; Preuss,
2017b). Studies comparing human and nonhuman brains show that the evolution of the human brain did not
involve just an increase in overall size. Most of this increase occurred in the neocortex (Finlay & Darlington,
1995; Rilling, 2006; Rilling & Insel, 1999; Verendeev & Sherwood, 2017) and came along with an extended
and delayed process of brain maturation (Hublin et al., 2015; Somel et al., 2009) and increased developmental
plasticity (Gómez-Robles et al., 2015, 2013). Studies on the evolution of the human brain have focused on
four main issues: the absolute and relative expansion of association areas, new patterns of long-distance corti-
cal connections, changes in the cellular structure of cortical areas, and changes in the expression of numerous
genes related to neural development and metabolism (Preuss, 2011; Verendeev & Sherwood, 2017).

Expansion of association areas


There is evidence of differential expansion of the inferior prefrontal, posterior parietal, and occipital cortex
of Homo erectus by 1.7–1.5  million years ago, which achieved its maximum in our own species between

272
Evolution of sensory valuation systems

Figure 13.8 Distr ibuted association zones are disproportionately expanded in humans. The estimated cortical expansion
in humans compared to macaques is shown on the left and in humans compared to chimpanzees on the
right. The colours indicate the scaling value required to achieve the size in the human brain. The primary
visual area (V1) is displayed in the inset. The figure shows that human cortical motor and primary sensory
areas have expanded very little (blue hues), whereas cortical association areas have expanded substantially
(orange-yellow hues). Reproduced with permission from Buckner and Krienen (2013).

100,000 and 35,000 years ago (Aldridge, 2011; Bruner, 2004; Neubauer et al., 2018; Ponce de León et al.,
2021). Given that cortical areas devoted to primary motor and sensory processing are about the same relative
size in humans and apes, it seems that most of the neocortical expansion in humans involved the association
cortex and secondary sensory cortex in the frontal, parietal, and temporal lobes, as can be seen in Figure 13.8
(Buckner & Krienen, 2013; Mars et al., 2017; Passingham et al., 2017; Preuss, 2017a). This means that the
human neocortex has more resources dedicated to processing nonprimary information than other primates.
For instance, throughout evolution, there were several functional segregations in the human parietal cortex
that enabled a finer-grained analysis of form and motion: whereas the intraparietal sulcus of monkeys has
two shape-sensitive regions, one representation of central vision, and one motion-sensitive area that is not
very sensitive to 3D motion, the intraparietal sulcus of humans has four regions sensitive to shape, three
representations of central space, and four motion-sensitive regions that are sensitive to 3D motion (Orban
et al., 2006). In the course of human evolution, there was a substantial increase in the granular areas of the
prefrontal cortex (Preuss & Wise, 2021). The proportion of grey and white matter volumes in the prefrontal
association cortex is substantially larger in humans than in nonhuman primates. The proportion of prefrontal
grey matter volume is 1.9 times greater in humans than in macaques and 1.2 times greater than in chimpan-
zees. The proportion of prefrontal subcortical white matter volume is 2.4 times greater in humans than in
macaques and 1.7 times greater than in chimpanzees (Donahue et al., 2018). These results show that dur-
ing the evolution of apes, and especially during human evolution, there was a disproportionate increase in
the number of neurons in the frontal association cortex, accompanied by an even greater disproportionate
increase in the connectivity of those neurons (Smaers et al., 2017).

Cortical connectivity
Although the structural connectivity and resting state functional connectivity of the human and nonhuman
primate neocortex are very similar, there appear to be human-specific structural and functional connectivity
patterns linking cortical association areas involved in language, social learning, and tool use (Hecht et al.,
2013; Rilling & Van Den Heuvel, 2018; van den Heuvel et al., 2016). The neocortex of humans is more
modularly structured than the cortex of other primates, such as macaques or chimpanzees, and the human

273
Esther Ureña and Marcos Nadal

temporal, lateral parietal, and inferior frontal multimodal association areas are more profusely interconnected
(Ardesch et al., 2019; Passingham et al., 2017). These results suggest that the evolution of the human brain
entailed a shift toward a greater modularization, which would have facilitated local functional specialization
and probably led to increased asymmetries in cortical area and thickness and an enhanced multimodal con-
nectivity, which would have improved the integration of higher-order multimodal information (Ardesch
et al., 2019; Mars et al., 2017; Xiang et al., 2020). Several studies have documented evolutionary changes
in the white matter tracts that connect the frontal, temporal, and parietal association cortices that awarded a
greater role to the inferior frontal gyrus in social learning, imitation, and tool use (Hecht et al., 2015, 2013).
There is also evidence of a substantial reorganization of the cortical terminations of the arcuate fasciculus,
which connects Broca and Wernicke’s areas, during human evolution. In humans, unlike in chimpanzees and
monkeys, the left hemisphere acuate fasciculus establishes strong connections between the frontal cortex and
a region in the middle and inferior temporal giri. This region is part of the extrastriate cortex in monkeys,
but in humans, it has enlarged substantially and represents word meaning (Preuss, 2017b; Rilling et al., 2008;
Rilling et al., 2012; Sousa et al., 2017).

Microstructure of the neocortex


The increase in connectivity throughout human evolution noted previously is reflected in changes in the
microstructure of the neocortex. For instance, Bianchi and colleagues (2013) found that, throughout the
cortex, the dendritic arbors of human pyramidal neurons are longer, are more branched, and have a higher
density of spines than those of chimpanzees and macaques. Changes in the cytoarchitecture of the neocortex
are clearest in the parietal, temporal, and frontal association areas. For instance, there have been changes in
the arrangements of minicolumns in the planum temporale, a transition area between the auditory associa-
tion cortex and the inferior parietal lobe that forms part of Wernicke’s area. Minicolumns in this region are
larger, contain more neuropil space, and pack more cells in humans than in other primates (Buxhoeveden &
Casanova, 2002; Buxhoeveden et  al., 2001), suggesting that neurons in this region are more extensively
interconnected in humans than in other primates (Buxhoeveden et al., 2001; Sherwood et al., 2017). In the
temporal lobe, the amount of white matter in humans is greater than predicted by primate allometric trends,
suggesting that temporal lobe connectivity patterns have undergone a certain amount of reorganization since
the appearance of the human lineage (Schenker et al., 2005). In the human frontopolar cortex and Broca’s
area, there is more neuropil, and neurons are more spaced than in other primates (Schenker et al., 2008;
Semendeferi et al., 2011), indicating that this region contains a greater proportion of dendrites, axons, syn-
apses, and glial cell processes, all contributing to an enhanced connectivity within the areas and with other
higher-order association areas. These results suggest that throughout human evolution, there was reorgani-
zation in the cytoarchitecture of certain cortical regions that involved a greater within- and between-area
connectivity and a greater proliferation of glial cells to meet the energy costs of the denser dendritic arbors
and increased long-range axons (Sherwood et al., 2017, 2006; Sousa, Meyer et al., 2017).

Gene expression
Recent studies have identified several uniquely human patterns of gene expression in different brain regions.
Such studies have found that most of the differences in gene expression patterns between humans and non-
human primates involve upregulation, leading to an increased expression of genes related to higher levels
of neural activity and metabolism, a pattern that is not observed in other tissues (Cáceres et  al., 2003).
For instance, throughout human evolution, there has been a substantial upregulation of genes related to
aerobic energy metabolism (Uddin et al., 2004) and to synaptogenesis and dendrite outgrowth in neurons
throughout the neocortex (Cáceres et al., 2007), to dopamine biosynthesis and signalling in populations of

274
Evolution of sensory valuation systems

interneurons in the neocortex and striatum (Sousa, Zhu et al., 2017), to synaptic function in neurons in the
prefrontal cortex (Berto et al., 2019), plus an increase in the complexity of gene coexpression networks in
frontopolar neurons (Konopka et al., 2012) and an accelerated evolution of gene expression in prefrontal
oligodendrocytes (Berto et al., 2019). The fact that most of the changes in gene expression in cortical tissue
have to do with facilitating axonal function, promoting synaptic transmission, plasticity, and energy metabo-
lism supports the notion that the evolution of the human brain entailed changes that significantly increased
neural and synaptic activity (Preuss, 2017b; Sousa, Meyer et al., 2017; Verendeev & Sherwood, 2017). It
is becoming increasingly clear that it is not only changes in coding regions of the genome that led to the
evolution of the specific features of the human brain but that gene regulation is also a driving force in brain
evolution (Florio et al., 2017; Tilot et al., 2021; Wilsch-Bräuninger et al., 2016; Won et al., 2019). Many
regulatory human accelerated regions—regions of the genome that show an extremely high rate of muta-
tions in humans and are indicative of fast evolution—are close to genes involved in the development of brain
cells and are, therefore, believed to affect proliferation of neural progenitor cells and their differentiation and
axogenesis (Liu et al., 2021). There is evidence that human-specific gene regulatory relations promoted the
evolution of the human brain by organizing various molecular programs at different stages in development
and in different cell types that led to cortical expansion, structural changes in the frontal cortex, and increases
in the connectivity within and between areas (Tilot et al., 2021; Won et al., 2019).

Evolution of the human amygdala and basal ganglia


Subcortical brain regions have traditionally been regarded as relatively conserved among mammals and
unchanged throughout human evolution and thus less subjected to evolutionary pressures (Lew & Semende-
feri, 2017). However, a more recent understanding of subcortical structures as part of integrated systems that
include cortical-subcortical circuits (Dixon & Dweck, 2021; Pessoa et al., 2022) has sparked an interest in
examining how they differ in humans. It is now known that, although regions such as the amygdala and basal
ganglia have not undergone such spectacular transformations in human evolution as the neocortex, there
have been changes in the size and connectivity of their nuclei that might be relevant to cognitive function
and dysfunction (Hardman et al., 2002; Raghanti et al., 2016; Stephenson et al., 2017).
There is evidence for structural specializations in the human amygdala. The human lateral nucleus is
substantially larger and contains almost 60% more neurons than expected for a human-sized ape brain,
whereas the basal and central nuclei are much smaller and contain fewer neurons than expected (Barger
et al., 2014, 2012, 2007). These patterns differ from what is observed in great apes, whose basal nucleus
contains the greatest number of neurons. The functional implications of these changes are unclear. However,
given the connectivity of the lateral nucleus with the temporal lobe, the increase in size and number of
neurons in the lateral nucleus of the amygdala might reflect the concerted expansion of the human temporal
association areas and the enhancement of evaluating multimodal information projected to the amygdala.
Conversely, given the connectivity of the central nucleus with the brainstem and the olfactory system, its
reduction in size might also be related to the decreasing reliance on olfaction along human evolution (Barger
et al., 2012, 2007).
There is also mounting evidence of structural and functional reorganization of the basal ganglia and their
connections. The human striatum is smaller—though not significantly—than expected for a human-sized
ape brain, suggesting that throughout human evolution, the striatum became somewhat smaller in relative
terms (Barger et  al., 2014). Despite this slight reduction in size, there are reasons to believe that human
evolution brought changes in the input, internal processing, and output of information to the basal ganglia.
Hardman and colleagues (2002) showed that humans have relatively more neurons than other primates in
two internal relay nuclei: the subthalamic nucleus and external globus pallidus. In addition, they found that
humans have relatively fewer dopaminergic substantia nigra neurons than other primates. Raghanti and

275
Esther Ureña and Marcos Nadal

Figure 13.9 Differences between humans and macaques in cortico-striatal resting-state functional connectivity of the
left and right dorsal caudate. The cortical regions with positive values (warm hues) have stronger functional
connectivity with the human dorsal caudate, whereas cortical regions with negative values (cool hues)
have stronger functional connectivity between macaque dorsal caudate and homologous cortical regions.
Reproduced from Liu and colleagues (2021), originally published and distributed under the terms of the
Creative Commons CC-BY license.

colleagues (2016) showed that humans have higher dopaminergic innervation within the medial caudate
nucleus, a convergence region for inputs from the dorsolateral, ventromedial, orbitofrontal, and dorsal ante-
rior cingulate regions of the prefrontal cortex, which might be related to speech production. Balsters and
colleagues (2020) showed that some regions of the human striatum have stronger functional connectivity to
the lateral prefrontal and frontopolar cortex than other mammals and primates. Finally, Liu and colleagues’
(2021) recent resting-state functional connectivity study found that the cortico-striatal connectivity profiles
of different regions of the human caudate and putamen are more complex than in other primates. Further-
more, they found that whereas the cortico-striatal resting-state functional connectivity of the rostral caudate
was similar in humans and other primates, there were differences in the cortico-striatal resting-state func-
tional connectivity of the dorsal caudate, a region involved in the prediction of action-outcome contingency
and learning of complex skills. In macaques, the dorsal caudate is most strongly connected to sensory and
motor regions, whereas in humans, it is most strongly connected to prefrontal regions (Figure 13.9).
Together, these results suggest that although human evolution produced no new structures in the basal
ganglia, there were changes in the intrinsic and extrinsic connectivity, which altered the degree of conver-
gence of prefrontal inputs, the extent to which substantia nigra dopamine regulates their activity, the intri-
cacy of internal processing, and the functional coupling with prefrontal association regions. The functional
implications of these changes are far from clear, but they seem to be related to changes in the prefrontal
association cortex linked to reward prediction, speech, and the acquisition of complex skills (Balsters et al.,
2020; Hardman et al., 2002; X. Liu et al., 2021; Raghanti et al., 2016).

Summary and conclusions


Neuroscience abounds with concepts born from novel observations. For instance, in 1889, Wilhelm His
coined dendrite to refer to the branching sort of processes that he saw emanating from the cellular bodies of
neurons, and in 1897, Charles Sherrington coined synapse to refer to the functional junction between neu-
rons he believed explained the one-way direction of neural transmission and his observation of the localized

276
Evolution of sensory valuation systems

nature of neural degeneration (Finger, 2000; Wickens, 2015). Concepts like dendrite and synapse were intro-
duced to refer to directly observable entities, such as parts of the neuron, or hypothetical entities that could
explain other observations, like the way neural impulses travelled or the way neural damage did not spread as
if throughout a network. But other concepts that are still commonly used in neuroscience, such as emotion,
motivation, and attention, were introduced centuries ago as part of philosophical or psychological discourses
based on deduction and introspection, not empirical evidence. In time, these concepts were accepted as the
names for mental processes and mental states, the constituting elements of the human mind (Danziger, 1997).
However, with the accumulation of psychological and neuroscientific evidence, it has become clear recently
that many of these terms, including motivation, emotion, and cognition, do not map onto the architecture
of the neural systems that make up the brain. It has become equally clear that continuing to treat them as
distinct entities is holding back psychology and neuroscience (Brick et al., 2021; Buzsáki, 2019; Pessoa et al.,
2022, 2019). Concepts with such a long history and an established position within psychology and neuro-
science as attention and emotion are currently being re-evaluated from this perspective (Anderson, 2011;
Barrett & Satpute, 2019; Fiske, 2020; Hommel et al., 2019).
Advances in neuroaesthetics during the last two decades have led to a similar re-evaluation of such fun-
damental concepts as aesthetic emotion, aesthetic pleasure, and aesthetic experience. There is little evidence
for special forms of perception, pleasure, or emotion that would lend support to the idea of a distinct kind of
aesthetic experience, aesthetic mode of processing, or aesthetic sense (Nadal & Skov, 2018; Skov & Nadal,
2018). What the evidence does show is that aesthetic valuation is rooted in neural systems for sensory valu-
ation. These systems are composed of nuclei distributed throughout the brain that work together to assess
the hedonic value of sensory information in the light of internal and external circumstances and of current
and anticipated states (Becker et al., 2019; Berridge & Kringelbach, 2015; Skov, 2019). Humans share with
many other classes of animals the basic constituents of these sensory valuation systems, which explains how
widespread animal preferences are. Like humans, birds, rodents, and nonhuman primates have preferences
for colours, curved contours, ordered and regular arrangements, and biological motion. These preferences
serve crucial functions during development: drawing attention to individually and socially relevant cues in
the environment, eliciting appropriate behaviours, and affording learning experiences (Kovach & Wilson,
1981; Miura & Matsushima, 2016; Nunes et al., 2020). There is no need to explain how and when colour
and form preferences appeared in human evolution, because colour and form preferences existed long before
humans. What we call “aesthetic preferences” and “aesthetic valuation” evolved from the preference sensory
valuation systems possessed by our primate, mammalian, and earlier ancestors. Many details of how this
evolution took place are unknown. But there is enough evidence to sketch the evolution of human sensory
valuation systems as a series of stages whereby innovations were added to pre-existing, often simpler systems.

Stage 1: Sensory valuation in single cells and simple neural networks


Sensory valuation began with complex molecular systems in unicellular organisms. These systems involve
receptors, ion channels, vesicular transporters, signalling molecules, and genetic regulatory circuits that ena-
ble these organisms to detect relevant substances in the environment, to trigger approach and avoidance, to
modify their behaviour depending on experience and internal states, and even to anticipate events. Close to
900 million years ago, colonies of such single-celled organisms evolved into multicellular organisms when
they had become so specialized that they needed each other to live. About 700 million years ago, some of
these specialized cells became able to transduce external stimulation into chemical and electrical signals they
conveyed to adjoining cells. These early neurons connected to light-, chemical-, and gravity-sensitive cells
and to each other, forming networks that provided animals with a targeted and fast system of intercellular
communication that could integrate exterior and interior information and orchestrate feeding and locomo-
tion movements. In the earliest animals with nervous systems, thus, sensory valuation tasks were divided up

277
Esther Ureña and Marcos Nadal

and consigned to specialized cells: some cells detected events and created signals, other cells conveyed these
signals, and others coordinated movements.

Stage 2: Sensory valuation in early bilaterians


About 580 million years ago, the body plan of some of those animals experienced a radical transformation,
giving rise to the anteroposterior axis of bilaterians. This change came with a reorganization of the nerv-
ous system that placed a sensory-neurosecretory system capable of orienting in relation to light, gravity, and
chemical gradients at the front, followed by nerve cords forming a mechanosensory griddle, and a whole
new suite of neural signalling systems. Some of these systems regulated neural development and plasticity,
while others were involved in neurotransmission and homeostatic regulation. With bilaterians, sensory valua-
tion systems became centralized and inextricably linked to physiological regulation through several signalling
pathways that relied on specialized neurotransmitters, such as dopamine and serotonin, and hormones, such
as adrenaline. These pathways enabled the targeted and general modulation of associative learning, approach
and avoidance movements in response to appetitive or aversive stimulation, exploratory movements and
reward seeking in response to homeostatic signals, and reproductive and activity cycles.

Stage 3: Sensory valuation in early vertebrates


The brain of early chordates, which appeared close to 560 million years ago, was segmented into diencephalic
forebrain, midbrain, hindbrain, hindbrain, and spinal cord. The brain of vertebrates, which appeared about
525 million years ago, evolved by elaborating this basic chordate scheme. Vertebrates developed true sensory
systems that processed visual, vestibular, mechanical, and electric stimuli and specialized, yet interconnected,
systems devoted to motor, motivational, and homeostatic control. The sensory valuation system of vertebrates
included a differentiated ventral striatum and amygdala. The ventral striatum evolved as a link between the
olfactory system and the motor system and hypothalamus, acquiring a fundamental role in releasing approach
and avoidance movements and regulating the activity of the endocrine and the autonomous nervous systems.
The capacity of the ventral striatum to establish associations between sensory information and actions and
their outcomes by relying on dopamine as reinforcement signals greatly expanded the capacity to learn and
create habits, turning it into a critical element of the sensory valuation system. The amygdala emerged as a
centre specialized in the affective tagging of odours by associating them with appetitive or aversive stimuli,
becoming crucial to the detection of the survival significance of perceived objects and events. In the course
of vertebrate evolution, the basal ganglia started to receive processed sensory information from the pallium
and developed a richer internal connectivity, which increased the flexibility of reward-based learning and the
repertory of value-based and experience-dependent behaviours. At the same time, more nuclei were added
to the amygdala, including some that connected directly to pallial regions that processed sensory information
other than olfaction, which enabled the integration of different kinds of information to organize responses
to affectively and socially relevant stimuli.

Stage 4: Sensory valuation in early mammals


Mammals appeared about 250  million years ago. They evolved new ways of moving their heads and of
breathing, smelling, chewing, and hearing, linked to important changes in their brains. The most striking
innovation was the appearance of the neocortex, which increased the accuracy of perceptual processing and
the precision of movement control while also creating convergence zones, such as the orbitofrontal cortex,
where olfactory, gustatory, and somatosensory information from the mouth came together. As the neocortex
evolved, it established stronger connections with the basal ganglia, creating integrative corticostriatal loops

278
Evolution of sensory valuation systems

involved in action selection, associative learning, and goal-directed behaviours. The neocortex also estab-
lished stronger connections with the amygdala, which became a centre for multisensory integration that
signalled the affective and social relevance of perceived objects and initiated the neuroendocrine, autonomic,
and behavioural components of approach-avoidance responses. The cortical-amygdala circuits were crucial
for the contextually dependent engagement of behaviours directed towards motivationally and socially rel-
evant goals, such as approach-avoidance learning, foraging, defence, and social signalling.

Stage 5: Sensory valuation in primates


Primates diverged from other mammals close to 85 million years ago. Their nocturnal and terminal branch-
dwelling lifestyle was made possible by changes in the motor, somatosensory, and visual cortex. As primates
evolved, their prefrontal cortex increased in size, enabling them to assess, compare, and update the value
of objects based on their attributes, current needs, and past experience and eventually to represent abstract
goals, learn and use abstract rules, and anticipate and plan into the future. In primates, the links between
the neocortex and the basal ganglia and the amygdala became stronger and, because of the abstract nature
of the information projected from association areas, these sensory valuation systems became uncoupled
from the here and now. They became able to assess and compare the reward value and affective and social
significance of objects that were anticipated, events that occurred in different contexts, and actions that
were planned.

Stage 6: Sensory valuation in humans


When humans appeared, about 7 million years ago, their brains were not much larger than chimpanzees’.
But about 2 million years ago, the size of the human brain began to grow disproportionally in relation to
the human body. This was due to the unprecedented increase in extension, specialization, and within and
between connectivity of neocortical association regions, which led to improvements in representation of
abstract goals and rules, planning, executive functions, social learning, imitation, tool use, and language.
The evolution of the neocortex occurred in tandem with the evolution of the structure and connectivity
of the amygdala and basal ganglia. The human amygdala became more involved in processing multimodal
information at the expense of olfactory information. The human basal ganglia became more elaborate in
their intrinsic and extrinsic connectivity and functionally more coupled to areas in the prefrontal cortex and
less coupled to sensory and motor areas. These changes could have led to the representation, assessment,
pleasurable experience, and anticipation of highly abstract rewards; improvements in reward prediction; and
the acquisition of language and complex skills.

Aesthetic valuation in the light of evolution


Neuroaesthetics has shown that aesthetic valuation arises from systems in our brain that assess the hedonic
value of sensory stimuli. The fact that other animals show similar preferences for perceptual attributes that
run on comparable systems suggests a common evolutionary history. This chapter has sketched what is
known about the origin and evolution of these sensory valuation systems. They originated in the com-
plex molecular interactions that enabled unicellular organisms to use internal and external information to
execute approach and avoidance behaviours and even to learn from their consequences. When multicellular
organisms appeared, sensory valuation tasks were divided and assigned to specialized cells. With the evolu-
tion of the CNS came a new way of regulating the internal milieu, one that relied on neurotransmitters
and hormones that modulated approach-avoidance behaviours depending on internal and external states
and allowed greater speed and flexibility of learning. In vertebrates, sensory valuation systems grew and

279
Esther Ureña and Marcos Nadal

diversified substantially, becoming crucial associative relays between sensation and behaviour, responsible
for learning and anticipating the consequences of actions, for signalling the affective and social relevance of
stimuli, and for setting in motion appropriate responses. In mammals, and later in primates, sensory valuation
systems became intertwined with the neocortex, most crucially, and with association areas and began receiv-
ing more abstract information, and they became more elaborate and specialized, capable of subtle regulation
of learning and of contextually dependent organization of approach-avoidance behaviours. The human sen-
sory valuation system evolved upon these foundations: it continued to diversify, specialize, and interact with
cortical association areas, creating abstract rewards and pleasures.
We like things and find them beautiful because of how our sensory valuation systems became wired
throughout evolution. We have colour, form, rhythmic, and melodic preferences because those systems
evolved to assess the hedonic value of certain perceptual attributes that signalled important objects in the
environment. We are pleased by liking and beauty because those systems evolved to use pleasure as an incen-
tive to approach things that were life promoting. Beauty can be an emotional experience because those sys-
tems evolved to mobilize the endocrine and autonomous nervous systems for action. We tend to like music
that combines a certain degree of predictability and surprise because the systems that create that enjoyment
evolved to anticipate events based on regular patterns. Music can bring back memories of events and places,
because those systems evolved to link affectively valenced objects and events to places. Familiarity, exposure,
and expertise influence what we like and find beautiful, because those systems evolved to depend on learning
and experience. We are able to enjoy the elegance of mathematical formulas because our sensory valuation
systems can assess the hedonic value of abstract concepts that are relayed from cortical association areas that
represent those formulas. Paraphrasing Dobzhansky (1973), everything in aesthetics makes sense in the light
of evolution.
The brain systems that allow us to feel pleasure and displeasure, and to like and dislike, did not evolve
because of the adaptive value of enjoying or creating opera, sculpture, dance, or fine calligraphy, just as
noses and ears did not evolve to support eyeglasses. Quite the other way around: the shape of eyeglasses was
designed to fit around ears and noses, just as cultural artefacts, such as music, dance, buildings, and gardens,
were designed to fit around the biological systems that generate pleasure and displeasure, liking and dislik-
ing. These systems evolved because of the selective advantage conferred by existing capacities to detect and
approach life-favouring substances, objects, and places; to detect and withdraw from life-threatening sub-
stances, objects, and places; and to modify behaviour based on such experiences. Thus, the attributes of the
biological systems underlying aesthetic liking, and of the cultural artefacts humans have invented to tickle
these systems, were shaped by their primordial function of organizing animal behaviour in adaptive ways:
approaching what’s good, avoiding what’s bad, and learning from this.

References
Aldridge, K. (2011). Patterns of differences in brain morphology in humans as compared to extant apes. Journal of Human
Evolution, 60(1), 94–105. https://doi.org/10.1016/j.jhevol.2010.09.007
Anderson, B. (2011, September). There is no such thing as attention. Frontiers in Psychology, 2, 246. https://doi.
org/10.3389/FPSYG.2011.00246/BIBTEX
Anderson, J. R., Kuwahata, H., Kuroshima, H., Leighty, K. A., & Fujita, K. (2005). Are monkeys aesthetists? Rensch
(1957) revisited. Journal of Experimental Psychology: Animal Behavior Processes, 31(1), 71–78. https://doi.org/10.1037/
0097-7403.31.1.71
Ardesch, D. J., Scholtens, L. H., Li, L., Preuss, T. M., Rilling, J. K., & van den Heuvel, M. P. (2019). Evolutionary expan-
sion of connectivity between multimodal association areas in the human brain compared with chimpanzees. Proceed-
ings of the National Academy of Sciences of the United States of America, 116(14), 7101–7106. https://doi.org/10.1073/
PNAS.1818512116/-/DCSUPPLEMENTAL
Arendt, D. (2021). Elementary nervous systems. Philosophical Transactions of the Royal Society B: Biological Sciences,
376(1821), 20200347. https://doi.org/10.1098/RSTB.2020.0347

280
Evolution of sensory valuation systems

Arendt, D., Urzainqui, I. Q., & Vergara, H. M. (2021). The conserved core of the nereid brain: Circular CNS, apical
nervous system and lhx6-arx-dlx neurons. Current Opinion in Neurobiology, 71, 178–187. https://doi.org/10.1016/j.
conb.2021.11.008
Ayala, F. J. (2017). Adaptive significance of ethics and aesthetics. In M. Tibayrenc & F. J. Ayala (Eds.), On human nature:
Biology, psychology, ethics, politics, and religion (pp. 601–623). Academic Press.
Bain, A. (1883). Mental and moral science (3rd ed.). Longmans, Green, and Co.
Balsters, J. H., Zerbi, V., Sallet, J., Wenderoth, N., & Mars, R. B. (2020). Primate homologs of mouse cortico-striatal
circuits. eLife, 9. https://doi.org/10.7554/eLife.53680
Bar, M.,  & Neta, M. (2006). Humans prefer curved visual objects. Psychological Science, 17(8), 645–648. https://doi.
org/10.1111/j.1467-9280.2006.01759.x
Bar, M., & Neta, M. (2007). Visual elements of subjective preference modulate amygdala activation. Neuropsychologia,
45(10), 2191–2200. https://doi.org/10.1016/j.neuropsychologia.2007.03.008
Bardi, L., Regolin, L., & Simion, F. (2011). Biological motion preference in humans at birth: Role of dynamic and con-
figural properties. Developmental Science, 14(2), 353–359. https://doi.org/10.1111/J.1467-7687.2010.00985.X
Barger, N., Hanson, K. L., Teffer, K., Schenker-Ahmed, N. M., & Semendeferi, K. (2014, May). Evidence for evolu-
tionary specialization in human limbic structures. Frontiers in Human Neuroscience, 8, 277. https://doi.org/10.3389/
fnhum.2014.00277
Barger, N., Stefanacci, L., Schumann, C. M., Sherwood, C. C., Annese, J., Allman, J. M., Buckwalter, J. A., Hof, P.
R.,  & Semendeferi, K. (2012). Neuronal populations in the basolateral nuclei of the amygdala are differentially
increased in humans compared with apes: A stereological study. Journal of Comparative Neurology, 520(13), 3035–3054.
https://doi.org/10.1002/CNE.23118
Barger, N., Stefanacci, L., & Semendeferi, K. (2007). A comparative volumetric analysis of the amygdaloid complex and
basolateral division in the human and ape brain. American Journal of Physical Anthropology, 134(3), 392–403. https://
doi.org/10.1002/AJPA.20684
Barrett, L. F., & Satpute, A. B. (2019). Historical pitfalls and new directions in the neuroscience of emotion. Neuroscience
Letters, 693, 9–18. https://doi.org/10.1016/J.NEULET.2017.07.045
Barron, A. B., Søvik, E., & Cornish, J. L. (2010, October). The roles of dopamine and related compounds in reward-
seeking behavior across animal phyla. Frontiers in Behavioral Neuroscience, 4, 163. https://doi.org/10.3389/FNBEH.
2010.00163/BIBTEX
Barsotti, E., Correia, A., & Cardona, A. (2021). Neural architectures in the light of comparative connectomics. Current
Opinion in Neurobiology, 71, 139–149. https://doi.org/10.1016/j.conb.2021.10.006
Barton, R. A. (2006). Primate brain evolution: Integrating comparative, neurophysiological, and ethological data. Evolu-
tionary Anthropology, 15(6), 224–236. https://doi.org/10.1002/evan.20105
Bartra, O., McGuire, J. T., & Kable, J. W. (2013). The valuation system: A coordinate-based meta-analysis of BOLD fMRI
experiments examining neural correlates of subjective value. NeuroImage, 76, 412–427. https://doi.org/10.1016/j.
neuroimage.2013.02.063
Becker, S., Bräscher, A. K., Bannister, S., Bensafi, M., Calma-Birling, D., Chan, R. C. K., Eerola, T., Ellingsen, D. M.,
Ferdenzi, C., Hanson, J. L., Joffily, M., Lidhar, N. K., Lowe, L. J., Martin, L. J., Musser, E. D., Noll-Hussong, M.,
Olino, T. M., Pintos Lobo, R., & Wang, Y. (2019). The role of hedonics in the human affectome. Neuroscience and
Biobehavioral Reviews, 102, 221–241. https://doi.org/10.1016/J.NEUBIOREV.2019.05.003
Berlyne, D. E. (1972). Ends and means of experimental aesthetics. Canadian Journal of Psychology, 26(4), (303–325).
https://doi.org/10.1037/h0082439
Berridge, K. C.,  & Kringelbach, M. L. (2015). Pleasure systems in the brain. Neuron, 86(3), 646–664. https://doi.
org/10.1016/j.neuron.2015.02.018
Bertamini, M., Palumbo, L., Gheorghes, T. N., & Galatsidas, M. (2016). Do observers like curvature or do they dislike
angularity? British Journal of Psychology, 107(1), 154–178. https://doi.org/10.1111/bjop.12132
Berto, S., Mendizabal, I., Usui, N., Toriumi, K., Chatterjee, P.,  & Douglas, C., Tamminga, C. A., Preuss, T. M.,
Yi, S. V.,  & Konopka, G. (2019). Accelerated evolution of oligodendrocytes in the human brain. Proceedings of
the National Academy of Sciences of the United States of America, 116(48), 24334–24342. https://doi.org/10.1073/
PNAS.1907982116/-/DCSUPPLEMENTAL
Bianchi, S., Stimpson, C. D., Bauernfeind, A. L., Schapiro, S. J., Baze, W. B., McArthur, M. J., Bronson, E., Hopkins,
W. D., Semendeferi, K., Jacobs, B., Hof, P. R., & Sherwood, C. C. (2013). Dendritic morphology of pyramidal
neurons in the chimpanzee neocortex: Regional specializations and comparison to humans. Cerebral Cortex, 23(10),
2429–2436. https://doi.org/10.1093/cercor/bhs239
Bingman, V. P., Rodríguez, F., & Salas, C. (2017). The hippocampus of nonmammalian vertebrates. In J. H. Kaas (Ed.),
Evolution of nervous systems (Vol. 1, 2nd ed., p. 479489). Academic Press.

281
Esther Ureña and Marcos Nadal

Bode, C., Helmy, M., & Bertamini, M. (2017). A cross-cultural comparison for preference for symmetry: Comparing
British and Egyptian nonexperts. Psihologija, 50(3), 383–402. https://doi.org/10.2298/PSI1703383B
Bray, D. (1995). Protein molecules as computational elements in living cells. Nature, 376(6538), 307–312. https://doi.
org/10.1038/376307a0
Brick, C., Hood, B., Ekroll, V., & de-Wit, L. (2021). Illusory essences: A bias holding back theorizing in psychological
science. Perspectives on Psychological Science. https://doi.org/10.1177/1745691621991838
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. NeuroImage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Bruner, E. (2004). Geometric morphometrics and paleoneurology: Brain shape evolution in the genus Homo. Journal of
Human Evolution, 47(5), 279–303. https://doi.org/10.1016/J.JHEVOL.2004.03.009
Brunet, M., Guy, F., Pilbeam, D., Mackaye, H. T., Likius, A., Ahounta, D., Beauvilain, A., Blondel, C., Bocherens, H.,
Boisserie, J. R., De Bonis, L., Coppens, Y., Dejax, J., Denys, C., Duringer, P., Eisenmann, V., Fanone, G., Fronty, P.,
Geraads, D.,.   .  . Zollikofer, C. (2002). A new hominid from the upper miocene of Chad, Central Africa. Nature,
418(6894), 145–151. http://doi.org/10.1038/nature00879
Buckner, R. L., & Krienen, F. M. (2013). The evolution of distributed association networks in the human brain. Trends
in Cognitive Sciences, 17(12), 648–665. https://doi.org/10.1016/j.tics.2013.09.017
Burkhardt, P., & Jékely, G. (2021). Evolution of synapses and neurotransmitter systems: The divide-and-conquer model
for early neural cell-type evolution. Current Opinion in Neurobiology, 71, 127–138. https://doi.org/10.1016/j.
conb.2021.11.002
Butler, A. B. (2000). Chordate evolution and the origin of craniates: An old brain in a new head. Anatomical Record,
261(3), 111–125. https://doi.org/10.1002/1097-0185(20000615)261:3<111::AID-AR6>3.0.CO;2-F
Butler, A. B., & Hodos, W. (2005). Comparative vertebrate neuroanatomy. Evolution and adaptation (2nd ed.). John Wiley &
Sons Inc.
Buxhoeveden, D. P., & Casanova, M. F. (2002). The minicolumn and evolution of the brain. Brain, Behavior and Evolution,
60(3), 125–151. https://doi.org/10.1159/000065935
Buxhoeveden, D. P., Switala, A. E., Litaker, M., Roy, E., & Casanova, M. F. (2001). Lateralization of minicolumns in
human planum temporale is absent in nonhuman primate cortex. Brain, Behavior and Evolution, 57(6), 349–358.
https://doi.org/10.1159/000047253
Buzsáki, G. (2019). The brain from inside out. Oxford University Press.
Cáceres, M., Lachuer, J., Zapala, M. A., Redmond, J. C., Kudo, L., Geschwind, D. H., Lockhart, D. J., Preuss, T. M., &
Barlow, C. (2003). Elevated gene expression levels distinguish human from non-human primate brains. Proceedings of
the National Academy of Sciences, 100(22), 13030–13035. https://doi.org/10.1073/PNAS.2135499100
Cáceres, M., Suwyn, C., Maddox, M., Thomas, J. W., & Preuss, T. M. (2007). Increased cortical expression of two syn-
aptogenic thrombospondins in human brain evolution. Cerebral Cortex, 17(10), 2312–2321. https://doi.org/10.1093/
cercor/bhl140
Cartmill, M. (1992). New views on primate origins. Evolutionary Anthropology: Issues, News, and Reviews, 1(3), 105–111.
https://doi.org/10.1002/EVAN.1360010308
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology,
29(23), 4084–4092.e4. https://doi.org/10.1016/j.cub.2019.09.067
Christensen, J. F. (2017). Pleasure junkies all around! Why it matters and why “the arts” might be the answer: A biopsy-
chological perspective. Proceedings of the Royal Society B: Biological Sciences, 284(1854), 20162837. https://doi.
org/10.1098/rspb.2016.2837
Chuquichambi, E. G., Palumbo, L., Rey, C., & Munar, E. (2021). Shape familiarity modulates preference for curvature
in drawings of common-use objects. PeerJ, 7, e11772. https://doi.org/10.7717/PEERJ.11772/SUPP-3
Cisek, P. (2022). Evolution of behavioural control from chordates to primates. Philosophical Transactions of the Royal Society
B, 377(1844), 20200522. https://doi.org/10.1098/RSTB.2020.0522
Corradi, G., Rosselló-Mir, J., Vañó, J., Chuquichambi, E., Bertamini, M., & Munar, E. (2019). The effects of presen-
tation time on preference for curvature of real objects and meaningless novel patterns. British Journal of Psychology,
110(4), 670–685. https://doi.org/10.1111/bjop.12367
Coyle, S. M., Flaum, E. M., Li, H., Krishnamurthy, D., & Prakash, M. (2019). Coupled active systems encode an emer-
gent hunting behavior in the unicellular predator Lacrymaria olor. Current Biology, 29(22), 3838–3850.e3. https://doi.
org/10.1016/J.CUB.2019.09.034
Danziger, K. (1997). Naming the mind. Sage Publications.
DeCasien, A. R., & Higham, J. P. (2019). Primate mosaic brain evolution reflects selection on sensory and cognitive
specialization. Nature Ecology and Evolution, 3(10), 1483–1493. https://doi.org/10.1038/s41559-019-0969-0

282
Evolution of sensory valuation systems

Delius, J. D., & Thompson, G. (1970). Brightness dependence of colour preferences in herring gull chicks. Zeitschrift Für
Tierpsychologie, 27(7), 842–849. https://doi.org/10.1111/j.1439-0310.1970.tb01905.x
Dexter, J. P., Prabakaran, S., & Gunawardena, J. (2019). A complex hierarchy of avoidance behaviors in a single-cell
eukaryote. Current Biology, 29(24), 4323–4329.e2. https://doi.org/10.1016/J.CUB.2019.10.059
Dixon, M. L.,  & Dweck, C. S. (2021). The amygdala and the prefrontal cortex: The co-construction of intelligent
decision-making. Psychological Review. https://doi.org/10.1037/REV0000339
Dobzhansky, T. (1962). Mankind evolving. The evolution of the human species. Yale University Press.
Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution. American Biology Teacher, 35(3),
125–129. https://doi.org/10.2307/4444260
Donahue, C. J., Glasser, M. F., Preuss, T. M., Rilling, J. K., & Van Essen, D. C. (2018). Quantitative assessment of pre-
frontal cortex in humans relative to nonhuman primates. Proceedings of the National Academy of Sciences of the United
States of America, 115(22), E5183–E5192. https://doi.org/10.1073/PNAS.1721653115/-/DCSUPPLEMENTAL
Dussutour, A. (2021). Learning in single cell organisms. Biochemical and Biophysical Research Communications, 564, 92–102.
https://doi.org/10.1016/J.BBRC.2021.02.018
Ebel, S. J., Kopp, K. S., & Liebal, K. (2020). Object preferences in captive Sumatran orang-utans (Pongo abelii). Behavioural
Processes, 170, 103993. https://doi.org/10.1016/J.BEPROC.2019.103993
Eysenck, H. J. (1941). A  critical and experimental study of colour preferences. American Journal of Psychology, 54(3),
385–394. https://doi.org/10.2307/1417683
Falk, D. (2015). Evolution of the primate brain. In W. Henke  & I. Tattersall (Eds.), Handbook of paleoanthropology 2
(pp. 1495–1525). Springer-Verlag.
Fantz, R. L. (1957). Form preferences in newly hatched chicks. Journal of Comparative and Physiological Psychology, 50(5),
422–430. https://doi.org/10.1037/h0044973
Fantz, R. L., & Miranda, S. B. (1975). Newborn-infant attention to form of contour. Child Development, 46, 224–228.
https://doi.org/10.2307/1128853
Feldman, E. B. (1971). Varieties of visual experience: Art as image and idea (2nd ed.). Prentice Hall/Harry N. Abrams.
Finger, S. (2000). Minds behind the brain: A history of the pioneers and their discoveries. Oxford University Press.
Fingerhut, J., & Prinz, J. J. (2020). Aesthetic emotions reconsidered. Monist, 103(2), 223–239. https://doi.org/10.1093/
monist/onz037
Finlay, B. L., & Darlington, R. B. (1995). Linked regularities in the development and evolution of mammalian brains.
Science, 268(5217), 1578–1584. https://doi.org/10.1126/SCIENCE.7777856
Fischer, G. J., Davis, S. J., & Nord, J. A. (1975). Prehatch color stimulation effects on color pecking preferences and
color discrimination learning in white leghorn chicks. Developmental Psychobiology, 8(6), 525–531. https://doi.
org/10.1002/dev.420080609
Fischer, G. J., Morris, G. L., & Ruhsam, J. P. (1975). Color pecking preferences in white leghorn chickens. Journal of
Comparative and Physiological Psychology, 88(1), 402–406. https://doi.org/10.1037/h0076227
Fiske, A. P. (2020). The lexical fallacy in emotion research: Mistaking vernacular words for psychological entities. Psycho-
logical Review, 127(1), 95–113. https://doi.org/10.1037/REV0000174
Fitch, W. T. (2021). Information and the single cell. Current Opinion in Neurobiology, 71, 150–157. https://doi.org/10.1016/
J.CONB.2021.10.004
Fleagle, J. G.,  & Seiffert, E. R. (2017). The phylogeny of primates. In Evolution of nervous systems (Vol. 3, 2nd ed.,
pp. 1–34). Academic Press.
Florio, M., Borrell, V., & Huttner, W. B. (2017). Human-specific genomic signatures of neocortical expansion. Current
Opinion in Neurobiology, 42, 33–44. https://doi.org/10.1016/J.CONB.2016.11.004
Formery, L., Schubert, M., & Croce, J. C. (2019). Ambulacrarians and the ancestry of deuterostome nervous systems. In
W. Tworzydlo & S. Bilinski (Eds.), Evo-devo: Non-model species in cell and developmental biology (pp. 31–59). Springer.
https://doi.org/10.1007/978-3-030-23459-1_3.
Fritzsch, B., & Glover, J. C. (2009). Structure of brains of primitive vertebrates (tunicates, amphioxus, lampreys) and
the basic features of the vertebrate brain. In J. H. Kaas (Ed.), Evolutionary neuroscience (pp. 123–146). Academic
Press.
Ginsburg, S., & Jablonka, E. (2021). Evolutionary transitions in learning and cognition. Philosophical Transactions of the
Royal Society B, 376(1821), 20190766. https://doi.org/10.1098/RSTB.2019.0766
Göhde, R., Naumann, B., Laundon, D., Imig, C., McDonald, K., Cooper, B. H., Varoqueaux, F., Fasshauer, D., & Bur-
khardt, P. (2021). Choanoflagellates and the ancestry of neurosecretory vesicles. Philosophical Transactions of the Royal
Society B, 376(1821), 20190759. https://doi.org/10.1098/RSTB.2019.0759
Gollwitzer, A., & Clark, M. S. (2019). Anxious attachment as an antecedent of people’s aversion towards pattern devi-
ancy. European Journal of Social Psychology, 49(6), 1206–1222. https://doi.org/10.1002/ejsp.2565

283
Esther Ureña and Marcos Nadal

Gollwitzer, A., Marshall, J., & Bargh, J. A. (2020). Pattern deviancy aversion predicts prejudice via a dislike of statistical
minorities. Journal of Experimental Psychology: General, 149(5), 828–854. https://doi.org/10.1037/xge0000682
Gómez-Puerto, G., Munar, E.,  & Nadal, M. (2016, January). Preference for curvature: A  historical and conceptual
framework. Frontiers in Human Neuroscience, 9, 712. https://doi.org/10.3389/fnhum.2015.00712
Gómez-Robles, A., Hopkins, W. D., Schapiro, S. J.,  & Sherwood, C. C. (2015). Relaxed genetic control of corti-
cal organization in human brains compared with chimpanzees. Proceedings of the National Academy of Sciences USA,
112(48), 14799–14804. https://doi.org/10.1073/pnas.1512646112
Gómez-Robles, A., Hopkins, W. D., & Sherwood, C. C. (2013). Increased morphological asymmetry, evolvability and
plasticity in human brain evolution. Proceedings of the Royal Society B: Biological Sciences, 280(1761), 20130575. https://
doi.org/10.1098/rspb.2013.0575
Goodale, M. A., & Milner, A. D. (1992). Separate visual pathways for perception and action. Trends in Neurosciences, 15(1),
20–25. https://doi.org/10.1016/0166-2236(92)90344-8
Goodale, M. A.,  & Westwood, D. A. (2004). An evolving view of duplex vision: Separate but interacting cortical
pathways for perception and action. Current Opinion in Neurobiology, 14(2), 203–211. https://doi.org/10.1016/j.
conb.2004.03.002
Goodwin, E. B., & Hess, E. H. (1969). Innate visual form preferences in the pecking behavior of young chicks. Behaviour,
34(4), 223–237. https://doi.org/10.1163/156853969x00134
Green, K. F., Moore, J. W., & Sargent, T. D. (1966). Color preference in squirrel monkeys (Saimiri sciureus). Psychonomic
Science, 4(11), 367–368. https://doi.org/10.3758/BF03342342
Grillner, S. (2021). Evolution of the vertebrate motor system—From forebrain to spinal cord. Current Opinion in Neuro-
biology, 71, 11–18. https://doi.org/10.1016/j.conb.2021.07.016
Grillner, S., & Robertson, B. (2016). The basal ganglia over 500 million years. Current Biology, 26(20), R1088–R1100.
https://doi.org/10.1016/J.CUB.2016.06.041
Grillner, S., & Robertson, B. (2017). Conserved features of the basal ganglia and related forebrain circuits—From lam-
preys to mammals. In J. H. Kaas (Ed.), Evolution of nervous systems (Vol. 1, pp. 321–326). Academic Press.
Haber, S. N., & Knutson, B. (2010). The reward circuit: Linking primate anatomy and human imaging. Neuropsychophar-
macology, 35(1), 4–26. https://doi.org/10.1038/npp.2009.129
Hardman, C. D., Henderson, J. M., Finkelstein, D. I., Horne, M. K., Paxinos, G., & Halliday, G. M. (2002). Compari-
son of the basal ganglia in rats, marmosets, macaques, baboons, and humans: Volume and neuronal number for the
output, internal relay, and striatal modulating nuclei. Journal of Comparative Neurology, 445(3), 238–255. https://doi.
org/10.1002/CNE.10165
Hecht, E. E., Gutman, D. A., Bradley, B. A., Preuss, T. M., & Stout, D. (2015). Virtual dissection and comparative con-
nectivity of the superior longitudinal fasciculus in chimpanzees and humans. NeuroImage, 108, 124–137. https://doi.
org/10.1016/J.NEUROIMAGE.2014.12.039
Hecht, E. E., Gutman, D. A., Preuss, T. M., Sanchez, M. M., Parr, L. A., & Rilling, J. K. (2013). Process versus prod-
uct in social learning: Comparative diffusion tensor imaging of neural systems for action execution—Observation
matching in macaques, chimpanzees, and humans. Cerebral Cortex, 23(5), 1014–1024. https://doi.org/10.1093/
CERCOR/BHS097
Heger, P., Zheng, W., Rottmann, A., Panfilio, K. A., & Wiehe, T. (2020). The genetic factors of bilaterian evolution.
eLife, 9, 1–45. https://doi.org/10.7554/eLife.45530
Hess, E. H. (1956). Natural preferences of chicks and ducklings for objects of different colors. Psychological Reports, 2(3),
477–483. https://doi.org/10.2466/pr0.1956.2.3.477
Hikosaka, O., Kim, H. F., Yasuda, M., & Yamamoto, S. (2014). Basal ganglia circuits for reward value—Guided behavior.
Annual Review of Neuroscience, 37, 289–306. https://doi.org/10.1146/ANNUREV-NEURO-071013-013924
Ho, C. L. A., Fichtel, C., & Huber, D. (2021). The gray mouse lemur (Microcebus murinus) as a model for early primate
brain evolution. Current Opinion in Neurobiology, 71, 92–99. https://doi.org/10.1016/j.conb.2021.09.012
Holland, L. Z. (2017). Invertebrate origins of vertebrate nervous systems. In J. H. Kaas (Ed.), Evolution of nervous systems
(Vol. 1, 2nd ed., pp. 3–23). Academic Press.
Holloway, R. L. (2015). The evolution of the hominid brain. In W. Henke & I. Tattersall (Eds.), Handbook of paleoanthro-
pology (pp. 1961–1988). Springer-Verlag.
Hommel, B., Chapman, C. S., Cisek, P., Neyedli, H. F., Song, J. H., & Welsh, T. N. (2019). No one knows what atten-
tion is. Attention, Perception and Psychophysics, 81(7), 2288–2303. https://doi.org/10.3758/s13414-019-01846-w
Hublin, J. J., Neubauer, S., & Gunz, P. (2015). Brain ontogeny and life history in Pleistocene hominins. Philosophical
Transactions of the Royal Society B, 370(1663), 20140062. https://doi.org/10.1098/rstb.2014.0062
Humphrey, N. K. (1971). Colour and brightness preferences in monkeys. Nature, 229(5287), 615–617. https://doi.
org/10.1038/229615a0

284
Evolution of sensory valuation systems

Humphrey, N. K. (1972). “Interest” and “pleasure”: Two determinants of a mnonkey’s visual preferences. Perception, 1(4),
395–416. https://doi.org/10.1068/p010395
Iyer, A. A., & Briggman, K. L. (2021). Amphibian behavioral diversity offers insights into evolutionary neurobiology.
Current Opinion in Neurobiology, 71, 19–28. https://doi.org/10.1016/j.conb.2021.07.015
Jadva, V., Hines, M., & Golombok, S. (2010). Infants’ preferences for toys, colors and shapes: Sex differences and similari-
ties. Archives of Sexual Behavior, 39, 1261–1273. https://doi.org/10.1007/s10508-010-9618-z
James, W. (1884). What is an emotion? Mind, 9, 188–205.
James, W. (1890). The principles of psychology. Henry Holt.
Kaas, J. H. (2017). The organization of neocortex in early mammals. In J. H. Kaas (Ed.), Evolution of nervous systems (2nd
ed., pp. 87–102). Academic Press.
Kaas, J. H., Qi, H. X., & Stepniewska, I. (2022). Escaping the nocturnal bottleneck, and the evolution of the dorsal and
ventral streams of visual processing in primates. Philosophical Transactions of the Royal Society B, 377(1844), 20210293.
https://doi.org/10.1098/RSTB.2021.0293
Kaas, J. H., & Stepniewska, I. (2016). Evolution of posterior parietal cortex and parietal-frontal networks for specific
actions in primates. Journal of Comparative Neurology, 524(3), 595–608. https://doi.org/10.1002/CNE.23838
Konopka, G., Friedrich, T., Davis-Turak, J., Winden, K., Oldham, M. C., Gao, F., Chen, L., Wang, G. Z., Luo, R.,
Preuss, T. M., & Geschwind, D. H. (2012). Human-specific transcriptional networks in the brain. Neuron, 75(4),
601–617. https://doi.org/10.1016/J.NEURON.2012.05.034
Kovach, J. K. (1971). Interaction of innate and acquired: Color preferences and early exposure learning in chicks. Journal
of Comparative and Physiological Psychology, 75(3), 386–398. https://doi.org/10.1037/h0030940
Kovach, J. K., & Hickox, J. E. (1971). Color preferences and early perceptual discrimination learning in comestic chicks.
Developmental Psychobiology, 4(3), 255–267. https://doi.org/10.1002/dev.420040304
Kovach, J. K., & Wilson, G. C. (1981). Behaviour and pleiotropy: Generalization of gene effects in the colour prefer-
ences of Japanese quail chicks (C.  coturnix japonica). Animal Behaviour, 29(3), 746–759. https://doi.org/10.1016/
S0003-3472(81)80008-5
Krubitzer, L. (2007). The magnificent compromise: Cortical field evolution in mammals. Neuron, 56(2), 201–208.
https://doi.org/10.1016/J.NEURON.2007.10.002
Lacalli, T. (2022). An evolutionary perspective on chordate brain organization and function: Insights from amphioxus,
and the problem of sentience. Philosophical Transactions of the Royal Society B, 377(1844), 20200520. https://doi.
org/10.1098/RSTB.2020.0520
Lan, G.,  & Tu, Y. (2016). Information processing in bacteria: Memory, computation, and statistical physics: A  key
issues review. Reports on Progress in Physics. Physical Society, 79(5), 052601. https://doi.org/10.1088/0034-4885/79/5/
052601
Larsch, J., & Baier, H. (2018). Biological motion as an innate perceptual mechanism driving social affiliation. Current
Biology, 28(22), 3523–3532.e4. https://doi.org/10.1016/j.cub.2018.09.014
Leder, H., Tinio, P. P. L., Brieber, D., Kröner, T., Jacobsen, T., & Rosenberg, R. (2019). Symmetry is not a universal law
of beauty. Empirical Studies of the Arts, 37(1), 104–114. https://doi.org/10.1177/0276237418777941
Levy, D. J., & Glimcher, P. W. (2012). The root of all value: A neural common currency for choice. Current Opinion in
Neurobiology, 22(6), 1027–1038. https://doi.org/10.1016/j.conb.2012.06.001
Lew, C. H., & Semendeferi, K. (2017). Evolutionary specializations of the human limbic system. In J. H. Kaas (Ed.),
Evolution of nervous systems (Vol. 4, 2nd ed., pp. 277–291). Academic Press.
Lichtneckert, R., & Reichert, H. (2009). Origin and evolution of the first nervous system. In J. H. Kaas (Ed.), Evolution-
ary neuroscience (pp. 51–77). Academic Press.
Liebeskind, B. J., Hofmann, H. A., Hillis, D. M., & Zakon, H. H. (2017). Evolution of animal neural systems. Annual Review of
Ecology, Evolution, and Systematics, 48(1), 377–398. https://doi.org/10.1146/ANNUREV-ECOLSYS-110316-023048
Liu, J., Mosti, F., & Silver, D. L. (2021). Human brain evolution: Emerging roles for regulatory DNA and RNA. Current
Opinion in Neurobiology, 71, 170–177. https://doi.org/10.1016/j.conb.2021.11.005
Liu, X., Eickhoff, S. B., Caspers, S., Wu, J., Genon, S., Hoffstaedter, F., Mars, R. B., Sommer, I. E., Eickhoff, C. R.,
Chen, J., Jardri, R., Reetz, K., Dogan, I., Aleman, A., Kogler, L., Gruber, O., Caspers, J., Mathys, C., & Patil, K.
R. (2021). Functional parcellation of human and macaque striatum reveals human-specific connectivity in the dorsal
caudate. NeuroImage, 235, 118006. https://doi.org/10.1016/J.NEUROIMAGE.2021.118006
Lopes, D. M. (2005). Painting. In B. Gaut & D. M. Lopes (Eds.), The Routledge companion to aesthetics (2nd ed., pp. 625–
637). Routledge.
Loui, P., Patterson, S., Sachs, M. E., Leung, Y., Zeng, T., & Przysinda, E. (2017, September). White matter correlates
of musical anhedonia: Implications for evolution of music. Frontiers in Psychology, 8, 1664. https://doi.org/10.3389/
fpsyg.2017.01664

285
Esther Ureña and Marcos Nadal

MacIver, M. A., & Finlay, B. L. (2022). The neuroecology of the water-to-land transition and the evolution of the verte-
brate brain. Philosophical Transactions of the Royal Society B, 377(1844), 2021. https://doi.org/10.1098/RSTB.2020.0523
Mallik, A., Chanda, M. L., & Levitin, D. J. (2017). Anhedonia to music and mu-opioids: Evidence from the administra-
tion of naltrexone. Scientific Reports, 7, 41952. https://doi.org/10.1038/srep41952
Manns, J. R., & Eichenbaum, H. (2009). The evolution of the hippocampus. In J. H. Kaas (Ed.), Evolutionary neuroscience
(pp. 603–627). Academic Press.
Marín, O., Smeets, W. J. A. J.,  & González, A. (1998). Evolution of the basal ganglia in tetrapods: A  new perspec-
tive based on recent studies in amphibians. Trends in Neurosciences, 21(11), 487–494. https://doi.org/10.1016/
S0166-2236(98)01297-1
Mars, R. B., Passingham, R. E., Neubert, F. X., Verhagen, L., & Sallet, J. (2017). Evolutionary specializations of the
human association cortex. In J. H. Kaas (Ed.), Evolution of nervous systems (Vol. 4, 2nd ed., pp. 185–205). Academic
Press.
Marshall, C. R. (2006). Explaining the Cambrian “explosion” of animals. Annual Review of Earth and Planetary Sciences,
34, 355–384. https://doi.org/10.1146/ANNUREV.EARTH.33.031504.103001
Martínez-García, F., Novejarque, A., & Lanuza, E. (2009). The evolution of the amygdala in vertebrates. In J. H. Kaas
(Ed.), Evolutionary neuroscience (pp. 313–392). Academic Press.
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J.,  & Marco-Pallarés, J. (2016). Neural
correlates of specific musical anhedonia. Proceedings of the National Academy of Sciences of the United States of America,
113(46), E7337–E7345. https://doi.org/10.1073/PNAS.1611211113/-/DCSUPPLEMENTAL
Mascalzoni, E., Regolin, L.,  & Vallortigara, G. (2010). Innate sensitivity for self-propelled causal agency in newly
hatched chicks. Proceedings of the National Academy of Sciences, USA, 107(9), 4483–4485. https://doi.org/10.1073/
pnas.0908792107
Mas-Herrero, E., Maini, L., Sescousse, G., & Zatorre, R. J. (2021, April 1). Common and distinct neural correlates of
music and food-induced pleasure: A coordinate-based meta-analysis of neuroimaging studies. Neuroscience and Biobe-
havioral Reviews, 123, 61–71. https://doi.org/10.1016/j.neubiorev.2020.12.008
Mayr, E. (1961). Cause and effect in biology. Science, 134(3489), 1501–1506. https://doi.org/10.1126/science.134.
3489.1501
McManus, I. C., Jones, A. L., & Cottrell, J. (1981). The aesthetics of colour. Perception, 10(6), 651–666. https://doi.
org/10.1068/p100651
Medina, L., Abellán, A., Vicario, A., Castro-Robles, B., & Desfilis, E. (2017). The amygdala. In J. H. Kaas (Ed.), Evolu-
tion of nervous systems (Vol. 1, 2nd ed., pp. 427–478). Academic Press.
Meier, N. C. (1942). Art in human affairs. McGraw-Hill.
Menninghaus, W., Wagner, V., Hanich, J., Wassiliwizky, E., Jacobsen, T.,  & Koelsch, S. (2017). The distancing—
Embracing model of the enjoyment of negative emotions in art reception. Behavioral and Brain Sciences, 40, e347.
https://doi.org/10.1017/S0140525X17000309
Menninghaus, W., Wagner, V., Wassiliwizky, E., Schindler, I., Hanich, J., Jacobsen, T., & Koelsch, S. (2019). What are
aesthetic emotions? Psychological Review, 126(2), 171–195. https://doi.org/10.1037/rev0000135
Miura, M., & Matsushima, T. (2016). Biological motion facilitates filial imprinting. Animal Behaviour, 116, 171–180.
https://doi.org/10.1016/J.ANBEHAV.2016.03.025
Moroz, L. L., Romanova, D. Y., & Kohn, A. B. (2021). Neural versus alternative integrative systems: Molecular insights
into origins of neurotransmitters. Philosophical Transactions of the Royal Society B, 376(1821), 20190762. https://doi.
org/10.1098/RSTB.2019.0762
Munar, E., Gómez-Puerto, G., Call, J., & Nadal, M. (2015). Common visual preference for curved contours in humans
and great apes. PLoS One, 10(11), e0141106. https://doi.org/10.1371/journal.pone.0141106
Murphy, W. J., Pevzner, P. A., & O’Brien, S. J. (2004). Mammalian phylogenomics comes of age. Trends in Genetics,
20(12), 631–639. https://doi.org/10.1016/j.tig.2004.09.005
Murray, E. A., & Fellows, L. K. (2021). Prefrontal cortex interactions with the amygdala in primates. Neuropsychopharma-
cology, 47(1), 163–179. https://doi.org/10.1038/s41386-021-01128-w
Nadal, M., Munar, E., Marty, G., & Cela-Conde, C. J. (2010). Visual complexity and beauty appreciation: Explaining the
divergence of results. Empirical Studies of the Arts, 28(2), 173–191. https://doi.org/10.2190/EM.28.2.d
Nadal, M., & Skov, M. (2018). The pleasure of art as a matter of fact. Proceedings of the Royal Society B: Biological Sciences,
285(1875). https://doi.org/10.1098/rspb.2017.2252
Nadal, M., & Ureña, E. (2021). One hundred years of empirical aesthetics: Fechner to Berlyne (1876–1976). In M.
Nadal & O. Vartanian (Eds.), The Oxford handbook of empirical aesthetics. Oxford University Press.
Nadal, M., & Vartanian, O. (2022). Empirical aesthetics: An overview. In M. Nadal & O. Vartanian (Eds.), The Oxford
handbook of empirical aesthetics. Oxford University Press.

286
Evolution of sensory valuation systems

Nakayasu, T., & Watanabe, E. (2014). Biological motion stimuli are attractive to medaka fish. Animal Cognition, 17(3),
559–575. https://doi.org/10.1007/s10071-013-0687-y
Neubauer, S., Hublin, J. J., & Gunz, P. (2018). The evolution of modern human brain shape. Science Advances, 4(1),
eaao5961. https://doi.org/10.1126/sciadv.aao5961
Nielsen, C., Brunet, T., & Arendt, D. (2018). Evolution of the bilaterian mouth and anus. Nature Ecology and Evolution,
2(9), 1358–1376. https://doi.org/10.1038/s41559-018-0641-0
Nieuwenhuys, R., Voogd, J., & van Huijzen, C. (2008). The human central nervous system (4th ed.). Springer.
Nudo, R. J., & Frost, S. B. (2009). The evolution of motor cortex and motor systems. In J. H. Kaas (Ed.), Evolutionary
neuroscience (pp. 727–749). Academic Press.
Nunes, A. R., Carreira, L., Anbalagan, S., Blechman, J., Levkowitz, G., & Oliveira, R. F. (2020). Perceptual mecha-
nisms of social affiliation in zebrafish. Scientific Reports, 10(1), 1–14, 3642. https://doi.org/10.1038/s41598-020-
60154-8
Orban, G. A., Claeys, K., Nelissen, K., Smans, R., Sunaert, S., Todd, J. T., Wardak, C., Durand, J. B., & Vanduffel,
W. (2006). Mapping the parietal cortex of human and non-human primates. Neuropsychologia, 44(13), 2647–2667.
https://doi.org/10.1016/J.NEUROPSYCHOLOGIA.2005.11.001
Palmer, S. E., Schloss, K. B., & Sammartino, J. (2013). Visual aesthetics and human preference. Annual Review of Psychol-
ogy, 64, 77–107. https://doi.org/10.1146/annurev-psych-120710-100504
Palumbo, L., & Bertamini, M. (2016). The curvature effect: A comparison between preference tasks. Empirical Studies of
the Arts, 34(1), 35–52. https://doi.org/10.1177/0276237415621185
Palumbo, L., Ruta, N., & Bertamini, M. (2015). Comparing angular and curved shapes in terms of implicit associations
and approach/avoidance responses. PloS One, 10(10), e0140043. https://doi.org/10.1371/journal.pone.0140043
Passingham, R. E., Smaers, J. B., & Sherwood, C. C. (2017). Evolutionary specializations of the human prefrontal cortex.
In J. H. Kaas (Ed.), Evolution of nervous systems (Vol. 4, 2nd ed., pp. 207–226). Academic Press.
Passingham, R. E., & Wise, S. P. (2012). The neurobiology of the prefrontal cortex: Anatomy, evolution, and the origin of insight.
Oxford University Press.
Pecchinenda, A., Bertamini, M., Makin, A. D. J., & Ruta, N. (2014). The pleasantness of visual symmetry: Always, never
or sometimes. PloS One, 9(3), e92685. https://doi.org/10.1371/journal.pone.0092685
Pessoa, L., Medina, L., & Desfilis, E. (2022). Refocusing neuroscience: Moving away from mental categories and towards
complex behaviours. Philosophical Transactions of the Royal Society B, 377(1844), 20200534. https://doi.org/10.1098/
RSTB.2020.0534
Pessoa, L., Medina, L., Hof, P. R., & Desfilis, E. (2019). Neural architecture of the vertebrate brain: Implications for
the interaction between emotion and cognition. Neuroscience and Biobehavioral Reviews, 107, 296–312. https://doi.
org/10.1016/J.NEUBIOREV.2019.09.021
Peterson, K. J.,  & Davidson, E. H. (2000). Regulatory evolution and the origin of the bilaterians. Proceedings of the
National Academy of Sciences, 97(9), 4430–4433. https://doi.org/10.1073/PNAS.97.9.4430
Pombal, M. A., & Megías, M. (2017). The nervous systems of jawless vertebrates. In J. H. Kaas (Ed.), Evolution of nervous
systems (Vol. 1, 2nd ed., pp. 37–59). Academic Press.
Ponce de León, M. S., Bienvenu, T., Marom, A., Engel, S., Tafforeau, P., Alatorre Warren, J. L. A., Lordkipanidze, D.,
Kurniawan, I., Murti, D. B., Suriyanto, R. A., Koesbardiati, T., & Zollikofer, C. P. E. (2021). The primitive brain of
early homo. Science, 372(6538), 165–171. https://doi.org/10.1126/science.aaz0032
Preuss, T. M. (2009). Primate brain evolution. In J. H. Kaas (Ed.), Evolutionary neuroscience 2 (pp. 793–825). Oxford:
Academic Press.
Preuss, T. M. (2011). The human brain: Rewired and running hot. Annals of the New York Academy of Sciences, 1225(S1),
E182–E191. https://doi.org/10.1111/J.1749-6632.2011.06001.X
Preuss, T. M. (2017a). An introduction to human brain evolutionary studies. In J. H. Kaas (Ed.), Evolution of nervous
systems (Vol. 4, 2nd ed., pp. 1–18). Academic Press.
Preuss, T. M. (2017b). The human brain: Evolution and distinctive features. In M. Tibayrenc & F. J. Ayala (Eds.), On
human nature: Biology, psychology, ethics, politics, and religion (pp. 125–149). Academic Press.
Preuss, T. M., & Wise, S. P. (2021). Evolution of prefrontal cortex. Neuropsychopharmacology, 47(1), 3–19. https://doi.
org/10.1038/s41386-021-01076-5
Raghanti, M. A., Edler, M. K., Stephenson, A. R., Wilson, L. J., Hopkins, W. D., Ely, J. J., Erwin, J. M., Jacobs, B., Hof,
P. R., & Sherwood, C. C. (2016). Human-specific increase of dopaminergic innervation in a striatal region associ-
ated with speech and language: A comparative analysis of the primate basal ganglia. Journal of Comparative Neurology,
524(10), 2117–2129. https://doi.org/10.1002/CNE.23937
Reiner, A. (2009). The evolution of the basal ganglia in mammals and other vertebrates. In J. H. Kaas (Ed.), Evolutionary
neuroscience (pp. 587–601). Academic Press.

287
Esther Ureña and Marcos Nadal

Reiner, A., Brauth, S. E., & Karten, H. J. (1984). Evolution of the amniote basal ganglia. Trends in Neurosciences, 7(9),
320–325. https://doi.org/10.1016/S0166-2236(84)80080-6
Reiner, A., Medina, L., & Veenman, C. L. (1998). Structural and functional evolution of the basal ganglia in vertebrates.
Brain Research Reviews, 28(3), 235–285. https://doi.org/10.1016/S0165-0173(98)00016-2
Rensch, B. (1957). Ästhetische Faktoren bei Farb- und Formbevorzugungen von Affen. Zeitschrift Für Tierpsychologie,
14(1), 71–99. https://doi.org/10.1111/j.1439-0310.1957.tb00526.x
Rensch, B. (1958). Die Wirksamkeit ästhetischer Faktoren bei Wirbeltieren. Zeitschrift Für Tierpsychologie, 15(4), 447–
461. https://doi.org/10.1111/j.1439-0310.1958.tb00575.x
Rilling, J. K. (2006). Human and nonhuman primate brains: Are they allometrically scaled versions of the same design?
Evolutionary Anthropology, 15(2), 65–77. https://doi.org/10.1002/evan.20095
Rilling, J. K., Glasser, M. F., Preuss, T. M., Ma, X., Zhao, T., Hu, X., & Behrens, T. E. J. (2008). The evolution of the
arcuate fasciculus revealed with comparative DTI. Nature Neuroscience, 11(4), 426–428. https://doi.org/10.1038/
nn2072
Rilling, J. K., Glasser, M., Jbabdi, S., Andersson, J.,  & Preuss, T. (2012, January). Continuity, divergence, and the
evolution of brain language pathways. Frontiers in Evolutionary Neuroscience, 3, 11. https://doi.org/10.3389/
FNEVO.2011.00011/BIBTEX
Rilling, J. K., & Insel, T. R. (1999). The primate neocortex in comparative perspective using magnetic resonance imag-
ing. Journal of Human Evolution, 37(2), 191–223. https://doi.org/10.1006/JHEV.1999.0313
Rilling, J. K., & Van Den Heuvel, M. P. (2018). Comparative primate connectomics. Brain, Behavior and Evolution, 91(3),
170–179. https://doi.org/10.1159/000488886
Ross, C. F., & Martin, R. D. (2009). The role of vision in the origin and evolution of primates. In J. H. Kaas (Ed.),
Evolutionary neuroscience (pp. 827–846). Academic Press.
Rowe, T. B. (2017). The emergence of mammals. In J. H. Kaas (Ed.), Evolution of nervous systems (Vol. 2, 2nd ed.,
pp. 1–52). Academic Press.
Rowe, T. B., & Shepherd, G. M. (2016). Role of ortho-retronasal olfaction in mammalian cortical evolution. Journal of
Comparative Neurology, 524(3), 471–495. https://doi.org/10.1002/cne.23802
Rudebeck, P. H., & Izquierdo, A. (2021). Foraging with the frontal cortex: A cross-species evaluation of reward-guided
behavior. Neuropsychopharmacology, 47(1), 134–146. https://doi.org/10.1038/s41386-021-01140-0
Ruta, N., Vañó, J., Pepperell, R., Corradi, G. B., Chuquichambi, E. G., Rey, C.,  & Munar, E. (2021). Preference
for paintings is also affected by curvature. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/
ACA0000395
Sachs, M. E., Ellis, R. J., Schlaug, G., & Loui, P. (2016). Brain connectivity reflects human aesthetic responses to music.
Social Cognitive and Affective Neuroscience, 11(6), 884–891. https://doi.org/10.1093/SCAN/NSW009
Sahgal, A., & Iversen, S. D. (1975). Colour preferences in the pigeon: A behavioural and psychopharmacological study.
Psychopharmacologia, 43(2), 175–179. https://doi.org/10.1007/BF00421021
Sahgal, A., Pratt, S. R., & Iversen, S. D. (1975). Response preferences of monkeys (Macaca mulatta) within wavelength
and lie-tilt dimensions. Journal of the Experimental Analysis of Behavior, 24(3), 377–381. https://doi.org/10.1901/
jeab.1975.24-377
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059
Salzen, E. A., Lily, R. E., & McKeown, J. R. (1971). Colour preference and imprinting in domestic chicks. Animal
Behaviour, 19(3), 542–547. https://doi.org/10.1016/S0003-3472(71)80109-4
Schaefer, H. H., & Hess, E. H. (1959). Color preferences in imprinting objects. Zeitschrift Für Tierpsychologie, 16, 161–172.
Schenker, N. M., Buxhoeveden, D. P., Blackmon, W. L., Amunts, K., Zilles, K., & Semendeferi, K. (2008). A compara-
tive quantitative analysis of cytoarchitecture and minicolumnar organization in Broca’s area in humans and great apes.
Journal of Comparative Neurology, 510(1), 117–128. https://doi.org/10.1002/CNE.21792
Schenker, N. M., Desgouttes, A. M., & Semendeferi, K. (2005). Neural connectivity and cortical substrates of cognition
in hominoids. Journal of Human Evolution, 49(5), 547–569. https://doi.org/10.1016/j.jhevol.2005.06.004
Schneider, G. E. (2014). Brain structure and its origins in development and in evolution of behavior and the mond. MIT Press.
Semendeferi, K., Teffer, K., Buxhoeveden, D. P., Park, M. S., Bludau, S., Amunts, K., Travis, K.,  & Buckwalter, J.
(2011). Spatial organization of neurons in the frontal pole sets humans apart from great apes. Cerebral Cortex, 21(7),
1485–1497. https://doi.org/10.1093/cercor/bhq191
Sescousse, G., Caldú, X., Segura, B., & Dreher, J. C. (2013). Processing of primary and secondary rewards: A quantita-
tive meta-analysis and review of human functional neuroimaging studies. Neuroscience and Biobehavioral Reviews, 37(4),
681–696. https://doi.org/10.1016/j.neubiorev.2013.02.002

288
Evolution of sensory valuation systems

Sherwood, C. C., Bauernfeind, A. L., Verendeev, A., Raghanti, M. A., & Hof, P. R. (2017). Evolutionary specializa-
tions of human brain microstructure. In J. H. Kaas (Ed.), Evolution of nervous systems (Vol. 4, 2nd ed., pp. 121–139).
Academic Press.
Sherwood, C. C., Stimpson, C. D., Raghanti, M. A., Wildman, D. E., Uddin, M., Grossman, L. I., Goodman, M.,
Redmond, J. C., Bonar, C. J., Erwin, J. M., & Hof, P. R. (2006). Evolution of increased glia-neuron ratios in the
human frontal cortex. Proceedings of the National Academy of Sciences, 103(37), 13606–13611. https://doi.org/10.1073/
PNAS.0605843103
Sifre, R., Olson, L., Gillespie, S., Klin, A., Jones, W.,  & Shultz, S. (2018). A  longitudinal investigation of preferen-
tial attention to biological motion in 2- to 24-month-old infants. Scientific Reports, 8(1), 1–10, 2527. https://doi.
org/10.1038/s41598-018-20808-0
Silvia, P. J., & Barona, C. M. (2009). Do people prefer curved objects? Angularity, expertise, and aesthetic preference.
Empirical Studies of the Arts, 27(1), 25–42. https://doi.org/10.2190/EM.27.1.b
Simion, F., Regolin, L., & Bulf, H. (2008). A predisposition for biological motion in the newborn baby. Proceedings of the
National Academy of Sciences, 105(2), 809–813. https://doi.org/10.1073/PNAS.0707021105
Skov, M. (2019). Aesthetic appreciation: The view from neuroimaging. Empirical Studies of the Arts, 37(2), 220–248.
https://doi.org/10.1177/0276237419839257
Skov, M., & Nadal, M. (2018). Art is not special: An assault on the last lines of defense against the naturalization of the
human mind. Reviews in the Neurosciences, 29(6), 699–702. https://doi.org/10.1515/revneuro-2017-0085
Skov, M., & Nadal, M. (2020). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Smaers, J. B., Gómez-Robles, A., Parks, A. N., & Sherwood, C. C. (2017). Exceptional evolutionary expansion of pre-
frontal cortex in great apes and humans. Current Biology, 27(5), 714–720. https://doi.org/10.1016/J.CUB.2017.01.020
Smeets, W. J. A. J., Marín, O.,  & González, A. (2000). Evolution of the basal ganglia: New perspectives through a
comparative approach. Journal of Anatomy, 196(4), 501–517. https://doi.org/10.1046/J.1469-7580.2000.19640501.X
Smith, S. J. (2021). Transcriptomic evidence for dense peptidergic networks within forebrains of four widely divergent
tetrapods. Current Opinion in Neurobiology, 71, 100–109. https://doi.org/10.1016/j.conb.2021.09.011
Somel, M., Franz, H., Yan, Z., Lorenc, A., Guo, S., Giger, T., Kelso, J., Nickel, B., Dannemann, M., Bahn, S., Webster,
M. J., Weickert, C. S., Lachmann, M., Pääbo, S., & Khaitovich, P. (2009). Transcriptional neoteny in the human brain.
Proceedings of the National Academy of Sciences, 106(14), 5743–5748. https://doi.org/10.1073/PNAS.0900544106
Sousa, A. M. M., Meyer, K. A., Santpere, G., Gulden, F. O., & Sestan, N. (2017). Evolution of the human nervous system
function, structure, and development. Cell, 170(2), 226–247. https://doi.org/10.1016/j.cell.2017.06.036
Sousa, A. M. M., Zhu, Y., Raghanti, M. A., Kitchen, R. R., Onorati, M., Tebbenkamp, A. T. N., Stutz, B., Meyer, K.
A., Li, M., Kawasawa, Y. I., Liu, F., Perez, R. G., Mele, M., Carvalho, T., Skarica, M., Gulden, F. O., Pletikos, M.,
Shibata, A., Stephenson, A. R.,.  . . Sestan, N. (2017). Molecular and cellular reorganization of neural circuits in the
human lineage. Science, 358(6366), 1027–1032. https://doi.org/10.1126/science.aan3456
Stephenson-Jones, M., Samuelsson, E., Ericsson, J., Robertson, B., & Grillner, S. (2011). Evolutionary conservation of
the basal ganglia as a common vertebrate mechanism for action selection. Current Biology, 21(13), 1081–1091. https://
doi.org/10.1016/j.cub.2011.05.001
Stephenson, A. R., Edler, M. K., Erwin, J. M., Jacobs, B., Hopkins, W. D., Hof, P. R., Sherwood, C. C., & Raghanti,
M. A. (2017). Cholinergic innervation of the basal ganglia in humans and other anthropoid primates. Journal of Com-
parative Neurology, 525(2), 319–332. https://doi.org/10.1002/CNE.24067
Striedter, G. F. (2005). Principles of brain evolution. Sinauer Associates, Inc.
Striedter, G. F., & Northcutt, R. G. (2020). Brains through time: A natural history of vertebrates. Oxford University Press.
Sully, J. (1892). The human mind: A textbook of psychology. Longmans, Green, and Co.
Suryanarayana, S. M., Pérez-Fernández, J., Robertson, B., & Grillner, S. (2021). The lamprey forebrain—Evolutionary
implications. Brain Behavior and Evolution, 1–16. https://doi.org/10.1159/000517492
Sussman, R. W. (1991). Primate origins and the evolution of angiosperms. American Journal of Primatology, 23(4), 209–
223. https://doi.org/10.1002/AJP.1350230402
Tang, S. K. Y., & Marshall, W. F. (2018). Cell learning. Current Biology, 28(20), R1180–R1184. https://doi.org/10.1016/j.
cub.2018.09.015
Tigges, M. (1963). Muster- und Farbbevorzugung bei Fischen und vögeln. Zeitschrift Für Tierpsychologie, 20(2), 129–142.
https://doi.org/10.1111/j.1439-0310.1963.tb01147.x
Tilot, A. K., Khramtsova, E. A., Liang, D., Grasby, K. L., Jahanshad, N., Painter, J., Colodro-Conde, L., Bralten, J.,
Hibar, D. P., Lind, P. A., Liu, S., Brotman, S. M., Thompson, P. M., Medland, S. E., Macciardi, F., Stranger, B. E.,
Davis, L. K., Fisher, S. E., & Stein, J. L. (2021). The evolutionary history of common genetic variants influencing
human cortical surface area. Cerebral Cortex, 31(4), 1873–1887. https://doi.org/10.1093/CERCOR/BHAA327

289
Esther Ureña and Marcos Nadal

Tinio, P. P. L., & Leder, H. (2009). Just how stable are stable aesthetic features? Symmetry, complexity, and the jaws of
massive familiarization. Acta Psychologica, 130(3), 241–250. https://doi.org/10.1016/j.actpsy.2009.01.001
Tosches, M. A., & Arendt, D. (2013). The bilaterian forebrain: An evolutionary chimaera. Current Opinion in Neurobiol-
ogy, 23(6), 1080–1089. https://doi.org/10.1016/J.CONB.2013.09.005
Uddin, M., Wildman, D. E., Liu, G., Xu, W., Johnson, R. M., Hof, P. R., Kapatos, G., Grossman, L. I., & Goodman,
M. (2004). Sister grouping of chimpanzees and humans as revealed by genome-wide phylogenetic analysis of brain
gene expression profiles. Proceedings of the National Academy of Sciences, 101(9), 2957–2962. https://doi.org/10.1073/
PNAS.0308725100
Valentine, C. W. (1962). The experimental psychology of beauty. Methuen and Co.
Vallortigara, G., & Regolin, L. (2006). Gravity bias in the interpretation of biological motion by inexperienced chicks.
Current Biology, 16(8), R279–R280. https://doi.org/10.1016/j.cub.2006.03.052
Vallortigara, G., Regolin, L., & Marconato, F. (2005). Visually inexperienced chicks exhibit spontaneous preference for
biological motion patterns. PLoS Biology, 3(7), e208. https://doi.org/10.1371/journal.pbio.0030208
van den Heuvel, M. P., Bullmore, E. T., & Sporns, O. (2016). Comparative conectomics. Trends in Cognitive Sciences,
20(5), 345–361. https://doi.org/10.1016/j.tics.2016.03.001
Van Geert, E., & Wagemans, J. (2019). Order, complexity, and aesthetic preferences for neatly organized compositions.
Psychology of Aesthetics, Creativity, and the Arts, 15(3), 484–504. https://doi.org/10.1037/aca0000276
Varoqueaux, F., & Fasshauer, D. (2017). Getting nervous: An evolutionary overhaul for communication. Annual Review
of Genetics, 51, 455–476. https://doi.org/10.1146/ANNUREV-GENET-120116-024648
Vartanian, O. (2022). Color. In M. Nadal & O. Vartanian (Eds.), Oxford handbook of empirical aesthetics. Oxford University
Press.
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov, M.
(2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceedings of the
National Academy of Sciences of the United States of America, 110(Suppl. 2), 10446–10453. https://doi.org/10.1073/
pnas.1301227110
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Rostrup, N., Skov, M., Corradi,
G., & Nadal, M. (2018). Preference for curvilinear contour in interior architectural spaces: Evidence from experts
and nonexperts. Psychology of Aesthetics, Creativity, and the Arts, 13(1), 110–116. https://doi.org/10.1037/aca0000150
Verendeev, A., & Sherwood, C. C. (2017). Human brain evolution. Current Opinion in Behavioral Sciences, 16, 41–45.
https://doi.org/10.1016/j.cobeha.2017.02.003
Washburn, S. L. (1970). Comment on “A possible evolutionary basis for aesthetic appreciation in men and apes”. Evolu-
tion and Human Behavior, 24(4), 824–825. https://doi.org/10.1111/j.1558-5646.1970.tb01818.x
Weichselbaum, H., Leder, H., & Ansorge, U. (2018). Implicit and explicit evaluation of visual symmetry as a function of
art expertise. i-Perception, 9(2), 2041669518761464. https://doi.org/10.1177/2041669518761464
Westerman, S. J., Gardner, P. H., Sutherland, E. J., White, T., Jordan, K., Watts, S.,  & Wells, S. (2012). Product
design: Preference for rounded versus angular design elements. Psychology and Marketing, 29, 595–605. https://doi.
org/10.1002/mar
Westphal-Fitch, G., Huber, L., Gómez, J. C., & Fitch, W. T. (2012). Production and perception rules underlying visual
patterns: Effects of symmetry and hierarchy. Philosophical Transactions of the Royal Society of London. Series B, Biological
Sciences, 367(1598), 2007–2022. https://doi.org/10.1098/rstb.2012.0098
Wickens, A. P. (2015). A history of the brain: From stone age surgery to modern neuroscience. Psychology Press.
Wilsch-Bräuninger, M., Florio, M., & Huttner, W. B. (2016). Neocortex expansion in development and evolution—
From cell biology to single genes. Current Opinion in Neurobiology, 39, 122–132. https://doi.org/10.1016/J.
CONB.2016.05.004
Winne, J., Teixeira, L., de Andrade Pessoa, J., Gavioli, E. C., Soares-Rachetti, V., André, E., & Lobão-Soares, B. (2015).
There is more to the picture than meets the rat: A study on rodent geometric shape and proportion preferences.
Behavioural Brain Research, 284, 187–195. https://doi.org/10.1016/j.bbr.2015.02.018
Wise, S. P. (2017). The evolution of the prefrontal cortex in early primates and anthropoids. In J. H. Kaas (Ed.), Evolution
of nervous systems (Vol. 3, 2nd ed., pp. 387–422). Academic Press.
Won, H., Huang, J., Opland, C. K., Hartl, C. L., & Geschwind, D. H. (2019). Human evolved regulatory elements
modulate genes involved in cortical expansion and neurodevelopmental disease susceptibility. Nature Communications,
10(1), 1–11, 2396. https://doi.org/10.1038/s41467-019-10248-3
Xiang, L., Crow, T. J., Hopkins, W. D., & Roberts, N. (2020). Comparison of surface area and cortical thickness asym-
metry in the human and chimpanzee brain. Cerebral Cortex. https://doi.org/10.1093/CERCOR/BHAA202
Zolman, J. F. (1969). Stimulus preferences and form discrimination learning in young chicks. Psychological Record, 19(3),
407–416. https://doi.org/10.1007/BF03393867

290
Evolution of sensory valuation systems

Zolman, J. F.,  & Lattin, W. J. (1972). Development of brightness preferences in young chicks: Effects on brightness
discrimination learning. Journal of Comparative and Physiological Psychology, 79(2), 271–283. https://doi.org/10.1037/
h0032565
Zolman, J. F., Pursley, D. G., Hall, J. A., & Sahley, C. L. (1975). Form preferences in successive discrimination learning
of young chicks. Journal of Comparative and Physiological Psychology, 89(10), 1180–1191. https://doi.org/10.1037/
h0077176

291
PART II

Art
14
PERCEPTION AND COGNITION IN
VISUAL ART EXPERIENCE
Rebecca Chamberlain

Experiences with visual art are complex, multi-faceted, contextually dependent phenomena. They integrate
a variety of personal and situational factors, as well as object and event-related attributes. Given the complex
behavioural nature of visual art experiences, it is with good reason that the neural basis of such phenomena
has been challenging to characterize and explain. Efforts to provide an understanding of the role neural
systems play in visual art experiences remain in their infancy. Nevertheless, research in the field of visual
neuroaesthetics has gathered apace in recent years. This chapter’s purpose is to outline the state of the knowl-
edge of the field at present. It will serve as a foundation for more in-depth discussions on the behavioural
and neural basis for individual components of visual art experience and responses to other visual stimuli in
other chapters in this volume. To provide a firm foundation to this topic, it will be necessary to first provide
some brief historical background with respect to psychological and neuroscientific research on experiences
with visual art. I will then outline some of the key psychological models that have been proposed to explain
some or all components of visual art experience. Following this, I will present more recent theoretical mod-
els that attempt to ground these psychological mechanisms in the workings of the brain. A helpful roadmap
is provided by one of the leading frameworks, the Aesthetic Triad (Chatterjee & Vartanian, 2014), which
characterizes visual art experiences as an integration of three interacting neural systems associated with:
emotion-valuation, sensory-motor and meaning-knowledge components of aesthetic experience. I will use
the structure of the Aesthetic Triad to delineate the various lines of empirical research that attempt to pin-
point and explain the neural basis of aspects of visual art experience. Finally, I will highlight areas of research
that are likely to be the focus of future investigations in the field and might serve to address some of the most
compelling gaps in our knowledge that remain.

Historical origins of empirical approaches to visual art experience


Empirical approaches to visual art experience can trace their origins back to Gustav Fechner’s (1876) semi-
nal work on aesthetic psychophysics, Vorschule der Aesthetik. Fechner’s aesthetics “from below” positioned
objective stimulus properties at the heart of the empirical aesthetic project, providing the foundation for
later efforts to establish lawful relationships between aspects of visual art and aesthetic preferences (Birkhoff,
1933; Eysenck, 1940). As such, early research into the nature of the visual art experience placed much of
their focus on the role of artwork properties such as colour, symmetry, proportion, contrast and contour, a
tradition which continues in empirical work today (e.g., Dijkstra & van Dongen, 2017; Ruta et al., 2021;

DOI: 10.4324/9781003008675-16 295


Rebecca Chamberlain

Specker et al., 2020). Fechner not only provided a theory of how certain visual stimuli come to be appreci-
ated, but he also pioneered methodological approaches for studying observers’ responses to artworks that
also remain in contemporary use including the method of production, the method of use and the method
of choice (Westphal-Fitch, 2019).
Much as Fechner’s theory and methodological approach informs contemporary research in the study of
experiences of art, the Gestalt movement of the early 20th century also holds influence over current theories
of visual art experience. Rudolf Arnheim (1965) put forward a psychological framework embedding Gestalt
principles in the study of art. Arnheim’s underlying theory was that Gestalt grouping principles (e.g., good
continuation, proximity, similarity, closure) can be formally analysed within works of art and drive an observ-
er’s aesthetic response to an artwork. Specifically, Arnheim linked aesthetic pleasure to Prägnanz (Koffka,
1935), a law that specifies that elements are perceptually organized according to the most parsimonious
structure, such that violations of Prägnanz result in unattractive stimuli. Furthermore, Arnheim argued
that expressive aspects of artworks could be derived quickly and spontaneously by the observer, as they are
embedded in the surface structure of an artwork. Arnheim’s theory therefore linked the visual structure of
an artwork (particularly the formation of groupings of artwork elements) to their aesthetic and expressive
properties. Later, Gestalt theory fell out of favour, due to a lack of empirical support for the posited mapping
between holistic activity in the brain and the structure of the visual world (Cupchik, 2007). However, the
notion that certain configurations of artwork elements give rise to different kinds of aesthetic experience
and that more harmonious groupings elicit aesthetic pleasure still prevails in contemporary thinking around
visual art and also design (Hekkert, 2006; Muth et al., 2013; Muth & Carbon, 2013; Van Geert & Wagemans,
2020).
In the 1970s Daniel Berlyne attempted to position a theoretical understanding of empirical aesthetics
firmly within biological mechanisms. Berlyne’s psychobiological aesthetics posited that the collative features
of a stimulus (e.g., novelty, complexity, uncertainty, etc.) influenced arousal levels of an organism, motivat-
ing behaviour and generating emotion via systems of reward and aversion (Berlyne, 1974). The relation-
ship between collative features and dependent variables such as liking were predicted to form an inverted
U-shape, a function of the interaction between reward and aversion systems as a particular collative property
increased. While the biological basis for Berlyne’s theory failed to gain empirical support and predicted
inverted-U relationships proved difficult to isolate in experimental conditions, Berlyne’s work represented
a crucial step toward establishing empirical aesthetics in the mainstream of experimental psychology. Addi-
tionally, much like Fechner and Arnheim’s work, the theory acknowledged the role of stimulus properties
in driving observers’ aesthetic responses but additionally accounted for the role of observer-specific factors
on the reception of a stimulus, such as prior experience. Later theory and research in the field also focused
on sensory and cognitive processing dynamics, modelling how observers respond to salient properties of the
stimulus through processing fluency mechanisms (Flavell et al., 2020; Reber et al., 2004) while at the same
time incorporating the sensory and cognitive history of the observer through effects of mere exposure and
preference for prototypical stimuli (Cutting, 2003; Winkielman et al., 2006; Zajonc, 1968). Contemporary
models of visual art experience attempt to capture many of the phenomena and perspectives set forth by
these early theories by integrating them into a series of information processing stages or modules (Chatter-
jee & Vartanian, 2014; Leder et al., 2004; Pelowski et al., 2017).
While Gestalt inspired accounts of visual art (Arnheim, 1965) and theories based in psychobiology (Ber-
lyne, 1974) put forward tentative predictions for the way in which aesthetic experience relies on the workings
of the brain, the ability to empirically validate such theories was limited until the wide availability of func-
tional neuroimaging technologies such as electroencephalography (EEG) and functional magnetic resonance
imaging (fMRI) in the latter stages of the 20th century. The first neuroaesthetic theories emerged in the
wake of these advances (see Chapter 1), with neurobiologist Semir Zeki and neuroscientist V. S. Ramachan-
dran positing connections between the functioning of the brain and art experience. The approaches of Zeki

296
Visual art experience

and Ramachandran have been classed as “descriptive neuroasthetics” (Chatterjee & Vartanian, 2014), as they
do not employ experimental techniques to test the role and function of brain regions in our experience of
visual art but instead make observations of how the structure and function of art maps onto the structure and
function of the brain. Their thesis is that there is a deep structure underlying and unifying our experience
of art. They posit that “all visual art must obey the laws of the visual system” (Zeki & Lamb, 1994, p. 607)
and that therefore artists have evolved specific pictorial techniques to “titillate the visual areas of the brain”
(p. 17, Ramachandran & Hirstein, 1999). Both Ramachandran and Hirstein (1999) and Zeki (1999) high-
light the artist’s ability to capture essential characteristics of a visual stimulus, in much the same way as the
visual system itself does. In this way artists are conceived as visual neuroscientists; they unconsciously reveal
the function and organization of the visual system (Zeki & Lamb, 1994). Ramachandran and Hirstein (1999)
go further and suggest that artists amplify particular visual features (such as form, colour, motion, contrast)
in order to produce a super stimulus, functioning via the peak shift principle to maximize motivational and
affective responses of the observer. Since Semir Zeki popularized the term neuroaesthetics in 1999, the neu-
roscientific study of visual art experience has proceeded apace. Before digging into the emergent findings
from this discipline, it will be helpful to outline prevailing psychological models of visual art experiences,
which can provide a framework for understanding how neural systems are involved in the experience of art.

Psychological models of the experience of visual art


Psychological models of visual art experience tend to place focus on different elements of the aesthetic pro-
cess, in turn reflecting relative focus on stimulus-based or observer-based features. Bottom-up approaches
tend to focus on early-stage processing of low- and mid-level sensory features, while top-down accounts
place emphasis on subjective factors that may shape incoming information. Different models also address
intervening mechanisms (i.e., how sensory input is translated into reward-based output) to varying degrees.
A relatively simple model that includes both early and late-stage processing is the mirror model (Tinio,
2013; Figure 14.1). This model captures the relationship between art creation and appreciation. The first
stage of aesthetic experience constitutes the final stage of artwork creation. While the observer engages in
early, automatic sensory processing at this stage, what is actually represented is the finalizing stage of artwork
creation, in which the surface-level information is added by the artist. The intermediate stage of the model
is one which engages memory-based processing in the observer, while the artist engages in expansion and
adaptation of earlier derived artistic themes. The final stage for the observer involves meaning-making and
the experience of aesthetic judgments and emotions, corresponding with initialization of the artistic ideas
contained within the artwork (Tinio, 2013). Tinio’s mirror model is the only model to acknowledge the
role of the artistic creative process in the experience of the observer. However, other models have provided
a much more in-depth description and explanation of the various stages of aesthetic processing undertaken
by the observer during an experience with visual art.
Locher et al.’s (2007, 2010) model focuses on earlier stage processing, arguing that viewing an artwork
involves translating an original gist-like representation into a focused stage of attention on elements forming
compositional units, which can then give rise to impressions of expressive and stylistic aspects (Locher, 2015).
The model serves to explain empirical effects of eye movements when viewing artworks, as well as qualita-
tive data based on observation of museum behaviour and observer descriptions of artwork viewing. In their
model, Locher and colleagues emphasize the role of both properties of the artwork but also the context of
the observer (Locher et al., 2010). By contrast, Silvia (2005) focuses on late-stage mechanisms for translat-
ing an experience of an artwork into an emotional response. Noting the deficiency of processing fluency
accounts of artwork experience (Reber et al., 2004) in capturing the range of emotional responses to art,
Silvia harnessed appraisal theory (Scherer, 2001) to explain why certain observers have certain emotional
reactions to different artworks. Silvia (2005) emphasizes the role of the evaluative process itself in generating

297
Rebecca Chamberlain

Figure 14.1 Tinio’s (2013) Mirror model of art. Reprinted with permission from Psychology of Aesthetics, Creativity, and
the Arts.

an emotional response to an artwork rather than any formal features of the artwork. From Silvia’s perspec-
tive, emotional responses to artworks result from a novelty check (to what extent the artwork coheres with
the observer’s existing schemas), an estimate of coping potential (the observer’s sense of their control over the
artwork viewing situation) and finally the importance of the artwork viewing situation to the observer’s sense
of self. More recently, the distancing-embracing model (Menninghaus et al., 2017) further elaborates on the
experience of negative emotions during art experiences (including but not limited to visual art). It posits that
distancing mechanisms activated by an artistic schema and implying safety and control interact with embrac-
ing mechanisms afforded by an artistic experience (meaning-making, aesthetic attributes of stimuli, mixed
emotions), which convert negative emotions into enjoyable aesthetic experiences. The models of Locher
et al. (2007, 2010), Silvia (2005) and Meninghaus et al. (2017) enable research in visual art experience to
encompass both low-level emotional processing reflected by fluency mechanisms as well as more complex
emotional responses to works of art.
The most comprehensive and influential model of experience with artworks is that of Leder et al. (2004;
Figure  14.2). Designed to model experiences with modern art, it encompasses both early and late-stage
processing, as well as elaborating on intervening mechanisms. It is a linear model which conceives of artwork
viewing as a series stages of information-processing, consisting of a pre-classification stage, followed by five
processing stages (perceptual analysis, implicit memory integration, explicit classification, cognitive mastery,
evaluation), resulting in both an aesthetic judgment and an aesthetic emotion. In a similar manner to the
model of Locher et al. (2010), Leder et al.’s (2004) model takes into account formal aspects of the artwork
(the focus of the perceptual analysis and explicit classification stages), as well as aspects that the observer
brings to the experience, such as meaning-making and integration with expertise and existing schema (the
focus of the implicit memory integration and cognitive mastery stages). It also distinguishes between two
potential outcomes of the artwork viewing process: aesthetic judgments and emotions, which arise from the
final evaluative stage. The model has been highly influential in the field, arguably due to its modular design,

298
Visual art experience
299

Figure 14.2 Leder et al.’s (2004) model of aesthetic appreciation and aesthetic judgements. Reprinted with permission from British Journal of Psychology.
Rebecca Chamberlain

which made the isolation of different aspects of the art viewing process possible to study empirically, as well
as having a level of complexity that allowed for the integration of processes that had previously been consid-
ered in a unitary manner (Leder & Nadal, 2014).
Pelowski et al. (2017) integrated Leder et al.’s (2004) model with the appraisal theory approach of Silvia
(2005) and an earlier iteration of their own model (Pelowski & Akiba, 2011) to create the Vienna integrated
model of top-down and bottom-up processes in art perception (VIMAP). The VIMAP represents the most
comprehensive model of artwork viewing to date, adding several behavioural outcomes to the various infor-
mation processing stages of the Leder et al. (2004) model. As a result, the VIMAP describes a much wider
range of artwork viewing phenomena than had been accounted for by previous models and theories. In
addition, the VIMAP is one of the first models to elaborate extensively on the potential neural substrates
of the various processing stages and resulting outcomes. I will briefly outline some of the brain regions and
functions implicated in this model in the following sections before going into more detail on the empirical
basis for these neural correlates later in this chapter.

The VIMAP and neural substrates of art viewing


Mapping on to Leder et al.’s (2004) model, the VIMAP’s (Pelowski et  al., 2017) processing stages begin
with a pre-classification stage (the observer’s state prior to their artwork viewing experience) which serves
to frame the observer’s experience through contextual factors (e.g., being in an art gallery context) and
personal factors (e.g., their expertise, personality, pre-existing emotional state, etc.). The neural substrates
implicated in this stage are largely based on empirical studies on framing effects, which point to the role of
the medial region of the orbitofrontal cortex (OFC) when observers view artworks within specific contexts
(e.g., Huang et al., 2011; Kirk et al., 2009).
The perceptual analysis stage (which takes place immediately following artwork onset) encompasses early
visual analysis (<100 ms post stimulus onset), low-level fluency effects (e.g., effects of contrast, contour, sym-
metry) and gist perception of artwork attributes such as colour and composition. This stage predominantly
recruits visual mechanisms, ranging from retinal responses, through to responses throughout the dorsal and
ventral visual pathways in the occipital cortex. The perceptual analysis stage is followed by the implicit
memory integration stage, which includes grouping functions of mid-level vision and detail-focused pro-
cessing of the kind alluded to in Locher et al.’s (2010) model. Higher-level fluency mechanisms may come
into play here based on the incoming expectations of the observer, in terms of the identity of aspects of the
artwork content. Such processing is likely to still largely involve visual mechanisms, albeit in higher regions
of the visual stream, which predominantly respond to the particular category of the represented stimulus
(e.g., regions such as the parahippocampal place area [PPA], fusiform face area [FFA], and lateral occipital
complex [LO]). In addition, regions associated with the integration of memory, such as the temporal lobes,
hippocampus and regions relating to processing fluency with emotional outcomes are likely to be engaged at
this stage (Pelowski et al., 2017). Following the implicit memory integration stage, the explicit classification
stage constitutes late-stage visual processing where more complex artistic attributes are computed, such as the
style and meaning of the artwork and fluency mechanisms based on conceptual fit. Such a stage likely starts
to recruit regions of the frontal lobes associated with higher-level cognitive processing and reward.
The final stage, cognitive mastery, involves the formation of some sense of coherent meaning surrounding
the artwork, arising from the observer’s schema and expectations and leading to distinct outcomes based on
processing checks which derive from the integration of appraisal theory (Silvia, 2005) into the model. The
outcomes depend on the level of schema congruency the observer experiences (e.g., how much does this
experience match my expectations?) and the relevance of the artwork viewing situation to the observer. Such
computations likely recruit regions of the frontal lobe including the dorsolateral prefrontal cortex (DLPFC),
the medial prefrontal cortex (MPFC) and OFC and link into other reward systems associated with different

300
Visual art experience

emotional outcomes. If there is high congruence and low self-relevance, the interaction with an artwork is
unlikely to have a great psychological impact on the observer and is likely to be brief. If there is low congru-
ence in the presence of low self-relevance, the observer is likely to have some sense of novelty and insight but
again without great psychological impact to the observer. However, if congruence is high and self-relevance
is high, the observer is likely to encounter some kind of flow or peak experience state, commensurate with
an extreme feeling of mastery (Pelowski et al., 2017).
Only if congruence is low and self-relevance is high for the observer is it necessary to proceed to further
stages of the model. Following this particular outcome comes a secondary control and coping check. At this
stage the observer assesses their coping potential: the degree to which it is necessary for them to continue
to engage with the artwork viewing experience. If the need for coping is low, then the observer will likely
reframe the context or situation or abandon the situation altogether, all of which tend to result in negative
outcomes of interactions with art. However, if the coping need is high, the observer will undergo a period
of self-reflection and potential schema change, which offers up the potential for transformative experience
with art (Pelowski et al., 2017).

The aesthetic triad


The benefit of the VIMAP is that it systematically explores a broad range of outcomes from art experi-
ences, spanning the facile to the transformative, where previous theories and models have tended to only
capture a subset of these phenomena (e.g., processing fluency for facile experiences (Reber et al., 2004),
the distancing-embracing model for negative emotions (Meninghaus et  al., 2017)). In doing so, it can
provide a broad framework for understanding the neuroscientific basis for artwork experiences. Chatterjee
and Vartanian’s (2014) widely used review of the emerging field of neuroaesthetics also sets out a putative
neuroscientific framework for aesthetic experience. In this framework, aesthetic experience is character-
ized as an emergent state, resulting from the interaction of three broadly distinguished neural systems:
sensory-motor, emotion-valuation and meaning-knowledge. The authors emphasize that the context and
nature of appraisals in aesthetic experience resulting from objects rather than outcome delineate aesthetic
experience from other kinds of evaluative experiences (Chatterjee  & Vartanian, 2014). They go on to
discuss putative neural circuits that underpin these three systems and pose ongoing questions for the field
that will be returned to at the conclusion of this chapter. These questions pertain to the role of individual
differences (e.g., gender and expertise), the mechanisms and neural instantiation of aesthetic experience
(e.g., where and how is value computed in aesthetic experience), the uniqueness of aesthetic emotions
(e.g., are the same regions implicated in both moral and aesthetic judgements) and the applications of aes-
thetics (e.g., therapeutic and long-term social benefits of aesthetic engagement). Before turning to these
pressing questions, we will consider empirical evidence for the role that each facet of the Aesthetic Triad
plays in experiences with artworks.

Sensory-motor
The sensory-motor system in Chatterjee and Vartanian’s (2014) review of neuroaesthetics encompasses sensa-
tion, perception and the motor system (Figure 14.3). Several low-level visual features are known to influence
aesthetic appraisal in non-artistic stimuli and are therefore likely to play a role in some artwork experiences.
There is also evidence that category-specific artworks differentially engage different parts of the ventral visual
stream, with portraits activating the FFA, scenes activating the PPA and sculpture activating fusiform body
areas as well as some motor regions (Boccia et al., 2016). The evidence for the neural basis for preference
for visual features in artworks and higher-level artwork content categories such as faces and bodies will now
be discussed.

301
Rebecca Chamberlain

Figure 14.3 The Aesthetic Triad (Chatterjee & Vartanian, 2014). Reprinted with permission from Trends in Cognitive
Sciences.

Symmetry is a powerful low-level stimulus feature that has been linked to a sustained posterior negativity
(SPN) in occipital electrodes in EEG studies ( Jacobsen & Höfel, 2003; Makin et al., 2016, 2018). The mag-
nitude of the SPN is tightly linked to preference for simple patterns, suggesting that some regions of visual
cortex code for aesthetic responses to symmetry via mechanisms of sensory processing fluency (Makin et al.,
2018). In an early fMRI study, Jacobsen et al. (2006) showed that different regions of the brain are recruited
when observers make symmetry judgments vs. aesthetic judgments. Aesthetic judgements uniquely activated
frontal and prefrontal regions associated with reward processing, whilst symmetry judgments uniquely acti-
vated regions of parietal and premotor cortex. The aforementioned studies use simple dot pattern stimuli;
however, symmetry has also been shown to be a predictor of preference for more complex and ecologically
valid stimuli such as faces, flowers and landscapes (Bertamini et al., 2019; Hůla & Flegr, 2016; Perrett et al.,
1999), leading researchers to posit that symmetry is an aesthetic primitive.
In addition to symmetry, stimulus curvature has been found to influence aesthetic preference for abstract
geometric shapes, real-life objects, environments and most recently painted artworks (Bar & Neta, 2006;
Palumbo et al., 2015, 2020; Ruta et al., 2021; Vartanian et al., 2013). The origin of a preference for curva-
ture remains a debate in the literature. Some authors suggest it derives from optimal stimulation of parts of
the visual system which respond to grouping principles such as good continuation in much the same way
as the SPN for symmetry (Bertamini et al., 2016). Other researchers argue that a preference for curvature
derives from an evolutionary adaptive avoidance of sharp stimuli (Bar & Neta, 2006), evidenced by func-
tional neuroimaging studies which show bilateral activation in the amygdala when participants view sharp-
contoured objects (Bar & Neta, 2007; Larson et al., 2009). In a study on neural responses to architecture,
curvature elicited activation in visual cortex when observers were making approach/avoidance judgments
(Vartanian et al., 2013).
Differences in hue, lightness and saturation of colour have been shown to drive aesthetic responses
(Palmer & Schloss, 2010). Ou et al. (2004, 2018) theorized that colour preferences are based on semantic

302
Visual art experience

associations with particular hues, and recent empirical research has found that colour associations (e.g., yel-
low = warm) are to some extent stable across observers for abstract artworks and artwork elements (Specker
et al., 2020). Ecological valence theory (EVT; Palmer & Schloss, 2010) posits that colour preferences are
determined by the emotional valence of objects associated with those colours, developing understanding of
how colour associations arise. Studies have also revealed that colour preferences could also be linked to asso-
ciations with abstract concepts, such as an observer’s university and political affiliation (Schloss et al., 2011;
Schloss & Palmer, 2014). The neuroscientific basis of the formation of colour associations in art has not yet
been explored; however, Hurlbert and Ling (2007) suggested that cone-opponent colour processing in reti-
nal cells could explain colour preferences, linking colour preference with very early stage colour processing
in the visual system.
The role of regions of visual cortex responsible for motion detection has been studied in relation to aes-
thetic experience of artworks. Indeed, this was the topic of an early theoretical treatment of the nascent field
of neuroaesthetics (Zeki & Lamb, 1994). In an fMRI study, Zeki and Stutters (2012) found experimental
evidence for a link between preference for particular non-artistic motion trajectories and activation in area
V5 of visual cortex, with preferred motion also activating the medial OFC. In relation to visual art, activa-
tion in V5 has been shown to correlate with implied motion in a range of artworks, from manga to cubism
(Di Dio et al., 2016; Kim & Blake, 2007; Osaka et al., 2010). Area V5 not only plays a correlational role in
aesthetic preference. It has been found that disrupting activity in V5 via transcranial magnetic stimulation
(TMS) has an impact on the perception of and liking for apparent motion in abstract paintings (Cattaneo
et al., 2017). Recent work with patients with Parkinson’s disease, a degenerative disorder that affects the basal
ganglia and substantia nigra and causes problems with movement, revealed that the patients gave lower judge-
ments of apparent motion in both high-motion action paintings and low-motion neoplastic paintings, with
preference for motion exhibiting a linear relationship for patients and a non-linear relationship for controls
(Humphries et al., 2021). Such work links the perception of implied motion in art with the functioning of
motor networks in the brain of the observer.
Consideration of the role of motor cortices in relation to visual art experience began with a theory paper
by Freedberg and Gallese (2007), who put forward the embodied simulation account of art. This account
asserts that motor regions of the brain (particularly those implicated in the simulation of observed actions)
respond to actions figuratively depicted in artworks, as well as the actions made by the artist when creating
the artwork. Subsequent empirical research has supported this theory, with studies revealing facilitation of
aesthetic processing when the observer’s gestures match those of the artist (Leder et al., 2012; Taylor et al.,
2012; Ticini et al., 2014). In addition, motor-related cortical activity has been shown in response to artistic
movements in the work of Lucio Fontana (Umiltà et  al., 2012) and Franz Kline (Sbriscia-Fioretti et  al.,
2013). Finally, there is emerging evidence of simulation of depicted facial expressions in paintings. Ardizzi et
al.(2020) found that aesthetic evaluations of painful facial expressions in Renaissance artworks were higher
when the observer’s facial muscle contraction matched the expression conveyed in the painting. Further-
more, activity in regions associated with empathy for pain correlated with participants’ aesthetic judgments
of painted depictions of pain (Ardizzi et al., 2021). While the evidence for the role of action observation
in visual art experiences is beginning to mount, other authors have argued that the involvement of motor
regions instead signals approach-avoidance responses, reflecting the observer’s preparation to move toward or
away from artworks that are liked or disliked (Ishizu & Zeki, 2011; Kawabata & Zeki, 2004).
Faces and bodies feature in many works of art. As previously discussed, in a meta-analysis (Boccia et al.,
2016) that included a set of empirical functional neuroimaging studies of artworks from different stimulus
categories (portraits, scenes, sculpture and abstract art), it was found that viewing portraits activated the FFA
and inferior occipital gyrus in addition to the amygdala, while paintings of bodies activated bilateral fusiform
areas, supplementary motor regions, inferior frontal gyri, hippocampi and insula among other regions. Such
work shows content-based specialization of sensory responses to works of art. In support of a causal role

303
Rebecca Chamberlain

for body-specific brain regions in aesthetic responses to bodies, Calvo-Merino et al. (2010) applied TMS
to the ventral primary motor cortex (vPMC) and extrastriate body area (EBA) in a paradigm that measured
the consistency of observers’ aesthetic response toward body postures. As a result, they found two routes to
aesthetic processing of bodies: one for local processing via EBA, which is linked to aesthetic sensitivity, and
the other global route via vPMC. To summarize, sensory and motor regions are strongly implicated in visual
art experience, though the nature of sensory and motor activation is very much dependent on the stimulus
class and medium represented in an artwork.

Knowledge-meaning
The knowledge-meaning component of the aesthetic triad relates predominantly to the effects of exper-
tise and framing which help to shape an observer’s experience with an artwork. In terms of expertise, it
is known that artists tend to focus on different aspects of an artwork to novices. Experts focus more on
relations between objects and generally make more global eye movements that are less drawn by local sali-
ent regions of a picture (Koide et al., 2015; Nodine et al., 1993; Pihko et al., 2011; Vogt, 1999; Vogt &
Magnussen, 2007; Zangemeister et al., 1995). These differences in the experience of visual art bear out
in neuroimaging studies. Greater art expertise is associated with lower event-related potential amplitudes
(ERPs) in right occipital electrodes, suggesting an attenuated response of visual cortical regions (Pang et al.,
2013). On the other hand, an EEG study by Else et al. (2015) revealed greater amplitude of the N1 com-
ponent (linked to attention and effort) and the P2 (linked to higher-order visual processing) in a group of
artists when compared to the response of non-artists, which was particularly salient for abstract artworks.
This aligns with behavioural findings that experts tend to prefer abstract artworks relative to non-artists (van
Paasschen et al., 2015). Finally, approximate entropy (indicating the rate of production of new informa-
tion in the EEG signal) was shown to be higher in artists in frontal brain regions during artwork viewing
and artwork imagery tasks (Shourie et al., 2014). Together these results indicate that artists undergo more
elaborate cognitive processing of visual artworks compared with novices, which is reflected in differentiated
neural responses.
The effect of being in an artistic context responding to artistic content can have marked effects on neural
responses to artworks. Kirk (2008) demonstrated that responses to objects in unusual contexts engaged an
aesthetic mode of processing that correlated with activation in OFC and the inferior frontal gyrus (IFG),
which is implicated in self-referential processing and processing ambiguous stimuli. Framing a stimulus as
originating from an artistic context modulates emotional appraisal for disgusting images (Wagner et  al.,
2014) and elicits medial OFC activation relative to stimuli framed as non-art (Kirk et al., 2009). A recent
study implicated the DLPFC in artistic framing by presenting participants with fear-inducing IAPS images
and labelling them as either artistic or documentary images (Kirk et al., 2020). The authors argue that the
DLPFC is involved in reappraising the fearful stimuli when observers believe they are intended to be art-
works (Kirk et al., 2020). In a similar manner, neuroimaging research shows that the OFC is activated when
observers believe they are viewing a real artwork in contrast to a duplicate (Huang et al., 2011). Across these
studies, we can see that the OFC and other regions of the PFC are engaged when observers believe them-
selves to be experiencing visual art relative to other stimulus kinds.

Emotion-valuation
Most of our experiences with visual art are marked by both rewarding emotional states and evaluative judg-
ments. In the context of visual art, experienced emotions can range from horror to elation, and there has
been much recent discussion about the possibility of a unique class of aesthetic emotions as distinct from
emotions generated in non-aesthetic contexts (Menninghaus et al., 2019; Skov & Nadal, 2020b).

304
Visual art experience

A network of regions associated with the experience of reward, including the ventral striatum (VS),
amygdala and frontal regions including the OFC, DLPFC and anterior cingulate cortex (ACC), are acti-
vated when observers experience artworks (Boccia et al., 2016; Lacey et al., 2011; Vartanian & Skov, 2014).
The VS includes the nucleus accumbens (NA) and the caudate nucleus and as such constitutes a key node
of the reward circuitry, implicated in reward prediction (Pagnoni et al., 2002). In support of its role in the
expectation of aesthetic reward, activity in the VS has been correlated with the experience of aesthetic chills
and positive emotional responses in research on music (Blood & Zatorre, 2001; Mueller et al., 2015; Salim-
poor et al., 2013; Trost et al., 2012). In the context of visual art, the engagement of the VS does not seem
to depend on particular stimulus features, as research has shown that merely labelling an image an artwork
results in greater activity in VS relative to the same stimulus labelled as “non-art” (Lacey et al., 2011).
As we have seen from discussions of sensory-motor and framing effects, the medial OFC appears to play
a key role in visual art experiences. The OFC has reliably been shown to be involved in stimulus evaluation
and reward computation, with the anterior OFC being implicated in the processing of secondary rewards,
while the posterior OFC is involved in the processing of primary rewards (Sescousse et al., 2010). Kawabata
and Zeki (2004) performed an fMRI study in which observers made beauty judgements about paintings
from a range of different content categories (still-life, landscape, abstract) and showed that when observers
viewed paintings they regarded as beautiful, the medial OFC showed increased activation relative to trials in
which observers viewed ugly paintings, regardless of the content category of the artwork. Still many more
studies have implicated the OFC in visual art experiences, especially those that are rewarding and involve
feelings of beauty (Ishizu & Zeki, 2011, 2017; Kirk, 2008; Kirk et al., 2009). While the OFC is reliably
recruited in neuroimaging studies that require observers to make judgments on visual art, there is speculation
that it represents a domain-general response to beauty and ugliness rather than one that is specific to visual
art per se (e.g. Zeki et al., 2014). A recent meta-analysis concluded that feelings of beauty toward faces and
artworks recruited different brain regions, with facial beauty being associated with the left ventral striatum,
while visual art activated the anterior medial PFC (Chuan-Peng et al., 2020), suggesting that there is at least
some parcellation of frontal cortex response for different aesthetic stimulus categories, albeit in regions out-
side of the OFC.
The causal role of the DLPFC in visual art experience has been shown in studies which have demon-
strated modulation of aesthetic preference for visual art through electrical stimulation of frontal regions
(Cattaneo, Lega, Flexas et al., 2014; Cattaneo, Lega, Gardelli et al., 2014). Elsewhere, the DLPFC has been
shown to have a number of functions, including working memory, response selection and emotional regu-
lation (Badre & Wagner, 2004; Brunoni & Vanderhasselt, 2014; Golkar et al., 2012). As such, it is likely
that the DLPFC determines how rewards are to be acted on, whereas the OFC determines the value of the
reward in the first place. Ticini (2017) posits that OFC is more likely to be recruited in tasks in which passive
visual art viewing is required, while the DLPFC comes online when observers are required to respond to the
stimulus with an action. Finally, the anterior cingulate cortex (ACC) has been associated with the adoption
of an aesthetic stance (Boccia et al., 2016), as it serves the role of connecting cognitive and emotional systems
necessary for the evaluation of sensory information (Pessoa, 2008).
Cela-Conde et al. (2013) argue that visual art experience takes place in two stages, implicating two dif-
ferent sets of neural networks. The first involves an initial appraisal of aesthetic qualities of the artwork and
mainly recruits occipital and frontal regions, while the latter stage involves appraisal of detailed aspects of
the aesthetic experience and recruits the default mode network (DMN). The DMN consists of a system
of brain regions (the precuneus, posterior and ventral anterior cingulate cortices, medial prefrontal cortex,
bilateral inferior parietal lobules, bilateral middle temporal gyri and left middle frontal gyrus) associated with
at-rest conditions or task-related decreases in activation in fMRI studies (Laird et al., 2009). Activation of the
DMN reflects a range of cognitive functions, including emotional and social cognition, introspection and
autobiographical memory retrieval (Laird et al., 2009). Vessel et al. (2012) were the first group of researchers

305
Rebecca Chamberlain

to implicate the DMN directly in visual art experience. In an fMRI study, participants were shown artworks
and asked to rate them on 1–4 scale based on how much the painting moved them, a gut-level response on
how “beautiful, compelling or powerful” the painting appeared to them. The most moving artworks (rated
4 on the scale) were associated with increased activation in the DMN, suggesting a qualitative difference in
response to the most moving artworks and those rated lower (1–3) on the scale. Vessel et al. (2012) concluded
that DMN activation signals the presence of a response which goes beyond mere liking: an experience that is
highly personally relevant and profound. In this context, artworks resonate strongly with an individual’s sense
of self. Vessel and colleagues link this result to earlier findings by Cupchik et al. (2009), who found a dif-
ference in brain activation in relation to taking a pragmatic or aesthetic mode toward art experience, which
corresponded with brain regions implicated in Vessel et al.’s (2012) study. In a follow-up study, Belfi et al.
(2019) found that for very pleasing art, activity in the DMN counteracted a suppressive effect that increased
with stimulus presentation. In addition, DMN activity returned to baseline as a function of stimulus offset
for very pleasing stimuli, indicating that activation of the DMN was tied to the artwork experience. Such
results, the authors argued, indicate that the DMN tracks the internal state of the observer during an experi-
ence with visual art (Belfi et al., 2019).
To summarize, for reward and evaluative regions including evolutionarily older regions (VS and NA)
and regions computing likely rewards, their value and the best actions to engage as a result of such rewards
(DLPFC and OFC) work together to shape our experiences with visual art. More recently, the DMN has
been implicated in moving experiences with art, connecting our sensorial experiences with inner self-
reflective mechanisms. The specificity of such experiences to visual art as a qualitatively separate stimulus
class remains an active debate in the field.

Looking to the future of the neuroscience of visual art experience


This review has hoped to capture the growing amount of empirical and theoretical work that elaborates
on psychological and neuroscientific aspects of visual art experience. We have seen how models of visual
art experience can be applied to our understanding of how brain mechanisms modulate these experiences
(Chatterjee & Vartanian, 2014; Pelowski et al., 2017). The outcome is that visual art experience is under-
pinned by a complex system of sensory, motor, memory and reward-based brain networks. However, there is
much that we still do not understand about visual art experience. To return to some of the questions posed
by Chatterjee and Vartanian (2014) in their review of neuroaesthetics, an emerging focus of research in the
field is the role of individual differences in shaping experiences with art. Degree of agreement between
observers in terms of stimulus preference has been shown to be highest for natural kinds like faces and land-
scapes and lowest for visual art (Vessel, 2010; Vessel et al., 2018), suggesting that there is a large amount of
between-subject variance in visual art experience that is yet to be accounted for. Chatterjee and Vartanian
(2014) point specifically toward gender and expertise as potentially important individual differences which
have received relatively little research attention to date. We now have robust evidence to suggest that women
are more likely to report feeling chills as a result of aesthetic experiences than men (Bignardi et al., 2021), but
the mechanisms for such an effect are as yet unexplored. We also have growing evidence that artistic expertise
modulates aesthetic judgments, including recruitment of regions such as the OFC in the service of aesthetic
experiences (Kirk et al., 2009). Explanatory accounts for expertise effects are not fully developed yet, but
more complex models such as the VIMAP (Pelowski et al., 2017) may be able to account for differential
experiences and preferences in experts compared with non-experts. Furthermore, recent work on aesthetic
sensitivity (Corradi et al., 2019, 2020) seeks to model individual observers’ tendencies to respond to certain
stimulus properties. However, the etiology of aesthetic sensitivity itself remains unknown. Aesthetic sensitiv-
ity for contour and symmetry was found to be weakly predicted by expertise but not by personality factors
such as openness to experience, suggesting that there is a large amount of sensitivity variation unaccounted

306
Visual art experience

for (Corradi et al., 2020). Whilst current work on aesthetic sensitivity is focused on the degree to which
stimulus factors (e.g., contour, symmetry) influence aesthetic preference, aesthetic sensitivity may also extend
to contextual factors like artistic framing, especially in the context of expertise, with experts potentially more
attuned to framing effects (Verpooten & Dewitte, 2017).
A fundamental question that is also addressed by Chatterjee and Vartanian (2014) is whether different
neural mechanisms play different roles in the generation of aesthetic rewards relative to rewards elicited by
other kinds of stimuli. The uniqueness of aesthetic experience and its distinction from other sensory-based,
rewarding experiences remains a highly contentious issue in the field (Skov & Nadal, 2020a). In their meta-
analysis of 93 neuroimaging studies of aesthetic appraisal, Brown et  al. (2011) concluded that aesthetic
processing represents no more than appraisal of the valence of objects. Similarly, recent research has demon-
strated overlap between brain regions that are implicated in social and moral judgements and those involved
in aesthetic appraisal (Workman et al., 2021), suggesting that judgements for aesthetic stimuli are not qualita-
tively different from judgement for non-aesthetic stimuli. There are yet more questions regarding the precise
role of neural mechanisms such as the DMN and sensory cortices in visual art experience that have yet to be
answered. Research in this review indicates that in some contexts, the functioning of sensory and motor cor-
tices are not just concurrent with but necessary for aesthetic appraisal (Calvo-Merino et al., 2010; Cattaneo
et al., 2017; Humphries et al., 2021). In addition, recent work involving the manipulation of somatosensory
information and its effects on aesthetic pleasure suggests that we may need to further investigate the role
bodily states play in governing our reactions to works of art (Haar et al., 2020). Such insights may in turn
shed new light on the evolutionary basis of our appreciation and creation of art (Panksepp, 1995). Bignardi
et al. (2021), suggest that variation in the experience of aesthetic chills is indeed linked to variation in the
genome, but very little is known about how variation at the genotypic level relates to phenotypic variation
in aesthetic experience. There is some indication that enhanced connectivity between sensory salience and
the DMN could be at play in individuals more prone to intense aesthetic experiences (Williams et al., 2018),
but the causal factors involved remain undetermined.
Finally, an emerging field of research focuses on the applications of visual art experience for well-being
and clinical therapeutic purposes (Gallo et al., 2021; Mastandrea et al., 2019). Research shows that taking
into account confounding factors such as age, socioeconomic status and gender, consuming art has a tangible
impact on self-reported well-being (Fancourt & Steptoe, 2019; Mak & Fancourt, 2019; Wang et al., 2020).
Whilst the neuroscientific underpinnings of such effects have yet to be explored, plausible mechanisms func-
tion via DMN engagement and connectivity (Luo et al., 2016; Shi et al., 2018; Williams et al., 2018) and
reduction in stress related hormones and associated physiological changes (Kaimal et al., 2016; Mastandrea
et al., 2019). Researchers are also beginning to elaborate on more esoteric emotional states in relation to
aesthetic experience, for example, in response to recreational horror (Andersen et al., 2020). Such emotional
states may have distinct representations in the brain. Ishizu and Zeki (2017) demonstrated that artistic images
rated as beautiful but associated with different emotional valence (sorrow vs. joy) both activated the medial
OFC but additionally activated other brain networks associated with specific emotional valence. Aside from
this example, little research on non-typical emotions has yet been directed toward visual art, but existing
research (especially in context of music; e.g., Juslin, 2013) can provide useful theories and frameworks to
enable us to understand what happens when we experience a work of art that saddens or even scares us.
To address these many outstanding questions concerning visual art experience, the field requires novel
methodological approaches and technologically advanced tools. With advanced tools for online testing,
researchers have access to much larger and diverse samples, crucial for studying individual differences.
Tools such as hyperscanning provide a paradigm for studying synchronicity between observers and between
observer and artist (Dumas et al., 2011; Liu et al., 2021), enabling researchers to capture the social aspects of
visual art experience. With the increased portability of functional neuroimaging, ecological validity is com-
ing to play a more fundamental role, with researchers focused on the role of authenticity in the visual art

307
Rebecca Chamberlain

experience (Specker et al., 2021) and neuroimaging research being conducted in situ in art galleries to cap-
ture the many contextual aspects which shape the gallery experience (Tröndle et al., 2014). Such techniques
will more fully explore the social and environmental features that are so crucial to the visual art experience.
To conclude, this chapter has surveyed psychological and neuroscientific modelling of experiences with
visual art. From Fechner to the present day, researchers have attempted to explain how features of the art-
work, observer and context contribute to our experiences with art. We can see that no one account captures
all aspects of visual art experience, an understandable limitation given the ever-changing nature of artworks
and the artworld within which they sit. However, inspired by Dennis Dutton’s (2009) approach to a defini-
tion of art which focuses on indisputable examples as a means to locating the centre of this phenomenon,
we can confidently describe psychological and neural features that are central to visual art experiences.
Central to our experience of art are sensory mechanisms for unpicking the form and meaning of an artwork,
with links into reward mechanisms that elicit a variety of emotional responses and behavioural motivations.
Wrapped around these sensory and reward mechanisms are expectations and modes of experiencing that are
shaped by the past experience of the observer as well as the artistic context. There is no doubt that in the
coming years, our understanding of these central mechanisms will become more nuanced, provide deeper
explanations and have measurable impacts on the evolving cultural landscape.

References
Andersen, M. M., Schjoedt, U., Price, H., Rosas, F. E., Scrivner, C., & Clasen, M. (2020). Playing with fear: A field
study in recreational horror. Psychological Science, 31(12), 1497–1510. https://doi.org/10.1177/0956797620972116
Ardizzi, M., Ferroni, F., Siri, F., Umiltà, M. A., Cotti, A., Calbi, M., Fadda, E., Freedberg, D., & Gallese, V. (2020).
Beholders’ sensorimotor engagement enhances aesthetic rating of pictorial facial expressions of pain. Psychological
Research, 84(2), 370–379. https://doi.org/10.1007/s00426-018-1067-7
Ardizzi, M., Ferroni, F., Umiltà, M. A., Pinardi, C., Errante, A., Ferri, F., Fadda, E., & Gallese, V. (2021). Visceromo-
tor roots of aesthetic evaluation of pain in art: An fMRI study. Social Cognitive and Affective Neuroscience, 16(11),
1113–1122. https://doi.org/10.1093/scan/nsab066
Arnheim, R. (1965). Art and visual perception: A psychology of the creative eye. Faber & Faber.
Badre, D., & Wagner, A. D. (2004). Selection, integration, and conflict monitoring: Assessing the nature and generality of
prefrontal cognitive control mechanisms. Neuron, 41(3), 473–487. https://doi.org/10.1016/s0896-6273(03)00851-1
Bar, M.,  & Neta, M. (2006). Humans prefer curved visual objects. Psychological Science, 17(8), 645–648. https://doi.
org/10.1111/j.1467-9280.2006.01759.x
Bar, M., & Neta, M. (2007). Visual elements of subjective preference modulate amygdala activation. Neuropsychologia,
45(10), 2191–2200. https://doi.org/10.1016/j.neuropsychologia.2007.03.008
Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., Pelli, D. G., & Starr, G. G. (2019).
Dynamics of aesthetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.
org/10.1016/j.neuroimage.2018.12.017
Berlyne, D. E. (1974). Studies in the new experimental aesthetics: Steps towards an objective psychology of aesthetic appreciation.
Hemisphere.
Bertamini, M., Palumbo, L., Gheorghes, T. N., & Galatsidas, M. (2016). Do observers like curvature or do they dislike
angularity? British Journal of Psychology, 107(1), 154–178. https://doi.org/10.1111/bjop.12132
Bertamini, M., Rampone, G., Makin, A. D. J., & Jessop, A. (2019). Symmetry preference in shapes, faces, flowers and
landscapes. PeerJ, 7, e7078. https://doi.org/10.7717/peerj.7078
Bignardi, G., Chamberlain, R., Kevenaar, S. T., Tamimy, Z., & Boomsma, D. I. (2021). On the etiology of aesthetic
chills: A behavioral genetic study. Frontiers in Neuroscience, 16.
Birkhoff, G. D. (1933). Aesthetic measure. Harvard University Press.
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences, 98(20), 11818–11823. https://doi.
org/10.1073/pnas.191355898
Boccia, M., Barbetti, S., Piccardi, L., Guariglia, C., Ferlazzo, F., Giannini, A. M., & Zaidel, D. W. (2016). Where does
brain neural activation in aesthetic responses to visual art occur? Meta-analytic evidence from neuroimaging studies.
Neuroscience and Biobehavioral Reviews, 60, 65–71. https://doi.org/10.1016/j.neubiorev.2015.09.009

308
Visual art experience

Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. NeuroImage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Brunoni, A. R., & Vanderhasselt, M. A. (2014). Working memory improvement with non-invasive brain stimulation of
the dorsolateral prefrontal cortex: A systematic review and meta-analysis. Brain and Cognition, 86, 1–9. https://doi.
org/10.1016/j.bandc.2014.01.008
Calvo-Merino, B., Urgesi, C., Orgs, G., Aglioti, S. M., & Haggard, P. (2010). Extrastriate body area underlies aesthetic
evaluation of body stimuli. Experimental Brain Research, 204(3), 447–456. https://doi.org/10.1007/s00221-010-2283-6
Cattaneo, Z., Lega, C., Flexas, A., Nadal, M., Munar, E.,  & Cela-Conde, C. J. (2014). The world can look better:
Enhancing beauty experience with brain stimulation. Social Cognitive and Affective Neuroscience, 9(11), 1713–1721.
https://doi.org/10.1093/scan/nst165
Cattaneo, Z., Lega, C., Gardelli, C., Merabet, L. B., Cela-Conde, C. J., & Nadal, M. (2014). The role of prefrontal and
parietal cortices in esthetic appreciation of representational and abstract art: A TMS study. NeuroImage, 99, 443–450.
https://doi.org/10.1016/j.neuroimage.2014.05.037
Cattaneo, Z., Schiavi, S., Silvanto, J., & Nadal, M. (2017). A TMS study on the contribution of visual area V5 to the
perception of implied motion in art and its appreciation. Cognitive Neuroscience, 8(1), 59–68. https://doi.org/10.108
0/17588928.2015.1083968
Cela-Conde, C. J., García-Prieto, J., Ramasco, J. J., Mirasso, C. R., Bajo, R., Munar, E., Flexas, A., del-Pozo, F., &
Maestú, F. (2013). Dynamics of brain networks in the aesthetic appreciation. Proceedings of the National Academy of Sci-
ences, 110(Suppl_2), 10454–10461. https://doi.org/10.1073/pnas.1302855110
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Chuan-Peng, H., Huang, Y., Eickhoff, S. B., Peng, K.,  & Sui, J. (2020). Seeking the “beauty center” in the brain:
A meta-analysis of fMRI studies of beautiful human faces and visual art. Cognitive, Affective and Behavioral Neuroscience,
20(6), 1200–1215. https://doi.org/10.3758/s13415-020-00827-z
Corradi, G., Belman, M., Currò, T., Chuquichambi, E. G., Rey, C., & Nadal, M. (2019). Aesthetic sensitivity to cur-
vature in real objects and abstract designs. Acta Psychologica, 197, 124–130. https://doi.org/10.1016/j.actpsy.2019.
05.012
Corradi, G., Chuquichambi, E. G., Barrada, J. R., Clemente, A., & Nadal, M. (2020). A new conception of visual aes-
thetic sensitivity. British Journal of Psychology, 111(4), 630–658. https://doi.org/10.1111/bjop.12427
Cupchik, G. C. (2007). A critical reflection on Arnheim’s Gestalt theory of aesthetics. Psychology of Aesthetics, Creativity,
and the Arts, 1(1), 16–24. https://doi.org/10.1037/1931-3896.1.1.16
Cupchik, G. C., Vartanian, O., Crawley, A., & Mikulis, D. J. (2009). Viewing artworks: Contributions of cognitive con-
trol and perceptual facilitation to aesthetic experience. Brain and Cognition, 70(1), 84–91. https://doi.org/10.1016/j.
bandc.2009.01.003
Cutting, J. E. (2003). Gustave Caillebotte, French Impressionism, and mere exposure. Psychonomic Bulletin and Review,
10(2), 319–343. https://doi.org/10.3758/BF03196493
Di Dio, C., Ardizzi, M., Massaro, D., Di Cesare, G., Gilli, G., Marchetti, A.,  & Gallese, V. (2016). Human, nature,
dynamism: The effects of content and movement perception on brain activations during the aesthetic judgment of
representational paintings. Frontiers in Human Neuroscience, 9, 705. https://doi.org/10.3389/fnhum.2015.00705
Dijkstra, K., & van Dongen, N. N. N. (2017). Moderate contrast in the evaluation of paintings is liked more but remem-
bered less than high contrast. Frontiers in Psychology, 8, 1507. https://doi.org/10.3389/fpsyg.2017.01507
Dumas, G., Lachat, F., Martinerie, J., Nadel, J., & George, N. (2011). From social behaviour to brain synchronization:
Review and perspectives in hyperscanning. IRBM, 32(1), 48–53. https://doi.org/10.1016/j.irbm.2011.01.002
Dutton, D. (2009). The art instinct: Beauty, pleasure, & human evolution. Oxford University Press.
Else, J. E., Ellis, J., & Orme, E. (2015). Art expertise modulates the emotional response to modern art, especially abstract:
An ERP investigation. Frontiers in Human Neuroscience, 9, 525. https://doi.org/10.3389/fnhum.2015.00525
Eysenck, H. (1940). The general factor in aesthetic judgements. British Journal of Psychology, 31, 94–102.
Fancourt, D., & Steptoe, A. (2019). Cultural engagement and mental health: Does socio-economic status explain the
association? Social Science and Medicine, 236, 112425. https://doi.org/10.1016/j.socscimed.2019.112425
Fechner, G. (1876). Vorschule der aesthetik (Vol. 1). Brietkopf & Härtel.
Flavell, J. C., Over, H., & Tipper, S. P. (2020). Competing for affection: Perceptual fluency and ambiguity solution. Journal
of Experimental Psychology: Human Perception and Performance, 46(3), 231–240. https://doi.org/10.1037/xhp0000702
Freedberg, D., & Gallese, V. (2007). Motion, emotion and empathy in esthetic experience. Trends in Cognitive Sciences,
11(5), 197–203. https://doi.org/10.1016/j.tics.2007.02.003
Gallo, L. M. H., Giampietro, V., Zunszain, P. A., & Tan, K. S. (2021). Covid-19 and mental health: Could visual art
exposure help? Frontiers in Psychology, 12, 650314. https://doi.org/10.3389/fpsyg.2021.650314

309
Rebecca Chamberlain

Golkar, A., Lonsdorf, T. B., Olsson, A., Lindstrom, K. M., Berrebi, J., Fransson, P., Schalling, M., Ingvar, M., & Öhman,
A. (2012). Distinct contributions of the dorsolateral prefrontal and orbitofrontal cortex during emotion regulation.
PLoS One, 7(11), e48107. https://doi.org/10.1371/journal.pone.0048107
Haar, A. J. H., Jain, A., Schoeller, F., & Maes, P. (2020). Augmenting aesthetic chills using a wearable prosthesis improves
their downstream effects on reward and social cognition. Scientific Reports, 10(1), 21603. https://doi.org/10.1038/
s41598-020-77951-w
Hekkert, P. (2006). Design aesthetics: Principles of pleasure in design. Psychology Science, 16.
Huang, M., Bridge, H., Kemp, M. J., & Parker, A. J. (2011). Human cortical activity evoked by the assignment of authen-
ticity when viewing works of art. Frontiers in Human Neuroscience, 5, 134. https://doi.org/10.3389/fnhum.2011.
00134
Hůla, M., & Flegr, J. (2016). What flowers do we like? The influence of shape and color on the rating of flower beauty.
PeerJ, 4, e2106. https://doi.org/10.7717/peerj.2106
Humphries, S., Rick, J., Weintraub, D., & Chatterjee, A. (2021). Movement in aesthetic experiences: What we can learn
from Parkinson disease. Journal of Cognitive Neuroscience, 33(7), 1329–1342. https://doi.org/10.1162/jocn_a_01718
Hurlbert, A. C., & Ling, Y. (2007). Biological components of sex differences in color preference. Current Biology, 17(16),
R623–R625. https://doi.org/10.1016/j.cub.2007.06.022
Ishizu, T., & Zeki, S. (2011). Toward a brain-based theory of beauty. PLoS One, 6(7), e21852. https://doi.org/10.1371/
journal.pone.0021852
Ishizu, T., & Zeki, S. (2017). The experience of beauty derived from sorrow. Human Brain Mapping, 38(8), 4185–4200.
https://doi.org/10.1002/hbm.23657
Jacobsen, T.,  & Höfel, L. (2003). Descriptive and evaluative judgment processes: Behavioral and electrophysiological
indices of processing symmetry and aesthetics. Cognitive, Affective and Behavioral Neuroscience, 3(4), 289–299. https://
doi.org/10.3758/CABN.3.4.289
Jacobsen, T., Schubotz, R. I., Höfel, L., & Cramon, D. Y. (2006). Brain correlates of aesthetic judgment of beauty. Neu-
roImage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010
Juslin, P. N. (2013). From everyday emotions to aesthetic emotions: Towards a unified theory of musical emotions. Physics
of Life Reviews, 10(3), 235–266. https://doi.org/10.1016/j.plrev.2013.05.008
Kaimal, G., Ray, K., & Muniz, J. (2016). Reduction of cortisol levels and participants’ responses following art making.
Art Therapy, 33(2), 74–80. https://doi.org/10.1080/07421656.2016.1166832
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Kim, C. Y., & Blake, R. (2007). Brain activity accompanying perception of implied motion in abstract paintings. Spatial
Vision, 20(6), 545–560. https://doi.org/10.1163/156856807782758395
Kirk, U. (2008). The neural basis of object-context relationships on aesthetic judgment. PLoS One, 3(11), e3754. https://
doi.org/10.1371/journal.pone.0003754
Kirk, U., Lilleholt, L., & Freedberg, D. (2020). Cognitive framing modulates emotional processing through dorsolateral
prefrontal cortex and ventrolateral prefrontal cortex networks: A functional magnetic resonance imaging study. Brain
and Behavior, 10(9), e01761. https://doi.org/10.1002/brb3.1761
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009). Modulation of aesthetic value by semantic con-
text: An fMRI study. NeuroImage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009
Koffka, K. (1935). Principles of Gestalt psychology (5th ed.). Routledge and Kegan Paul.
Koide, N., Kubo, T., Nishida, S., Shibata, T., & Ikeda, K. (2015). Art expertise reduces influence of visual salience on
fixation in viewing abstract-paintings. PLoS One, 10(2), e0117696. https://doi.org/10.1371/journal.pone.0117696
Lacey, S., Hagtvedt, H., Patrick, V. M., Anderson, A., Stilla, R., Deshpande, G., Hu, X., Sato, J. R., Reddy, S.,  &
Sathian, K. (2011). Art for reward’s sake: Visual art recruits the ventral striatum. NeuroImage, 55(1), 420–433. https://
doi.org/10.1016/j.neuroimage.2010.11.027
Laird, A. R., Eickhoff, S. B., Li, K., Robin, D. A., Glahn, D. C., & Fox, P. T. (2009). Investigating the functional hetero-
geneity of the default mode network using coordinate-based meta-analytic modeling. Journal of Neuroscience, 29(46),
14496–14505. https://doi.org/10.1523/JNEUROSCI.4004-09.2009
Larson, C. L., Aronoff, J., Sarinopoulos, I. C., & Zhu, D. C. (2009). Recognizing threat: A simple geometric shape acti-
vates neural circuitry for threat detection. Journal of Cognitive Neuroscience, 21(8), 1523–1535. https://doi.org/10.1162/
jocn.2009.21111
Leder, H., Bär, S., & Topolinski, S. (2012). Covert painting simulations influence aesthetic appreciation of artworks.
Psychological Science, 23(12), 1479–1481. https://doi.org/10.1177/0956797612452866
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811

310
Visual art experience

Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Liu, T., Duan, L., Dai, R., Pelowski, M., & Zhu, C. (2021). Team-work, Team-brain: Exploring synchrony and team
interdependence in a nine-person drumming task via multiparticipant hyperscanning and inter-brain network topol-
ogy with fNIRS. NeuroImage, 237, 118147. https://doi.org/10.1016/j.neuroimage.2021.118147
Locher, P. (2015). The aesthetic experience with visual art “at first glance”. In P. F. Bundgard & F. Sjternfelt (Eds.),
Investigations into the phenomenology and the ontology of the work of art: What are artworks and how do we experience them?
(pp. 75–88). Springer.
Locher, P., Krupinski, E. A., Mello-Thoms, C., & Nodine, C. F. (2007). Visual interest in pictorial art during an aesthetic
experience. Spatial Vision, 21(1–2), 55–77. https://doi.org/10.1163/156856807782753868
Locher, P., Overbeeke, K.,  & Wensveen, S. (2010). Aesthetic interaction: A  framework. Design Issues, 26(2), 70–79.
https://doi.org/10.1162/DESI_a_00017
Luo, Y., Kong, F., Qi, S., You, X., & Huang, X. (2016). Resting-state functional connectivity of the default mode net-
work associated with happiness. Social Cognitive and Affective Neuroscience, 11(3), 516–524. https://doi.org/10.1093/
scan/nsv132
Mak, H. W., & Fancourt, D. (2019). Longitudinal associations between ability in arts activities, behavioural difficulties
and self-esteem: Analyses from the 1970 British cohort study. Scientific Reports, 9(1), 14236. https://doi.org/10.1038/
s41598-019-49847-x
Makin, A. D. J., Helmy, M., & Bertamini, M. (2018). Visual cortex activation predicts visual preference: Evidence from
Britain and Egypt. Quarterly Journal of Experimental Psychology, 71(8), 1771–1780. https://doi.org/10.1080/1747021
8.2017.1350870
Makin, A. D. J., Wright, D., Rampone, G., Palumbo, L., Guest, M., Sheehan, R., Cleaver, H., & Bertamini, M. (2016).
An electrophysiological index of perceptual goodness. Cerebral Cortex, 26(12), 4416–4434. https://doi.org/10.1093/
cercor/bhw255
Mastandrea, S., Fagioli, S., & Biasi, V. (2019). Art and psychological well-being: Linking the brain to the aesthetic emo-
tion. Frontiers in Psychology, 10, 739. https://doi.org/10.3389/fpsyg.2019.00739
Menninghaus, W., Wagner, V., Hanich, J., Wassiliwizky, E., Jacobsen, T., & Koelsch, S. (2017). The distancing-embracing
model of the enjoyment of negative emotions in art reception. Behavioral and Brain Sciences, 40, e347. https://doi.
org/10.1017/S0140525X17000309
Menninghaus, W., Wagner, V., Wassiliwizky, E., Schindler, I., Hanich, J., Jacobsen, T., & Koelsch, S. (2019). What are
aesthetic emotions? Psychological Review, 126(2), 171–195. https://doi.org/10.1037/rev0000135
Mueller, K., Fritz, T., Mildner, T., Richter, M., Schulze, K., Lepsien, J., Schroeter, M. L., & Möller, H. E. (2015).
Investigating the dynamics of the brain response to music: A central role of the ventral striatum/nucleus accumbens.
NeuroImage, 116, 68–79. https://doi.org/10.1016/j.neuroimage.2015.05.006
Muth, C., & Carbon, C. C. (2013). The aesthetic aha: On the pleasure of having insights into Gestalt. Acta Psychologica,
144(1), 25–30. https://doi.org/10.1016/j.actpsy.2013.05.001
Muth, C., Pepperell, R., & Carbon, C.-C. (2013). Give me gestalt! Preference for cubist artworks revealing high detect-
ability of objects. Leonardo, 46(5), 488–489. https://doi.org/10.1162/LEON_a_00649
Nodine, C. F., Locher, P. J., & Krupinski, E. A. (1993). The role of formal art training on perception and aesthetic judg-
ment of art compositions. Leonardo, 26(3), 219. https://doi.org/10.2307/1575815
Osaka, N., Matsuyoshi, D., Ikeda, T., & Osaka, M. (2010). Implied motion because of instability in Hokusai manga acti-
vates the human motion-sensitive extrastriate visual cortex: An fMRI study of the impact of visual art. NeuroReport,
21(4), 264–267. https://doi.org/10.1097/WNR.0b013e328335b371
Ou, L.-C., Luo, M. R., Woodcock, A.,  & Wright, A. (2004). A  study of colour emotion and colour preference.
Part I: Colour emotions for single colours. Color Research and Application, 29(3), 232–240. https://doi.org/10.1002/
col.20010
Ou, L. C., Yuan, Y., Sato, T., Lee, W.-Y., Szabó, F., Sueeprasan, S., & Huertas, R. (2018). Universal models of colour
emotion and colour harmony. Color Research and Application, 43(5), 736–748. https://doi.org/10.1002/col.22243
Pagnoni, G., Zink, C. F., Montague, P. R., & Berns, G. S. (2002). Activity in human ventral striatum locked to errors of
reward prediction. Nature Neuroscience, 5(2), 97–98. https://doi.org/10.1038/nn802
Palmer, S. E., & Schloss, K. B. (2010). An ecological valence theory of human color preference. Proceedings of the National
Academy of Sciences, 107(19), 8877–8882. https://doi.org/10.1073/pnas.0906172107
Palumbo, L., Rampone, G., Bertamini, M., Sinico, M., Clarke, E., & Vartanian, O. (2020). Visual preference for abstract
curvature and for interior spaces: Beyond undergraduate student samples. Psychology of Aesthetics, Creativity, and the
Arts. https://doi.org/10.1037/aca0000359

311
Rebecca Chamberlain

Palumbo, L., Ruta, N., & Bertamini, M. (2015). Comparing angular and curved shapes in terms of implicit associations
and approach/avoidance responses. PLoS One, 10(10), e0140043. https://doi.org/10.1371/journal.pone.0140043
Pang, C. Y., Nadal, M., Müller-Paul, J. S., Rosenberg, R., & Klein, C. (2013). Electrophysiological correlates of looking
at paintings and its association with art expertise. Biological Psychology, 93(1), 246–254. https://doi.org/10.1016/j.
biopsycho.2012.10.013
Panksepp, J. (1995). The emotional sources of “chills” induced by music. Music Perception, 13(2), 171–207. https://doi.
org/10.2307/40285693
Pelowski, M., & Akiba, F. (2011). A model of art perception, evaluation and emotion in transformative aesthetic experi-
ence. New Ideas in Psychology, 29(2), 80–97. https://doi.org/10.1016/j.newideapsych.2010.04.001
Pelowski, M., Markey, P. S., Forster, M., Gerger, G., & Leder, H. (2017). Move me, astonish me . . . delight my eyes
and brain: The Vienna integrated model of top-down and bottom-up processes in art perception (VIMAP) and cor-
responding affective, evaluative, and neurophysiological correlates. Physics of Life Reviews, 21, 80–125. https://doi.
org/10.1016/j.plrev.2017.02.003
Perrett, D. I., Burt, D. M., Penton-Voak, I. S., Lee, K. J., Rowland, D. A., & Edwards, R. (1999). Symmetry and human
facial attractiveness. Evolution and Human Behavior, 20(5), 295–307. https://doi.org/10.1016/S1090-5138(99)00014-8
Pessoa, L. (2008). On the relationship between emotion and cognition. Nature Reviews Neuroscience, 9(2), 148–158.
https://doi.org/10.1038/nrn2317
Pihko, E., Virtanen, A., Saarinen, V. M., Pannasch, S., Hirvenkari, L., Tossavainen, T., Haapala, A., & Hari, R. (2011).
Experiencing art: The influence of expertise and painting abstraction level. Frontiers in Human Neuroscience, 5, 94.
https://doi.org/10.3389/fnhum.2011.00094
Ramachandran, V. S., & Hirstein, W. (1999). The science of art: A neurological theory of aesthetic experience. Journal
of Consciousness Studies, 6(6–7), 15–51.
Reber, R., Schwarz, N.,  & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the per-
ceiver’s processing experience? Personality and Social Psychology Review, 8(4), 364–382. https://doi.org/10.1207/
s15327957pspr0804_3
Ruta, N., Vañó, J., Pepperell, R., Corradi, G. B., Chuquichambi, E. G., Rey, C.,  & Munar, E. (2021). Preference
for paintings is also affected by curvature. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/
aca0000395
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059
Sbriscia-Fioretti, B., Berchio, C., Freedberg, D., Gallese, V., & Umiltà, M. A. (2013). ERP modulation during observa-
tion of abstract paintings by Franz Kline. PLoS One, 8(10), e75241. https://doi.org/10.1371/journal.pone.0075241
Scherer, K. R. (2001). Appraisal: The evolution of an idea. In K. R. Scherer, A. Schorr & T. Johnstone (Eds.), Appraisal
processes in emotion: Theory, methods, research (pp. 20–34). Oxford University Press.
Schloss, K. B., & Palmer, S. E. (2014). The politics of color: Preferences for Republican red versus Democratic blue.
Psychonomic Bulletin and Review, 21(6), 1481–1488. https://doi.org/10.3758/s13423-014-0635-0
Schloss, K. B., Poggesi, R. M.,  & Palmer, S. E. (2011). Effects of university affiliation and “school spirit” on color
preferences: Berkeley versus Stanford. Psychonomic Bulletin and Review, 18(3), 498–504. https://doi.org/10.3758/
s13423-011-0073-1
Sescousse, G., Redouté, J., & Dreher, J. C. (2010). The architecture of reward value coding in the human orbitofrontal
cortex. Journal of Neuroscience, 30(39), 13095–13104. https://doi.org/10.1523/JNEUROSCI.3501-10.2010
Shi, L., Sun, J., Wu, X., Wei, D., Chen, Q., Yang, W., Chen, H., & Qiu, J. (2018). Brain networks of happiness: Dynamic
functional connectivity among the default, cognitive and salience networks relates to subjective well-being. Social
Cognitive and Affective Neuroscience, 13(8), 851–862. https://doi.org/10.1093/scan/nsy059
Shourie, N., Firoozabadi, M., & Badie, K. (2014). Analysis of EEG signals related to artists and nonartists during visual
perception, mental imagery, and rest using approximate entropy. BioMed Research International, 2014, 764382. https://
doi.org/10.1155/2014/764382
Silvia, P. J. (2005). Cognitive appraisals and interest in visual art: Exploring an appraisal theory of aesthetic emotions.
Empirical Studies of the Arts, 23(2), 119–133. https://doi.org/10.2190/12AV-AH2P-MCEH-289E
Skov, M., & Nadal, M. (2020a). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Skov, M., & Nadal, M. (2020b). There are no aesthetic emotions: Comment on Menninghaus et al. (2019). Psychological
Review, 127(4), 640–649. https://doi.org/10.1037/rev0000187
Specker, E., Fekete, A., Trupp, M. D., & Leder, H. (2021). Is a “real” artwork better than a reproduction? A meta-analysis
of the genuineness effect. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/aca0000399

312
Visual art experience

Specker, E., Forster, M., Brinkmann, H., Boddy, J., Immelmann, B., Goller, J., Pelowski, M., Rosenberg, R., & Leder,
H. (2020). Warm, lively, rough? Assessing agreement on aesthetic effects of artworks. PLoS One, 15(5), e0232083.
https://doi.org/10.1371/journal.pone.0232083
Taylor, J. E. J., Witt, J. K., & Grimaldi, P. J. (2012). Uncovering the connection between artist and audience: Viewing
painted brushstrokes evokes corresponding action representations in the observer. Cognition, 125(1), 26–36. https://
doi.org/10.1016/j.cognition.2012.06.012
Ticini, L. F. (2017). The role of the orbitofrontal and dorsolateral prefrontal cortices in aesthetic preference for art. Behav-
ioral Sciences, 7(2), 31. https://doi.org/10.3390/bs7020031
Ticini, L. F., Rachman, L., Pelletier, J.,  & Dubal, S. (2014). Enhancing aesthetic appreciation by priming canvases
with actions that match the artist’s painting style. Frontiers in Human Neuroscience, 8, 391. https://doi.org/10.3389/
fnhum.2014.00391
Tinio, P. P. L. (2013). From artistic creation to aesthetic reception: The mirror model of art. Psychology of Aesthetics, Crea-
tivity, and the Arts, 7(3), 265–275. https://doi.org/10.1037/a0030872
Tröndle, M., Greenwood, S., Kirchberg, V., & Tschacher, W. (2014). An integrative and comprehensive methodology for
studying aesthetic experience in the field: Merging movement tracking, physiology, and psychological data. Environ-
ment and Behavior, 46(1), 102–135. https://doi.org/10.1177/0013916512453839
Trost, W., Ethofer, T., Zentner, M., & Vuilleumier, P. (2012). Mapping aesthetic musical emotions in the brain. Cerebral
Cortex, 22(12), 2769–2783. https://doi.org/10.1093/cercor/bhr353
Umiltà, M. A., Berchio, C., Sestito, M., Freedberg, D., & Gallese, V. (2012). Abstract art and cortical motor activation:
An EEG study. Frontiers in Human Neuroscience, 6, 311. https://doi.org/10.3389/fnhum.2012.00311
Van Geert, E., & Wagemans, J. (2020). Order, complexity, and aesthetic appreciation. Psychology of Aesthetics, Creativity,
and the Arts, 14(2), 135–154. https://doi.org/10.1037/aca0000224
van Paasschen, J., Bacci, F., & Melcher, D. P. (2015). The influence of art expertise and training on emotion and prefer-
ence ratings for representational and abstract artworks. PLoS One, 10(8), e0134241. https://doi.org/10.1371/journal.
pone.0134241
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov, M.
(2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceedings of the
National Academy of Sciences, 110(Suppl_2), 10446–10453. https://doi.org/10.1073/pnas.1301227110
Vartanian, O., & Skov, M. (2014). Neural correlates of viewing paintings: Evidence from a quantitative meta-analysis
of functional magnetic resonance imaging data. Brain and Cognition, 87, 52–56. https://doi.org/10.1016/j.bandc.
2014.03.004
Verpooten, J., & Dewitte, S. (2017). The conundrum of modern art: Prestige-driven coevolutionary aesthetics trumps
evolutionary aesthetics among art experts. Human Nature, 28(1), 16–38. https://doi.org/10.1007/s12110-016-
9274-7
Vessel, E. A. (2010). Beauty and the beholder: Highly individual taste for abstract, but not real-world images. Journal of
Vision, 10(2), 1–14. https://doi.org/10.1167/10.2.18
Vessel, E. A., Maurer, N., Denker, A. H., & Starr, G. G. (2018). Stronger shared taste for natural aesthetic domains than
for artifacts of human culture. Cognition, 179, 121–131. https://doi.org/10.1016/j.cognition.2018.06.009
Vessel, E. A., Starr, G. G., & Rubin, N. (2012). The brain on art: Intense aesthetic experience activates the default mode
network. Frontiers in Human Neuroscience, 6, 66. https://doi.org/10.3389/fnhum.2012.00066
Vogt, S. (1999). Looking at paintings: Patterns of eye movements in artistically naïve and sophisticated subjects. Leonardo,
32(4), 325–325. https://doi.org/10.1162/002409499553325
Vogt, S., & Magnussen, S. (2007). Expertise in pictorial perception: Eye-movement patterns and visual memory in artists
and laymen. Perception, 36(1), 91–100. https://doi.org/10.1068/p5262
Wagner, V., Menninghaus, W., Hanich, J., & Jacobsen, T. (2014). Art schema effects on affective experience: The case of
disgusting images. Psychology of Aesthetics, Creativity, and the Arts, 8(2), 120–129. https://doi.org/10.1037/a0036126
Wang, S., Mak, H. W., & Fancourt, D. (2020). Arts, mental distress, mental health functioning and life satisfaction: Fixed-
effects analyses of a nationally representative panel study. BMC Public Health, 20(1), 208. https://doi.org/10.1186/
s12889-019-8109-y
Westphal-Fitch, G. (2019). Revisiting Fechner’s methods. In M. Nadal & O. Vartanian (Eds.), The Oxford handbook of
empirical aesthetics. Oxford University Press.
Williams, P. G., Johnson, K. T., Curtis, B. J., King, J. B., & Anderson, J. S. (2018). Individual differences in aesthetic
engagement are reflected in resting-state fMRI connectivity: Implications for stress resilience. NeuroImage, 179, 156–
165. https://doi.org/10.1016/j.neuroimage.2018.06.042
Winkielman, P., Halberstadt, J., Fazendeiro, T., & Catty, S. (2006). Prototypes are attractive because they are easy on the
mind. Psychological Science, 17(9), 799–806. https://doi.org/10.1111/j.1467-9280.2006.01785.x

313
Rebecca Chamberlain

Workman, C. I., Humphries, S., Hartung, F., Aguirre, G. K., Kable, J. W., & Chatterjee, A. (2021). Morality is in the
eye of the beholder: The neurocognitive basis of the “anomalous-is-bad” stereotype. Annals of the New York Academy
of Sciences, 1494(1), 3–17. https://doi.org/10.1111/nyas.14575
Zajonc, R. B. (1968). Attitudinal effects of mere exposure. Journal of Personality and Social Psychology, 9, 1–6.
Zangemeister, W. H., Sherman, K.,  & Stark, L. (1995). Evidence for a global scanpath strategy in viewing abstract
compared with realistic images. Neuropsychologia, 33(8), 1009–1025. https://doi.org/10.1016/0028-3932(95)00014-t
Zeki, S. (1999). Art and the brain. Journal of Consciousness Studies, 6(6–7), 76–96.
Zeki, S.,  & Lamb, M. (1994). The neurology of kinetic art. Brain, 117(3), 607–636. https://doi.org/10.1093/
brain/117.3.607
Zeki, S., Romaya, J. P., Benincasa, D. M. T., & Atiyah, M. F. (2014). The experience of mathematical beauty and its
neural correlates. Frontiers in Human Neuroscience, 8, 68. https://doi.org/10.3389/fnhum.2014.00068
Zeki, S., & Stutters, J. (2012). A brain-derived metric for preferred kinetic stimuli. Open Biology, 2(2), 120001. https://
doi.org/10.1098/rsob.120001

314
15
THE MUSIC SYSTEM
Amy M. Belfi and Psyche Loui

While the study of music perception and cognition (from a psychological perspective) spans several decades,
the cognitive neuroscience of music has grown substantially in recent years. Like the field of neuroaesthetics
more broadly, music cognition is a highly interdisciplinary field, with researchers from diverse backgrounds
including music theory, neuroscience, cognitive science, and music therapy, among others. In recent years,
music cognition has become a fairly well-established subfield within cognitive psychology and neurosci-
ence more broadly, with its own academic societies (Society for Music Perception and Cognition, Neuro-
sciences of Music). Perhaps as a result, music cognition researchers may not consider themselves part of the
neuroaesthetics community. Thus there is some, but not complete, overlap between music cognition and
neuroaesthetics. Another reason for this separation has to do with the fact that “aesthetics” may come from
an academic tradition that is often considered to refer more specifically to the visual arts or visual stimuli
(see, e.g., Arnheim, 1966; Berlyne, 1971). Additionally, some aspects of music cognition focus more on basic
perceptual functions, such as pitch and rhythm, which some may consider have little to do with aesthetics.
Nevertheless, listeners certainly do have aesthetic responses to music, and such responses are one of the major
reasons individuals choose to listen to music—people listen to music because they like it (Sanflippo et al.,
2020).
However, before discussing such aesthetic responses to music, it is important to consider the “building
blocks” of music perception. In this chapter, we break music down into its component parts and focus on
neuroscientific findings related to these. In the following sections, we discuss foundational and recent work
investigating cognitive and neural systems underlying pitch, tonality, timbre, and rhythm and conclude with
a final section on affective responses to music. While discussing music in this way is inherently reductionist
(see our final section on “Major Challenges, Goals, and Suggestions”), understanding the basic components
of music, and how listeners respond to them, is important for work seeking to understand listeners’ aesthetic
responses to music.

Pitch
The diverse musical systems around the world share certain perceptual and acoustic properties; of these, pitch
is one of the most basic perceptual attributes. The attribute of pitch ranges along a perceptual spectrum from
low to high. That is, the perceptual property of pitch is dependent on the acoustic property of frequency:
Higher-frequency sounds tend to be perceived as higher pitched, and lower-frequency sounds tend to be

DOI: 10.4324/9781003008675-17 315


Amy M. Belfi and Psyche Loui

perceived as lower pitched. However, there are different configurations of energy along the frequency spec-
trum that can give rise to the same percept of pitch, as long as they share the same fundamental frequency
(F0, the greatest common factor in frequency of periodic sounds). In other words, the percept of pitch can
be invariant across different spectral configurations but corresponds most closely to the physical attribute of
the F0.
While different harmonics (higher frequencies present in the sound) with the same F0 can have different
timbres (i.e., sound qualities; see more on timbre in the following), they nevertheless share the same sensation
of pitch. That is, the note C4 sounds like the same note whether it is played on a piano, a violin, or a synthe-
sizer. In the most extreme case, a pitch can still be heard even when there is no energy at the F0 altogether,
as long as energy is present at multiple harmonics. This phenomenon, known as virtual pitch, has been both
a challenge and a test case for neural models of pitch coding: a neural mechanism that codes for pitch must
yield an output similar to the F0 even when no energy is present at the F0, as long as harmonics of the F0
are present. At the same time, neural models of pitch must also account for psychological phenomena of how
pitch is perceived to be organized at a higher cognitive level. Pitch systems in many different cultures are
organized around the octave, which is a doubling in frequency (Sethares, 2004). Within each octave, pitches
are perceived as being organized chromatically in a continuum that is psychologically independent from the
height of the octave. Western listeners tend to perceive pitches with corresponding F0s that are related to
each other in doubled ratios (octaves) as being similar, a phenomenon known as octave equivalence. For
example, C4 (“middle C” on a piano keyboard) and C5 are perceived as similar, because they are one octave
apart. Octave equivalence is observed less frequently in populations with less exposure to Western culture
( Jacoby et al., 2019) suggesting at least some influence of culture on this perceptual phenomenon. These two
independent dimensions, pitch chroma (i.e., the position of a pitch within the octave) and pitch height (i.e.,
the perceived “highness” of the pitch), are derived from perceptual ratings of musical pitch from Western lis-
teners (Shepard, 1980). Musical training seems to hone a finer-grained ability to discriminate between small
differences in pitch chroma: classical musicians can have frequency discrimination thresholds that are up to
six times smaller than that of those without formal musical training (Micheyl et al., 2006).

Neural correlates of pitch


In light of the physical and psychological characteristics of pitch reviewed previously, a theory of neural
coding of pitch must account for virtual pitch, the psychological organization of pitch chroma and height,
and observed effects of formal musical training while being physiologically plausible given the constraints of
neurons within the auditory system. In the latter regard, two peripheral constraints in the auditory system
are the tonotopic (frequency-dependent) organization of the basilar membrane in the cochlea and the firing
of the auditory nerve. These constraints give rise to the two classes of theories of pitch coding: place coding
and time coding.
Place coding theories are predicted upon the fact that the basilar membrane contains channels of hair
cells that are sensitive to high frequencies up to 10,000 Hz at the basal end and low frequencies down to
~50 Hz at the apical end (Von Bekesy, 1960), with neighboring channels with center frequencies roughly
a third of an octave apart (Moore, 1985). Human hearing, however, ranges from around 20 to 20,000 Hz,
with the most sensitive region for pitch corresponding to F0s ranging from 100 to 5,000 Hz, and humans
can distinguish pitches that are much less than a third of an octave apart (Micheyl et al., 2006). Therefore,
the tonotopic mapping of the basilar membrane, known as the spatial coding theory, is insufficient to explain
the perception of pitch.
Another candidate mechanism for pitch coding, the temporal coding theory, comes from the firing rate
of the auditory nerve fibers. Specifically, this refers to the autocorrelation of auditory nerve firing patterns
that are phase locked to the acoustic stimulus (Licklider, 1951). Temporal coding may explain some aspects

316
The music system

of pitch, as the phase locking from auditory nerve fibers is accurate up to approximately 4,000 Hz ( Joris &
Verschooten, 2013). However, the rates of phase locking are not preserved between the auditory nerve fibers
and the auditory cortex, and furthermore, nerve fibers cannot fire up to the upper limit of the audible fre-
quency range (up to 20,000 Hz); thus, the temporal coding account cannot explain cortical or higher-level
pitch mechanisms, nor can it explain auditory perception at high frequencies (Oxenham, 2013).
Due to the invasiveness of recording from specific waystations of the auditory system, the majority of
neurophysiological (single or multiple-unit) data to understand pitch coding has relied on animal models.
Recording from the auditory cortex of marmoset monkeys, Bendor and Wang (2005) found that a small
population of neurons in the anterolateral border of the primary auditory cortex responded selectively to
both the F0 and harmonic complexes of overtones above the F0, thus providing a neural correlate of pitch
in a manner that is independent of frequency. Follow-up studies showed that this pitch constancy is likely
accomplished using a combination of spectral and temporal cues (Bendor et al., 2012).
At the level of the auditory cortex and beyond, human fMRI experiments have been effective at iden-
tifying the higher-level organization of our mental representations of pitch. Building on these spectral and
temporal cues, and then recombining pitches hierarchically to form melodies and harmonies (Warren et al.,
2003), the neural coding of pitch provides the basis for higher-order coding of musical structure in the
brain. Spectral vs. temporal cues are found to be specialized on opposite hemispheres (Albouy et al., 2020).
The idea that spectral and temporal cues must combine to code for the sensation of pitch is one of several
dual-coding or dual-processing mechanisms that have garnered support in auditory neuroscience. One such
model postulates that different processing streams from the auditory cortex code for location and identity
(“where” vs. “what”) information (Rauschecker & Tian, 2000). This is supported by the findings that ante-
rior and posterior superior temporal areas, which abut the auditory cortex, are sensitive to different acoustic
features. While the anterior superior temporal lobe is more sensitive to sound features that enable object
recognition, such as frequency comodulation (Zatorre et al., 2004), the posterior superior temporal lobe is
more sensitive to spatial information and is active slightly earlier than the anterior temporal lobe (Lomber &
Malhotra, 2008). This principle is related to the dual-stream anatomy proposed in speech and language
research (Hickok & Poeppel, 2007) which suggests a distinction between the posterior and dorsal pathway
and the anterior and ventral pathway: the posterior and dorsal pathway mediates articulation, whereas the
anterior ventral pathway supports comprehension. Similarly, a dual-stream neuroanatomical model has been
described for musical behavior, specifically singing (Loui, 2015): while the dorsal pathway enables auditory-
motor integration, the ventral pathway may enable melody recognition or the linking of a sound to its
meaning and its context.

Amusia
Dual-stream models can be particularly informative when they offer specific predictions about neuropsycho-
logical populations. One such population is people with congenital amusia, who have impaired pitch percep-
tion and/or production (for review, see Peretz, 2016). Congenital amusia, also known as tone-deafness, is a
lifelong musical disorder that prevents otherwise normal individuals from developing skills in pitch percep-
tion and production (Ayotte et al., 2002). It has known genetic correlates (Drayna et al., 2001) and is most
commonly identified in a neuropsychological test known as the Montreal Battery for Evaluation of Amusia
(MBEA; Peretz et al., 2003); however, many individuals who perform at normal levels on the MBEA never-
theless self-identify as being tone-deaf (Cuddy et al., 2005). People with congenital amusia, who could not
consciously report the directions of pitched intervals, could nevertheless produce (by singing) pairs of tones
with above-chance accuracy in pitch interval direction (Loui et al., 2008). This dissociation between percep-
tion and production ability suggests that congenital amusics may have a selective impairment in one of the
streams of auditory processing. Although there has been some support for a disrupted “where” pathway in

317
Amy M. Belfi and Psyche Loui

the form of impaired visuospatial processing in congenital amusics (Douglas & Bilkey, 2007), this finding was
not replicated by other groups (Williamson et al., 2011). The variable results from congenital amusia seem to
suggest that there are other sources of variance that come with amusia: on one hand, there may be multiple
subgroups of amusics, some that have more memory deficits, some that have more spatial deficits, and others
that have more of a lack of awareness of pitch and how it translates to space. On the other hand, the presence
of amusia may be highly confounded or comorbid with other impairments, such as mild forms of speech and
language impairments (dyslexia, specific language impairment; Loui, Kroog et al., 2011), cognitive impair-
ments (attention deficits, learning disabilities, memory deficits), and socio-emotional deficits (e.g., autism
spectrum disorders; Thompson et al., 2012). Thus, it remains unclear what the profile of a “pure” congenital
amusic would be.
Studies using multiple neuroimaging modalities have shown differences in auditory and auditory-motor
brain processes that are linked to congenital amusia. Using structural brain morphometry and cortical thick-
ness measures, Hyde and colleagues (Hyde et al., 2006, 2007) observed differences in the superior temporal
gyrus (STG) and the inferior frontal gyrus (IFG) of amusics. These findings are bolstered from the correla-
tive to the causal level by noninvasive neurostimulation studies (Hohmann et al., 2018) which showed that
applying transcranial direct current stimulation to STG and IFG led to disrupted pitch matching. As the STG
is heavily involved in pitch processing and is a hub of the where/what pathway distinction, a disruption in
this region may give rise to an inability to form memories, access mental representations, or otherwise have
conscious access of pitch direction or melodic information. Furthermore, the simultaneous disruption of
frontal lobe regions, specifically the IFG, may suggest further difficulties in memory, learning, or auditory-
motor integration of pitch information. A  parsimonious explanation for these simultaneously observed
deficits in multiple cortical regions was that white matter connectivity between the STG and IFG could be
disrupted in congenital amusics, leading to abnormal neuronal development or migration at the endpoints.
This was observed in a diffusion tensor imaging study (Loui et al., 2009), in which the arcuate fasciculus,
which connects the superior temporal and inferior frontal gyri, was smaller in volume in people with con-
genital amusia; furthermore, the volume of the arcuate fasciculus was inversely correlated with the size of
the pitch discrimination threshold across subjects. Further studies have employed MEG in combination with
structural MRI to provide further support for a right-hemisphere frontotemporal deficit in connectivity
(Albouy et al., 2013).
Event-related potentials (ERP) studies linked the deficit in congenital amusia to relatively late stages
of the auditory perception-cognition pathway (Peretz et al., 2009), suggesting that congenital amusia may
be fundamentally an issue with the lack of awareness of pitch information. This lack of awareness may
explain additional observations of pitch memory deficits in amusics (Tillmann et al., 2016) and explains the
perception-production mismatch (Loui et al., 2008), as discussed previously. Further studies have extended
these findings from congenital amusia into the linguistic domain, specifically showing that intonation pro-
cessing is impaired in amusics, even in speech perception (Liu et al., 2010; Lolli et al., 2015). Taken together,
results from congenital amusia provide support for a crucial role of the frontotemporal pathway, probably
right-lateralized, that enables the tight coupling between pitch perception and production.
While congenital amusia refers to the lifelong deficit of musical abilities, acquired amusia refers to a loss
of musical ability that results from brain damage. Studies in lesion patients indicate that deficits in various
components of pitch perception are associated with damage to locations along the auditory pathway, which
tend to be right-lateralized (Särkämö et al., 2009; Stewart et al., 2006). Furthermore, individuals with lesions
covering the inferior fronto-occipital fasciculus and the arcuate fasciculus are least likely to recover from
acquired amusia after a stroke, whereas those who recover from acquired amusia are more likely to have dam-
age in only one of these pathways (Sihvonen et al., 2016, 2017). These findings provide insight into possible
targets for neurorehabilitation after a stroke or other brain injury and provide evidence to the necessity of
these cortical structures for the processes underlying pitch perception.

318
The music system

Absolute pitch
While amusia is a deficit in perception of music, specifically in pitch, absolute pitch (AP) seems ostensibly to
be the opposite. People with AP have the ability to identify the pitch class of musical notes without an exter-
nal reference (Ward, 1999). In addition to being more common in musically trained individuals, especially
those who started musical training before the age of seven, AP runs in families and is more common among
those with East Asian descent. It is especially common in those of East Asian heritage who speak tone lan-
guages fluently, suggesting that the ability is associated with early language experience (Baharloo et al., 2000;
Deutsch et al., 2006). Furthermore, among people with AP, there is a range of pitch identification ability:
while some AP possessors are able to name any note in any timbre, other individuals have AP only for spe-
cific instruments, commonly their instruments of choice, and the accuracy of their pitch identification ability
falls off at very high and very low pitch ranges (Athos et al., 2007; Miyazaki, 1989). While some research-
ers interpreted this range of behavior as bimodal (Athos et al., 2007), others view it more as a continuum
of abilities (Miyazaki, 1989). Still others view the results as a dichotomy between distinct neurocognitive
processes, some of which are more sensitive to learning and memory, while others are more possibly innate
(Ross et al., 2005). Due to these intriguing interactions between genetic and environmental factors, AP has
been described as an ideal model for understanding the influences of genes and the environment (Baharloo
et al., 1998; Zatorre, 2003).
A classic finding in the AP literature is that when given frequent and infrequent tone stimuli in an ERP
study, most non-AP listeners show a P300 (positive waveform around 300 ms) after the infrequent tone; in
contrast, AP listeners show no such P300, suggesting that they did not require working memory to update
their continuous mental representations of these tone stimuli (Klein et al., 1984). This readily accessible and
seemingly effortlessly stored mental template for pitch categories may come from enhanced encoding and
categorical representation of pitch from a young age. The neural substrates that enable this automatic pitch
categorization likely come from the planum temporale, a region within the STG, which is exceptionally
larger in the left hemisphere of AP musicians (Schlaug et al., 1995). In addition to being larger in volume, the
left STG is also higher in white matter connectivity in AP musicians relative to their non-AP counterparts,
and pitch categorization accuracy is correlated with tract volume (Loui, Li et al., 2011). While these results
point to a selective increase in left-lateralized connectivity in the temporal lobe, functional MRI results point
to a distributed network of enhanced activity throughout the brain, albeit with results centering around the
left STG (Loui et al., 2012). Thus, it appears that both specific brain structure and general network-level brain
functioning are special in AP possessors, and these results are observed even after controlling for possible dif-
ferences in the length and start age of musical training, which are usually longer and earlier in AP musicians.
While AP is presumed to be rare, affecting less than 1% of the general population (Ward, 1999), most lis-
teners possess some absolute memory for pitch, as shown by being able to produce familiar songs at the right
starting pitch after repeated listening (Levitin, 1994). However, the enhanced categorization ability seems
relatively rare and specific to a unique population. This has led some researchers to ask whether the AP pos-
sessors’ tendency to categorize pitch might be thought of as a savant-like ability, as seen in some individuals
with autism (Mottron et al., 2013). In that regard, musicians with and without AP were tested on subclinical
traits of autism using the autism spectrum quotient (Dohn et al., 2012). Results showed that while AP pos-
sessors scored higher than non-AP counterparts in the autism quotient, they were still well lower than would
meet criteria for autism spectrum disorders. Crucially, AP possessors showed some higher scores in imagina-
tion but no differences from controls in social and communicative subscales. Taken together, AP could be
considered an enhanced perceptual categorization ability, likely subserved by a network of regions centering
around the structurally altered superior temporal lobe. Whether this brain network and its supported func-
tions can be trained in the lab, or in the practice room, remains an active area of both psychological and
pedagogical research.

319
Amy M. Belfi and Psyche Loui

Tonality
Tonality describes the perceived relationships between pitches in a particular context, or musical key. For
example, the key of C major consists of the following notes: C, D, E, F, G, A, B. Pitches other than these
are called nondiatonic, as they are not part of the key of C. Within each key, the notes are arranged in a
hierarchical manner, such that every pitch has a perceived level of “stability” within the key. The tonic is
considered the most central tone of this scale and subsequently is perceived as the most stable: in the key of C
major, the tonic note is C. The other notes of the scale are hierarchically organized based on their relation-
ships with the tonic (and therefore their perceived stability). To continue with the example of C major, the
note C is the tonic, and the notes constituting the major triad—E (the third note of the scale) and G (the
fifth)—are considered closely related to the tonic and therefore also highly stable notes. Therefore, tonality
constitutes the context in which each pitch is heard in a piece of music and designates certain pitches as more
or less likely to occur within this context.
Foundational studies of tonality perception confirmed that listeners, even those without formal musical
training, do perceive these tonal hierarchies. Early evidence for the psychological existence of such tonal
systems came from “probe-tone” studies: In these experiments, a melodic context is presented (e.g., a series
of notes), followed by a final “probe-tone.” The melodic context sets up a particular musical key—for
example, the context may consist of the ascending tones of the C major scale. Participants are then asked to
rate how well the final probe-tone fits within the preceding context. Early work using this paradigm indi-
cated that listeners were highly consistent in identifying which notes best fit a tonal context. The tonic was
rated as the best fitting, followed by other tones of the major chord and then other tones within the scale
(Krumhansl & Shepard, 1979). Using multidimensional scaling of participants’ ratings, Krumhansl and Kes-
sler (1982) derived a spatial map representing the mental distances between musical keys. Geometrically, this
map can be represented in the space of a torus. This toroidal space, therefore, effectively maps the psychological
distance between musical keys, as opposed to a purely music theoretic account of tonal space (Krumhansl &
Kessler, 1982).
While this work successfully provided a psychological map of tonal space, two decades later, researchers
sought to identify a topographic map of musical tonality in the brain. Using fMRI, researchers sought to
investigate the neural correlates of musical tonality ( Janata et al., 2002). Janata and colleagues (2002) cre-
ated a piece of music which systematically modulated across the surface of the toroidal space representing
the perception of tonality. Participants listened to this musical piece while neural responses were recorded
using fMRI. Results indicated that the medial prefrontal cortex (mPFC) was the key region in which neural
activity tracked movement through the tonal space of the music. That is, the blood-oxygen-level-dependent
responses in voxels within the mPFC correlated with movement of the location of the melody on the tonal
torus surface, with certain voxels being more sensitive to certain positions on the torus. This work has there-
fore provided both a psychological and a neuroanatomical map of tonality as it applies to Western music.

Harmonic and melodic expectancy


Tonality influences listeners’ perception of both melody and harmony: melodies consist of pitches presented
sequentially over time, while harmony is the simultaneous presentation of pitches. As described previously,
even untrained listeners learn to anticipate the likelihood of certain melodic and harmonic musical events,
based on the tonal context. The perception of both harmony and melody relies on the perception of tonal
contexts—that is, a certain pitch or chord may be perceived as harmonic in one context but not in another.
Early theories of melody perception also proposed that listeners form expectations about which pitches will
come next in a melody (e.g., the implication-realization model; Narmour, 1990, 1992).

320
The music system

One frequent approach to understanding the neural correlates of melodic and harmonic expectancy has
been to investigate what happens in the brain when such expectations are violated. For example, researchers
may set up a tonal context by presenting a series of chord progressions. When presented with chords that
deviate from what one would expect given this tonal context (for example, chords that contain out-of-key
notes), researchers have identified a specific electrophysiological response in the brain, termed the early right
anterior negativity, or ERAN (Koelsch et al., 2000), the neural generators of which are located in the infe-
rior frontal gyrus (IFG). While the ERAN tends to be localized to the right hemisphere, evidence indicates
that even left-hemisphere structures are critical for this response. For example, patients with focal damage to
the left hemisphere IFG showed abnormal ERAN responses and behavioral deficits in detecting harmonic
violations (Sammler et al., 2011).
The ERAN in response to harmonic violations has been replicated many times, in many situations
(although it has sometimes been found bilaterally and is referred to as the EAN; e.g., Koelsch et al., 2013;
Loui et al., 2005; Zhang et al., 2018). For example, the ERAN is generated automatically but is sensitive to
the demands of attention (Loui et al., 2005; Lee et al., 2019) and is found in a similar magnitude even when
participants are aware of an upcoming irregular chord (Guo & Koelsch, 2016). A magnetic counterpart to
the ERAN has also been identified using magnetoencephalography (MEG) (Kim et al., 2011). These find-
ings suggest that the ERAN is a robust response to violations of musical tonality. Despite this robustness of
the ERAN, other musical features have been found to modulate this effect. For example, frequent changes
in musical timbre have been shown to reduce ERAN amplitude (Fiveash et al., 2018), and ERAN ampli-
tude has been shown to be larger in musicians than non-musicians (Brattico et al., 2013; Tervaniemi et al.,
2012). Additionally, individuals with congenital amusia do not exhibit an ERAN, suggesting that deficits in
pitch perception lead to subsequent abnormalities in the perception and neural responses to harmony (Sun
et al., 2018). In sum, this work suggests that perception of melody and harmony interact with other musical
features and individual differences in music perceptual abilities.

Consonance and dissonance


An important concept in the discussion of tonality is that of consonance and dissonance, which describe the
relationships between either simultaneous or subsequent pitches. Dissonance is generally considered unpleas-
ant, while consonance tends to be perceived as pleasant. For example, dissonant chords evoke a stronger
ERAN than unexpected but consonant chords (Pagès-Portabella & Toro, 2020). While there is no univer-
sally accepted definition of consonance or dissonance, generally, the perception of dissonance is the result
of perceived roughness or tension (sometimes referred to as sensory consonance/dissonance). While sensory
consonance and dissonance are suggested to be universal percepts due to the structure of the human auditory
system, listeners can perceive dissonance even in the absence of sensory roughness. Recent work has sug-
gested that the perception of dissonance is not universal—while listeners who have a lifetime of exposure to
Western music tend to find dissonant intervals unpleasant, listeners who have no exposure to Western music
do not find dissonance unpleasant (McDermott et al., 2016).
Despite the fact that dissonance is not always displeasing (nor consonance pleasing; Schafer  & Eerola,
2020), musical dissonance has frequently been used as a proxy for “unpleasant” music. In this way, rela-
tively low-level musical features can result in higher-order responses to music, such as reward and affect.
Recent work has indicated that individual differences in preference for dissonance are associated with neu-
ral responses in low-level auditory regions, such as the inferior colliculus (Kim et al., 2017). Differential
responses to musical consonance and dissonance have been found in other auditory structures, such as the
STG (Foo et al., 2016). Related to the idea that musical dissonance is perceived as unpleasant, dissonance has
also been associated with activity in emotion-related brain regions such as the mPFC (Bravo et al., 2020).

321
Amy M. Belfi and Psyche Loui

Consonant music has also been used as a musical “reward” in a reinforcement learning context, and activity
in the nucleus accumbens was found to correlate with reward prediction errors (Gold et al., 2019). As these
works suggest, acoustic properties of music such as harmony can influence higher-order responses to music,
such as emotion and reward.

Timbre
While the sensation of pitch is correlated with the F0 of periodic sounds, the sensation of timbre is an emer-
gent property of spectral and temporal characteristics of sounds. Due to its complex nature and poorly defined
characteristics, historically timbre has been affectionately dubbed the “auditory wastebasket” (Siedenburg &
McAdams, 2017). Nevertheless, the study of timbre has benefited from sophisticated developments in tech-
nical tools for sound description developed in recent years.
The temporal envelope of a sound, especially the time between its onset and its peak amplitude (“attack
time”), is a strong determinant of the perceptual attribute of “bite.” While attack time is a feature of the
temporal envelope, spectral centroid is a feature of the spectral envelope: spectral centroid is computed as the
weighted average of the frequency of all harmonics present; this gives rise to its “brightness.” Classic studies
have used multidimensional scaling of pairwise similarity ratings between sounds of different timbres and
found that spectral centroid and attack time are two orthogonal dimensions that account for much of the
variance in judgments in sound quality (Wessel, 1979). A third most salient dimension is both spectral and
temporal in nature: it is spectral flux, which is the change in spectral centroid over time (McAdams, 2013).
Because these dimensions of sound are clearly defined and orthogonal to each other, both perceptually and
in terms of their underlying acoustic attributes, they are useful both for musical expression and for creating
well-controlled experimental stimuli in order to understand how the brain processes timbre. An ERP study
compared the three dimensions of timbre (attack time, spectral centroid, and spectral flux) directly using the
mismatch negativity (MMN), a negative frontally centered waveform around 200 ms after the onset of any
small deviations from the context. The MMNs elicited by the different dimensions were indeed separate in
their underlying sources, suggesting separate dimensions of processing in auditory memory and providing
convergent results with the psychophysical data (Caclin et al., 2006).
fMRI data showed that the bilateral auditory cortices were sensitive to both pitch and spectral changes,
but modulations in spectral shape additionally elicited right-lateralized brain activity in the superior temporal
sulcus (STS; Warren et al., 2005). This right-lateralized STS activity supports the idea that right hemisphere
is relatively tuned to spectral changes, whereas the left hemisphere is more tuned to fine-grained temporal
structure (Flinker et al., 2019). Since timbre is an important cue towards the perception of auditory objects,
the study of “dystimbria,” or the impairment of spectral and temporal analysis without loudness or pitch pro-
cessing difficulty (Griffiths et al., 2007), can provide a window into how the brain perceives objects from the
world of sound (cf. Bregman, 1994). On the other hand, people with exceptionally fine-grained training in
listening to timbres, such as piano tuners, have enhanced grey matter in the superior temporal structures as
well as the hippocampal complex, which is crucial for learning and memory (Teki et al., 2012). These grey
matter differences are correlated with the duration of one’s career in piano tuning and not with actual age,
suggesting that there are training-related adaptations in grey matter that come with advanced musical experi-
ence, even within the specific task of listening for timbre in order to tune a musical instrument.

Rhythm
The terms rhythm, beat, meter, and their associated constructs have been used in different ways. Here, we
will define rhythm as a pattern of time intervals between stimuli. Listeners may feel a sense of beat when
listening to rhythmic auditory patterns. Meter refers to the hierarchical organization of musical beats, which

322
The music system

can be denoted by the time signature of a musical piece. Classic work on the perception of musical rhythm
suggests that the meter serves as a generative grammar, which can be subdivided into bars and further subdi-
vided into beats (Longuet-Higgins & Lee, 1982). The meter of a piece also designates certain beats as more
likely to be accented—for example, a “waltz” meter consists of groupings of three beats, of which the first
is accented.
As with the perception of melody and harmony, an important component of rhythm perception is
expectation. It has been proposed that after hearing the first two notes in a melodic sequence, the listener
forms expectations about subsequent notes (Longuet-Higgins & Lee, 1982). In their classic model of rhythm
perception, Longuet-Higgins and Lee (1982) propose that if a listeners’ expectation for the timing of the
third note is confirmed, they then move up the metrical hierarchy by grouping the heard units into a single
unit. Later, Desain and Honing (2003) found that expert listeners consistently perceive groups of rhythmic
categories, which are influenced by the perceived metrical context. Without being given a metrical context,
listeners tend to perceive a duple meter (and subsequently categorize the rhythm as falling into such). Other
foundational work on the perception of rhythm indicated that certain patterns induce listeners to generate an
“internal clock,” which leads to a stronger ability to reproduce the heard sequences (Povel & Essens, 1985).
This work suggests that listeners have pre-existing expectations about musical rhythms. More recently,
Jacoby and McDermott (2017) sought to investigate listeners’ priors (i.e., previously held biases) when
evaluating rhythmic sequences and tested whether these were consistent across cultures. Importantly, they
developed a method for testing rhythmic perception in non-expert listeners. In this task, participants were
asked to reproduce a rhythm by tapping; this original “seed” rhythm was replaced with the participant’s
reproduction in an iterative fashion. The authors found that the tapping patterns tended to converge on to
simple integer ratios but that the specific ratios favored were those represented in Western music. Listeners
unfamiliar with Western music exhibited priors for different integer ratios, which reflected those most com-
mon in their musical traditions.
In terms of the neural structures underlying the perception of rhythm and its associated constructs, much
prior work has focused on the involvement of the motor system. Studies using fMRI to investigate the neural
correlates of rhythm perception consistently indicate motor-related brain structures, such as the motor cortex,
cerebellum, and basal ganglia (Chen et al., 2008). One recent meta-analysis identified consistent patterns of activ-
ity in motor cortical regions while listening to music (Gordon et al., 2018). There is even some evidence that the
aesthetic appreciation of musical rhythms influences motor cortex activity: Rhythmic patterns deemed beautiful
by listeners were found to be associated with greater activity in the premotor cortex and cerebellum compared
to rhythms that were judged as not beautiful (Kornysheva et al., 2010). This result is reminiscent of findings from
fMRI studies on dance, in which an action observation network consisting of sensorimotor areas (including the
premotor cortex) are consistently active during aesthetic appreciation of dance movements (Cross et al., 2011).
In addition to findings illustrating the role of the motor system, more recent evidence suggests that it is
interactions between the auditory and motor systems that underlie rhythm perception. Both structural and
functional connectivity between auditory and motor structures may reflect individual differences in rhythm
perceptual abilities (Chen et al., 2006; Grahn & Rowe, 2009). Recent work has found that, at rest, musicians
show greater auditory-sensorimotor functional connectivity than non-musicians (Tanaka & Kirino, 2018).
The volume of the arcuate fasciculus, the prominent auditory-motor pathway discussed earlier in the con-
text of pitch and amusia, was found to correlate also with the ability to synchronize with a musical rhythm
(Vaquero et al., 2018).

Beat
Beat is the basic building block of musical rhythm which allows for the synchronization of musical events.
Beat not only synchronizes musical events to one another but also synchronizes our brains to the rhythm of

323
Amy M. Belfi and Psyche Loui

musical stimuli. When presented with a beat, EEG recordings show rhythmic activity at the beat frequency
(Nozaradan et al., 2011) and show phase-locking with the beat frequency (Vanden Bosch der Nederlanden
et al., 2020). This neural sense of beat is produced through the coupled oscillation of auditory and motor
pathways, including the premotor cortex (PMC) and the supplementary motor area (SMA; Grahn & Brett,
2007). The coupling between stimuli and oscillatory neural responses, known as neuronal entrainment, is
stronger for attended stimuli (Lakatos et al., 2008). Entrainment of neuronal oscillations could be responsible
for the ability to predict a pulse even in situations where there is no spectral energy produced on the beat
itself (Large et al., 2015). Additional roles for beat perception have also been identified in the basal ganglia
(BG) and the striatum (Grahn & Rowe, 2009). SMA, PMC, and BG contributions to beat perception are
particularly apparent in classical musicians and in individuals with Parkinson’s disease (PD). Perceiving beat-
based rhythms, which have simple integer ratios in duration and recur with a predictable metric accent,
activates the PMC and SMA when compared to rhythms without simple integer ratios (Grahn  & Brett,
2007). In addition, persons with PD (which primarily affects BG) show a deficit in discriminating beat-based
rhythms but not rhythms that lack an underlying beat, suggesting an important role of BG in the detection
and generation of internal beats (Grahn & Brett, 2009).
Time-sensitive measures in the brain have provided additional insight into the mechanisms underlying
rhythm perception. For example, in auditory cortex, beta rhythms measured by MEG are thought to underlie
external auditory stimuli processing, whereas gamma rhythms from the same regions are thought to be involved
with processing of the underlying beat (Fujioka et al., 2009). Furthermore, ERP findings have indicated that
healthy individuals show a larger mismatch negativity when presented with deviations from an isochronous
beat from a low-pitched voice as compared to a high-pitched voice, suggesting that humans are more sensitive
to beat information at lower pitches. The superior time perception in low-pitched sounds may explain why
low-pitched instruments more commonly signal the rhythm in musical ensembles (Hove et al., 2014).
The coupling between auditory and motor systems when confronted with beat-based stimuli might explain
why we feel “groove,” which is defined as the urge to move to music (Stupacher et al., 2013). Music that is rated
as high in groove elicits larger motor-evoked potentials (measured from motor neurons in the hand and arm) in
musicians than low-groove music or noise (Stupacher et al., 2013). One feature that determines the sensation of
groove is syncopation, which can be defined as a slight violation of an expected, beat-based rhythm. However,
this is not to say that higher syncopation necessarily means a higher sense of groove. In fact, an inverse U-shaped
relationship between groove and syncopation has been observed, with medium-syncopated music tending to
elicit a stronger desire to move than either high- or low-syncopated music (Witek et al., 2014).
While a better understanding of musical beat processing is useful in its own right, it can also aid in both
normal development and in symptom management for certain disease populations. In a sample of 14-month-
old infants, parents holding the infant faced an experimenter and then proceeded to move either synchro-
nously (on-beat) or asynchronously (off-beat) from one another. Infants proceeded to show significantly
higher prosocial behavior towards the experimenter after the synchronous condition as compared to the
asynchronous condition, thus demonstrating the social value of synchronizing to a beat even during infancy
(Cirelli et al., 2014). Moreover, music-based movement therapy has proven to be effective for ameliorating
gait detriments in persons with PD, presumably by allowing them to entrain to external beat stimuli. For
example, a better understanding of groove could aid in gait therapies for PD, as high-groove music has been
shown to be more effective than low-groove music for such therapies, particularly among individuals with a
large deficit in beat perception (Leow et al., 2014).

Meter
Meter refers to the hierarchical organization of beats into recurring groups. Meter usually consists of group-
ings of two, three, or four beats, though in some cases rarer meters that contain five, seven, or even larger or

324
The music system

variable sets of beats may occur. Regardless of the number of beats, there is a tendency to perceive an accent
on the first beat of a given measure (often with secondary accents later on for meters containing four or more
beats). This tendency in humans to accentuate metrical beats is so strong that EEG recordings have shown
an enhancement of the beat frequency associated with the “first” beat of an imagined meter, even with an
unaccented isochronous beat stimulus (Nozaradan et al., 2011). Perceptions of meter can even be extended as
far back as birth: In an ERP study comparing a standard rock drum stimulus to two deviant stimuli in which
metrically important beats were excluded, sleeping newborns showed a negativity response in cases where
metrical expectations were violated (Winkler et al., 2009).

Deficits in rhythm perception


While amusia has typically been defined as deficits in pitch perception, deficits in rhythm perception have
been referred to by various names, such as beat deafness (Phillips-Silver et al., 2011), dysrhythmia (Launay
et al., 2014), arrythmia (Patel, 2006), or simply “rhythm processing deficits” (Tranchant & Vuvan, 2015).
Such rhythm processing deficits have also been considered a sub-type of amusia more broadly (Pfeifer &
Hamann, 2015). Recent work has indicated that “pitch deafness” and “beat deafness” tend to co-occur:
in one study, 6/10 individuals with pitch perception deficits also had deficits in rhythm perception and
production (Lagrois & Peretz, 2019). Despite this high co-occurrence of both pitch and rhythm processing
disorders, these results also indicate that the two processes are distinct. Rhythm processing remains intact in
some individuals with typical pitch-based amusia (Lagrois & Peretz, 2019; Phillips-Silver et al., 2013), and
pitch-processing remains intact in some individuals with rhythm processing deficits (Launay et al., 2014).
Just as disorders of pitch and rhythm perception appear to be dissociated, there also may be multiple types
of rhythm processing deficits. For example, some individuals perform poorly on rhythm perception tasks
(e.g., identifying whether a metronome is aligned to a beat) but not rhythm production tasks (e.g., tapping to
a beat; Bégel et al., 2017; Fujii & Schlaug, 2013). This supports evidence that rhythm production is distinct
from rhythm perception (Tierney & Kraus, 2015). Interestingly, one case study identified who could effec-
tively synchronize to the beat of a metronome but not to the beat of a piece of music (Phillips-Silver et al.,
2011). This suggests that rhythm perception and production may vary based on the complexity of the rhyth-
mic stimulus. Neuroimaging research has indicated that structures in the auditory and motor system likely
underlie individual differences in rhythmic perception and production abilities (Grahn & Schuit, 2012).

Affect
Emotional responses to music engage a wide range of neural regions which also underlie the basic components of
music perception. In discussing affective responses to music, it is first important to make the distinction between
emotion recognition, or identifying particular emotions conveyed by music, versus actual feelings experienced in
the listener in response to the music. That is, music can both convey and evoke emotions, and these conveyed
and felt emotions are not always congruent. For example, the song “Happy,” by Pharrell Williams, may convey
feelings of happiness which are easily recognized by a listener, but that same listener may feel sad when they hear
this song, perhaps because it reminds them of a sad moment from their life or another host of reasons (for a
review of the potential mechanisms underlying music-evoked emotions, see Juslin, 2013). Here, we will briefly
review the neural substrates underlying both musical emotion recognition and felt emotional responses to music.

Emotion recognition
It appears that the ability to recognize basic emotions (such as happiness and sadness) in music is, in part,
related to music perception abilities such as pitch perception. Individuals with congenital amusia were found

325
Amy M. Belfi and Psyche Loui

to have impaired recognition of musical emotions but not recognition of emotions in facial expressions
(Lévêque et al., 2018). Recent work has replicated this finding in individuals with brain damage: Persons
with damage to the left hemisphere scored lower on the MBEA than healthy comparisons (indicating poorer
music perception) and also exhibited deficits in recognizing emotions in music (Pralus et al., 2020). Perhaps
unsurprisingly, then, prior work also indicates that the auditory cortex plays a role in emotion recognition in
music. For example, fMRI work has identified increased activity in the auditory cortex for joyful music, as
compared to fearful music, and this difference in activity was still present even after accounting for acoustical
differences in the two categories of stimuli (Koelsch, Skouras et al., 2013). Furthermore, the auditory cortex
may represent emotion in a domain-general manner, such that regions in the auditory cortex encode for
emotional content of music regardless of instrument timbre (Paquette et al., 2018).
When considering the neural structures underlying music emotion recognition, much of prior work
has focused on the role of medial temporal lobe structures, particularly the amygdala and hippocampus
(for review, see Frühholz et al., 2014). Neuropsychological studies of persons with amygdala damage (both
bilateral and unilateral) have identified that these individuals are selectively impaired at recognizing threat in
music, while identification of other basic emotions (happiness, sadness) is preserved (Gosselin et al., 2005,
2007, 2011). In addition to communicating basic emotions, such as happiness, sadness, or fear, evidence
also indicates that music can communicate more complex emotions or intentions (Steinbeis  & Koelsch,
2009). For example, both musically trained and untrained listeners are able to accurately identify intentions
expressed by music, such as “domineering,” “insolent,” or “caring” (Aucouturier & Canonne, 2017; Cowen
et al., 2020). Individuals with frontotemporal dementia were impaired at identifying such complex emotions
or intentions in music (Downey et al., 2013), and performance on this task was associated with gray matter
volume in the ventromedial prefrontal cortex (vmPFC).

Felt emotions
Some of the earliest neuroimaging studies investigating felt emotional responses to music suggested the
involvement of subcortical reward regions, such as the ventral striatum (Blood et al., 1999; Blood & Zatorre,
2001; Menon & Levitin, 2005). More recently, evidence has distinguished between reward-related regions
involved in the anticipation versus the experience of musical chills, a particularly strong emotional and aes-
thetic response to music. Salimpoor and colleagues (2011) identified that activity in the nucleus accumbens
(NAcc) increased during the peak emotional response to music (i.e., musical chills), while activity in the
caudate was greatest during the anticipation of musical chills (Salimpoor et al., 2011). Music may recruit
similar neural systems as monetary rewards, such that the nucleus accumbens responds to reward prediction
errors in music (Gold et al., 2019). Converging with these neuroimaging results, deep brain stimulation in
the NAcc was shown to increase specific musical preferences in a case study, again supporting the role of the
NAcc in musical pleasure (Mantione et al., 2014).
The prefrontal cortex has also been implicated in felt emotions in response to music. As with the amyg-
dala and ventral striatum, activity in the mPFC has been shown to correspond with the intensity of musical
chills (Blood & Zatorre, 2001). Sad music, more than happy music, was shown to activate the vmPFC and
other structures in the “default mode network” (Taruffi et al., 2017). In particular, the mPFC may play an
important role in the aesthetic appreciation of music (Kim et al., 2019). This finding is paralleled in works
on visual aesthetics (Belfi et al., 2019; Vessel et al., 2019), suggesting that the mPFC and other structures in
this network may play a particularly important role in aesthetic emotional responses, regardless of the modal-
ity of the stimulus. Similarly, other work has identified that the vmPFC responds selectively to low-arousal
emotions in music, both basic and complex (Trost et al., 2012).
While substantial prior research indicates the involvement of auditory and emotion-related regions in
musical emotions, more recently, work has indicated that connectivity between these regions underlies

326
The music system

emotional responses to music. For example, increased connectivity between the auditory cortex and reward-
related regions has been found while listening to joyful music (Koelsch, Skouras et al., 2013) and fearful
music (Koelsch et al., 2018). Similarly, the reward value of a piece of music can be predicted by the strength
of functional connectivity between the NAcc and STG (Salimpoor et  al., 2013), and individuals with a
strong tendency to experience musical pleasure also display stronger connectivity between the NAcc and
STG while listening to music (Shany et al., 2019). This suggests that not only musical pleasure during listen-
ing, but individual differences in musical pleasure, are associated with connectivity between auditory and
emotion-related regions.
In addition to the functional connectivity analyses described previously, structural imaging provides con-
verging evidence suggesting that individual differences in musical reward are related to connectivity between
auditory and reward-related regions. For example, individuals who more frequently experience musical chills
have higher tract volume between the pSTG and emotion-related regions, including the medial prefrontal
cortex (Sachs et al., 2016). Brain stimulation studies also support the idea that connectivity between audi-
tory and reward structures is a critical neural substrate of musical reward (Mas-Herrero et al., 2018). And,
finally, studies of individuals with reduced emotional responses to music (called “musical anhedonia”) have
also indicated a role for connections between auditory and reward regions. That is, individuals with musical
anhedonia have been shown to have both reduced functional (Martínez-Molina et al., 2016) and structural
(Belfi et al., 2017; Loui et al., 2017; Satoh et al., 2016) connectivity between such regions.

Major challenges, goals, and suggestions


The past several decades have seen rapidly increasing interest in music perception and cognition, from
perspectives spanning music theory (Meyer, 1956) and music performance (Bernstein, 1973) to cognitive
psychology (Krumhansl & Shepard, 1979) and more recently in cognitive neuroscience (Blood et al., 1999;
Schlaug et al., 1995). Since these early studies, the field has grown substantially, particularly in recent years
(Cheever et al., 2018). Here, we briefly discuss what we see as the most important challenges and directions
for future research in the field. In particular, we consider the current emphasis on examining music in its
“naturalistic” form and the importance of balancing such approaches with more traditional methods. Impor-
tantly, we also consider the need to broaden the field beyond studying traditional Western music listeners
and styles of music.

Reductionist vs. naturalistic approaches


As presented in this chapter, we describe the various “components” of music—pitch, tonality, timbre, and
rhythm—and describe the neural systems and structures underlying the perception of these musical proper-
ties. Subdividing music this way, into its component parts, perhaps reflects the most traditional approach to
studying music perception. As it stands, researchers tend to specialize in one domain of music, and studies
tend to focus on one particular aspect or musical feature. This reductionist approach is not unlike that in
other fields, for example, researchers who study particular aspects of cognitive or perceptual processes such as
language or visual perception. However, music poses a unique challenge to this type of approach: Any stud-
ies that break music down into its component parts tend to retain a high degree of experimental control but
may distort essential aspects of the listening experience being studied. As can be seen in the present chapter,
it is a challenge to even talk about music in a reductionist way—basic “building blocks” of music, such as
pitch perception, influence higher-order perception of harmony and dissonance, which in turn influences
the perception of musical tension and affect.
More broadly in the field of cognitive neuroscience, and also within neuroaesthetics specifically, there is
rapidly growing interest in studying perception in “real world” settings, such as classrooms (Dikker et al.,

327
Amy M. Belfi and Psyche Loui

2017), movie theaters (Fröber & Thomaschke, 2019), or art museums (Brieber et al., 2014; Herrera-Arcos
et al., 2017). Similarly, music cognition researchers have begun to study music in the context of live con-
certs (Belfi et al., 2021; Coutinho & Scherer, 2017; Egermann et al., 2013; Merrill et al., 2020; Upham &
McAdams, 2018). Although these types of studies are still rare, more commonly researchers are investigating
neural responses to naturalistic music listening in the context of recorded music. While traditionally, music
perception researchers create stimuli for the purposes of manipulating certain musical features, this natural-
istic approach seeks to investigate neural responses to “real” music (Alluri et al., 2012, 2013; Brauchli et al.,
2020; Cong et al., 2014; Omigie et al., 2020; Sachs et al., 2020). This approach has the benefit of maintain-
ing the integrity of the music itself and a typical listening experience but also results in a lack of experimental
control. While we feel that the push towards more naturalistic listening conditions is an important and useful
approach as the field moves forward, both approaches are useful and can provide complementary evidence
to one another.

Diversity and representation


Music perception and cognition is not immune to the issues present within the broader fields of cognitive
psychology and neuroscience. In particular, the use of WEIRD populations (Western, educated, industrial-
ized, rich, democratic) limits the generalizability of much of the field’s work (Henrich et al., 2010). On one
hand, some musical features may map consistently onto psychological constructs across cultures, as has been
shown in perceived emotions (Fritz et al., 2009). On the other hand, some findings in music perception and
cognition that are well replicated in Western listeners may not translate cross-culturally. For example, non-
Western listeners may not perceive dissonance as less preferable than consonance (McDermott et al., 2016),
exhibit different rhythmic priors ( Jacoby  & McDermott, 2017), and do not exhibit octave equivalence
( Jacoby et al., 2019). In addition to the fact that the field lacks diversity in its participant samples, perhaps
one unique aspect of music research is that attention should also be paid to the diversity of the music studied.
Much of the present chapter focuses on perceptual constructs that are prevalent in listeners familiar with
Western tonal music, but for a fuller review of issues and suggestions for conducting cross-cultural research
in music cognition, see ( Jacoby et al., 2020).

Conclusions
The present chapter seeks to provide a broad overview of classic and recent work investigating the neuroana-
tomical and neurobiological underpinnings of music perception and cognition. Neuroimaging, neurophysi-
ological, and neuropsychological studies have each provided important insights into the neural underpinnings
of music perception. Listening to music engages a host of cognitive and neural systems, ranging from low-
level auditory perception to higher-order cognitive functions. The study of music and the brain is a rapidly
growing field, and our hope is that this work can shed insights into the aesthetic experience more generally
while continuing to move towards rigorous yet ecologically valid methods and approaches that embrace
more diverse participant groups and study materials.

References
Albouy, P., Benjamin, L., Morillon, B.,  & Zatorre, R. J. (2020). Distinct sensitivity to spectrotemporal modula-
tion supports brain asymmetry for speech and melody. Science, 367(6481), 16. https://doi.org/10.1126/science.
1252826
Albouy, P., Mattout, J., Bouet, R., Maby, E., Sanchez, G., Aguera, P. E., Daligault, S., Delpuech, C., Bertrand, O., Cac-
lin, A., & Tillmann, B. (2013). Impaired pitch perception and memory in congenital amusia: The deficit starts in the
auditory cortex. Brain, 136(5), 1639–1661. https://doi.org/10.1093/brain/awt082

328
The music system

Alluri, V., Toiviainen, P., Jääskeläinen, I. P., Glerean, E., Sams, M., & Brattico, E. (2012). Large-scale brain networks
emerge from dynamic processing of musical timbre, key and rhythm. NeuroImage, 59(4), 3677–3689. https://doi.
org/10.1016/j.neuroimage.2011.11.019
Alluri, V., Toiviainen, P., Lund, T. E., Wallentin, M., Vuust, P., Nandi, A. K., Ristaniemi, T., & Brattico, E. (2013). From
Vivaldi to Beatles and back: Predicting lateralized brain responses to music. NeuroImage, 83, 627–636. https://doi.
org/10.1016/j.neuroimage.2013.06.064
Arnheim, R. (1966). Toward a psychology of art. University of California Press.
Athos, E. A., Levinson, B., Kistler, A., Zemansky, J., Bostrom, A., Freimer, N., & Gitschier, J. (2007). Dichotomy and
perceptual distortions in absolute pitch ability. Proceedings of the National Academy of Sciences of the United States of
America, 104(37), 14795–14800. https://doi.org/10.1073/pnas.0703868104
Aucouturier, J. J., & Canonne, C. (2017). Musical friends and foes: The social cognition of affiliation and control in
improvised interactions. Cognition, 161, 94–108. https://doi.org/10.1016/j.cognition.2017.01.019
Ayotte, J., Peretz, I., & Hyde, K. (2002). Congenital amusia. A group study of adults afflicted with a music-specific dis-
order. Brain, 125(2), 238–251. https://doi.org/10.1093/brain/awf028
Baharloo, S., Johnston, P. A., Service, S. K., Gitschier, J., & Freimer, N. B. (1998). Absolute pitch: An approach for
identification of genetic and nongenetic components. American Journal of Human Genetics, 62(2), 224–231. https://
doi.org/10.1086/301704
Baharloo, S., Service, S. K., Risch, N., Gitschier, J., & Freimer, N. B. (2000). Familial aggregation of absolute pitch.
American Journal of Human Genetics, 67(3), 755–758. https://doi.org/10.1086/303057
Bégel, V., Benoit, C. E., Correa, A., Cutanda, D., Kotz, S. A., & Dalla Bella, S. (2017). “Lost in time” but still moving
to the beat. Neuropsychologia, 94, 129–138. https://doi.org/10.1016/j.neuropsychologia.2016.11.022
Belfi, A. M., Evans, E., Heskje, J., Bruss, J., & Tranel, D. (2017). Musical anhedonia after focal brain damage. Neuropsy-
chologia, 97, 29–37. https://doi.org/10.1016/j.neuropsychologia.2017.01.030
Belfi, A. M., Samson, D. W., Crane, J.,  & Schmidt, N. L. (2021). Aesthetic judgments of live and recorded music:
Effects of congruence between musical artist and piece. Frontiers in Psychology, 12, 618025. https://doi.org/10.3389/
fpsyg.2021.618025
Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., Pelli, D. G., & Starr, G. G. (2019).
Dynamics of aesthetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.
org/10.1016/j.neuroimage.2018.12.017
Bendor, D., Osmanski, M. S., & Wang, X. (2012). Dual-pitch processing mechanisms in primate auditory cortex. Journal
of Neuroscience, 32(46), 16149–16161. https://doi.org/10.1523/JNEUROSCI.2563-12.2012
Bendor, D., & Wang, X. (2005). The neuronal representation of pitch in primate auditory cortex. Nature, 436(7054),
1161–1165. https://doi.org/10.1038/nature03867
Berlyne, D. E. (1971). Aesthetics and psychobiology. Appleton-Century-Crofts.
Bernstein, L. (1973). The unanswered question. Cambridge, MA: Harvard University Press.
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Blood, A. J., Zatorre, R. J., Bermudez, P., & Evans, A. C. (1999). Emotional responses to pleasant and unpleasant music
correlate with activity in paralimbic brain regions. Nature Neuroscience, 2(4), 382–387. https://doi.org/10.1038/7299
Brattico, E., Tupala, T., Glerean, E., & Tervaniemi, M. (2013). Modulated neural processing of Western harmony in folk
musicians. Psychophysiology, 50(7), 653–663. https://doi.org/10.1111/psyp.12049
Brauchli, C., Leipold, S., & Jäncke, L. (2020). Diminished large-scale functional brain networks in absolute pitch dur-
ing the perception of naturalistic music and audiobooks. NeuroImage, 216, 116513. https://doi.org/10.1016/j.
neuroimage.2019.116513
Bravo, F., Cross, I., Hopkins, C., Gonzalez, N., Docampo, J., Bruno, C., & Stamatakis, E. A. (2020). Anterior cingulate
and medial prefrontal cortex response to systematically controlled tonal dissonance during passive music listening.
Human Brain Mapping, 41(1), 46–66. https://doi.org/10.1002/hbm.24786
Bregman, A. S. (1994). Auditory scene analysis: The perceptual organization of sound. MIT Press.
Brieber, D., Nadal, M., Leder, H., & Rosenberg, R. (2014). Art in time and space: Context modulates the relation
between art experience and viewing time. PLoS One, 9(6), e99019. https://doi.org/10.1371/journal.pone.0099019
Caclin, A., Brattico, E., Tervaniemi, M., Näätänen, R., Morlet, D., Giard, M. H., & McAdams, S. (2006). Separate neu-
ral processing of timbre dimensions in auditory sensory memory. Journal of Cognitive Neuroscience, 18(12), 1959–1972.
https://doi.org/10.1162/jocn.2006.18.12.1959
Cheever, T., Taylor, A., Finkelstein, R., Edwards, E., Thomas, L., Bradt, J., Holochwost, S. J., Johnson, J. K., Limb,
C., Patel, A. D., Tottenham, N., Iyengar, S., Rutter, D., Fleming, R.,  & Collins, F. S. (2018). NIH/Kennedy

329
Amy M. Belfi and Psyche Loui

center workshop on music and the brain: Finding harmony. Neuron, 97(6), 1214–1218. https://doi.org/10.1016/j.
neuron.2018.02.004
Chen, J. L., Penhune, V. B., & Zatorre, R. J. (2008). Listening to musical rhythms recruits motor regions of the brain.
Cerebral Cortex, 18(12), 2844–2854. https://doi.org/10.1093/cercor/bhn042
Chen, J. L., Zatorre, R. J., & Penhune, V. B. (2006). Interactions between auditory and dorsal premotor cortex during syn-
chronization to musical rhythms. NeuroImage, 32(4), 1771–1781. https://doi.org/10.1016/j.neuroimage.2006.04.207
Cirelli, L. K., Einarson, K. M., & Trainor, L. J. (2014). Interpersonal synchrony increases prosocial behavior in infants.
Developmental Science, 17(6), 1003–1011. https://doi.org/10.1111/desc.12193
Cong, F., Puoliväli, T., Alluri, V., Sipola, T., Burunat, I., Toiviainen, P., Nandi, A. K., Brattico, E., & Ristaniemi, T.
(2014). Key issues in decomposing fMRI during naturalistic and continuous music experience with independent
component analysis. Journal of Neuroscience Methods, 223, 74–84. https://doi.org/10.1016/j.jneumeth.2013.11.025
Coutinho, E., & Scherer, K. R. (2017). The effect of context and audio-visual modality on emotions elicited by a musical
performance. Psychology of Music, 45(4), 550–569. https://doi.org/10.1177/0305735616670496
Cowen, A. S., Fang, X., Sauter, D., & Keltner, D. (2020). What music makes us feel: At least 13 dimensions organize
subjective experiences associated with music across different cultures. Proceedings of the National Academy of Sciences of
the United States of America, 117(4), 1924–1934. https://doi.org/10.1073/pnas.1910704117
Cross, E. S., Kirsch, L., Ticini, L. F., & Schütz-Bosbach, S. (2011). The impact of aesthetic evaluation and physical ability
on dance perception. Frontiers in Human Neuroscience, 5, 102. https://doi.org/10.3389/fnhum.2011.00102
Cuddy, L. L., Balkwill, L. L., Peretz, I.,  & Holden, R. R. (2005, March). Musical difficulties are rare: A  study of
“tone deafness” among university students. Annals of the New York Academy of Sciences, 1060, 311–324. https://doi.
org/10.1196/annals.1360.026
Desain, P., & Honing, H. (2003). The formation of rhythmic categories and metric priming. Perception, 32, 341–365.
Deutsch, D., Henthorn, T., Marvin, E., & Xu, H. (2006). Absolute pitch among American and Chinese conservatory
students: Prevalence differences, and evidence for a speech-related critical period. Journal of the Acoustical Society of
America, 119(2), 719–722. https://doi.org/10.1121/1.2151799
Dikker, S., Wan, L., Davidesco, I., Kaggen, L., Oostrik, M., McClintock, J., Rowland, J., Michalareas, G., Van Bavel,
J. J., Ding, M., & Poeppel, D. (2017). Brain-to-brain synchrony tracks real-world dynamic group interactions in the
classroom. Current Biology, 27(9), 1375–1380. https://doi.org/10.1016/j.cub.2017.04.002
Dohn, A., Garza-Villarreal, E. A., Heaton, P., & Vuust, P. (2012). Do musicians with perfect pitch have more autism
traits than musicians without perfect pitch? an empirical study. PLoS One, 7(5), e37961. https://doi.org/10.1371/
journal.pone.0037961
Douglas, K. M., & Bilkey, D. K. (2007). Amusia is associated with deficits in spatial processing. Nature Neuroscience, 10(7),
915–921. https://doi.org/10.1038/nn1925
Downey, L. E., Blezat, A., Nicholas, J., Omar, R., Golden, H. L., Mahoney, C. J., Crutch, S. J.,  & Warren, J. D.
(2013). Mentalising music in frontotemporal dementia. Cortex, 49(7), 1844–1855. https://doi.org/10.1016/j.
cortex.2012.09.011
Drayna, D., Manichaikul, A., De Lange, M., Snieder, H., & Spector, T. (2001). Genetic correlates of musical pitch rec-
ognition in humans. Science, 291(5510), 1969–1972. https://doi.org/10.1126/science.291.5510.1969
Egermann, H., Pearce, M. T., Wiggins, G. A., & McAdams, S. (2013). Probabilistic models of expectation violation
predict psychophysiological emotional responses to live concert music. Cognitive, Affective and Behavioral Neuroscience,
13(3), 533–553. https://doi.org/10.3758/s13415-013-0161-y
Fiveash, A., Thompson, W. F., Badcock, N. A., & McArthur, G. (2018, May). Syntactic processing in music and lan-
guage: Effects of interrupting auditory streams with alternating timbres. International Journal of Psychophysiology, 129,
31–40. https://doi.org/10.1016/j.ijpsycho.2018.05.003
Flinker, A., Doyle, W. K., Mehta, A. D., Devinsky, O., & Poeppel, D. (2019). Spectrotemporal modulation provides a uni-
fying framework for auditory cortical asymmetries. Nature Human Behaviour, 3, 393–405. https://doi.org/10.1016/j.
physbeh.2017.03.040
Foo, F., King-Stephens, D., Weber, P., Laxer, K., Parvizi, J., & Knight, R. T. (2016, April). Differential processing of
consonance and dissonance within the human superior temporal gyrus. Frontiers in Human Neuroscience, 10, 1–12.
https://doi.org/10.3389/fnhum.2016.00154
Fritz, T., Jentschke, S., Gosselin, N., Sammler, D., Peretz, I., Turner, R., Friederici, A. D., & Koelsch, S. (2009). Uni-
versal recognition of three basic emotions in music. Current Biology, 19(7), 573–576. https://doi.org/10.1016/j.
cub.2009.02.058
Fröber, K., & Thomaschke, R. (2019). In the dark cube: Movie theater context enhances the valuation and aesthetic
experience of watching films. Psychology of Aesthetics, Creativity, and the Arts, 15(3), 528–544. https://doi.org/10.1037/
aca0000295

330
The music system

Frühholz, S., Trost, W., & Grandjean, D. (2014). The role of the medial temporal limbic system in processing emotions
in voice and music. Progress in Neurobiology, 123, 1–17. https://doi.org/10.1016/j.pneurobio.2014.09.003
Fujii, S.,  & Schlaug, G. (2013, November). The Harvard beat assessment test (H-BAT): A  battery for assessing beat
perception and production and their dissociation. Frontiers in Human Neuroscience, 7, 771. https://doi.org/10.3389/
fnhum.2013.00771
Fujioka, T., Trainor, L. J., Large, E. W., & Ross, B. (2009). Beta and gamma rhythms in human auditory cortex dur-
ing musical beat processing. Annals of the New York Academy of Sciences, 1169, 89–92. https://doi.org/10.1111/
j.1749-6632.2009.04779.x
Gold, B. P., Mas-Herrero, E., Zeighami, Y., Benovoy, M., Dagher, A., & Zatorre, R. J. (2019). Musical reward predic-
tion errors engage the nucleus accumbens and motivate learning. Proceedings of the National Academy of Sciences, 116(8),
3310–3315. https://doi.org/10.1073/PNAS.1809855116
Gordon, C. L., Cobb, P. R., & Balasubramaniam, R. (2018). Recruitment of the motor system during music listening:
An ALE meta-analysis of fMRI data. PLoS One, 13(11), e0207213. https://doi.org/10.1371/journal.pone.0207213
Gosselin, N., Peretz, I., Hasboun, D., Baulac, M.,  & Samson, S. (2011). Impaired recognition of musical emotions
and facial expressions following anteromedial temporal lobe excision. Cortex, 47(9), 1116–1125. https://doi.
org/10.1016/j.cortex.2011.05.012
Gosselin, N., Peretz, I., Johnsen, E., & Adolphs, R. (2007). Amygdala damage impairs emotion recognition from music.
Neuropsychologia, 45(2), 236–244. https://doi.org/10.1016/j.neuropsychologia.2006.07.012
Gosselin, N., Peretz, I., Noulhiane, M., Hasboun, D., Beckett, C., Baulac, M., & Samson, S. (2005). Impaired recogni-
tion of scary music following unilateral temporal lobe excision. Brain, 128(3), 628–640. https://doi.org/10.1093/
brain/awh420
Grahn, J. A., & Brett, M. (2007). Rhythm and beat perception in motor areas of the brain. Journal of Cognitive Neurosci-
ence, 19(5), 893–906. https://doi.org/10.1162/jocn.2007.19.5.893
Grahn, J. A., & Brett, M. (2009). Impairment of beat-based rhythm discrimination in Parkinson’s disease. Cortex, 45(1),
54–61. https://doi.org/10.1016/j.cortex.2008.01.005
Grahn, J. A., & Rowe, J. B. (2009). Feeling the beat: Premotor and striatal interactions in musicians and nonmusicians
during beat perception. Journal of Neuroscience, 29(23), 7540–7548. https://doi.org/10.1523/JNEUROSCI.2018-
08.2009
Grahn, J. A., & Schuit, D. (2012). Individual differences in rhythmic ability: Behavioral and neuroimaging investigations.
Psychomusicology: Music, Mind, and Brain, 22(2), 105–121. https://doi.org/10.1037/a0031188
Griffiths, T. D., Kumar, S., Warren, J. D., Stewart, L., Stephan, K. E., & Friston, K. J. (2007). Approaches to the cortical
analysis of auditory objects. Hearing Research, 229(1–2), 46–53. https://doi.org/10.1016/j.heares.2007.01.010
Guo, S., & Koelsch, S. (2016, January). Effects of veridical expectations on syntax processing in music: Event-related
potential evidence. Scientific Reports, 6, 19064. https://doi.org/10.1038/srep19064
Henrich, J., Heine, S. J.,  & Norenzayan, A. (2010). The weirdest people in the world? Behavioral and Brain Sciences,
33(2–3), 61–83; discussion 83. https://doi.org/10.1017/S0140525X0999152X
Herrera-Arcos, G., Tamez-Duque, J., Acosta-De-Anda, E. Y., Kwan-Loo, K., de-Alba, M., Tamez-Duque, U.,
Contreras-Vidal, J. L., & Soto, R. (2017, November). Modulation of neural activity during guided viewing of visual
art. Frontiers in Human Neuroscience, 11, 581. https://doi.org/10.3389/fnhum.2017.00581
Hickok, G., & Poeppel, D. (2007). The cortical organization of speech processing. Nature Neuroscience, 8(5), 393–402.
https://doi.org/10.1038/nrn2113
Hohmann, A., Loui, P., Li, C. H., & Schlaug, G. (2018, January). Reverse engineering tone-deafness: Disrupting pitch-
matching by creating temporary dysfunctions in the auditory-motor network. Frontiers in Human Neuroscience, 12, 9.
https://doi.org/10.3389/fnhum.2018.00009
Hove, M. J., Marie, C., Bruce, I. C., & Trainor, L. J. (2014). Superior time perception for lower musical pitch explains
why bass-ranged instruments lay down musical rhythms. Proceedings of the National Academy of Sciences, 111(28),
10383–10388. https://doi.org/10.1073/pnas.1402039111
Hyde, K. L., Lerch, J. P., Zatorre, R. J., Griffiths, T. D., Evans, A. C., & Peretz, I. (2007). Cortical thickness in con-
genital amusia: When less is better than more. Journal of Neuroscience, 27(47), 13028–13032. https://doi.org/10.1523/
JNEUROSCI.3039-07.2007
Hyde, K. L., Zatorre, R. J., Griffiths, T. D., Lerch, J. P., & Peretz, I. (2006). Morphometry of the amusic brain: A two-
site study. Brain, 129(10), 2562–2570. https://doi.org/10.1093/brain/awl204
Jacoby, N., Margulis, E. H., Clayton, M., Hannon, E., Honing, H., Iversen, J., Klein, T. R., Mehr, S. A., Pearson, L.,
Peretz, I., Perlman, M., Polak, R., Ravignani, A., Savage, P. E., Steingo, G., Stevens, C. J., Trainor, L., Trehub, S.,
Veal, M., & Wald-Fuhrmann, M. (2020). Cross-cultural work in music cognition. Music Perception, 37(3), 185–195.
https://doi.org/10.1525/MP.2020.37.3.185

331
Amy M. Belfi and Psyche Loui

Jacoby, N.,  & McDermott, J. H. (2017). Integer ratio priors on musical rhythm revealed cross-culturally by iterated
reproduction. Current Biology, 27(3), 359–370. https://doi.org/10.1016/j.cub.2016.12.031
Jacoby, N., Undurraga, E. A., McPherson, M. J., Valdés, J., Ossandón, T.,  & McDermott, J. H. (2019). Universal
and non-universal features of musical pitch perception revealed by singing. Current Biology, 29(19), 3229–3243.e12.
https://doi.org/10.1016/j.cub.2019.08.020
Janata, P., Birk, J. L., Van Horn, J. D., Leman, M., Tillmann, B.,  & Bharucha, J. J. (2002). The cortical topogra-
phy of tonal structures underlying Western music. Science, 298(5601), 2167–2170. https://doi.org/10.1126/science.
1076262
Joris, P. X., & Verschooten, E. (2013). On the limit of neural phase locking to fine structure in humans. In B. Moore,
R. Patterson, I. Winter, R. Carlyon, & H. Gockel (Eds.), Basic aspects of hearing, advances in experimental medicine and
biology. Springer.
Juslin, P. N. (2013). From everyday emotions to aesthetic emotions: Towards a unified theory of musical emotions. Physics
of Life Reviews, 10(3), 235–266. https://doi.org/10.1016/j.plrev.2013.05.008
Kim, S. G., Kim, J. S., & Chung, C. K. (2011). The effect of conditional probability of chord progression on brain
response: An MEG study. PloS One, 6(2), e17337. https://doi.org/10.1371/journal.pone.0017337
Kim, S. G., Lepsien, J., Fritz, T. H., Mildner, T., & Mueller, K. (2017). Dissonance encoding in human inferior col-
liculus covaries with individual differences in dislike of dissonant music. Scientific Reports, 7(1), 5726. https://doi.
org/10.1038/s41598-017-06105-2
Kim, S. G., Mueller, K., Lepsien, J., Mildner, T., & Fritz, T. H. (2019). Brain networks underlying aesthetic appreciation
as modulated by interaction of the spectral and temporal organisations of music. Scientific Reports, 9(1), 19446. https://
doi.org/10.1038/s41598-019-55781-9
Klein, M., Coles, M. G. H., & Donchin, E. (1984). People with absolute pitch process tones without producing a P300.
Science, 223(4642), 1306–1309. https://doi.org/10.1126/science.223.4642.1306
Koelsch, S., Gunter, T., Friederici, A. D., & Schröger, E. (2000). Brain indices of music processing: “nonmusicians”
are musical. Journal of Cognitive Neuroscience, 12(3), 520–541. https://doi.org/10.1162/089892900562183, PubMed:
10931776
Koelsch, S., Rohrmeier, M., Torrecuso, R.,  & Jentschke, S. (2013). Processing of hierarchical syntactic structure in
music. Proceedings of the National Academy of Sciences of the United States of America, 110(38), 15443–15448. https://doi.
org/10.1073/pnas.1300272110
Koelsch, S., Skouras, S., Fritz, T., Herrera, P., Bonhage, C., Küssner, M. B., & Jacobs, A. M. (2013). The roles of super-
ficial amygdala and auditory cortex in music-evoked fear and joy. NeuroImage, 81, 49–60. https://doi.org/10.1016/j.
neuroimage.2013.05.008
Koelsch, S., Skouras, S., & Lohmann, G. (2018). The auditory cortex hosts network nodes influential for emotion pro-
cessing: An fMRI study on music-evoked fear and joy. PLoS One, 13(1), e0190057. https://doi.org/10.1371/journal.
pone.0190057
Kornysheva, K., Von Cramon, D. Y., Jacobsen, T., & Schubotz, R. I. (2010). Tuning-in to the beat: Aesthetic apprecia-
tion of musical rhythms correlates with a premotor activity boost. Human Brain Mapping, 31(1), 48–64. https://doi.
org/10.1002/hbm.20844
Krumhansl, C. L.,  & Kessler, E. J. (1982). Tracing the dynamic changes in perceived tonal organization in a spatial
representation of musical keys. Psychological Review, 89(4), 334–368. https://doi.org/10.1037/0033-295X.89.4.334,
PubMed: 7134332
Krumhansl, C. L.,  & Shepard, R. N. (1979). Quantification of the hierarchy of tonal functions within a dia-
tonic context. Journal of Experimental Psychology: Human Perception and Performance, 5(4), 579–594. https://doi.
org/10.1037//0096-1523.5.4.579
Lagrois, M. É., & Peretz, I. (2019). The co-occurrence of pitch and rhythm disorders in congenital amusia. Cortex, 113,
229–238. https://doi.org/10.1016/j.cortex.2018.11.036
Lakatos, P., Karmos, G., Mehta, A. D., Ulbert, I., & Schroeder, C. E. (2008). Entrainment of neuronal oscillations as a
mechanism of attentional selection. Science, 320(5872), 110–113. https://doi.org/10.1126/science.1154735
Large, E. W., Herrera, J. A., & Velasco, M. J. (2015, November). Neural networks for beat perception in musical rhythm.
Frontiers in Systems Neuroscience, 9, 1–14. https://doi.org/10.3389/fnsys.2015.00159
Launay, J., Grube, M., & Stewart, L. (2014, February). Dysrhythmia: A specific congenital rhythm perception deficit.
Frontiers in Psychology, 5, 1–8. https://doi.org/10.3389/fpsyg.2014.00018
Lee, D. J., Jung, H., & Loui, P. (2019). Attention modulates electrophysiological responses to simultaneous music and
language syntax processing. Brain Sciences, 9(11). https://doi.org/10.3390/brainsci9110305
Leow, L. A., Parrott, T., & Grahn, J. A. (2014, October). Individual differences in beat perception affect gait responses to
low- and high-groove music. Frontiers in Human Neuroscience, 8, 1–12. https://doi.org/10.3389/fnhum.2014.00811

332
The music system

Lévêque, Y., Teyssier, P., Bouchet, P., Bigand, E., Caclin, A.,  & Tillmann, B. (2018). Musical emotions in congeni-
tal amusia: Impaired recognition, but preserved emotional intensity. Neuropsychology, 32(7), 880–894. https://doi.
org/10.1037/neu0000461
Levitin, D. J. (1994). Absolute memory for musical pitch: Evidence from the production of learned melodies. Perception
and Psychophysics, 56(4), 414–423. https://doi.org/10.3758/BF03206733
Licklider, J. C. R. (1951). A duplex theory of pitch perception. The Journal of the Acoustical Society of America, 23, 147–
147. https://doi.org/10.1007/BF02156143
Liu, F., Patel, A. D., Fourcin, A., & Stewart, L. (2010). Intonation processing in congenital amusia: Discrimination, iden-
tification and imitation. Brain, 133(6), 1682–1693. https://doi.org/10.1093/brain/awq089
Lolli, S. L., Lewenstein, A. D., Basurto, J., Winnik, S., & Loui, P. (2015, September). Sound frequency affects speech
emotion perception: Results from congenital amusia. Frontiers in Psychology, 6, 1–10. https://doi.org/10.3389/
fpsyg.2015.01340
Lomber, S. G., & Malhotra, S. (2008). Double dissociation of “what” and “where” processing in auditory cortex. Nature
Neuroscience, 11(5), 609–616. https://doi.org/10.1038/nn.2108
Longuet-Higgins, H. C., & Lee, C. S. (1982). The perception of musical rhythms. Perception, 11(2), 115–128. https://
doi.org/10.1068/p110115
Loui, P. (2015). A  dual-stream neuroanatomy of singing. Music Perception, 32(3), 232–241. https://doi.org/10.1525/
mp.2015.32.3.232
Loui, P., Alsop, D., & Schlaug, G. (2009). Tone deafness: A new disconnection syndrome? Journal of Neuroscience: the Offi-
cial Journal of the Society for Neuroscience, 29(33), 10215–10220. https://doi.org/10.1523/JNEUROSCI.1701-09.2009
Loui, P., Grent-’t-Jong, T., Torpey, D., & Woldorff, M. (2005). Effects of attention on the neural processing of harmonic
syntax in Western music. Cognitive Brain Research, 25(3), 678–687. https://doi.org/10.1016/j.cogbrainres.2005.08.019
Loui, P., Guenther, F. H., Mathys, C., & Schlaug, G. (2008). Action-perception mismatch in tone-deafness. Current Biol-
ogy, 18(8), R331–R332. https://doi.org/10.1016/j.cub.2008.02.045
Loui, P., Kroog, K., Zuk, J., Winner, E., & Schlaug, G. (2011, May). Relating pitch awareness to phonemic awareness
in children: Implications for tone-deafness and dyslexia. Frontiers in Psychology, 2, 1–5. https://doi.org/10.3389/
fpsyg.2011.00111
Loui, P., Li, H. C. C., Hohmann, A., & Schlaug, G. (2011). Enhanced cortical connectivity in absolute pitch musicians.
Journal of Cognitive Neuroscience, 23(4), 1015–1026. https://doi.org/10.1162/jocn.2010.21500
Loui, P., Patterson, S., Sachs, M. E., Leung, Y., Zeng, T., & Przysinda, E. (2017, September). White matter correlates
of musical anhedonia: Implications for evolution of music. Frontiers in Psychology, 8, 1–10. https://doi.org/10.3389/
fpsyg.2017.01664
Loui, P., Zamm, A., & Schlaug, G. (2012). Enhanced functional networks in absolute pitch. NeuroImage, 63(2), 632–640.
https://doi.org/10.1016/j.neuroimage.2012.07.030
Mantione, M., Figee, M.,  & Denys, D. (2014, May). A  case of musical preference for Johnny Cash following deep
brain stimulation of the nucleus accumbens. Frontiers in Behavioral Neuroscience, 8, 152. https://doi.org/10.3389/
fnbeh.2014.00152
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J.,  & Marco-Pallarés, J. (2016). Neu-
ral correlates of specific musical anhedonia. Proceedings of the National Academy of Sciences, 201611211. https://doi.
org/10.1073/PNAS.1611211113
Mas-Herrero, E., Dagher, A., & Zatorre, R. J. (2018). Modulating musical reward sensitivity up and down with tran-
scranial magnetic stimulation. Nature Human Behaviour, 2(1), 27–32. https://doi.org/10.1038/s41562-017-0241-z
McAdams, S. (2013). Musical timbre perception. In The Psychology of music. Elsevier.
McDermott, J. H., Schultz, A. F., Undurraga, E. A.,  & Godoy, R. A. (2016). Indifference to dissonance in native
Amazonians reveals cultural variation in music perception. Nature, 535(7613), 547–550. https://doi.org/10.1038/
nature18635
Menon, V., & Levitin, D. J. (2005). The rewards of music listening: Response and physiological connectivity of the mes-
olimbic system. NeuroImage, 28(1), 175–184. https://doi.org/10.1016/j.neuroimage.2005.05.053
Merrill, J., Czepiel, A., Fink, L. T., Toelle, J., & Wald-Fuhrmann, M. (2020). The aesthetic experience of live con-
certs: Self-reports and psychophysiology. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/
aca0000390
Meyer, L. B. (1956). Emotion and meaning in music. University of Chicago Press.
Micheyl, C., Delhommeau, K., Perrot, X., & Oxenham, A. J. (2006). Influence of musical and psychoacoustical training
on pitch discrimination. Hearing Research, 219(1–2), 36–47. https://doi.org/10.1016/j.heares.2006.05.004
Miyazaki, K. (1989). Absolute pitch identification: Effects of timbre and pitch region. Music Perception, 7(1), 1–14.
https://doi.org/10.2307/40285445

333
Amy M. Belfi and Psyche Loui

Moore, B. C. (1985). Frequency selectivity and temporal resolution in normal and hearing-impaired listeners. British
Journal of Audiology, 19(3), 198–201 https://doi.org/10.3109/03005368509078973
Mottron, L., Bouvet, L., Bonnel, A., Samson, F., Burack, J. A., Dawson, M., & Heaton, P. (2013). Veridical mapping in
the development of exceptional autistic abilities. Neuroscience and Biobehavioral Reviews, 37(2), 209–228. https://doi.
org/10.1016/j.neubiorev.2012.11.016
Narmour, E. (1990). The analysis and cognition of basic melodic structures: The implication-realization model. Chicago: Univer-
sity of Chicago Press.
Narmour, E. (1992). The analysis and cognition of melodic complexity: The implication-realization model. Chicago: University
of Chicago Press.
Nozaradan, S., Peretz, I., Missal, M., & Mouraux, A. (2011). Tagging the neuronal entrainment to beat and meter. Journal
of Neuroscience, 31(28), 10234–10240. https://doi.org/10.1523/JNEUROSCI.0411-11.2011
Omigie, D., Lehongre, K., Navarro, V., Adam, C., & Samson, S. (2020). Neuro-oscillatory tracking of low- and high-
level musico-acoustic features during naturalistic music listening: Insights from an intracranial electroencephalography
study. Psychomusicology: Music, Mind, and Brain, 30(1), 37–51. https://doi.org/10.1037/pmu0000249
Oxenham, A. J. (2013). Revisiting place and temporal theories of pitch. Acoustical Science and Technology, 34(6), 388–396.
https://doi.org/10.1250/ast.34.388
Pagès-Portabella, C., & Toro, J. M. (2020). Dissonant endings of chord progressions elicit a larger ERAN than ambiguous
endings in musicians. Psychophysiology, 57(2), e13476. https://doi.org/10.1111/psyp.13476
Paquette, S., Takerkart, S., Saget, S., Peretz, I., & Belin, P. (2018). Cross-classification of musical and vocal emotions
in the auditory cortex. Annals of the New York Academy of Sciences, 1423(1), 329–337. https://doi.org/10.1111/nyas.
13666
Patel, A. D. (2006). Musical rhythm, linguistic rhythm, and human evolution. Music Perception, 24(1), 99–104. https://
doi.org/10.1525/mp.2006.24.1.99
Peretz, I. (2016). Neurobiology of congenital amusia. Trends in Cognitive Sciences, 20(11), 857–867. https://doi.
org/10.1016/j.tics.2016.09.002
Peretz, I., Brattico, E., Järvenpää, M., & Tervaniemi, M. (2009). The amusic brain: In tune, out of key, and unaware.
Brain, 132(5), 1277–1286. https://doi.org/10.1093/brain/awp055
Peretz, I., Champod, A. S., & Hyde, K. (2003). Varieties of musical disorders: The Montreal battery of evaluation of
amusia. Annals of the New York Academy of Sciences, 999, 58–75. https://doi.org/10.1196/annals.1284.006
Pfeifer, J., & Hamann, S. (2015). Revising the diagnosis of congenital amusia with the Montreal battery of evaluation of
amusia. Frontiers in Human Neuroscience, 9, 1–15. https://doi.org/10.3389/fnhum.2015.00161
Phillips-Silver, J., Toiviainen, P., Gosselin, N., & Peretz, I. (2013). Amusic does not mean unmusical: Beat perception
and synchronization ability despite pitch deafness. Cognitive Neuropsychology, 30(5), 311–331. https://doi.org/10.108
0/02643294.2013.863183
Phillips-Silver, J., Toiviainen, P., Gosselin, N., Piché, O., Nozaradan, S., Palmer, C.,  & Peretz, I. (2011). Born to
dance but beat deaf: A new form of congenital amusia. Neuropsychologia, 49(5), 961–969. https://doi.org/10.1016/j.
neuropsychologia.2011.02.002
Povel, D. J.,  & Essens, P. (1985). Perception of temporal patterns. Music Perception, 2(4), 411–440. https://doi.
org/10.2307/40285311
Pralus, A., Belfi, A., Hirel, C., Lévêque, Y., Fornoni, L., Bigand, E., Jung, J., Tranel, D., Nighoghossian, N., Tillmann,
B., & Caclin, A. (2020). Recognition of musical emotions and their perceived intensity after unilateral brain damage.
Cortex, 130, 78–93. https://doi.org/10.1016/j.cortex.2020.05.015
Rauschecker, J. P., & Tian, B. (2000). Mechanisms and streams for processing of “what” and “where” in auditory cor-
tex. Proceedings of the National Academy of Sciences of the United States of America, 97(22), 11800–11806. https://doi.
org/10.1073/pnas.97.22.11800
Ross, D. A., Gore, J. C., & Marks, L. E. (2005). Absolute pitch: Music and beyond. Epilepsy and Behavior, 7(4), 578–601.
https://doi.org/10.1016/j.yebeh.2005.05.019
Sachs, M. E., Ellis, R. J., Schlaug, G., & Loui, P. (2016). Brain connectivity reflects human aesthetic responses to music.
Social Cognitive and Affective Neuroscience, 11(6), 884–891. https://doi.org/10.1093/scan/nsw009
Sachs, M. E., Habibi, A., Damasio, A., & Kaplan, J. T. (2020). Dynamic intersubject neural synchronization reflects affec-
tive responses to sad music. NeuroImage, 218, 116512. https://doi.org/10.1016/j.neuroimage.2019.116512
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059

334
The music system

Sammler, D., Koelsch, S., & Friederici, A. D. (2011). Are left fronto-temporal brain areas a prerequisite for normal music-
syntactic processing? Cortex, 47(6), 659–673. https://doi.org/10.1016/j.cortex.2010.04.007
Sanfilippo, K. R. M., Spiro, N., Molina-Solana, M., & Lamont, A. (2020). Do the shuffle: Exploring reasons for music
listening through shuffled play. PLoS One, 15(2), e0228457. https://doi.org/10.1371/journal.pone.0228457
Särkämö, T., Tervaniemi, M., Soinila, S., Autti, T., Silvennoinen, H. M., Laine, M., & Hietanen, M. (2009). Cognitive
deficits associated with acquired amusia after stroke: A neuropsychological follow-up study. Neuropsychologia, 47(12),
2642–2651. https://doi.org/10.1016/j.neuropsychologia.2009.05.015
Satoh, M., Kato, N., Tabei, K. I., Nakano, C., Abe, M., Fujita, R., Kida, H., Tomimoto, H., & Kondo, K. (2016). A case
of musical anhedonia due to right putaminal hemorrhage: A disconnection syndrome between the auditory cortex
and insula. Neurocase, 22(6), 518–525. https://doi.org/10.1080/13554794.2016.1264609
Schafer, K., & Eerola, T. (2020). How listening to music and engagement with other media provide a sense of belonging: An
exploratory study of social surrogacy. Psychology of Music, 48(2), 232–251. https://doi.org/10.1177/0305735618795036
Schlaug, G., Jäncke, L., Huang, Y., & Steinmetz, H. (1995). In vivo evidence of structural brain asymmetry in musicians.
Science, 267(5198), 699–701. https://doi.org/10.1126/science.7839149
Sethares, W. (2004). Tuning timbre Spetrum scale. Springer.
Shany, O., Singer, N., Gold, B. P., Jacoby, N., Tarrasch, R., Hendler, T., & Granot, R. (2019). Surprise-related activation
in the nucleus accumbens interacts with music-induced pleasantness. Social Cognitive and Affective Neuroscience, 14(4),
459–470. https://doi.org/10.1093/SCAN/NSZ019
Shepard, R. N. (1980). Multidimensional scaling, tree-fitting, and clustering. Science, 210(4468), 390–398. https://doi.
org/10.1126/science.210.4468.390
Siedenburg, K., & McAdams, S. (2017, October). Four distinctions for the auditory “wastebasket” of timbre. Frontiers in
Psychology, 8, 6–9. https://doi.org/10.3389/fpsyg.2017.01747
Sihvonen, A. J., Ripollés, P., Leo, V., Rodríguez-Fornells, A., Soinila, S.,  & Särkämö, T. (2016). Neural basis of
acquired amusia and its recovery after stroke. Journal of Neuroscience, 36(34), 8872–8881. https://doi.org/10.1523/
JNEUROSCI.0709-16.2016
Sihvonen, A. J., Ripollés, P., Särkämö, T., Leo, V., Rodríguez-Fornells, A., Saunavaara, J., Parkkola, R., & Soinila, S.
(2017). Tracting the neural basis of music: Deficient structural connectivity underlying acquired amusia. Cortex, 1–19.
https://doi.org/10.1016/j.cortex.2017.09.028
Steinbeis, N.,  & Koelsch, S. (2009). Understanding the intentions behind man-made products elicits neural activity
in areas dedicated to mental state attribution. Cerebral Cortex, 19(3), 619–623. https://doi.org/10.1093/cercor/
bhn110
Stewart, L., von Kriegstein, K., Warren, J. D., & Griffiths, T. D. (2006). Music and the brain: Disorders of musical listen-
ing. Brain, 129(Pt 10), 2533–2553. https://doi.org/10.1093/brain/awl171
Stupacher, J., Hove, M. J., Novembre, G., Schütz-Bosbach, S.,  & Keller, P. E. (2013). Musical groove modulates
motor cortex excitability: A  TMS investigation. Brain and Cognition, 82(2), 127–136. https://doi.org/10.1016/j.
bandc.2013.03.003
Sun, Y., Lu, X., Ho, H. T., Johnson, B. W., Sammler, D., & Thompson, W. F. (2018, May). Syntactic processing in music
and language: Parallel abnormalities observed in congenital amusia. NeuroImage: Clinical, 19, 640–651. https://doi.
org/10.1016/j.nicl.2018.05.032
Tanaka, S.,  & Kirino, E. (2018, November). The parietal opercular auditory-sensorimotor network in musicians:
A resting-state fMRI study. Brain and Cognition, 120(2017), 43–47. https://doi.org/10.1016/j.bandc.2017.11.001
Taruffi, L., Pehrs, C., Skouras, S., & Koelsch, S. (2017). Effects of sad and happy music on mind-wandering and the
default mode network. Scientific Reports, 7(1), 1–10. https://doi.org/10.1038/s41598-017-14849-0
Teki, S., Kumar, S., von Kriegstein, K., Stewart, L., Lyness, C. R., Moore, B. C. J., Capleton, B., & Griffiths, T. D.
(2012). Navigating the auditory scene: An expert role for the hippocampus. Journal of Neuroscience, 32(35), 12251–
12257. https://doi.org/10.1523/JNEUROSCI.0082-12.2012
Tervaniemi, M., Tupala, T., & Brattico, E. (2012). Expertise in folk music alters the brain processing of Western harmony.
Annals of the New York Academy of Sciences, 1252(1), 147–151. https://doi.org/10.1111/j.1749-6632.2011.06428.x
Thompson, W. F., Marin, M. M., & Stewart, L. (2012). Reduced sensitivity to emotional prosody in congenital amusia
rekindles the musical protolanguage hypothesis. Proceedings of the National Academy of Sciences of the United States of
America, 109(46), 19027–19032. https://doi.org/10.1073/pnas.1210344109
Tierney, A., & Kraus, N. (2015). Evidence for multiple rhythmic skills. PLoS One, 10(9), 1–14. https://doi.org/10.1371/
journal.pone.0136645
Tillmann, B., Lévêque, Y., Fornoni, L., Albouy, P., & Caclin, A. (2016). Impaired short-term memory for pitch in con-
genital amusia. Brain Research, 1640(B), 251–263. https://doi.org/10.1016/j.brainres.2015.10.035
Tranchant, P., & Vuvan, D. T. (2015, May). Current conceptual challenges in the study of rhythm processing deficits.
Frontiers in Neuroscience, 9, 8–10. https://doi.org/10.3389/fnins.2015.00197

335
Amy M. Belfi and Psyche Loui

Trost, W., Ethofer, T., Zentner, M., & Vuilleumier, P. (2012). Mapping aesthetic musical emotions in the brain. Cerebral
Cortex, 22(12), 2769–2783. https://doi.org/10.1093/cercor/bhr353
Upham, F., & McAdams, S. (2018). Activity analysis and coordination in continuous responses to music. Music Perception,
35(3), 253–294. https://doi.org/10.1525/mp.2018.35.3.253
Vanden Bosch der Nederlanden, C. M., Joanisse, M. F., & Grahn, J. A. (2020, March). Music as a scaffold for listening
to speech: Better neural phase-locking to song than speech. NeuroImage, 214, 116767. https://doi.org/10.1016/j.
neuroimage.2020.116767
Vaquero, L., Ramos-Escobar, N., François, C., Penhune, V., & Rodríguez-Fornells, A. (2018, June). White-matter struc-
tural connectivity predicts short-term melody and rhythm learning in non-musicians. NeuroImage, 181, 252–262.
https://doi.org/10.1016/j.neuroimage.2018.06.054
Vessel, E. A., Isik, A. I., Belfi, A. M., Stahl, J. L., & Starr, G. G. (2019). The default-mode network represents aesthetic
appeal that generalizes across visual domains. Proceedings of the National Academy of Sciences of the United States of America,
116(38), 19155–19164. https://doi.org/10.1073/pnas.1902650116
Von Bekesy, G. (1960). Experiments in hearing. New York: McGraw-Hill.
Ward, W. D. (1999). Absolute pitch. Psychology of Music, 265–298.
Warren, J. D., Jennings, A. R., & Griffiths, T. D. (2005). Analysis of the spectral envelope of sounds by the human brain.
NeuroImage, 24(4), 1052–1057. https://doi.org/10.1016/j.neuroimage.2004.10.031
Warren, J. D., Uppenkamp, S., Patterson, R. D., & Griffiths, T. D. (2003). Analyzing pitch chroma and pitch height in the
human brain. Annals of the New York Academy of Sciences, 999(17), 212–214. https://doi.org/10.1196/annals.1284.032
Wessel, D. L. (1979). Timbre space as a musical control structure. Computer Music Journal, 3(2), 45. https://doi.
org/10.2307/3680283
Williamson, V. J., Cocchini, G., & Stewart, L. (2011). The relationship between pitch and space in congenital amusia.
Brain and Cognition, 76(1), 70–76. https://doi.org/10.1016/j.bandc.2011.02.016
Winkler, I., Háden, G. P., Ladinig, O., Sziller, I., & Honing, H. (2009). Newborn infants detect the beat in music. Pro-
ceedings of the National Academy of Sciences of the United States of America, 106(7), 2468–2471. https://doi.org/10.1073/
pnas.0809035106
Witek, M. A. G., Clarke, E. F., Wallentin, M., Kringelbach, M. L., & Vuust, P. (2014). Syncopation, body-movement
and pleasure in groove music. PLoS One, 9(4), e94446. https://doi.org/10.1371/journal.pone.0094446
Zatorre, R. J. (2003). Absolute pitch: A model for understanding the influence of genes and development on neural and
cognitive function. Nature Neuroscience, 6(7), 692–695. https://doi.org/10.1038/nn1085
Zatorre, R. J., Bouffard, M., & Belin, P. (2004). Sensitivity to auditory object features in human temporal neocortex.
Journal of Neuroscience, 24(14), 3637–3642. https://doi.org/10.1523/JNEUROSCI.5458-03.2004
Zhang, J., Zhou, X., Chang, R., & Yang, Y. (2018, December). Effects of global and local contexts on chord process-
ing: An ERP study. Neuropsychologia, 109(2017), 149–154. https://doi.org/10.1016/j.neuropsychologia.2017.12.016

336
16
WATCHING AND ENGAGING
IN DANCE
Beatriz Calvo-Merino

Most research in neuroaesthetics has focused on the affective, perceptual and cognitive mechanisms that
participate in the appreciation of painting, music and architecture. Interest in dance and other art forms has
developed more prominently at the beginning of the 21st century. The last two decades have seen a steady
growth of publications focussing on the relationship between the dance and the observer and on the interplay
between the subjective, the cognitive and the neural mechanisms in aesthetic appreciation of this performing
art. Dance reaches the human brain as a coordinated collage of body parts perceived as a human figure that
undergoes displacement in time and space. This chapter presents what is known about the way the observer’s
brain perceives this information, integrates it and engages aesthetically and emotionally with it.

Observing dance
This section describes experimental studies that have outlined the neural, cognitive, physiological and behav-
ioural mechanisms involved in the observation of dance. We will adopt a narrow approach to observing
dance and focus primarily on the visual perception of bodily movements within the performing arts con-
text, executed by a motor expert (i.e., a dancer). Although this may seem overly reductionist, it reflects the
conceptual and methodological approach of the studies that brought dance, dancers and neuroaesthetics of
performing arts into sight for researchers in psychology, philosophy and neuroscience.
The first functional magnetic resonance imaging (fMRI) study of brain activity while observing dance
took place in central London almost 20 years ago. Researchers were experimental psychologists and neuro-
scientists who wished to understand how motor familiarity with an observed action modulated the way our
brain perceives that action (Calvo-Merino et al., 2005). This question was crucial at that time, as it critically
helped shape the working mechanism underlying the mirror neuron theory (Rizzolatti & Craighero, 2004).
The mirror neuron theory was proposed by Giacomo Rizzolatti and his collaborators to provide a framework
for the research evidence suggesting that underlying our observation of other people’s actions are the same
neural correlates we use to perform those actions ourselves (di Pellegrino et al., 1992; Gallese et al., 1996;
Rizzolatti et al., 1996).
Before we concentrate on studies related to the perception of dance, we will briefly revise the main con-
cepts and results of these early discoveries, which changed the way we understand how we see other people’s
actions. In the 90s, in a laboratory in Parma (Italy), Rizzolatti and his collaborators described, for the first
time, a population of neurons in the premotor cortex of the macaque brain that responded to the observation

DOI: 10.4324/9781003008675-18 337


Beatriz Calvo-Merino

of a hand action. The neurons of the premotor cortex were active when the primate performed an action
(the role of the premotor cortex in motor execution is well known; Graziano et al., 2002), but importantly,
the same neurons responded when the monkey observed the experimenter performing the same movement
(di Pellegrino et al., 1992). This early experiment was followed by a series of additional experiments that
replicated these results and explored in detail the characteristics and properties of the premotor area (and
other regions in the parietal cortex) beyond action execution and established the foundations for the concept
of mirror neuron (Gallese et al., 1996; Rizzolatti et al., 1996; Rizzolatti & Craighero, 2004). A mirror neuron
was defined as a neuron that participates in action execution (therefore, we can say it has motor properties),
and also in action observation (it has additional visual properties).
The idea of sensorimotor regions of the brain, normally known for their role in execution, now playing
a role in perception should not be held lightly. This discovery was revolutionary at the time, and since the
beginning of the 21st century, studies set out to determine whether the human brain had neurons in the sen-
sorimotor regions with similar properties. Studies using functional magnetic resonance (Decety et al., 1997;
Grèzes et al., 2001; Grossman et al., 2000), transcranial magnetic stimulation (Fadiga et al., 1995) and elec-
troencephalography (Pineda, 2005) measured brain activity during different tasks related to the perception of
movements and provided evidence of a mechanism with similar sensorimotor resonance in the human brain
(Rizzolatti & Sinigaglia, 2016).
Experimental evidence plays a necessary role in science by providing descriptions of phenomena. In the
case of research on mirror neurons, studies described a shared neural mechanism for performing and seeing
actions (Rizzolatti & Craighero, 2004). Interest increases substantially when scientists provide interpretations
and explanations for the observed data. The potential role of mirror neurons in humans has been contro-
versial for years (Cook et al., 2014). Amongst the more solid potential functions, we find action recogni-
tion, action understanding and action imitation (Oztop et al., 2013; Rizzolatti & Sinigaglia, 2016). When
we learn to perform a motor skill or an action, we acquire vast information that we store in neural regions,
forming a motor representation. The mirror neuron theory suggests that when we observe others’ actions,
we activate our own motor representations, and this helps us understand how it feels or what the observed
action means. If this last statement were true, we would expect differential activation in the mirror neuron
areas when observing actions that we have learnt (and therefore stored in our motor repertoire), as opposed
to observing actions we have not learnt motorically. This is precisely the prediction that set off the first study
on action observation of dance.
We used fMRI to measure the brain activity of participants with different types of acquired motor skills
(expert ballet dancers, expert capoeiristas and non-experts) while they observed videos of ballet and capoeira
movements. The results showed stronger activation in those brain areas attributed to the mirror neuron
system while the observers watched the familiar movement (i.e., when the ballet dancers watched ballet
and the capoeiristas watched capoeira) than when participants watched non-familiar movement. Within
the network of brain regions that were specifically activated when observing familiar movements were the
ventral and dorsal premotor cortex and intraparietal sulcus and superior parietal lobe in the parietal cortex
(Calvo-Merino et al., 2005). These regions are part of what has been described as the action representation
system or action representation network (Decety et al., 1997; Grafton et al., 1996; Grèzes & Decety, 2001;
Cross et al., 2009). These results provided experimental evidence in support of the mirror neuron theory,
which proposed that we activate our own action information and representations when observing others’
actions. In other words, we match what we see to what we have previously learnt and stored in the form of a
motor or action memory in sensorimotor regions. The stronger the match between the observed action and
the stored memory, the stronger the response of the sensorimotor mechanism supporting that representa-
tion. The content of an action representation includes the specific motor commands to perform the action,
sensory and proprioceptive information related to how it feels that performance and semantic information
associated with that action, amongst other types of information ( Jeannerod, 1997).

338
Watching and engaging in dance

The concept of an active sensorimotor engagement during action observation was by no means the con-
clusion of a single study. Other neuroimaging and neurophysiological studies using expertise paradigms in
which dancers observed familiar movements before and after learning them (Cross et al., 2006, 2009) or in
which dancers watched known actions while having their brain activity recorded (Orgs et al., 2008) reached
similar conclusions. Moreover, dissociating the motor from the visual representation that is engaged during
observation has been explored by comparing male and female ballet dancers that observed gender-specific
movements (in classical ballet, many movements are performed by both genders, but there are also some
gender-specific movements, which only women or men perform). All dancers would be expected to have
visual experience with all the movements because they train and work together. However, they would only
have motoric experience for the movements they actually have learnt to perform (Calvo-Merino et  al.,
2006). This result suggested that we activate an internal motor simulation when observing an action, over
and above potential visual and semantic resonances.
This raises the question of whether an expert dancer and nonexpert observer see the same when watch-
ing dance. Studies have examined whether the ability to activate motor representations during observation
modulates not only how brain areas respond but also how we perceive the action at the cognitive and behav-
ioural levels. In other words, can dance expertise change what the observer sees? Experimental psychology
tries to understand how each feature is processed in isolation and as part of a whole. When exploring how
we perceive dance, we need to consider at least two intertwined elements: the body of the dancer and the
movement itself. Both elements have been independently studied in cognitive psychology and neurosci-
ence, and specialized brain regions for body processing (such as the extrastriate body area, Downing et al.,
2001; the fusiform body area, Downing & Peelen, 2011) and movement (superior temporal sulcus, Yovel &
O’Toole, 2016) have been described. These regions also participate in the perception of body dance pos-
tures (Urgesi, Calvo-Merino et al., 2007) and learnt dance movements (Di Nota et al., 2016; Calvo-Merino
et al., 2005).
The dancer’s body can be easily depicted as a still image. But how do we represent the dance movement
without the dancer’s body? Johansson created a technique called point-light display precisely to investigate
motion perception in the absence of body form information ( Johansson, 1973). The point light technique
is the older sister of what we know now as motion capture. Johansson attached dots of light to the main
joints of an actor (ankles, knees, hips, elbows, wrist, shoulders and head) and recorded video clips while they
walked and performed actions in a dark room. The videos depicted just 12 white dots moving in a black
background, and yet, surprisingly, observers watching these videos (and others in similar studies) were able
to identify the action being performed ( Johansson, 1973), the gender of the actor (Cutting & Kozlowski,
1977), the identity of the person performing the action when the actors were familiar to the observers (Loula
et al., 2005), and a significant level of emotion information about the state of the performer, such as if a per-
son walking was in a sad mood (Dittrich et al., 1996). The amount of information with social connotations
that our brain is able to integrate from just the perception of dynamic dots is extraordinary (Troje, 2002).
Returning to our initial question, do dancers see more of the dance because of their previous dance expe-
rience with dance execution? The short answer is yes. Research showed that dancers observing point-light
displays of ballet movements were able to distinguish when these movements (e.g., pas de chat represented
with points of light) were performed by different dancers in a visual discrimination task, while controls with
no experience in dance were unable to tell dancers apart (Calvo-Merino et  al., 2010). Interestingly, this
visual performance advantage was only present when the dance movement was shown in its usual orienta-
tion (upright). When the point lights depicting the movement were presented in an inverted orientation,
this advantage disappeared, and the task became equally difficult for dancers and non-dancers. These results
suggest that expertise can increase visual sensitivity when perceiving stimuli in the same form that we have
learned. Other studies that focused on the importance of visual and motor familiarity in visual motion
perception have devised interesting alternative paradigms. One of the most memorable studies in this area

339
Beatriz Calvo-Merino

compared performance in a visual task depicting biological motion via point-light displays of movements
depicting strangers, friends and the participants themselves (Loula et al., 2005). Interestingly, people are very
good at identifying videos of themselves, strengthening the hypothesis supporting that knowledge about our
learnt actions modulates perception.
This section has described how we observe dance and how being a dancer and acquiring the specific
sensorimotor knowledge that comes along with practice affects dance perception. It has also provided some
experimental evidence supporting this hypothesis. The following sections go beyond mere observation of
dance and explore the neural underpinnings of engagement with dance in the aesthetic domain and in the
emotional domain.

Engaging with dance aesthetically


Dance is a performing art, and, as such, observation of dance is intrinsically linked to the concept of aes-
thetics. Aesthetically engaging with art requires the contribution and interaction of multiple processes that
lay within different brain regions and cognitive mechanisms (Leder et  al., 2004; Leder  & Nadal, 2014;
Chatterjee & Vartanian, 2014). Aesthetically engaging with dance will be no different conceptually from
engaging with other visual art, with one exception. The observation of dance is almost always tied to the
observation of a dancer’s body executing the dance. The dancer’s body can be thought of as the instrument
or the container delivering the dance. This instrument (the body) has a close correspondence in format to
the observer’s body. We are contained in our bodies, which come in different shapes and are trained in differ-
ent skills. However, beyond the obvious visual and abilities differences, the basic physiology and structure of
our bodies is quite similar, and we all have multiple extensions of our bodies distributed in our brains in the
shape of body representations (Longo et al., 2010). We could say that there is an important correspondence
between the observers and the dance container (the body), which in return should influence how we observe
and engage with dance (Calvo-Merino et al., 2010; Christensen & Calvo-Merino, 2013).
We noted previously that neuroscientists have described brain regions particularly tuned to process bodily
information (Downing et al., 2001; Downing & Peelen, 2011; Peelen & Downing, 2005; Urgesi, Calvo-
Merino et al., 2007). Among these, it is worth highlighting two of them that lie in different visual pathways.
First, the extrastriate body area, located in the intersection between occipital and temporal cortices in the
ventral visual pathway, has been associated with processing body form information (Urgesi, Candidi et al.,
2007; Downing et al., 2006). Other regions in the frontal lobe, like the premotor cortex, towards the end
of the dorsal visual pathway, respond to the body movement (Binkofski & Buccino, 2006; Urgesi, Candidi
et  al., 2007). A  simple question that emerges from this knowledge is: Do these body-interested regions
engage aesthetically when watching dance? This was answered in a study using a technique called transcranial
magnetic stimulation (TMS). This technique allows evoking a temporal virtual lesion over a brain area while
participants engage with a task (Miniussi & Ruzzoli, 2013). The study showed that aesthetic evaluation of
dance body postures was affected by the virtual interruption of processing in the premotor and the EBA
areas, suggesting that regions that participate in the visual perception of the body also play a role during the
aesthetic evaluation of these bodies when depicting a dance posture (Calvo-Merino et al., 2010). The role
of these areas in aesthetic body perception has been confirmed in subsequent studies (Cazzato et al., 2012),
overall providing these body-interested brain regions with a fascinating role in aesthetic perception.
Dance is of course more than dance postures. Dance is dynamic and flows in time and space. The obser-
vation of whole dance movements has been explored using behavioural and neuroimaging studies. Calvo-
Merino and colleagues (2008) showed that observing dance movements we like (in contrast to movements
we like less) evokes stronger activation in brain regions related to the body and movement processing, such
as the premotor cortex (in the dorsal area). The premotor cortex is part of the mirror neuron system and
action observation network, and its function during observation is, among others, to create an internal

340
Watching and engaging in dance

simulation of the perceived movement. These results suggest an interesting relationship between simulat-
ing motorically the observed action and preferences. Other studies led by Cross and Kirsh and colleagues
strengthen this conclusion with neural and behavioural evidence. Cross et al. (2011) showed that the stronger
the liking and the perceived difficulty of the dance, the stronger the brain response in posterior regions of
the action observation network, such as occipitotemporal and parietal cortices. Kirsch and colleagues (2013)
used a dance training paradigm to show that behavioural ratings of positive aesthetic evaluations of dance
movements increased with motor familiarity. Overall, experimental evidence indicates an unequivocal rela-
tionship between aesthetic engagement and internal sensorimotor simulation of the perceived movement.
This sensorimotor approach to the aesthetic of dance fits well with Freedberg and Gallese’s (2007) theory of
embodied aesthetics.

Engaging with dance emotionally


Dance and emotion are closely intertwined. Dancers perform movements with a high level of sophistication
that sparks a combination of emotions. These can take shape in positive and negative valence, mixed with
high or low levels of arousal. For example, excitement can carry high arousal values as well as indicating a
positive valence, while despair may evoke similar high arousal values but point toward a much more negative
valence. While other fields of visual cognitive neuroscience have developed sets of stimuli for faces, bod-
ies and movement research on emotion, dance research has only started recently to produce stimuli where
these variables are controlled. Christensen, Nadal and colleagues (2014) published the first normalized video
library that describes a series of genuine classical ballet movements normalized over a series of elements that
include valence, arousal and liking, amongst others. Subsequently, other dance libraries using dance move-
ments (beyond classical ballet) depicting emotions have been made available (Christensen et al., 2019). These
libraries facilitate research on dance observation and emotion perception, allowing different laboratories to
work with similar stimuli, reducing the number of different variables that may affect the process of interest.
At the behavioural level, it has been shown that observers can discriminate the emotion intended to
be transmitted by the dance relatively well (Christensen, Gaigg et al., 2014; Christensen et al., 2016). The
perception of the emotion is enhanced when dance is accompanied by music representing the same valence
(i.e., a movement illustrating a happy emotion accompanied by music depicting the same emotion) (Chris-
tensen, Gaigg et al., 2014). Interestingly, the ability to discriminate emotion in dance is also larger when
observers have expertise with the dance movements they are observing (i.e., dancers judging emotion in
dance movements they know, Christensen et al., 2016). Montero (2012) also proposed that expert dancers
have a better understanding of the observed performance, as they additionally engage with details that may
go unnoticed to the non-dancer untrained eye, such as grace, precision or power. This fits with the idea
proposed earlier suggesting that expertise allows seeing more (Calvo-Merino et al., 2010).
It feels natural to expand the idea that, if expertise allows you to see more when watching dance, it may
also allow you to emotionally engage more when watching dance. To evaluate this statement, we need to move
from describing neural activations and behavioural responses to a measurement intrinsically linked to the
concept of feeling, emotion and arousal: the physiological response. This response has been widely employed
in emotion perception research by noting variations of the skin conductance during the observations of dif-
ferent stimuli that vary on their emotional content. Christensen and colleagues (2016) measured the galvanic
response—skin conductance—of a group of dancers and a group of non-dancers while they observed and
rated the emotions of dance movements. Results show significant differences in the physiological response of
the experts while they observed dance movements that were perceived as happy, as compared with those that
were perceived as sad (i.e., the difference between the perception of positive- and negative-valanced dance
movements). Interestingly, this effect was not found in the control non-expert group and was also not found
in the control video condition (when dance videos were played backwards to keep the same kinematic but

341
Beatriz Calvo-Merino

disrupt the form in which the movement was learnt by the dancers) (Christensen et al., 2016). A general
conclusion from this and similar work (Christensen et al., 2021), in combination with previous studies inves-
tigating how dance expertise modulates different aspects of the perception of dance, suggests that expertise
modulates what we see in the dance and how we feel it.

From observation to embodiment


The investigation of dance observation has grown in parallel to research in action observation and the role
of familiarity in perception. Neuroimaging and cognitive studies describe an internal motor simulation
automatically engaged during action observation and, by inclusion, during dance observation. This motor
representation plays a special role in dance observation, as it lies at the heart of the observer’s expertise. It
has been shown that dance expertise modulates: (1) how our brain responds to dance, (2) how much we see
of the dance and (3) how we feel the dance. Expertise effects in visual perception are not unusual but are
mainly focused on the visual expertise of the observer (Gauthier & Tarr, 1997). The highlight of the dance
and neuroscience studies described in this chapter is that they go beyond visual experience and focus on how
motor expertise influences perception.
The concept of motor simulation, while strongly popular at the beginning of the 21st century, has been
slowly replaced by a more general term that includes other aspects of simulation (beyond motor). This is
the concept of embodiment. Earlier descriptions of this term identified the body as a critical player in any
interaction with the world (Shapiro, 2010). In the domain of observing others, the concept relates to a re-
enactment of the visual, motor and visceral reactions of the observed situation. A clear example is found in
embodied emotions, where a re-enactment of the observed emotion is present in the observer at multiple
levels (Niedenthal, 2007). The concept of embodied emotions could be easily adopted in the emotional
engagement of dance. If we re-interpret the studies we described earlier, one could say experts show stronger
embodiment than non-dancers during emotion recognition of dance. Similarly, we have been entertaining
the idea of sensorimotor aesthetics of performing art. If we adapt this concept to the new terminology, we
land in a field called embodied aesthetics, which, similarly to the idea of embodied emotions, includes a re-
acting of bodily sensations during art perception.

Conclusion: major challenges, goals and suggestions


Neuroscientific and psychological research on dance perception is about to take another important leap. This
is driven mainly by three aspects. First, the advance of mobile technology is finding its way into theatres,
allowing data collection beyond the laboratory. Some studies had already tried to increase the ecological
validity of the studies by collecting data while participants sat in the theatre or observed live dance perfor-
mances, as opposed to videos on a screen ( Jola et al., 2012; Jola & Grosbras, 2013). The results point towards
the same direction as those performed in the laboratory. Portable devices to record electroencephalography
or physiology with high precision are key not only to confirm laboratory data in an ecological setting but to
allow developing new questions related to synchronous watching between observers and synchronous brain
recording between dancer (performer) and observer or to understand the development of learning a new
dance sequence and its potential implications in therapy (e.g., Barnstaple et al., 2020).
The second element refers to the advance of computational analysis able to explore larger and dynamic
datasets. Currently, studies mainly average the neural and behavioural responses to short dance sequences
registered during the experiment. Recent research has highlighted the importance of investigating the
dynamics of subjective experiences over time using a combination of continuous ratings and recordings
(Isik  & Vessel, 2019). This would be crucial to dance perception studies. Dance, as a performing art, is
dynamic, and observers’ engagement is likely to evolve over time. Continuous ratings can provide a reliable

342
Watching and engaging in dance

window into the temporal dynamics of dance perception without interfering with the observer’s experiences
(Isik & Vessel, 2019).
And, finally, research in dance needs to look back at dance. Neuroscientists often make the mistake of
not considering dance movements as individual gems but rather as a large category that can be collapsed to
investigate general questions. Some studies have now explored the kinematics of dance (i.e., Torrents et al.,
2013), how the brain may be tuned to certain kinematic parameters and how modelling of biomechanical
metrics can explain biological determinants of aesthetic perception in dance (Bronner & Shippen, 2015).
Some of the later neuroimaging studies have attempted to look back at the stimulus space and determine the
motor properties of the movements that were preferred by observers and elicited stronger sensorimotor reso-
nance. For example, these studies suggest stronger motor resonance while observing movements with large
displacement or with a certain perceived level of complexity (Calvo-Merino et al., 2010; Cross et al., 2011).
However, we still need a more systematic investigation of how individual kinematic parameters modulate
dance perception.

References
Barnstaple, R., Protzak, J., DeSouza, J. F. X., & Gramann, K. (2020). Mobile brain/body Imaging in dance: A dynamic
transdisciplinary field for applied research. European Journal of Neuroscience, 16. https://doi.org/10.1111/ejn.14866
[Epub ahead of print]. PubMed: 32544262
Binkofski, F., & Buccino, G. (2006). The role of ventral premotor cortex in action execution and action understanding.
Journal of Physiology, Paris, 99(4–6), 396–405. https://doi.org/10.1016/j.jphysparis.2006.03.005. Epub May 24 2006.
PubMed: 16723210
Bronner, S., & Shippen, J. (2015). Biomechanical metrics of aesthetic perception in dance. Experimental Brain Research,
233(12), 3565–3581. https://doi.org/10.1007/s00221-015-4424-4, PubMed: 26319546
Calvo-Merino, B., Glaser, D. E., Grèzes, J., Passingham, R. E., & Haggard, P. (2005). Action observation and acquired
motor skills: An fMRI study with expert dancers. Cerebral Cortex, 15(8), 1243–1249. https://doi.org/10.1093/
cercor/bhi007
Calvo-Merino, B., Grèzes, J., Glaser, D. E., Passingham, R. E., & Haggard, P. (2006). Seeing or doing: Influence of
visual and motor familiarity in action observation. Current Biology, 16(19), 1905–1910. https://doi.org/10.1016/j.
cub.2006.07.065
Calvo-Merino, B., Jola, C., Glaser, D. E., & Haggard, P. (2008). Towards a sensorimotor aesthetics of performing art.
Consciousness and Cognition, 17(3), 911–922. https://doi.org/10.1016/j.concog.2007.11.003
Calvo-Merino, B., Urgesi, C., Orgs, G., Aglioti, S. M., & Haggard, P. (2010). Extrastriate body area underlies aesthetic
evaluation of body stimuli. Experimental Brain Research, 204(3), 447–456. https://doi.org/10.1007/s00221-010-2283-6
Cazzato, V., Siega, S., & Urgesi, C. (2012). “What women like”: Influence of motion and form on esthetic body percep-
tion. Frontiers in Psychology, 3, 235. http://doi.org/10.3389/fpsyg.2012.00235
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. http://doi.
org/10.1016/j.tics.2014.03.003
Christensen, J. F., Azevedo, R. T., & Tsakiris, M. (2021). Emotion matters: Different psychophysiological responses to
expressive and non-expressive full-body movements. Acta Psychologica (Amst), 212, 103215. https://doi.org/10.1016/j.
actpsy.2020.103215
Christensen, J. F., & Calvo-Merino, B. (2013). Dance as a subject for empirical aesthetics. Psychology of Aesthetics, Creativ-
ity, and the Arts, 7(1), 76–88. https://doi.org/10.1037/a0031827
Christensen, J. F., Gaigg, S. B., Gomila, A., Oke, P., & Calvo-Merino, B. (2014). Enhancing emotional experiences to
dance through music: The role of valence and arousal in the cross-modal bias. Frontiers in Human Neuroscience, 8, 757.
https://doi.org/10.3389/fnhum.2014.00757
Christensen, J. F., Gomila, A., Gaigg, S. B., Sivarajah, N.,  & Calvo-Merino, B. (2016). Dance expertise modulates
behavioral and psychophysiological responses to affective body movement. Journal of Experimental Psychology. Human
Perception and Performance, 42(8), 1139–1147. https://doi.org/10.1037/xhp0000176
Christensen, J. F., Lambrechts, A.,  & Tsakiris, M. (2019). The Warburg dance movement library—The WADAMO
library: A validation study. Perception, 48(1), 26–57. https://doi.org/10.1177/0301006618816631
Christensen, J. F., Nadal, M., & Cela-Conde, C. J. (2014). A norming study and library of 203 dance movements. Percep-
tion, 43(2–3), 178–206. http://doi.org/10.1068/p7581

343
Beatriz Calvo-Merino

Cook, R., Bird, G., Catmur, C., Press, C., & Heyes, C. (2014, April). Mirror neurons: From origin to function. Behav-
ioral and Brain Sciences, 37(2), 177–192. https://doi.org/10.1017/S0140525X13000903, PubMed: 24775147
Cross, E. S., Hamilton, A. F., & Grafton, S. T. (2006). Building a motor simulation de novo: Observation of dance by
dancers. NeuroImage, 31(3), 1257–1267. https://doi.org/10.1016/j.neuroimage.2006.01.033
Cross, E. S., Hamilton, A. F., Kraemer, D. J. M., Kelley, W. M., & Grafton, S. T. (2009). Dissociable substrates for body
motion and physical experience in the human action observation network. European Journal of Neuroscience, 30(7),
1383–1392. https://doi.org/10.1111/j.1460-9568.2009.06941.x
Cross, E. S., Kirsch, L., Ticini, L. F., & Schütz-Bosbach, S. (2011). The impact of aesthetic evaluation and physical ability
on dance perception. Frontiers in Human Neuroscience, 5, 102. http://doi.org/10.3389/fnhum.2011.00102
Cutting, J. E., & Kozlowski, L. T. (1977). Recognizing friends by their walk—Gait perception without familiarity cues.
Bulletin of the Psychonomic Society, 9(5), 353–356. https://doi.org/10.3758/BF03337021
Decety, J., Grèzes, J., Costes, N., Perani, D., Jeannerod, M., Procyk, E., Grassi, F., & Fazio, F. (1997). Brain activity dur-
ing observation of actions: Influence of action content and subject’s strategy. Brain, 120(10), 1763–1777. https://doi.
org/10.1093/brain/120.10.1763
Di Nota, P. M., Levkov, G., Bar, R., & DeSouza, J. F. X. (2016, July). Lateral occipitotemporal cortex (LOTC) activity
is greatest while viewing dance compared to visualization and movement: Learning and expertise effects. Experimental
Brain Research, 234(7), 2007–2023. https://doi.org/10.1007/s00221-016-4607-7. Epub March  9 2016. PubMed:
26960739
di Pellegrino, G., Fadiga, L., Fogassi, L., Gallese, V., & Rizzolatti, G. (1992). Understanding motor events: A neurophysi-
ological study. Experimental Brain Research, 91(1), 176–180. https://doi.org/10.1007/BF00230027
Dittrich, W. H., Troscianko, T., Lea, S. E. G., & Morgan, D. (1996). Perception of emotion from dynamic point-light
displays represented in dance. Perception, 25(6), 727–738. https://doi.org/10.1068/p250727
Downing, P. E., Jiang, Y., Shuman, M., & Kanwisher, N. (2001). A cortical area selective for visual processing of the
human body. Science, 293(5539), 2470–2473. https://doi.org/10.1126/science.1063414
Downing, P. E.,  & Peelen, M. V. (2011). The role of occipitotemporal body-selective regions in person perception.
Cognitive Neuroscience, 2(3–4), 186–203. http://doi.org/10.1080/17588928.2011.582945
Downing, P. E., Peelen, M. V., Wiggett, A. J., & Tew, B. D. (2006). The role of the extrastriate body area in action per-
ception. Social Neuroscience, 1(1), 52–62. https://doi.org/10.1080/17470910600668854, PubMed: 18633775
Fadiga, L., Fogassi, L., Pavesi, G., & Rizzolatti, G. (1995). Motor facilitation during action observation: A magnetic
stimulation study. Journal of Neurophysiology, 73(6), 2608–2611. https://doi.org/10.1152/jn.1995.73.6.2608
Freedberg, D., & Gallese, V. (2007). Motion, emotion and empathy in esthetic experience. Trends in Cognitive Sciences,
11, 198–203.
Gallese, V., Fadiga, L., Fogassi, L., & Rizzolatti, G. (1996). Action recognition in the premotor cortex. Brain, 119(2),
593–609. https://doi.org/10.1093/brain/119.2.593
Gauthier, I., & Tarr, M. J. (1997). Becoming a “Greeble” expert: Exploring mechanisms for face recognition. Vision
Research, 37(12), 1673–1682. https://doi.org/10.1016/s0042-6989(96)00286-6
Grafton, S. T., Arbib, M. A., Fadiga, L., & Rizzolatti, G. (1996). Localization of grasp representations in humans by posi-
tron emission tomography. 2. Observation compared with imagination. Experimental Brain Research, 112, 103–111.
Graziano, M. S., Taylor, C. S., Moore, T., & Cooke, D. F. (2002). The cortical control of movement revisited. Neuron,
36(3), 349–362. https://doi.org/10.1016/s0896-6273(02)01003-6
Grèzes J., & Decety, J. (2001). Functional anatomy of execution, mental simulation, observation, and verb generation of
actions: a meta-analysis. Human Brain Mapping, 12, 1–19.
Grèzes, J., Fonlupt, P., Bertenthal, B., Delon-Martin, C., Segebarth, C., & Decety, J. (2001). Does perception of bio-
logical motion rely on specific brain regions? NeuroImage, 13(5), 775–785. https://doi.org/10.1006/nimg.2000.0740
Grossman, E., Donnelly, M., Price, R., Pickens, D., Morgan, V., Neighbor, G., & Blake, R. (2000). Brain areas involved
in perception of biological motion. Journal of Cognitive Neuroscience, 12(5), 711–720. https://doi.org/10.1162/
089892900562417
Isik, A. I.,  & Vessel, E. A. (2019, October  25). Continuous ratings of movie watching reveal idiosyncratic dynam-
ics of aesthetic enjoyment. PLoS One, 14(10), e0223896. https://doi.org/10.1371/journal.pone.0223896, PubMed:
31652277, PubMed Central: PMC6814238
Jeannerod, M. (1997). The cognitive neuroscience of action. Wiley-Blackwell.
Johansson, G. (1973). Visual perception of biological motion and a model for its analysis. Perception and Psychophysics,
14(2), 201–211. https://doi.org/10.3758/BF03212378
Jola, C., Abedian-Amiri, A., Kuppuswamy, A., Pollick, F. E., & Grosbras, M. H. (2012). Motor simulation without motor
expertise: Enhanced corticospinal excitability in visually experienced dance spectators. PLoS One, 7(3), e33343.
https://doi.org/10.1371/journal.pone.0033343

344
Watching and engaging in dance

Jola, C., & Grosbras, M. H. (2013). In the here and now: Enhanced motor corticospinal excitability in novices when
watching live compared to video recorded dance. Cognitive Neuroscience, 4(2), 90–98. https://doi.org/10.1080/1758
8928.2013.776035
Kirsch, L. P., Drommelschmidt, K. A., & Cross, E. S. (2013). The impact of sensorimotor experience on affective evalu-
ation of dance. Frontiers in Human Neuroscience, 7, 521. https://doi.org/10.3389/fnhum.2013.00521
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Longo, M. R., Azañón, E., & Haggard, P. (2010, February). More than skin deep: Body representation beyond primary
somatosensory cortex. Neuropsychologia, 48(3), 655–668. https://doi.org/10.1016/j.neuropsychologia.2009.08.022.
Epub August 29 2009. PubMed: 19720070
Loula, F., Prasad, S., Harber, K., & Shiffrar, M. (2005). Recognizing people from their movement. Journal of Experimental
Psychology. Human Perception and Performance, 31(1), 210–220. https://doi.org/10.1037/0096-1523.31.1.210
Miniussi, C., & Ruzzoli, M. (2013). Transcranial stimulation and cognition. Handbook of Clinical Neurology, 116, 739–
750. https://doi.org/10.1016/B978-0-444-53497-2.00056-5, PubMed: 24112935
Montero, B. (2012). Practice makes perfect: The effect of dance training on the aesthetic judge. Phenomenology and the
Cognitive Sciences, 11(1), 59–68. https://doi.org/10.1007/s11097-011-9236-9
Niedenthal, P. M. (2007). Embodying emotion. Science, 316(5827), 1002–1005. https://doi.org/10.1126/science.1136930
Orgs, G., Dombrowski, J. H., Heil, M., & Jansen-Osmann, P. (2008). Expertise in dance modulates alpha/beta event-
related desynchronization during action observation. European Journal of Neuroscience, 27(12), 3380–3384. https://doi.
org/10.1111/j.1460-9568.2008.06271.x
Oztop, E., Kawato, M., & Arbib, M. A. (2013). Mirror neurons: Functions, mechanisms and models. Neuroscience Letters,
540, 43–55. https://doi.org/10.1016/j.neulet.2012.10.005. Epub October 11, 2012. PubMed: 23063951
Peelen, M. V., & Downing, P. E. (2005). Selectivity for the human body in the fusiform gyrus. Journal of Neurophysiology,
93(1), 603–608. https://doi.org/10.1152/jn.00513.2004, PubMed: 15295012
Pineda, J. A. (2005, December 1). The functional significance of mu rhythms: Translating “seeing” and “hearing” into
“doing”. Brain Research. Brain Research Reviews, 50(1), 57–68. https://doi.org/10.1016/j.brainresrev.2005.04.005.
Epub May 31, 2005. PubMed: 15925412
Rizzolatti, G., & Craighero, L. (2004). The mirror-neuron system. Annual Review of Neuroscience, 27, 169–192. https://
doi.org/10.1146/annurev.neuro.27.070203.144230
Rizzolatti, G., Fadiga, L., Matelli, M., Bettinardi, V., Paulesu, E., Perani, D., & Fazio, F. (1996). Localization of grasp
representations in humans by PET: 1. Observation versus execution. Experimental Brain Research, 111(2), 246–252.
https://doi.org/10.1007/BF00227301
Rizzolatti, G., & Sinigaglia, C. (2016, December). The mirror mechanism: A basic principle of brain function. Nature
Reviews: Neuroscience, 17(12), 757–765. https://doi.org/10.1038/nrn.2016.135. Epub October 20, 2016. PubMed:
27761004
Shapiro, L. (2010). Embodied cognition (2nd ed.). Routledge. https://doi.org/10.4324/9781315180380
Torrents, C., Castañer, M., Jofre, T., Morey, G., & Reverter, F. (2013). Kinematic parameters that influence the aesthetic
perception of beauty in contemporary dance. Perception, 42(4), 447–458. https://doi.org/10.1068/p7117, PubMed:
23866557
Troje, N. F. (2002). Decomposing biological motion: A framework for analysis and synthesis of human gait patterns.
Journal of Vision, 2(5), 371–387. https://doi.org/10.1167/2.5.2, PubMed: 12678652
Urgesi, C., Calvo-Merino, B., Haggard, P., & Aglioti, S. M. (2007). Visual and sensorimotor processing of human bodies.
Journal of Neuroscience, 27(30), 8023–8030. https://doi.org/10.1523/JNEUROSCI.0789-07.2007
Urgesi, C., Candidi, M., Ionta, S., & Aglioti, S. M. (2007). Representation of body identity and body actions in extras-
triate body area and ventral premotor cortex. Nature Neuroscience, 10(1), 30–31. https://doi.org/10.1038/nn1815
Yovel, G., & O’Toole, A. J. (2016, May). Recognizing people in motion. Trends in Cognitive Sciences, 20(5), 383–395.
https://doi.org/10.1016/j.tics.2016.02.005. Epub March 24 2016. PubMed: 27016844

345
17
MAKING SENSE OF SPACE
The neuroaesthetics of architecture

Zakaria Djebbara, Lars Brorson Fich and Giovanni Vecchiato

Of all the topics studied by neuroaesthetics, the experience of architecture is perhaps the most challenging
and complex for several reasons. Consider the following four complications. First, space is three-dimensional
and therefore difficult to investigate using flat screens. Since architecture is not just perceived from one
vantage point but is something we move through, around, and past, its multi-dimensionality is a crucial fea-
ture. Second, architecture is multisensory. It is composed of visual, acoustic, and climatic properties: It feels
like something when we touch it and, moving through it, engages the kinesthetic sense (Holl et al., 2006;
Pallasmaa, 2012). Third, and perhaps the most difficult, it involves culture and history. We relate to certain
spaces with specific actions—for example, rituals in temples, churches, mosques, or synagogues—that are
arguably difficult, if not impossible, to restrain. How much of our architectural experience has to do with
what we have been trained to do through our history and culture? Fourth, the fact that architecture is part of
everyday life and, to a varying degree, restrains our potential behaviour makes it more challenging to study
in laboratory experiments than in practice. The advantage of laboratory experiments is that a single task and
environmental variable can be isolated, but that does not reveal much about how we solve everyday tasks
on the fly with numerous impressions on sensory systems. Furthermore, different lab methods come with
inherent limitations. Neuroscience studies need to consider both what to present as architecture—that is, a
façade, an urban setting, an interior, or a landscape—and how to present architectural stimuli—that is, the
technological limitations of presenting visual, auditory, thermal, or tactile stimuli.
These problems do not make it is impossible to conduct meaningful neuroaesthetics studies of architec-
ture; they just mean that it is difficult to design studies that can capture all aspects of the architecture experi-
ence while at the same time being able to provide accurate measures of brain processes. It is important to
realize that single studies cannot stand alone and that any study must be interpreted with critical attention
to the technological limitations set by choice of method and the task conditions employed by the research
design. Therefore, we start by discussing the main neuroimaging techniques used in neuroaesthetics studies
of architecture, highlighting their most significant limitations. Hereafter, we review findings that highlight
the neural mechanisms involved in behaviour associated with scenic or architectural contexts. By reviewing
empirical studies that vary architectural design features, we try to provide an overview of the neural regions
responsive to scenic and architectural changes. Although these brain regions are spatially distributed, they
are functionally linked by multiple networks across the brain that continuously share information, suggesting
that built environments have a widespread impact on cognition and behaviour. We argue that the experience
of built environments relies on the activation of network dynamics rather than on localized activity in specific

346 DOI: 10.4324/9781003008675-19


Making sense of space

regions. We propose a model of this large-scale neural network that may be of interest for future studies on
the aesthetics of scene perception and architecture.

Neuroimaging
Studies of architecture experiences have primarily used two neuroimaging techniques: electroencephalography
(EEG) and functional magnetic resonance imaging (fMRI). Therefore, our introduction to the methods used to
study responses to built environments focuses on these two techniques. It is important to note that EEG
measures a different aspect of neural activity than fMRI and therefore serves a different methodological
purpose. We briefly present the underlying mechanisms measured by the two methods and their advantages
and disadvantages.

Electroencephalography
Hans Berger (1873–1941) is usually credited with making the first systematic recordings of brain oscillations
using electrodes placed on the scalp (Gloor, 1969). Today, we know that these electric signals reflect synaptic
activity (Engel et al., 2001; Friston, 2010; Singer, 1999; F. Varela et al., 2001). This activity stems from a
series of charges in and between the neurons: When a pre-synaptic neuron triggers an action potential, it gets
excited and releases a neurotransmitter within the synaptic cleft, in turn modifying the post-synaptic neu-
ron potential and so on among neurons (Figure 17.1). EEG can pick up different kinds of electrical charges
depending on the distance and the sum of the synaptic activity.
Action potentials are generally much shorter (a millisecond) than post-synaptic potentials, and because
of the cortical depth of their occurrence, action potentials are impossible to measure using EEG. The
post-synaptic potential endures for hundreds of milliseconds and functions through temporal and spatial

Figure 17.1 The anatomy of a neuron. Dendrites emanate from the soma and are connected to other neuronal axons.
The pre-synaptic neuron can release different neurotransmitters during synaptic transmission, picking
back up parts of the released neurotransmitter as a control. The target neuron also picks up the released
neurotransmitters.

347
Djebbara, Fich and Vecchiato

summation. Summation is the process of sufficient excitatory post-synaptic potentials (EPSP) and stems from
other neurons firing simultaneously, resulting in the post-synaptic neuron firing an action potential. The
inverse is also possible. The summation can be a mixture of EPSP and inhibitory post-synaptic potentials
(IPSP), thereby avoiding firing an action potential. GABAergic activity is the dominating inhibitor, while
glutamate is the dominating exciter1 (Buzsáki, 2006; Kandel, 2013; Luck, 2005). During the synaptic activ-
ity, depending on whether it is an excitatory or inhibitory neurotransmitter, an electrical current flows and
creates a dipole that can be measured as either positively (EPSP) or negatively (IPSP) charged. Due to the
short duration of the action potential spike, it is more difficult to compete with the enduring EPSP and IPSP
when measuring the EEG. In fact, event-related potentials (ERPs) are generated by an assembly of post-
synaptic potentials (Luck, 2005).
Modern EEG consists of a specific number of electrodes (channels) and an amplifier that inflates the
recorded synaptic activity. The oscillations that the EEG produces are voltage (potential) differences over
time compared to a reference channel. Since the amplifiers can easily sample data at 1 kHz (1000 data points
per second), the EEG easily picks up the temporal processes of the synaptic development throughout the
brain. With the temporal resolution as a key advantage, the spatial resolution (source-localization) is highly
limited and requires several signal processing operations to bring about an estimation (Palmer et al., 2011).
This issue relates to the recorded signal from the channels. The signal reflects voltage changes in the entire
environment of the electrode itself, which may be intermixed with electrically charged non-brain sources,
for example, signals from computer screens or neck-muscle activity. In addition, the volume conduction
issue makes the single-channel information depend on the whole underlying cortical activity. Such problems
can, to some degree, be solved through algorithms considering physical and anatomical constraints. The final
neuroimaging solution is essentially based on a statistical decomposition of the signal that identifies nodes
from which the signal originated (Palmer et al., 2011).

Functional magnetic resonance imaging


The spatial resolution of fMRI stems from the traditional MRI technology, which is grounded in detecting
magnetic resonance energy from hydrogen atoms of water in a volume of tissue. The MRI scanner is a large,
tube-like device where the human participant lies still as she is being scanned. By generating a high magnetic
field, the MRI scanner functions by aligning the dipoles of the protons in our hydrogen atoms from random
directions to the same direction (Figure 17.2). A short pulse of radio-frequency (RF) energy wave is then
used to perturb the protons to align to the side. The protons return to their aligned directions as the pulse

Figure 17.2 Magnetic resonance imaging. (A) The protons are randomly positioned before exposing the brain to the
magnetic field. (B) Applying the magnetic field will align the protons in the same direction. (C) An emitted
short RF pulse perturbs the protons. (D) The process of returning to the aligned position releases a small
amount of energy that can be detected and amplified using a receiver coil.

348
Making sense of space

ends, still under the magnetic field. Crucially, while the protons return to the aligned direction, they emit
energy that is detected by the MRI sensors (Gore, 2003; Mills et al., 2017; Raichle, 2003).
In order to study functional brain activity, fMRI uses the MRI method to measure differences between
oxygenated haemoglobin and deoxygenated haemoglobin in localized parts of the brain (Buzsáki, 2006,
pp. 92–93). Experiments have shown that this so-called “blood-oxygenation-level-dependent” (BOLD) sig-
nal reflects neuronal activity (Logothetis, 2003): When neurons are active, they consume energy and there-
fore require replenishment of oxygen. The BOLD method entails calculating the difference in oxygenized
blood flow in localized areas of the brain, revealing where neural activity occurs in the brain. The temporal
constraint in fMRI stems from emitting the short RF pulse and collecting the re-orientation energy in each
proton. This process is usually referred to as repetition time (TR), which defines the cycle time between two
points in a series of pulses and is typically no faster than 1 second.
In sum, continuous EEG measures the summation of post-synaptic potentials of pyramidal assemblies
within the cortical layers intermixed with the electrical activity of deeper brain structures. In contrast, fMRI
measures the difference in oxygenated haemoglobin in specific brain regions with a specific TR interval.

Methodological limitations
In contrast to EEG, fMRI has exemplary spatial resolution but poor temporal resolution. In terms of brain
dynamics, the fMRI is limited in the frequency with which it can retrieve information about brain activity.
Cognitive processes, such as decision-making, navigation, and memory recall, occur at a much faster pace
than the fMRI can record. For instance, the dynamics underlying decision-making in navigation as a winner-
takes-all system (Cisek, 2019; Maranesi et al., 2019; Pezzulo & Cisek, 2016) is too fast for detection by use
of fMRI. Neuronal networks generally function at an impressive speed, and given the web of connections
between neurons (Sporns, 2011), it takes merely a few steps to reach from one module/region of the brain to
another. For this reason, analysing the temporal development of the neuronal synaptic activity is important to
illuminate cognitive functions and large-scale neuronal communication. However, fMRI remains a popular
method in cognitive neuroscience, as it provides an impressive spatial resolution of brain activity, allowing
experiments to identify important neuronal hubs and nodes under specific circumstances. For instance, the
involvement of the hippocampus in navigation has been demonstrated in multiple fMRI studies (Doeller &
Burgess, 2008; Hartley et al., 2003; Murias et al., 2019). As such, fMRI provides a spatial understanding of
brain activity that excels at locating important nodes (Table 17.1).

Table 17.1 A brief overview of the main disadvantages and advantages of fMRI and EEG that
the reader must consider when further reading through the different studies.

fMRI EEG

Disadvantages
Indirect neuronal activity Poor signal-to-noise ratio
Low temporal resolution Low spatial resolution
Low ecological validity High inter-subject variability
Absence of mobility
Advantages
High spatial resolution High temporal resolution
Good signal-to-noise ratio Mobile solutions
High spatial resolution Enables 3D interactive setups

349
Djebbara, Fich and Vecchiato

On the other hand, EEG remains a limited neuroimaging method due to its poor signal-to-noise ratio.
This issue stems from the sensors placed on the scalp being highly susceptible to noise, for example, from
nearby electromagnetic fields, muscle activity, and friction. The acquired data must undergo algorithmic
cleaning and refinement processes to improve their signal-to-noise ratio. Consequently, statistical processing
and adjusting of the signal may complicate test-retest reliability and thereby add to the so-called “reproduc-
ibility crisis” (Höller et  al., 2017). Additionally, to improve the signal-to-noise ratio, classical laboratory
experiments are designed for a high number of trials, restricting the types of processes that one can measure
given the brain’s adaptability.
However, EEG is a light instrument that allows for both outdoor and virtual reality (VR) uses. As such,
it can be used to study freely moving participants under more natural circumstances (Gramann et al., 2014).
Mobile brain/body imaging (MoBI) techniques involve the integration of a mobile EEG setup with in actio
participants, allowing experiments to go beyond stationary laboratory set-ups (Gramann et al., 2011; Makeig
et al., 2009). Such MoBI approaches make experiments that can link mind, brain, and movement behaviour
possible (Makeig et al., 2009; Parada, 2018; Parada & Rossi, 2020). Indeed, MoBI has already shown itself
promising for understanding neural activity associated with movements that take place in built environments
(Djebbara et al., 2019).
To sum up, it is currently not possible to conduct magnetic resonance studies in the actual built environ-
ment. Instead, studies that use VR and a joystick to control movement while the participant is lying in the
scanner are gaining momentum. Virtual navigation is an increasingly popular tool in fMRI studies (Migo
et al., 2016; Murias et al., 2019; Pine et al., 2002; Spiers & Maguire, 2006), as it allows participants to visually
navigate 3D spaces displayed on 2D screens without the involvement of the body. However, this approach
has received criticism for numerous reasons. A central critique is that since navigation and spatial orientation
rely on proprioceptive, motor, and vestibular systems, the measured brain activity may not accurately reflect
mechanisms of navigation and spatial orientation (Taube et al., 2013). The disregard of the body has other
consequences as well. For example, recent frameworks of human cognition, including navigation and spatial
orientation, have argued that architecture experience depends on the embodied integration of body, brain,
and environment (Cisek  & Pastor-Bernier, 2014; Clark, 2013; Gallagher, 2017; Pastor-Bernier  & Cisek,
2011; Pezzulo & Cisek, 2016; Thompson & Varela, 2001).

Experimental studies of architecture experiences


Having discussed the two methods that have been applied to the study of architecture experiences, we go
on to review studies that have examined the role of (1) landscapes and urban spaces, (2) architectural design
forms, and (3) the body in shaping neural responses to built environments. Based on this review, we seek to
identify the brain regions and networks involved in different forms of architecture experience. Furthermore,
we aim to identify how these neural correlates are associated with specific behavioural events.

Scenes and landscapes


Views of nature and green spaces are central to studies of landscapes. The pioneering studies by Roger Ulrich
suggested that views of nature result in more positive emotional states and aesthetic experiences (1983, 1981)
and that such views may accelerate recovery after surgery (1984). Combined with Kaplan’s attention restora-
tion theory (Kaplan & Kaplan, 1989; S. Kaplan, 1995), this led to the idea of biophilic design (Browning &
Ryan, 2020), the notion that architecture should incorporate parts of nature in its design solutions. However,
the numerous studies of green areas versus urban built environments have been criticized for using sensory
stimuli that are not systematically varied, making it difficult to ascertain how different aspects of the environ-
ment affect measured differences.

350
Making sense of space

To address this issue, Chang and colleagues (2021) conducted an fMRI study (n = 44, 22 women) with a
particular interest in the presentation of visual stimuli. Forty-five static images from videos capturing nature
scenes were extracted, and the number of pixels associated with trees, bushes, and the grass was calculated.
Images varied in green density from 0% to 70% and were clustered in three levels. Two weeks before the
behavioural measures, the participants viewed each clustered green density image as either intact or shuffled
(block-scrambled). Then participants rated their perceived stress and landscape preferences on the day of
scanning. Analysis revealed that the involvement of the ventral posterior cingulate cortex (vPCC) varied as
a function of green density and co-varied with behavioural measures of avoidance. Additionally, the cuneus,
inferiorly bounded by the calcarine sulcus, was reported to be involved in the visual processing of the stimuli.
Effective connectivity analysis further revealed that vPCC responses modulated attentional regions primarily
in a feedforward modality. Chang and colleagues interpreted the involvement of vPCC as an interference
with the neuroendocrine system, the purpose of which is to adjust stress levels through the hormone cortisol.
In sum, the study suggested that the activity of vPCC as measured using fMRI co-varies with the degree to
which images of nature scenes display forms of greenery such as trees, bushes, or grass.
A limitation of Chang and colleagues’ study is its use of static images. No animate stimuli were included.
Are there reasons to believe that animate stimuli cause different brain responses? Zhao and colleagues (2020)
set out to understand the difference in how aesthetic judgments of static and dynamic landscapes affect
neural activation (see also Chapter 9). Using fMRI (n = 22, 11 women), participants were exposed to a
set of 118 dynamic (animated) landscapes, static landscapes (images), and grey squares for luminance judg-
ment. The luminance judgment served as a baseline for when the participants had to decide whether the
presented landscape was judged aesthetically pleasing. Based on the aesthetic judgment of a separate group
of participants (n = 20), only beautiful and neutral landscapes were included. Results indicated that dynamic
landscapes were judged more beautiful than static landscapes. Contrasting the dynamic landscapes with the
static landscapes revealed a higher involvement of the middle temporal gyrus and the hippocampus for the
former condition than for the latter. Aware of the sensitivity of the middle temporal gyrus (MT/V5) to visual
motion, Zhao and colleagues (2020) suggested that a general increase of neuronal activity in visual areas
might enhance aesthetic appraisal. This finding suggests that movement plays an important role in modulat-
ing aesthetic judgments (Battaglia et al., 2011; Cattaneo et al., 2017; Di Dio et al., 2016).
While both experiments aim to understand responses to natural scenery and landscapes, neither involves
actual body movement, the primary source of animation of the environment. In fact, to act is a medium of
collecting sensory information about the built environment that, in turn, is shaped by affordances.2 An impor-
tant fMRI study of scenes and landscapes relative to affordances is the one by Bonner and Epstein (2017).
With a predefined interest in the parahippocampal place area (PPA), early visual cortex (EVC), retrosplenial
cortex (RSC), and occipital place area (OPA), the researchers sought to understand how affordances are
reflected in the brain during both natural and built environments. The fMRI study (n = 12, 8 women) first
used highly controllable virtual spaces where doors appeared in an increasing fashion, from “no doors” to
doors on the left, right, and centre walls and later introduced natural complex images. A total of 18 different
images with varying affordances were presented. To correct for early visual stimuli, half of the images in each
condition included paintings in the location of doors and were approximately the same size as the doors. To
the authors’ surprise, the only region that reflected any sensitivity to architecture affordances was the OPA—
in fact, the OPA is deeply involved in identifying architectural affordances. Although prior studies had dem-
onstrated the involvement of PPA and RSC in the perception of scenes and spatial layouts (Bar & Aminoff,
2003; Epstein & Kanwisher, 1998; Epstein et al., 2003), Bonner and Epstein observed no systematic change
within the PPA or the RSC. Following up on these findings, in a separate study of natural environments, a
new group of participants (n = 16, 8 women) were exposed to 50 images of indoor environments with clear
paths as represented by heat maps. When the heat maps were compared to model representations, comprising
the product of angular-histogram vectors and tuning curves, the OPA (similar to the first experiment) was

351
Djebbara, Fich and Vecchiato

discovered to display the strongest effect for architectural affordances. Although PPA did reach significance in
the new study, the OPA showed comparatively stronger coding for affordances, possibly reflecting the PPA’s
sensitivity to place identity. It is possible that the dorsal visual stream harbours a perceptual mechanism that
processes visuospatial information highly sensitive to the potential actions underlying navigational function
in the environment.
Together, these studies demonstrate that brain responses are tightly related to the history of human biol-
ogy. The response to green density, dynamic visual stimuli, and the potential to act in the environment are
all examples of biologically important functions to maintain homeostasis that in turn motivate and ensure
survival. Similar points have been put forward by the comprehensive meta-analysis demonstrating the impor-
tance of visceral perception in aesthetic appraisal (Brown et al., 2011).

Architectural forms
A group of interdisciplinary researchers, comprising architects and experimental psychologists, investigated
whether the contour of space affects aesthetic judgment and approach-avoidance decisions (Vartanian et al.,
2013). The fMRI study (n  =  18, 12 women) was designed to investigate whether curvilinear rooms are
judged more beautiful than rectilinear rooms and whether this difference in form can elicit differences in
approach and avoidance behaviour. The architects selected the rooms to reflect curvilinearity, openness, and
ceiling height variance. During exposure to the images, the participants first answered whether they con-
sidered the room beautiful and then whether they would enter the space. Immediately post-scanning, the
participants rated the same images again according to pleasantness and beauty using a 5-point Likert scale.
As such, three behavioural dimensions were gathered: beauty and approach-avoidance judgments during the
fMRI scanning and pleasantness post-scan. As expected, curvilinear spaces positively affected beauty judg-
ments but did not affect approach-avoidance decisions. However, spaces that were considered approachable
were viewed for significantly longer periods. Additionally, pleasantness as scored post-scan was found to be a
significant predictor of both beauty judgment and approach-avoidance decisions. Exploring linear relation-
ships on a neural level, Vartanian et al. (2013) found that the frontopolar cortex superior frontal gyrus, globus
pallidus, precuneus, parahippocampus, and middle occipital gyrus co-varied with beauty ratings, while the
precuneus, middle frontal gyrus, and anterior cingulate cortex (ACC) co-varied with pleasantness. In the case
of approach-avoidance, the fMRI contrast of curvilinear-rectilinear revealed that the left lingual gyrus and
right calcarine were significantly more active for curvilinear spaces. Interestingly, the fMRI contrast between
curvilinear and rectilinear spaces during beauty judgment revealed significant activation of the ACC alone.
The ACC has been demonstrated to contribute to reward and emotion processing (Kringelbach & Rolls,
2004; Yu et al., 2011), highlighting the emotional dimension in judging the beauty of the built environment.
The same research group also collected measures of perceived room-height and closure (Vartanian et al.,
2015). Building on existing studies, for example, the study on ceiling height and preference by Baird et al.
(1978) and the study by Fich and colleagues (2014) on cortisol levels and perceived closure, Vartanian and
colleagues (2015) hypothesized that spaces with higher ceilings would be judged as more beautiful, and
similarly for open spaces. Open spaces are defined according to Stamps’ (2005) idea that open rooms afford
greater perceived visual and locomotive permeability. Furthermore, they predicted that brain regions related
to emotion, pleasure, and reward would be involved in processing spaces with higher ceilings. Similar to their
first study, the behavioural measures involved both judgments of beauty and approach-avoidance. Therefore,
Vartanian et al. (2015) also hypothesized that rooms with higher ceilings would be approached more than
rooms with lower ceilings, and similarly for open spaces as opposed to closed spaces. Analysing the data,
the authors confirmed that rooms with higher ceilings are judged to be more beautiful. Yet they were not
approached more than rooms with lower ceilings, in contrast to predictions. The analysis also revealed that
open rooms were judged more beautiful, while enclosed rooms were more likely to be avoided. These results

352
Making sense of space

confirmed the authors’ initial hypotheses as well as the existing literature. In terms of the fMRI results, the
study revealed an increased involvement of the precuneus and left middle frontal gyrus during judgments of
beauty for high-ceilinged rooms, whereas an increased involvement of the left middle temporal gyrus and
right superior temporal gyrus were observed when participants appraised the beauty of open rooms. In con-
trast, enclosed spaces elicited greater activity in the anterior midcingulate cortex when participants consid-
ered approach-avoidance decisions. To conclude, spaces with higher ceilings were judged as more beautiful,
similarly for open spaces, while the authors’ prediction based on their first study that brain regions associated
with emotion, pleasure, and reward would co-vary with beauty was not met. Instead, Vartanian et al. (2015)
identified the precuneus, which is associated with sensorimotor and visual processes (Kitada et al., 2014), and
the left middle frontal gyrus to be increasingly involved.
Bermudez and colleagues (2017) sought to understand how the environment can elicit contemplative
states by correlating self-reports and fMRI data from a subject group of architects (n =12, no women). The
externally induced contemplative state is a state that is closely related to the aesthetic experience, albeit not
identical. To assess the participants’ level of immersion, Bermudez and colleagues (2017) conducted face-to-
face interviews in the spirit of micro-phenomenology after each session. The experience’s depth was assessed
using questionnaires and interviews based on six criteria: positive emotional response, sense of self, immer-
sion, appreciation (beauty), experiential intensity, and connectedness. While the architects were looking at
both traditional buildings (control) and contemplative buildings (experimental condition), fMRI data of
the architects were collected. Hereafter, the architects engaged in the interviews and filled in the question-
naires. To the authors’ surprise, analysis of the fMRI data found the depth of experience to correlate with
the engagement of sensorimotor regions, whereas an inverse relation was found for prefrontal regions. In
other words, the architects with the strongest immersive experiences were those with a deactivation of frontal
regions (often thought to be associated with internally induced meditative states and a sense of agency and
self ) and increased activity in sensorimotor and integrative regions. Thus, Bermudez and colleagues (2017)
provide support for an embodied and hedonic sensory valuation aspect of architecture where the hedonic
valuation emerges from the effortless commitment to the environment. According to the authors, beauty
and appreciation may be better understood as a non-evaluative experience as opposed to a critical judgment.
While still limited in number of fMRI studies of responses to architectural forms, specific design forms
have a significant effect on movement and the motivation for movement, as indexed by both brain activity
and behavioural measures. Animated landscapes, approach-avoidance, and affordances appear to affect the
brain differently from static, in-affordable, and low-ceiling spaces. With recent technological advances, tak-
ing the experience of architecture to animate beings is no longer a challenge. In return, studies begin to
approach more ecological research designs that consider the body an active contributor to brain activity—a
critical aspect of cognition (Gramann et al., 2011; Makeig et al., 2009).

Embodiment in motion
Theories that take the dynamic relation between brain, body, and environment as constitutive of cognition
are gaining traction (Clark, 2015; Gallagher, 2017; Pezzulo et al., 2015). These theories posit that affordances
can be conceived as predictions of how possible interactions can unfold and, consequently, what percep-
tions are possible. In short, they predict that architectural affordances affect sensorimotor brain dynamics
associated with perceptions and actions. Djebbara, Fich, Petrini and Gramann (2019) set out to test the
hypothesis that perceptual processes are altered due to architectural affordances. To do so, they manipulated
the architectural transitions between two spaces and used a combination of virtual reality and mobile EEG to
measure neural activity as their participants (n =19, 9 women) walked through the passage with the task of
fetching a red circle in the subsequent space. Additionally, an emotional questionnaire, the Self-Assessment
Manikin (Bradley & Lang, 1994), was filled out after each trial. The research design followed a systematic

353
Djebbara, Fich and Vecchiato

ERP motor-priming paradigm, meaning that two cues were presented in each trial. The first cue revealed
which transition to pass, while the second cue signalled whether to go. The transitions varied in width so
that one was not passible, as it was too narrow, while the two others were passable. As hypothesized, Djeb-
bara et al. (2019) found affordances to be reflected in early perceptual processes within sensorimotor regions
upon perceiving the transition, suggesting that our action possibilities colour perception from early on. The
narrow transitions were reported to be experienced as more dominating, negative, and less arousing than the
passable transitions.
There exist numerous reasons the visual and motor systems are co-dependent. Visually guided move-
ment, for instance, demands attention from both systems, which more appropriately should be considered a
single system. Djebbara et al. (2021) set out to re-analyse the datasets from the previous study to assess how
increased demand of attention in approaching the transitions was reflected in brain activity. They hypoth-
esized that since the attentional mechanisms directing the gating function of sensory signals share neuronal
resources with motor-related processes, the phase between the Go-cue and the passing through the transition
would result in stronger alpha-band desynchronization over sensorimotor regions. Their hypothesis was also
based on studies linking parieto-occipital alpha to anticipatory attention (Arnal & Giraud, 2012; Bastiaansen
et al., 2001) and inhibition of irrelevant visual information (Engel et al., 2001; Jensen et al., 2002). Surpris-
ingly, they found that while perceiving the type of transition and waiting for the Go/NoGo-cue, the alpha
desynchronization over left occipital and right parahippocampal regions co-varied with the architectural
affordances. Once again, poor affordances, that is, impassable transitions, were processed differently than
passable transitions. Participants approached the impassable transitions significantly slower than the passable
transitions when given the Go-cue. Moreover, they found the narrow condition to be processed significantly
differently, with stronger alpha desynchronization over both the posterior parietal cortex and right supple-
mentary motor area correlating with the approach phase. Djebbara and colleagues (2021) have argued that
the suppression reflected in the alpha-band is inherently related to thalamo-cortical inhibition of irrelevant
sensory information that facilitates attention being directed towards the transition. Furthermore, they inter-
pret involvement of parahippocampal and occipital regions as related to scene processing, as demonstrated by
numerous studies (Epstein & Baker, 2019; Epstein et al., 2003). Djebbara et al. (2021) concluded that since
brain activity resonates with architectural affordances, which are predictions, the experience of space is based
on continuous sensory and motor predictions.

Movement in interior spaces


As reviewed previously, visual appearances of interior spaces, such as whether they take the form of curvilin-
ear or rectilinear contour, can affect the ACC (Vartanian et al., 2013). Building on this finding, Banaei and
colleagues (2017) applied virtual reality and mobile EEG to allow participants a fuller experience of such
spaces (n =15, 8 women). Based on a data-driven approach to architectural styles and emotional responses,
the researchers set out to create 68 rooms (4 rooms ⨯ 17 features) to identify which brain regions correlated
with the strongest emotional responses. Each participant wore a 128-channel EEG setup combined with
head-mounted displays and walked a pre-determined path involving moving forward and turning both
directions. By the end of the walk, the participants filled out a virtual Self-Assessment Manikin (Bradley &
Lang, 1994) and conducted a Stroop test (Stroop, 1935). The study demonstrated the involvement of the
ACC, precentral gyrus, posterior cingulate cortex, and middle temporal gyrus both from stimulus onset and
in different room perspectives. Additionally, they found that the ACC correlated with rooms that were rated
higher in terms of pleasure and arousal.
Similarly, Vecchiato and colleagues investigated how rooms varying in curvilinear and rectilinear design
affected different types of responses (Vecchiato, Jelic et al., 2015; Vecchiato, Tieri et al., 2015). A combina-
tion of a 24-channel EEG and a Cave Automatic Virtual Environment was used to immerse the participants

354
Making sense of space

(n =12, all women) in three different spaces. The environments included an empty room, a room of rec-
tilinear furniture, and a room of curvilinear furniture. Subjective reports regarding pleasantness, novelty,
familiarity, comfort, arousal, and presence were recorded on a 9-point rating scale after the exposure of the
spaces. Based on the sensor-level time-frequency analysis, the authors’ analysis revealed the suppression of
µ-rhythm in rooms reported as pleasant and comfortable and the involvement of the frontoparietal network
in the perception of places.

Neural mechanisms involved in architecture experiences


As scenes and architecture affect an assemblage of operations in the brain, it is not our aim to identify a
small-scale network responsible for processing scenes and architecture. Instead, we aim to propose a large-
scale network that represents an overlap of several small-scale networks revealing how scenes and architec-
ture may affect the general brain dynamics. Based on the studies reviewed, it can be suggested that certain
brain regions are devoted to processing architectural scenes, while other regions seem to be involved in the
processing of animated and action-oriented perception. A  table of the involved regions, culled from the
experimental reports reviewed previously, summarizes the associated tasks (Table 17.2). In the following, we
discuss how, based on these findings, humans make sense of space.
As numerous regions reoccur in the reviewed studies, these seem to be critical to the process of making
sense of space. These regions (Figure 17.3) make up a large-scale network that includes the anterior cingulate
cortex, parahippocampal place area, prefrontal cortex (PFC), occipital cortex (OCC), retrosplenial cortex,
and posterior parietal cortex (PPC). Some of these areas have also been identified as reoccurring regions in
studies of scene perception (Djebbara et al., 2021; Epstein et al., 2003, 2007; Epstein & Baker, 2019), as well
as in studies of preference and beauty (Flexas et al., 2014; Jacobsen et al., 2006). Viewing these regions from
a broader perspective, it appears that the sense-making of space is inherently related to the perception of
place, sensorimotor integration, and affordances. However, it is difficult to make such an otherwise attractive
conclusion based only on the reviewed papers. Not only is the number of reviewed papers insufficient, each
study only concerns one distinctive task.3 We would like to emphasize the importance of the involvement of
multiple networks in behaviour.
Accumulating evidence suggests that several processes occur in parallel within the brain, affecting the
brain as a whole (Anderson et al., 2013; Cisek & Kalaska, 2005; Gallivan et al., 2016, 2018; Pessoa, 2014).
Thus, it is difficult to understand a single process without considering other processes. For instance, Chang
and colleagues (2021) demonstrated that the density of greenery shown in static images co-varied with
posterior parietal cortex activity, an area also sensitive to both beauty judgments and action-oriented affor-
dances (Djebbara et al., 2021; Vartanian et al., 2013). Similarly, Zhao and colleagues (2020) found the middle
occipital gyrus to be involved in the experience of animated landscapes as opposed to static ones, an area also
demonstrated to be involved in beauty judgments by Vartanian and colleagues (2013) and in affordances by
Djebbara and colleagues (2021). Although there are neuronal correlates of beauty judgment and aesthetic
experiences, the idea that such an experience resides in a specific small-scale network seems counter-intuitive
based on the current evidence. Instead, different neuronal systems must be involved in the computation of
aesthetic experiences. Thus, it may be more valuable and advantageous to cast the experience of aesthetics as
an achievable brain-dynamic in various networks, that is, as an attainable dynamic by diverse networks. We
are careful not to portray aesthetic experiences as states of networks, but rather as dynamics, that is, changes in
energy and activity between various networks (Anderson et al., 2013; Camazine et al., 2001; Thompson &
Varela, 2001; Varela et al., 2001, 1974). This is because networks in the brain reflect human capacities (the
unfolding of functions; Sporns, 2011, Chapter 4) rather than outcomes like aesthetic experiences. Such a
position suggests two things. First, neuroaesthetics of the built environment must be considered a function
of network dynamics rather than localized brain regions. Second, there may be different kinds of aesthetic

355
Table 17.2 An overview of the positively involved brain regions as reported by the reviewed studies. * Multiple Talairach coordinates were reported; thus, a mean
of the Talairach coordinates is presented. ** Talairach coordinates were not provided; thus, a coordinate is retr ieved from a Talairach Atlas by Lancaster
et al. (2000).

Talairach coordinates

Study Task Brain regions BA Hemisphere X Y Z

(Chang et al., Perceiving static Ventral poster ior par ietal cortex BA23 R/L ±14 −51 12
2021) images of varying Dorsal poster ior par ietal cortex BA31 R/L ±8 −57 33
green density. Super ior temporal gyrus BA39 R/L ±45 −51 16
Super ior paretal lobule BA7 R/L ±29 −51 42
Middle frontal gyrus BA6 R/L ±29 −4 52
Precentral gyrus BA6 R/L ±36 0 30
Medial dorsal nucleus (thalamus) – R/L ±10 −18 10
Cuneus BA17 R/L ±15 −89 2

Djebbara, Fich and Vecchiato


(Zhao et al., Compar ing static Fusifor m gyrus – R 30 −48 −15
2020) images of landscape Middle occipital gyrus BA18 L −33 −90 0
and animated videos Thalamus – R* 9 −11 0
of the landscape. – L −21 −18 6
356

Insula BA47 L −33 18 −3


BA47 R 33 21 −3
Supplementary motor area BA6 L* −3 21 54
Cerebellum – L −33 −63 −42
Infer ior frontal operculum/ BA44 R* 46 20 19
Infer ior frontal tr iangle
Infer ior frontal operculum/ BA44 L* −45 14 22
Infer ior frontal tr iangle
Middle cingulate cortex BA32 R/L* 5 24 38
BA32 L −6 27 33
Infer ior occipital gyr i BA18 R* 30 −89 −5
BA18 L −27 −93 −3
Lingual gyrus BA18 L* −20 −72 −6
Calcar ine BA30 R 21 −54 12
Orbitofrontal cortex BA47 L* −38 39 −12
Middle temporal gyrus BA21 R 45 −63 3
BA21 L −51 −69 6
(Bonner & Fixate on a cross and Occipital place area BA19 R** 34 −77 21
Epstein, perceive various Parahippocampal place area BA30 R** 20 −39 −5
2017) static images. Retrosplenial cortex BA29 R** 7 −43 17
Early visual cortex – – – – –
(Vartanian Judging beauty, Frontopolar cortex BA10 L −14 58 −3
et al., 2013) pleasantness, and Superior frontal gyrus BA6 L −25 25 54
approach-avoidance. Globus pallidus – R 15 −6 −1
Precuneus BA7 L −28 −70 42
Parahippocampus BA27 L −25 −32 2
Middle occipital gyrus BA19 L* −36 −79 17
Middle frontal gyrus BA9/46 R 33 40 11
BA9/46 L −37 28 14
Anterior cingulate cortex BA32 L −18 41 11
(Vartanian Judging beauty, Precuneus BA19 L −36 −78 42
et al., 2015) approach-avoidance. Middle frontal gyrus BA6 L −32 18 60
Middle temporal gyrus BA21 L −52 −2 −20

Making sense of space


Superior temporal gyrus BA22 R 58 −14 −4
Anterior midcingulate cortex BA24 R/L 10 30 26
−84
357

(Bermudez Perceiving regular Middle occipital gyrus BA19 R 38 8


et al., 2017) and contemplative Fusiform gyrus BA37 L −40 −62 −10
architecture. Precentral gyrus BA6 R 44 −6 32
BA6 L −38 −2 40
Middle frontal gyrus BA6 L −20 2 54
Superior frontal gyrus BA6 L −4 6 62
(Djebbara Self-assessing emotion, Precuneus BA 31 R 1 −69 29
et al., 2021) a motor-priming Middle occipital gyrus BA18 L −24 −78 2
transition between Parahippocampal gyrus BA18 R 35 −62 7
spaces. Precentral gyrus BA6 L 38 −5 39
(Banaei et al., Self-assessing emotion Anterior cingulate cortex BA24 R/L 0 1 26
2017) and exploration of Precentral gyrus BA4 R 32 −20 53
space. BA4 L −32 −18 54
Posterior cingulate cortex BA23 L −3 −58 16
Middle temporal gyrus BA39 R 40 −58 24
BA39 L −63 −40 0
Superior frontal gyrus BA6 R 9 7 56
Djebbara, Fich and Vecchiato

Figure 17.3 The proposed network relevant to the identification and interaction with the built environment is based on
the reviewed studies. The orange area represents the cingulum, the blue area represents the thalamus, and
the red area represents a part of the entorhinal region.

experiences such as beauty, appraisal, pleasure, appreciation, or liking that are modulated by different degrees of
activity across the involved networks. In this view, different kinds of aesthetic experience would depend on
the specific involvement of individual networks, while the degree of aesthetic experience would depend on
the dynamics of the involvement in a broader sense.4
In understanding the environment, the role of the thalamus is particularly interesting as the only source
of sensory information about the body and environment, projected from the sense-specific receptors
to the neo-cortex through a large collection of neurons (nuclei).5 The thalamic nuclei receive far more
downward projections than it projects upwards, suggesting that top-down information modulated sensory
representations—as proposed by the predictive brain hypothesis (Friston, 2010; Wolff et al., 2020). Based on
several studies investigating the connectivity of the thalamus (Sherman, 2017; Theyel et al., 2010; Viaene
et al., 2011), it is clear that the function of the thalamus goes beyond mere forward relay. Instead, several
thalamic nuclei are organized so that they afford to relay information between cortical areas using feedback
modulations from the cortex in parallel with direct cortico-cortical pathways (Saalmann et al., 2012; Schmid
et al., 2012). Therefore, it appears to be taking an active part in many dynamic processes such as visual and
motor processes in parallel with cortico-cortical information transmission (Sherman & Guillery, 1996, 2002;
Sherman, 2016). As new information in the thalamus accumulates, its function and pathways appear to be
increasingly complex. For instance, a study on the posteromedial cortex (PMC) of a macaque by Parvizi
and colleagues (2006) revealed that the PCC receives far more posterior thalamic projections than any other
area in the PMC. Additionally, the PCC is highly interconnected with the parahippocampal formation that
is also included in our network model. With the bilateral parahippocampal areas being anatomically close to

358
Making sense of space

the entorhinal cortex and the hippocampus—where the grid cells and place cells reside—it is appealing to
interpret them as interrelated and collectively engaged in processing spatial information.
Such a view suggests that the PCC takes a world-engaging role where spatial information plays an impor-
tant role. For instance, the PPC is the first region to display activity in patients when regaining conscious-
ness (Laureys et al., 1999) and also the region to display deactivation during sleep (Maquet et al., 1997) and
propofol-induced anesthesia (Fiset et al., 1999)—all of which are scenarios that require some notion of space.
The PPC is furthermore also highly involved in sensorimotor functions, including actions, motor planning,
and decision-making, as it also serves as a critical node for sensorimotor integration (Andersen & Buneo,
2002; Andersen & Cui, 2009; Cisek & Kalaska, 2005; Rizzolatti et al., 1988). Given that architecture is a
spatial experience, it is not surprising that the identified regions have previously been involved in sensorimo-
tor functions relevant to spatial interaction. This statement should not be confused with the identification of
a spatial awareness network but rather a space-sensitive complex large-scale network (Figure 17.3).

Challenges ahead
Considering the difficulty of investigating the four-dimensionality of architecture, the study of full-blown
architecture experiences requires a fully mobile setup that does not detach the agent from the architecture.
Thus, the first challenge to neuroaesthetics of architecture is to develop a mobile, lightweight instrument
that allows highly detailed neuroimaging and a high temporal resolution of brain dynamics during the
engagement with architecture. The second challenge concerns the development of a device allowing for an
ecological presentation of the built environment. Although the use of computer screens and virtual reality, as
demonstrated by the reviewed studies, can reveal how architecture affects our brain, a less virtual device that
allows for more natural settings and is compatible with neuroimaging techniques is highly desirable. Con-
sidering that studies of architecture experiences have gone from using static pictures as stimuli to interactive
virtual reality, the future of architecture studies is very promising (Presti et al., 2021). Embracing movement
in architecture studies is especially important, since current results suggest that “statically perceiving” as com-
pared to actual “engagement” with the built environment are fundamentally different phenomena (Thibault
et al., 2014, 2016).
On a neurobiological level, the challenge ahead is to advance our understanding of how the identified
large-scale network is involved in everyday human capacities. For example, how is the state of PCC involved
in any specific human capacity? As emotions are typically associated with an inward perception, that is,
interoception, one may ask if built environments can affect our sensitivity to emotions, and if so, how? Can
it enhance reaction times? Can it alter our self-image, including our sense of identity? As many mental pro-
cesses apply trans-thalamic pathways in parallel with direct pathways, an even more pressing question is: to
what extent can the processing of sensations of the built environment impact human mental functions, that
is, memory, attention, judgment, decision-making, and so on?
The reviewed studies expose the diversity of investigated cognitive processes and the increasingly detailed
depictions of the built environment. As such, the field is in constant expansion. Some studies focused on the
sensory valuation of architecture, while the higher-order critical judgment of beauty occupied other studies.
In itself, the diversity of experimental studies proves no challenge, but the diversity of interpretations may
be problematic. The lower-order studies dealing with the embodiment in architecture seem to be guided by
biological and evolutionary theories, whereas cognitive theories guide higher-order approaches. Reviewing
the currently existing studies, it appears the premise is that, regardless of the approach, results will eventu-
ally converge into a unified understanding of the impact of architecture. A challenge is thus developing a
framework able to unify the different levels. This challenge appears to be an intractable issue with the current
knowledge. We have only commenced the collection of information about how the designed world affects
the brain, and it is thus difficult to provide any proper conclusion besides calling for extensive research into

359
Djebbara, Fich and Vecchiato

behavioural and neuronal responses to the built environment. Indeed, multiple attempts at a unifying theory
are currently in competition, hopefully harvesting even more experimental studies that act as the bond in
our arguments and theories.

Notes
1 Glutamate is an excitatory neurotransmitter that all cortical pyramidal cells can release and is associated with depo-
larization and discharge in the target neurons. Conversely, GABA is inhibitory and associated with hyperpolarization
(Buzsáki, 2004, 2006, p. 88).
2 The concept of “affordances,” as coined by James J. Gibson, is defined as that which the environment offers the animal,
that is, what it provides or furnishes, either for good or ill (Gibson, 1986, p. 127). An articulated version by Andy
Clark is perhaps more useful. Clark states that affordances reflect “the possibilities for use, intervention and action
which the physical world offers a given agent and are determined by the ‘fit’ between the agent’s physical structure,
capacities and skills and the action-related properties of the environment itself ” (1999).
3 Different tasks give rise to different network activities in the brain (Buzsáki, 2006, 2019; Sporns, 2011).
4 Although our view does not disagree with Skov’s view on aesthetic experience as the outcome of mesolimbic involve-
ment (2019), we suggest instead that different kinds of aesthetic experiences depend on the involvement of other
systems and thus have a more inclusive view; some aesthetic experiences are not rooted in mesolimbic involvement.
While numerous studies show the involvement of the mesolimbic system to reflect addictive behaviour, or more
generally modulation for adaptive behaviour in terms of learning and motivation (Berke, 2018; Hamid et al., 2016;
Shohamy & Adcock, 2010), we find it difficult to exclusively associate its involvement with the experience of aesthet-
ics or even exclusively pleasure.
5 The following is based on studies conducted on humans, primates, rodents, and felines.

References
Andersen, R. A., & Buneo, C. A. (2002). Intentional maps in posterior parietal cortex. Annual Review of Neuroscience,
25(1), 189–220. https://doi.org/10.1146/annurev.neuro.25.112701.142922
Andersen, R. A., & Cui, H. (2009). Intention, action planning, and decision making in parietal-frontal circuits. Neuron,
63(5), 568–583. https://doi.org/10.1016/j.neuron.2009.08.028
Anderson, M. L., Kinnison, J., & Pessoa, L. (2013). Describing functional diversity of brain regions and brain networks.
NeuroImage, 73, 50–58. https://doi.org/10.1016/j.neuroimage.2013.01.071
Arnal, L. H., & Giraud, A. L. (2012). Cortical oscillations and sensory predictions. In Trends in cognitive sciences (Vol. 16,
Issue 7, pp. 390–398). Elsevier Current Trends. https://doi.org/10.1016/j.tics.2012.05.003
Baird, J. C., Cassidy, B., & Kurr, J. (1978). Room preference as a function of architectural features and user activities.
Journal of Applied Psychology, 63(6), 719–727. https://doi.org/10.1037/0021-9010.63.6.719
Banaei, M., Hatami, J., Yazdanfar, A.,  & Gramann, K. (2017). Walking through architectural spaces: The impact of
interior forms on human brain dynamics. Frontiers in Human Neuroscience, 11(417), 477. https://doi.org/10.3389/
fnhum.2017.00477
Bar, M., & Aminoff, E. (2003). Cortical analysis of visual context. Neuron, 38(2), 347–358. https://doi.org/10.1016/
s0896-6273(03)00167-3
Bastiaansen, M. C. M., Böcker, K. B. E., Brunia, C. H. M., De Munck, J. C., & Spekreijse, H. (2001). Event-related
desynchronization during anticipatory attention for an upcoming stimulus: A comparative EEG/MEG study. Clinical
Neurophysiology, 112(2), 393–403. https://doi.org/10.1016/S1388-2457(00)00537-X
Battaglia, F., Lisanby, S. H., & Freedberg, D. (2011). Corticomotor excitability during observation and imagination of
a work of art. Frontiers in Human Neuroscience, 5. http://www.frontiersin.org/article/10.3389/fnhum.2011.00079.
https://doi.org/10.3389/fnhum.2011.00079
Berke, J. D. (2018). What does dopamine mean? Nature Neuroscience, 21(6), 787–793. https://doi.org/10.1038/
s41593-018-0152-y
Bermudez, J., Krizaj, D., Lipschitz, D. L., Bueler, C. E., Rogowska, J., Yurgelun-Todd, D., & Nakamura, Y. (2017).
Externally-induced meditative states: An exploratory fMRI study of architects’ responses to contemplative architec-
ture. Frontiers of Architectural Research, 6(2), 123–136. https://doi.org/10.1016/j.foar.2017.02.002
Bonner, M. F.,  & Epstein, R. A. (2017). Coding of navigational affordances in the human visual system. Proceedings
of the National Academy of Sciences of the United States of America, 114(18), 4793–4798. https://doi.org/10.1073/
pnas.1618228114

360
Making sense of space

Bradley, M. M., & Lang, P. J. (1994). Measuring emotion: The self-assessment manikin and the semantic differential. I.
Journal of Behavior Therapy and Experimental Psychiatry, 25(1), 49–59. https://doi.org/10.1016/0005-7916(94)90063-9
Brown, S., Gao, X., Tisdelle, L., Eickhoff, S. B., & Liotti, M. (2011). Naturalizing aesthetics: Brain areas for aesthetic
appraisal across sensory modalities. NeuroImage, 58(1), 250–258. https://doi.org/10.1016/j.neuroimage.2011.06.012
Browning, B., & Ryan, C. (2020). Nature inside: A biophilic design guide. RIBA Publishing.
Buzsáki, G. (2004). Large-scale recording of neuronal ensembles. In Nature neuroscience (Vol. 7, Issue 5, pp. 446–451).
Nature Publishing Group. https://doi.org/10.1038/nn1233
Buzsáki, G. (2006). Rhythms of the brain. Oxford University Press.
Buzsáki, G. (2019). The brain from inside out. Oxford University Press.
Camazine, S., Deneubourg, J.-L., Franks, N., Sneyd, J., Theraulaz, G., & Bonabeau, E. (2001). Self-organization in bio-
logical systems (P. Anderson, J. Epstein, D. Foley, S. Levin, & M. Nowak, Eds., 1st ed.). Princeton University Press.
Cattaneo, Z., Schiavi, S., Silvanto, J., & Nadal, M. (2017). A TMS study on the contribution of visual area V5 to the
perception of implied motion in art and its appreciation. Cognitive Neuroscience, 8(1), 59–68. https://doi.org/10.108
0/17588928.2015.1083968
Chang, D. H. F., Jiang, B., Wong, N. H. L., Wong, J. J., Webster, C., & Lee, T. M. C. (2021). The human posterior
cingulate and the stress-response benefits of viewing green urban landscapes. NeuroImage, 226, 117555. https://doi.
org/10.1016/j.neuroimage.2020.117555
Cisek, P. (2019). Resynthesizing behavior through phylogenetic refinement. Attention, Perception and Psychophysics, 81(7),
2265–2287. https://doi.org/10.3758/s13414-019-01760-1
Cisek, P.,  & Kalaska, J. F. (2005). Neural correlates of reaching decisions in dorsal premotor cortex: Specification
of multiple direction choices and final selection of action. Neuron, 45(5), 801–814. https://doi.org/10.1016/j.
neuron.2005.01.027
Cisek, P., & Pastor-Bernier, A. (2014). On the challenges and mechanisms of embodied decisions. Philosophical Transac-
tions of the Royal Society of London. Series B, Biological Sciences, 369(1655). https://doi.org/10.1098/rstb.2013.0479
Clark, A. (1999). An embodied cognitive science? Trends in Cognitive Sciences, 3(9), 345–351. https://doi.org/10.1016/
S1364-6613(99)01361-3
Clark, A. (2013). Whatever next? Predictive brains, situated agents, and the future of cognitive science. Behavioral and
Brain Sciences, 36(3), 181–204. https://doi.org/10.1017/S0140525X12000477
Clark, A. (2015). Surfing uncertainty: Prediction, action and the embodied mind. Oxford University Press.
Di Dio, C., Ardizzi, M., Massaro, D., Di Cesare, G., Gilli, G., Marchetti, A., & Gallese, V. (2016). Human, nature, dyna-
mism: The effects of content and movement perception on brain activations during the aesthetic judgment of represen-
tational paintings. Frontiers in Human Neuroscience, 9. http://www.frontiersin.org/article/10.3389/fnhum.2015.00705
Djebbara, Z., Fich, L. B., & Gramann, K. (2019). Understanding perceptual experience of art using mobile brain/body
imaging. In A. Nijholt (Ed.), Brain art: Brain–computer interfaces for artistic expression (1st ed.). Springer Nature. https://
doi.org/10.1007/978-3-030-14323-7_9
Djebbara, Z., Fich, L. B., & Gramann, K. (2021). The brain dynamics of architectural affordances during transition.
Scientific Reports, 11(1), 2796. https://doi.org/10.1038/s41598-021-82504-w
Djebbara, Z., Fich, L. B., Petrini, L., & Gramann, K. (2019). Sensorimotor brain dynamics reflect architectural affor-
dances. Proceedings of the National Academy of Sciences of the United States of America, 116(29), 14769–14778. https://
doi.org/10.1073/pnas.1900648116
Doeller, C. F., & Burgess, N. (2008). Distinct error-correcting and incidental learning of location relative to landmarks
and boundaries. Proceedings of the National Academy of Sciences of the United States of America, 105(15), 5909–5914.
https://doi.org/10.1073/pnas.0711433105
Engel, A. K., Fries, P., & Singer, W. (2001). Dynamic predictions: Oscillations and synchrony in top—Down processing.
Nature Reviews. Neuroscience, 2(10), 704–716. https://doi.org/10.1038/35094565
Epstein, R. A.,  & Baker, C. I. (2019). Scene perception in the human brain. Annual Review of Vision Science, 5(1),
373–397. https://doi.org/10.1146/annurev-vision-091718-014809
Epstein, R. A., Graham, K. S., & Downing, P. E. (2003). Viewpoint-specific scene representations in human parahip-
pocampal cortex. Neuron, 37(5), 865–876. https://doi.org/10.1016/S0896-6273(03)00117-X
Epstein, R. A., & Kanwisher, N. (1998). A cortical representation of the local visual environment. Nature, 392(6676),
598–601. https://doi.org/10.1038/33402
Epstein, R. A., Parker, W. E.,  & Feiler, A. M. (2007). Where am I  now? Distinct roles for parahippocampal and
retrosplenial cortices in place recognition. Journal of Neuroscience, 27(23), 6141–6149. https://doi.org/10.1523/
JNEUROSCI.0799-07.2007
Fich, L. B., Jönsson, P., Kirkegaard, P. H., Wallergård, M., Garde, A. H., & Hansen, Å. (2014). Can architectural design
alter the physiological reaction to psychosocial stress? A  virtual TSST experiment. Physiology and Behavior, 135,
91–97. https://doi.org/10.1016/j.physbeh.2014.05.034

361
Djebbara, Fich and Vecchiato

Fiset, P., Paus, T., Daloze, T., Plourde, G., Meuret, P., Bonhomme, V., Hajj-Ali, N., Backman, S. B.,  & Evans, A.
C. (1999). Brain mechanisms of propofol-induced loss of consciousness in humans: A  positron emission tomo-
graphic study. Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 19(13), 5506–5513. https://doi.
org/10.1523/JNEUROSCI.19-13-05506.1999
Flexas, A., Rosselló, J., de Miguel, P., Nadal, M., & Munar, E. (2014). Cognitive control and unusual decisions about
beauty: An fMRI study. Frontiers in Human Neuroscience, 8, 520. https://doi.org/10.3389/fnhum.2014.00520
Friston, K. (2010). The free-energy principle: A  unified brain theory? Nature Reviews. Neuroscience, 11(2), 127–138.
https://doi.org/10.1038/nrn2787
Gallagher, S. (2017). Enactivist interventions: Rethinking the mind (1st ed.). Oxford University Press.
Gallivan, J. P., Chapman, C. S., Wolpert, D. M., & Flanagan, J. R. (2018). Decision-making in sensorimotor control.
In Nature reviews neuroscience (Vol. 19, Issue 9, pp.  519–534). Nature Publishing Group. https://doi.org/10.1038/
s41583-018-0045-9
Gallivan, J. P., Logan, L., Wolpert, D. M.,  & Flanagan, J. R. (2016). Parallel specification of competing sensorimo-
tor control policies for alternative action options. Nature Neuroscience, 19(2), 320–326. https://doi.org/10.1038/
nn.4214
Gibson, J. (1986). The ecological approach to visual perception. Psychology Press, Taylor & Francis Group.
Gloor, P. (1969). Hans Berger on electroencephalography. American Journal of EEG Technology, 9(1), 1–8. https://doi.org/
10.1080/00029238.1969.11080728
Gore, J. C. (2003). Principles and practice of functional MRI of the human brain. Journal of Clinical Investigation, 112(1),
4–9. https://doi.org/10.1172/JCI19010
Gramann, K., Ferris, D. P., Gwin, J.,  & Makeig, S. (2014). Imaging natural cognition in action. International Jour-
nal of Psychophysiology: Official Journal of the International Organization of Psychophysiology, 91(1), 22–29. https://doi.
org/10.1016/j.ijpsycho.2013.09.003
Gramann, K., Gwin, J. T., Ferris, D. P., Oie, K., Jung, T. P., Lin, C. T., Liao, L. D., & Makeig, S. (2011). Cognition in
action: Imaging brain/body dynamics in mobile humans. Reviews in the Neurosciences, 22(6), 593–608. https://doi.
org/10.1515/RNS.2011.047
Hamid, A. A., Pettibone, J. R., Mabrouk, O. S., Hetrick, V. L., Schmidt, R., Vander Weele, C. M., Kennedy, R. T.,
Aragona, B. J., & Berke, J. D. (2016). Mesolimbic dopamine signals the value of work. Nature Neuroscience, 19(1),
117–126. https://doi.org/10.1038/nn.4173
Hartley, T., Maguire, E. A., Spiers, H. J., & Burgess, N. (2003). The well-worn route and the path less traveled: Dis-
tinct neural bases of route following and wayfinding in humans. Neuron, 37(5), 877–888. https://doi.org/10.1016/
S0896-6273(03)00095-3
Holl, S., Pallasmaa, J., & Pérez-Gómez, A. (2006). Questions of perception: Phenomenology of architecture (1st ed.). William
Stout Publishers.
Höller, Y., Uhl, A., Bathke, A., Thomschewski, A., Butz, K., Nardone, R., Fell, J., & Trinka, E. (2017). Reliability of
EEG measures of interaction: A paradigm shift is needed to fight the reproducibility crisis. Frontiers in Human Neurosci-
ence, 11, 441. https://doi.org/10.3389/fnhum.2017.00441
Jacobsen, T., Schubotz, R. I., Höfel, L., & v Cramon, D. Y. (2006). Brain correlates of aesthetic judgment of beauty.
Neuroimage, 29(1), 276–285. https://doi.org/10.1016/j.neuroimage.2005.07.010
Jensen, O., Gelfand, J., Kounios, J., & Lisman, J. E. (2002). Oscillations in the alpha band (9–12 Hz) increase with mem-
ory load during retention in a short-term memory task. Cerebral Cortex, 12(8), 877–882. https://doi.org/10.1093/
cercor/12.8.877
Kandel, E. R. (2013). Principles of neural science ( J. Schwarts, T. Jessell, S. Siegelbaum, & A. J. Hudspeth, Eds., 5th ed.).
McGrow-Hill.
Kaplan, R.,  & Kaplan, S. (1989). The experience of nature: A  psychological perspective. In The experience of nature:
A psychological perspective. Cambridge University Press.
Kaplan, S. (1995). The restorative benefits of nature: Toward an integrative framework. Journal of Environmental Psychology,
15(3), 169–182. https://doi.org/10.1016/0272-4944(95)90001-2
Kitada, R., Sasaki, A. T., Okamoto, Y., Kochiyama, T., & Sadato, N. (2014). Role of the precuneus in the detection of
incongruency between tactile and visual texture information: A functional MRI study. Neuropsychologia, 64, 252–262.
https://doi.org/10.1016/j.neuropsychologia.2014.09.028
Kringelbach, M. L.,  & Rolls, E. T. (2004). The functional neuroanatomy of the human orbitofrontal cortex: Evi-
dence from neuroimaging and neuropsychology. Progress in Neurobiology, 72(5), 341–372. https://doi.org/10.1016/j.
pneurobio.2004.03.006
Lancaster, J. L., Woldorff, M. G., Parsons, L. M., Liotti, M., Freitas, C. S., Rainey, L., Kochunov, P. V., Nickerson, D.,
Mikiten, S. A., & Fox, P. T. (2000). Automated Talairach atlas labels for functional brain mapping. Human Brain Map-
ping, 10(3), 120–131. https://doi.org/10.1002/1097-0193(200007)10:3<120::aid-hbm30>3.0.co;2-8

362
Making sense of space

Laureys, S., Goldman, S., Phillips, C., Van Bogaert, P., Aerts, J., Luxen, A., Franck, G., & Maquet, P. (1999). Impaired
effective cortical connectivity in vegetative state: Preliminary investigation using PET. NeuroImage, 9(4), 377–382.
https://doi.org/10.1006/nimg.1998.0414
Logothetis, N. K. (2003). The underpinnings of the BOLD functional magnetic resonance imaging signal. Journal of
Neuroscience, 23(10), 3963–3971. https://doi.org/10.1523/JNEUROSCI.23-10-03963.2003
Luck, S. J. (1963 [2005]). An introduction to the event-related potential technique (2nd ed.). MIT Press.
Makeig, S., Gramann, K., Jung, T. P., Sejnowski, T. J., & Poizner, H. (2009). Linking brain, mind and behavior. Interna-
tional Journal of Psychophysiology, 73(2), 95–100. https://doi.org/10.1016/J.IJPSYCHO.2008.11.008
Maquet, P., Degueldre, C., Delfiore, G., Aerts, J., Péters, J. M., Luxen, A.,  & Franck, G. (1997). Functional neu-
roanatomy of human slow wave sleep. Journal of Neuroscience, 17(8), 2807–2812. https://doi.org/10.1523/
JNEUROSCI.17-08-02807.1997
Maranesi, M., Bruni, S., Livi, A., Donnarumma, F., Pezzulo, G.,  & Bonini, L. (2019). Differential neural dynamics
underling pragmatic and semantic affordance processing in macaque ventral premotor cortex. Scientific Reports, 9(1),
11700. https://doi.org/10.1038/s41598-019-48216-y
Migo, E. M., O’Daly, O., Mitterschiffthaler, M., Antonova, E., Dawson, G. R., Dourish, C. T., Craig, K. J., Simmons,
A., Wilcock, G. K., McCulloch, E., Jackson, S. H. D., Kopelman, M. D., Williams, S. C. R., & Morris, R. G. (2016).
Investigating virtual reality navigation in amnestic mild cognitive impairment using fMRI. Aging, Neuropsychology, and
Cognition, 23(2), 196–217. https://doi.org/10.1080/13825585.2015.1073218
Mills, A. F., Sakai, O., Anderson, S. W., & Jara, H. (2017). Principles of quantitative MR imaging with illustrated review
of applicable modular pulse diagrams. In Radiographics (Vol. 37, Issue 7, pp.  2083–2105). Radiological Society of
North America Inc. https://doi.org/10.1148/rg.2017160099
Murias, K., Slone, E., Tariq, S., & Iaria, G. (2019). Development of spatial orientation skills: An fMRI study. Brain Imag-
ing and Behavior, 13(6), 1590–1601. https://doi.org/10.1007/s11682-018-0028-5
Pallasmaa, J. (2012). The eyes of the skin: Architecture and the senses. Wiley.
Palmer, J. A., Kreutz-Delgado, K., & Makeig, S. (2011). AMICA: An adaptive mixture of independent component analyzers
with shared components. Technical Report, Swartz Center for Computational Neuroscience.
Parada, F. J. (2018). Understanding natural cognition in everyday settings: 3 pressing challenges. Frontiers in Human Neu-
roscience, 12, 386. https://doi.org/10.3389/fnhum.2018.00386
Parada, F. J., & Rossi, A. (2020). Perfect timing: Mobile brain/body imaging scaffolds the 4E-cognition research program.
European Journal of Neuroscience, 54(12), 8081–8091. https://doi.org/10.1111/ejn.14783
Parvizi, J., Van Hoesen, G. W., Buckwalter, J., & Damasio, A. (2006). Neural connections of the posteromedial cor-
tex in the macaque. Proceedings of the National Academy of Sciences, 103(5), 1563–1568. https://doi.org/10.1073/
pnas.0507729103
Pastor-Bernier, A., & Cisek, P. (2011). Neural correlates of biased competition in premotor cortex. Journal of Neuroscience,
31(19), 7083–7088. https://doi.org/10.1523/JNEUROSCI.5681-10.2011
Pessoa, L. (2014). Understanding brain networks and brain organization. Physics of Life Reviews, 11(3), 400–435. https://
doi.org/10.1016/j.plrev.2014.03.005
Pezzulo, G., & Cisek, P. (2016). Navigating the affordance landscape: Feedback control as a process model of behavior
and cognition. Trends in Cognitive Sciences, 20(6), 414–424. https://doi.org/10.1016/j.tics.2016.03.013
Pezzulo, G., Rigoli, F., & Friston, K. (2015). Active Inference, homeostatic regulation and adaptive behavioural control.
In Progress in neurobiology (Vol. 134, pp. 17–35). Elsevier Ltd. https://doi.org/10.1016/j.pneurobio.2015.09.001
Pine, D. S., Grun, J., Maguire, E. A., Burgess, N., Zarahn, E., Koda, V., Fyer, A., Szeszko, P. R., & Bilder, R. M. (2002).
Neurodevelopmental aspects of spatial navigation: A virtual reality fMRI study. NeuroImage, 15(2), 396–406. https://
doi.org/10.1006/nimg.2001.0988
Presti, P., Ruzzon, D., Avanzini, P., Caruana, F., Rizzolatti, G., & Vecchiato, G. (2021). Dynamic experience of archi-
tectural forms affects arousal and valence perception in virtual environments. Preprint. https://doi.org/10.21203/
rs.3.rs-910384/v1.
Raichle, M. E. (2003). Functional brain imaging and human brain function. In Journal of neuroscience (Vol. 23, Issue 10,
pp. 3959–3962). Society for Neuroscience. https://doi.org/10.1523/JNEUROSCI.23-10-03959.2003
Rizzolatti, G., Camarda, R., Fogassi, L., Gentilucci, M., Luppino, G., & Matelli, M. (1988). Functional organization
of inferior area 6 in the macaque monkey. Experimental Brain Research, 71(3), 491–507. https://doi.org/10.1007/
BF00248742
Saalmann, Y. B., Pinsk, M. A., Wang, L., Li, X., & Kastner, S. (2012). The pulvinar regulates information transmis-
sion between cortical areas based on attention demands. Science (New York, NY), 337(6095), 753–756. https://doi.
org/10.1126/science.1223082
Schmid, M. C., Singer, W., & Fries, P. (2012). Thalamic coordination of cortical communication. Neuron, 75(4), 551–
552. https://doi.org/10.1016/j.neuron.2012.08.009

363
Djebbara, Fich and Vecchiato

Sherman, S. M. (2017). Functioning of circuits connecting thalamus and cortex. In Comprehensive physiology (Vol. 7, Issue
2, pp. 713–739). John Wiley & Sons, Inc. https://doi.org/10.1002/cphy.c160032
Sherman, S. M. (2016). Thalamus plays a central role in ongoing cortical functioning. Nature Neuroscience, 19(4), 533–
541. https://doi.org/10.1038/nn.4269
Sherman, S. M., & Guillery, R. W. (1996). Functional organization of thalamocortical relays. Journal of Neurophysiology,
76(3), 1367–1395. https://doi.org/10.1152/jn.1996.76.3.1367
Sherman, S. M., & Guillery, R. W. (2002). The role of the thalamus in the flow of information to the cortex. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 357(1428), 1695–1708. https://doi.org/10.1098/
rstb.2002.1161
Shohamy, D., & Adcock, R. A. (2010). Dopamine and adaptive memory. Trends in Cognitive Sciences, 14(10), 464–472.
https://doi.org/10.1016/j.tics.2010.08.002
Singer, W. (1999). Neuronal synchrony: A  versatile code for the definition of relations? In Neuron (Vol. 24, Issue 1,
pp. 49–65). Cell Press. https://doi.org/10.1016/S0896-6273(00)80821-1
Skov, M. (2019). Aesthetic appreciation: The view from neuroimaging. Empirical Studies of the Arts, 37(2), 220–248.
https://doi.org/10.1177/0276237419839257
Spiers, H. J., & Maguire, E. A. (2006). Thoughts, behaviour, and brain dynamics during navigation in the real world.
NeuroImage, 31(4), 1826–1840. https://doi.org/10.1016/j.neuroimage.2006.01.037
Sporns, O. (2011). Networks of the brain (1st ed.). MIT Press.
Stamps, A. E. (2005). Visual permeability, locomotive permeability, safety, and enclosure. Environment and Behavior, 37(5),
587–619. https://doi.org/10.1177/0013916505276741
Stroop, J. R. (1935). Studies of interference in serial verbal reactions. Journal of Experimental Psychology, 18(6), 643–662.
https://doi.org/10.1037/h0054651
Taube, J. S., Valerio, S., & Yoder, R. M. (2013). Is navigation in virtual reality with fMRI really navigation? Journal of
Cognitive Neuroscience, 25(7), 1008–1019. https://doi.org/10.1162/jocn_a_00386
Theyel, B. B., Llano, D. A., & Sherman, S. M. (2010). The corticothalamocortical circuit drives higher-order cortex in
the mouse. Nature Neuroscience, 13(1), 84–88. https://doi.org/10.1038/nn.2449
Thibault, R. T., Lifshitz, M., Jones, J. M., & Raz, A. (2014). Posture alters human resting-state. Cortex, 58, 199–205.
https://doi.org/10.1016/j.cortex.2014.06.014
Thibault, R. T., Lifshitz, M., & Raz, A. (2016). Body position alters human resting-state: Insights from multi-postural
magnetoencephalography. Brain Imaging and Behavior, 10(3), 772–780. https://doi.org/10.1007/s11682-015-9447-8
Thompson, E., & Varela, F. J. (2001). Radical embodiment: Neural dynamics and consciousness. Trends in Cognitive Sci-
ences, 5(10), 418–425. https://doi.org/10.1016/S1364-6613(00)01750-2
Ulrich, R. S. (1981). Natural versus urban scenes: Some psychophysiological effects. Environment and Behavior, 13(5),
523–556. https://doi.org/10.1177/0013916581135001
Ulrich, R. S. (1983). Aesthetic and affective response to natural environment. Human Behavior and Environment, 6,
85–125.
Ulrich, R. S. (1984). View through a window may influence recovery from surgery. Science, 224(4647), 420–421.
https://doi.org/10.1126/science.6143402
Varela, F. G., Lachaux, J. P., Rodriguez, E., & Martinerie, J. (2001). The brainweb: Phase synchronization and large-scale
integration. Nature Reviews Neuroscience, 2(4), 229–239. https://doi.org/10.1038/35067550
Varela, F. G., Maturana, H. R., & Uribe, R. (1974). Autopoiesis: The organization of living systems, its characterization
and a model. Currents in Modern Biology, 5(4), 187–196. https://doi.org/10.1016/0303-2647(74)90031-8
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Gonzalez-Mora, J. L., Leder, H., Modroño, C., Nadal, M.,
Rostrup, N., & Skov, M. (2015). Architectural design and the brain: Effects of ceiling height and perceived enclosure
on beauty judgments and approach-avoidance decisions. Journal of Environmental Psychology, 41, 10–18. https://doi.
org/10.1016/j.jenvp.2014.11.006
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., Nadal, M., Rostrup, N., & Skov,
M. (2013). Impact of contour on aesthetic judgments and approach-avoidance decisions in architecture. Proceed-
ings of the National Academy of Sciences of the United States of America, 110, 10446–10453. https://doi.org/10.1073/
pnas.1301227110
Vecchiato, G., Jelic, A., Tieri, G., Maglione, A. G., De Matteis, F., & Babiloni, F. (2015). Neurophysiological correlates of
embodiment and motivational factors during the perception of virtual architectural environments. Cognitive Processing,
16(1), 425–429. https://doi.org/10.1007/s10339-015-0725-6
Vecchiato, G., Tieri, G., Jelic, A., De Matteis, F., Maglione, A. G., & Babiloni, F. (2015). Electroencephalographic cor-
relates of sensorimotor integration and embodiment during the appreciation of virtual architectural environments.
Frontiers in Psychology, 6, 1944. https://doi.org/10.3389/fpsyg.2015.01944

364
Making sense of space

Viaene, A. N., Petrof, I., & Sherman, S. M. (2011). Properties of the thalamic projection from the posterior medial
nucleus to primary and secondary somatosensory cortices in the mouse. Proceedings of the National Academy of Sciences
of the United States of America, 108(44), 18156–18161. https://doi.org/10.1073/pnas.1114828108
Wolff, M., Morceau, S., Martin-Cortecero, J., Folkard, R., & Groh, A. (2020). A thalamic bridge from sensory percep-
tion to cognition. Neuroscience and biobehavioral Reviews. https://doi.org/10.1016/j.neubiorev.2020.11.013
Yu, C., Zhou, Y., Liu, Y., Jiang, T., Dong, H., Zhang, Y., & Walter, M. (2011). Functional segregation of the human
cingulate cortex is confirmed by functional connectivity based neuroanatomical parcellation. NeuroImage, 54(4),
2571–2581. https://doi.org/10.1016/j.neuroimage.2010.11.018
Zhao, X., Wang, J., Li, J., Luo, G., Li, T., Chatterjee, A., Zhang, W., & He, X. (2020). The neural mechanism of aes-
thetic judgments of dynamic landscapes: An fMRI study. Scientific Reports, 10(1), 20774. https://doi.org/10.1038/
s41598-020-77658-y

365
18
LITERATURE AND POETRY
Arthur M. Jacobs

When the literary critic I.A. Richards (1929) had 100 students write an essay about 13 poems of presumed dif-
ferential poetic quality, his subjective conclusion after a qualitative analysis of these essays was that he got “a hun-
dred verdicts from a hundred readers” (Martindale & Dailey, 1995). However, when using the same 13 poems
as Richards about 70 years later while applying quantitative methods, Martindale and Dailey (1995, p. 307)
showed that “whether using rating scales or writing essays, people agree quite well about the connotative mean-
ing of poetry.” These two studies represent quite well the basic tension in the field of theoretical and empirical
studies of literature stretched by many controversies such as those between close vs. distant reading (Herrmann
et  al., 2021), direct vs. indirect methods (Dixon  & Bortolussi, 2016), feature-driven vs. subjective feelings
approaches (Belfi et al., 2018; Hoffstaedter, 1987), or natural vs. manipulated texts designs ( Jacobs, 2015c).
This chapter focuses on ways to overcome these seemingly irreconcilable controversies within the emergent
neurocognitive poetics perspective ( Jacobs, 2011, 2015a, 2015b, 2015c; Schrott & Jacobs, 2011; Willems &
Jacobs, 2016). In a nutshell, this perspective adds models and methods of cognitive neuroscience, machine
learning, and data science to those of classical and modern (cognitive) poetics, as pioneered by Aristotle, Jakob-
son, and others. Guided by a comprehensive model of literary reading ( Jacobs, 2011, 2015a, 2015b, 2021;
Jacobs & Willems, 2019; Nicklas & Jacobs, 2017), it attempts to integrate qualitative and quantitative methods
in a spiral approach in which “subjective” experiential data are not considered simply as the exploratory foreplay
of quantitative, “objective” behavioural or neuroimaging data but both contribute to hypothesis testing and
model development in their own right, ideally in designs combining them ( Jacobs, 2015c, 2016).

Neuroaesthetics and verbal art


There are some persistent obstacles that appear to be inherent in the empirical studies of art and
literature . . .: i) there is no single essence to the quality of aesthetic objects and to aesthetic, ii) it
is difficult to consciously identify and articulate the components of aesthetic experience, and iii)
there is a lack of sophisticated theories and mini theories.
(Graesser et al., 1996)

While research on the neuroaesthetics of nonverbal art has rapidly and continuously developed since the early
works of the 90s (for review, see Chatterjee & Vartanian, 2016; Nadal & Chatterjee, 2019), verbal art was

366 DOI: 10.4324/9781003008675-20


Literature and poetry

long ignored by cognitive neuroscientists and psychologists alike, and despite a productive intermezzo around
the middle of the last decade, practically continues to be so. This neglect has to do with (psycho)linguists’
long-term focus on non-literary language as study material and a concomitant indifferentism in affective and
aesthetic processes associated with language and reading. It also did not help that from the beginnings of psy-
chology and empirical aesthetics, psychological theories of emotion and aesthetics simply were not interested
in language or literature reception. Finally, even my own field, experimental reading research, almost totally
neglected verbal art as study material until quite recently with the emergence of the neurocognitive poetics
perspective ( Jacobs, 2015a; Schrott & Jacobs, 2011).
A reason that is often put forward for this wanting situation is that poetic texts may be the most complex
(i.e., high-dimensional) stimulus materials experimental scientists may have to deal with and that the prac-
tically unlimited number of potentially intervening features associated with the (con)text-author-reader/
listener nexus involved in verbal art reception hinders proper scientific investigation. However, this argument
conceals another crucial aspect: in contrast to much of nonverbal art stimuli, language is a highly structured
stimulus, and formal written language as it comes in most poetic texts is even more structured. Words, sen-
tences, and poems not only follow transparent rules, but many of their features, such as syllables, syntactic
simplicity, or metre can easily be quantified.
Indeed, already in the 1950s, it was proposed that prose texts can be statistically analysed in analogy
with a gas or fluid mixture and poetry with a mixed crystal, the words playing the role of atomic molecules
(Fucks, 1952). Thanks to the amazing technological advances in machine learning and data science in the
last decade, even hundreds or thousands of dynamically interacting (text) features are no longer a prob-
lem for quantification facilitating experimental psychology or cognitive neuroscience studies of literature
(see Jacobs & Willems, 2019; Willems & Jacobs, 2016). Thus, extremely powerful neural networks, like
GPT3 (https://openai.com/blog/openai-api/), with its dozens of layers and billions of free parameters,
that are trained with millions of papers or books can nowadays generate scientific essays, stories, or poems
hardly distinguishable from human-made ones, a development which only five years ago was deemed
impossible.
Since quantification—or at least categorical/typological specification allowing experimental
manipulation—of text features and predictive modelling based on machine learning methods is indispensable
for (neuro)scientific studies of literature, a short excursion on work in this field is in order (see the following).
Such methods enormously facilitate informed answers (based on quantitative predictions) to the basic ques-
tion of neurocognitive poetics: By which means do authors bring readers/listeners to shed tears, experience
unforgettable blissful moments, or even change their lives? Or, stated differently in the words of van Peer
(1986): “which kind of literary effect is caused in the mind of the reader by which kind of stylistic devices?”
The first step in answering questions about how a literary text achieves its affective and aesthetic effects at
the neuronal, behavioural, and experiential levels of observation is quantitative narrative analysis (QNA) or
its twin, qualitative-quantitative narrative analysis (Q2NA, Jacobs, 2018b; Gambino et al., 2020), which is
indispensable wherever texts offer figurative language using devices like metaphor or chiasmus that still resist
error-free automatized detection or quantification and thus require expert evaluation.

Neurocognitive studies on literature reception


There’s an undeniable scarcity of studies using standard neuroscientific methods (i.e., electrophysiological
and neuroimaging techniques) to investigate the reading of or listening to literature in general or poetry in
particular (i.e., entire novels or poems). Therefore, here I take a somewhat broader perspective on both the
stimulus materials and the methods side. First, I’ll include studies using stimuli that can be subsumed under
the term “poetic language” in a loose sense. Second, I’ll also include studies using methods that strictly

367
Arthur M. Jacobs

speaking are not neuroscientific but closely related and thus can inform about the neurobiology underlying
the appreciation of literature, such as eye tracking or heart rate variability.

Poetic language processing


The question whether a text was processed poetically or not was decided on the basis of the sub-
jects’ judgements on a scale from 1 (= non-poetic) to 5 (= poetic).
(Hoffstaedter, 1987, p. 77)

First, as regards linguistics and literature theory, there seems to be no absolute criterion to separate literary
from non-literary or poetic from non-poetic language (e.g., Jacobs  & Kinder, 2018). Second, whether a
reader or reader group has processed a text generally admitted to be poetic—such as Shakespeare’s sonnets—
in a “poetic” way is a matter of education, vocabulary, language skills, motivation, context, or personality
factors, to name just a few intervening variables. Moreover, even “expert” readers may discover special stylis-
tic features of a text only during a second reading while still missing others (e.g., obscure metaphors). Thus,
selecting studies on “poetic language” processing based on the fact that they used texts labelled as poetry is
no guarantee, since uninformed or disengaged readers may very well miss most or all of the “poetic” features
of the texts, at least during the first reading. For empirical studies of literature to be promising, it is therefore
useful, if not necessary, to apply methods allowing qualification and quantification of both the poetic poten-
tial of the texts and the reception skills of the readers.
The pragmatic approach to the issue of poeticity and poetic language processing I choose here integrates
the two main practices found in the field: (1) studies focusing on text features in the tradition of modern
or cognitive poetics, which typically do not collect any reader response data (e.g., Jakobson & Lévi-Strauss,
1962; Stockwell, 2005), and (2) studies focusing on reader responses, which typically do not apply QNA or
Q2NA to texts (e.g., Belfi et al., 2018; Hoffstaedter, 1987; Miall & Kuiken, 1994). Thus, if a text has a high
poetic potential, that is, contains a certain density of formal features associated with the genre poetry at the
four levels proposed by Jakobson, and if reader responses to this text reveal aesthetic appreciation—ideally in
a way predicted by these features—for example, via machine learning–assisted predictive modelling, then it
is safe to assume that the study offers data relevant to poetic language processing.
Reflecting two distinct methodological traditions in psychology (i.e., the experimental and correlational),
as well as two schools of thought in the field of empirical studies of literature (cf. Bortolussi & Dixon, 2003),
most studies on poetic text materials used either an experimental procedure including text manipulations
aimed at discovering causal relations or a procedure using natural, unmanipulated texts producing correla-
tional data. Each has its strengths and weaknesses (cf. Jacobs, 2015c), but, as was already shown in a pioneer-
ing study by Hoffstaedter (1987; see also Dixon et al., 1993), both approaches can also be usefully combined.
Studies that use natural texts can further be divided into ones that include a qualitative or quantitative text
feature analysis and those which don’t.

Doctored text materials


The pioneering studies on poetry reception by van Peer (1983, 1986) made use of what the author called
“doctored poems,” text versions in which certain words of the originals (poems by Emily Dickinson, E.E.
Cummings, and Dylan Thomas) were changed to decrease the poetic potential and induce affective-aesthetic
effects. An example is the manipulation of parallelism by the change of Dickinson’s line The Brain is deeper
than the sea into: The Brain is calmer than the sea, thereby removing the assonance of deeper-sea and poten-
tially making the line less poetic. Based on foregrounding theory (for review, see van Peer et al., 2021), van

368
Literature and poetry

Peer predicted that such a “deactivation” or removal of stylistic features would decrease e.g., poeticity ratings
of the poems. The results by and large confirmed the general prediction (with variations across poems and
thus stylistic devices). The experimental design of this early study did not allow testing of how much each
of the chosen stylistic features contributed to the perceived poeticity of a poem, nor did it allow testing of
parametric effects of the degree of foregrounding. However, in a later study, van Peer et al. (2007) created
six versions—in decreasing order of poetic potential—of a single line of the poem “Espectros” by the post-
Romantic Portuguese poet Antero de Quental. The results were mixed, but grosso modo indicated that the
degree of foregrounding affected readers’ aesthetic appreciation in the predicted direction.
Various studies following in the footsteps of van Peer used doctored texts (e.g., Hanauer, 1998; Menning-
haus et al., 2014), but only a handful also applied neurocognitive methods. Two pioneering studies from the
same year (Hoorn, 1996; Pynte et al., 1996) used electrophysiological methods (i.e., event-related potentials;
ERPs) to study metaphor comprehension (the metaphors used in both studies were rather simplistic and non-
poetic, though). Curiously enough, in the same year, the German poet Durs Grünbein (1996) had introduced
the term ‘‘factor N400’’ (named after the N400 component of ERP known to signal semantic anomalies) as
a general proxy for foregrounding features as “brainphysiological attention catchers” (cf. Jacobs, 2015b). He
speculated that metaphors like “urne and uterus” or “term of endearment and cruelty” cause neurolinguistic
clashes and called for a poetry full of images rich in factor N400. The results of both neurocognitive stud-
ies fit with the idea that the N400 component somehow reflects metaphoric processing but—according to
Hoorn—rather than being part of special metaphor processes, it would inform readers about the literal status
of an expression.
Only about 20 years later, Bambini et al. (2018) were presumably the first to study ERPs generated by
natural poetic metaphors in their original context. The stimuli were 62 genitival metaphorical constructions
of the “A of B” form (e.g., grass of velvet, sky of pearl, eyes of steel) embedded in their original context
(i.e., excerpts from either Italian poems or novels from the 19th and 20th centuries) as compared with lit-
eral expressions (e.g., throne of velvet). The results showed more negative ERPs for metaphors, unfolding
in an N400 followed by a sustained negativity over frontal sites, which the authors interpreted as reflecting
the manipulation of multiple meanings in working memory, possibly responsible for the poetic effect. This
late negativity effect was driven by familiarity, with a more negative response for less familiar metaphors.
The fact that Bambini et al.’s (2018) study stands out because of its use of poetic metaphors (i.e., metaphors
used in poetic texts) reflects the wanting state of the art in the field. While there are plenty of behavioural
and neurocognitive studies looking at the processing of non-poetic metaphors (e.g., Forgács et al., 2012),
only very few behavioural studies so far investigated the comprehension of poetic metaphors (for review, see
Jacobs & Kinder, 2017).
Another pioneering study using “doctored” materials by Thierry et al. (2008) was dedicated to Shake-
speare’s artful playing with part of speech, that is, so-called functional shifts. Thus, the word conversion in
“lip something loving in my ear” instead of “whisper something loving in my ear” works against the laws of grammar
but—according to the authors—elicits a positive mental activity which should be revealed by means of brain
electrical methods. The stimuli were 40 original sentences containing the three most common types found
in Shakespeare’s works (noun-to-verb, verb-to-noun, and adjective-to-verb conversions) and 80 “doctored”
sentences in which the critical word was replaced by (1) a correct grammatical equivalent that could be
semantically expected, (2) a grammatically correct but semantically incongruent word, and (3) a semantically
and grammatically unsuitable word (e.g., I know you don’t want to speak, but lip [whisper/honey/bake] something
loving in my ear). In a meaningfulness decision task, the Shakespearean functional shift condition elicited sig-
nificant late anterior negativity/LAN and P600 modulations but failed to modulate the N400 component.
Thus, words which had their functional status changed triggered both an early syntactic evaluation process
thought to be mainly automatic and a delayed re-evaluation or repair process that is more controlled, but
semantic integration required no additional processing. The authors proposed that this dissociation between

369
Arthur M. Jacobs

syntactic and semantic evaluation enabled Shakespeare to create dramatic effects without diverting his public
away from meaning.
Five years later, Keidel et al. (2013) published a follow-up study on the poetic effects of functional shifts,
this time using the fMRI method. The results showed that sentences featuring functional shifts elicited
significant activation beyond brain regions classically activated by typical language tasks, including the
left caudate nucleus, the right inferior frontal gyrus, and the right inferior temporal gyrus. The authors’
interpretation was that these findings show how Shakespeare’s grammatical exploration forces the listener
to take a more active role in integrating the meaning of what is said. However, in the light of a particularly
revealing study by Bohrn et al. (2013) on the processing of German proverbs and antiproverbs, Keidel et
al.’s (2013) finding of ventral striatal activation (i.e., caudate nucleus) for artfully deviating sentences may
also indicate that listeners enjoyed a greater pleasure when hearing foregrounded language in the form of
functional shifts.
A study by Vaughan-Evans et al. (2016) investigated event-related brain potentials elicited by the final
word of sentences written in Cynghanedd (“harmony” in Welsh), an ancient poetic form that requires precise
consonantal repetition (and/or internal rhyme) in conjunction with distinct stress patterns. The results pro-
vided evidence for the idea that before readers even consider literal meaning, the musical (sound) properties
of poetry speak to the human mind in ways that escape consciousness, an idea that has inspired poets since
the beginning of poetry (Schrott & Jacobs, 2011) and that is confirmed by recent behavioural and neuro-
cognitive research on the role of phonological iconicity in word recognition and poetry reception (Aryani
et al., 2016, 2018a, 2018b, 2019; Aryani & Jacobs, 2018). Further brain-electric results by Chen et al. (2016)
suggested that top-down prosodic expectations regarding rhyme scheme can modulate early phonological
processing during silent reading.
In another neuroimaging study on the neural bases of literary awareness (i.e., the capacity to consider,
manipulate, and derive meaning in complex texts involving a more flexible situation model building process
for accommodating varying related meaning threads), O’Sullivan et al. (2015) used four-line texts, either
poetic or prosaic, presented line after line, in the scanner. Their results suggested that posterior parietal cor-
tex (PCC) activation was related to the extent to which a situation model has been updated, that anterior
temporal pole (ATL) activity was related to storing the narrative of a situation model, and that dorsomedial
prefrontal cortex (dmPFC) activity was related to force attention to settle on a narrative for a particular
(mental) simulation. Moreover, activation of the temporo-parietal junction (TPJ) and surrounding ventro-
lateral parietal areas was suggested to indicate reflexive updating of situation models in line with informa-
tion retrieved from memory, whereas the left inferior frontal gyrus (IFG) likely maintained the contextual
separation between representations that are similar, such as in metaphors. Texts with evolving meaning were
supposed to activate the ventromedial prefrontal cortex (vmPFC)—likely reflecting the motivational sig-
nificance of the developing meaning—as well as lateral anterior PFC, thought to be involved in construing
relationships between less directly related words/meaning threads.

Natural texts with feature analysis


Perhaps the exemplary behavioural study on reading natural literary texts including a feature analysis is the
one by Miall and Kuiken (1994). The authors used three short stories, divided into segments, and first
applied an intuitive text analysis to count the number of foregrounding features at the phonetic (e.g., allit-
eration), grammatical (e.g., number of subphrases), and semantic (e.g., metaphor) levels per segment. As
a proxy for the degree of foregrounding, segment feature density was then computed as the frequency of
each feature category per syllable. The authors also combined an indirect online measure (reading time as
determined by the self-paced reading technique during first reading) with a direct online measure (ratings
of strikingness and affect during rereading). Their main prediction was that the reader response measures

370
Literature and poetry

should all be positively correlated with feature density, for example, that text segments with higher densities
produce longer reading times and higher affect ratings. Overall, this prediction was confirmed, indicating
that foregrounding as a technique of prolonging processing by defamiliarization, with regard to everyday
communication, was correct.
Various studies followed Miall and Kuiken’s (1994) approach of determining text features—in a qualita-
tive, quantitative, or combined fashion—and then reporting correlations between feature type, frequency,
or value on the one hand and direct, indirect, online, or offline reader response measures on the other
(e.g., Jacobs et al., 2016; Xue et al., 2019, 2020). Since feature analysis methodology has seen a drastic pro-
gress in the last years and is indispensable for advancing neurocognitive poetics, the next paragraph presents
a short summary.
Machine Learning–Assisted QNA. Curiously enough, the presumably first works on computational
stylometry/QNA of literary texts, such as from Coledridge or Rilke, were published in Biometrika, dealing
with basic features such as sentence length (Yule, 1938) or word length distributions (Fucks, 1952). The
advent of computerized methods of text analysis starting in the 60s (Stone et al., 1966) together with the
onset of empirical studies of literature boosted by the creation of journals like Poetics or Journal of Literary
Semantics in the 70s led to a wealth of insights about the basic question of neurocognitive poetics, albeit
not yet at the neurocognitive level. Thus, using his Regressive Imagery Dictionary—a word list–based
automatized text analysis tool using 2900 words distributed across 29 categories such as sensation or social
behaviour—(Martindale, 1973), the author could show that literary change (e.g., from metaphysical to non-
metaphysical English poetry) contributes to how much pleasure had been induced by whatever readers had
understood or interpreted of a text (Martindale, 1984). Seven quantifiable text features were identified as
being important in this process of pleasurable reading, for example, variation in word or phrase length and
polarity (negative or positive valence of words), all co-determining the arousal potential of texts. More recent
computational stylistics studies using Q(2)NA and/or advanced sentiment analyses have identified a myriad
of features that may drive the aesthetic success of literary texts and poetry in particular (e.g., Delmonte,
2016; Gambino et al., 2020; Jockers & Witten, 2010; Kim et al., 2017; Moretti, 2000; Simonto, 1989, 1990;
Jacobs, 2018a, 2018b; for review, see Herrmann et al., 2021).

Natural texts without feature analysis


A main reason for discarding a feature-analytic approach was expressed by Carminati et al.’s (2006) study on
rhyme scheme in poetry:

When investigating authentic poetry, however, it is obviously impossible to exercise such control on
the materials: not only do poets differ in style and tone, but also sections taken from the same poem
and written by the same author can vary greatly in such features as foregrounding, enjambment,
punctuation, semantic shifts and rhyming (in)consistencies, all features that may affect reading time.

Although the early study by Miall and Kuiken (1994) already showed that Carminati et al.’s view does
not hold in general, data from behavioural and neurocognitive studies renouncing on feature analysis can
be informative and provide useful constraints for model development in neurocognitive poetics. Thus, an
extensive study by Lüdtke et al. (2014) on reading a poetry of mood corpus of 24 German poems from the
18th, 19th, and 20th centuries (including authors such as Hölderlin or Heym) tested the neurocognitive
poetics model (NCPM) of literary reading ( Jacobs, 2011, 2015a, 2015b, 2021; see Figure 18.1). In line with
the model’s postulate that backgrounding elements facilitate emotional involvement, while foreground-
ing features promote aesthetic evaluation, they identified different predictors for both processes: familiarity
and situational embedding were the main factors mediating mood empathy, and aesthetic liking was best

371
Arthur M. Jacobs

predicted from foregrounding features like style and form. (It is of note that in a later study, the data were
reanalysed in a feature-analytic approach; Jacobs et al., 2016.)
An early neuroimaging study by Zeman et  al. (2013) compared prose to poetry reading, finding that
areas specifically related to poetry reception were the right posterior/mid-cingulate, parahippocampal, and
left superior temporal gyrus, as well as the bilateral hippocampus. Interestingly, the dorsomedial prefrontal
cortex has been shown to exhibit reduced activation during reading of poetic pieces compared to the read-
ing of prosaic pieces (O’Sullivan et al., 2015). Similarly, the dorsolateral prefrontal cortex showed greater
deactivation in experts (compared to novices) during the generation of poetry, suggesting the experts’ capac-
ity to more readily suspend cognitive control and enter a state conducive to creative improvisation (Liu
et al., 2015). Finally, augmenting neuroimaging with inventive psychophysiology, Wassiliwizky et al. (2017)
showed that recited poetry can act as a powerful stimulus for eliciting peak emotional responses, including
chills and objectively measurable goosebumps that engage the primary reward circuitry in the brain. While
these responses to poetry were largely analogous to those found for music, their neural underpinnings also
showed differences in the role of the nucleus accumbens.
Although these studies suggest an awakening interest in the neuroscience of poetry reception and produc-
tion compared to the mainstream of cognitive neuroscience research, such research still is a very rare bird.
Thus, sadly enough, 25 years later, Graesser et al.’s (1996) previous citation still appears to be valid. The third
reason for this scarcity given in the citation—the lack of theoretical models—surely remains a major obstacle.
Thus, despite its first publication in 2011 ( Jacobs), the NCPM still seems pretty much alone in the space of
comprehensive neurocognitive models of literary and poetic language processing. The next section discusses
the most recent version of this model.

Neurocognitive models of poetic language processing


Figure 18.1 shows a simplified sketch of the NCPM, merging aspects from previous versions ( Jacobs, 2011,
2015a, 2015b; Jacobs & Willems, 2019; Nicklas & Jacobs, 2017). In a nutshell, the NCPM hypothesizes a
dual-route processing of literary texts. The first, depicted in the upper half of Figure 18.1, is a fast, automatic
route for (implicit) text processing dominantly—but not exclusively—operative during the reading of prose,
that is, texts in which background (BG) elements are much more likely than foreground (FG) elements and
thus have a relatively high BG/FG ratio (cf. Jacobs, 2015b). Those elements (e.g., familiar words with low
surprise values) inform readers about the facts of a story, allowing them to easily build up and update situa-
tion models and thus make meaning out of the series of sentences. When embedded in appropriate narra-
tive structures such as the universal suspense scheme (Brewer & Lichtenstein, 1982) supported by optimal
descriptive density, action density, and perceptual-motor enactment (cf. Jacobs & Lüdtke, 2017), those ele-
ments facilitate immersion; the experiencing of narrative or fiction emotions, such as sympathy, suspense, or
vicarious fear and hope; and finally easy meaning making.
As shown in the left upper panel of Figure 18.1 (A), the ventral and dorsal reading paths of the left hemi-
sphere are dominantly active here, including the ventral occipital cortex (yellow), medial temporal gyrus
(red), inferior frontal gyrus (blue), angular gyrus (green), superior parietal lobe (cyan), and supramarginal
gyrus (magenta). The right upper panel (B) shows neural networks important for (easy) situation model
building and updating, that is, the anterior temporal pole (blue), anterior and posterior cingulate cortex
(green, turquoise), and precuneus (red). Together with the amygdala (magenta), the anterior and medial
cingulate cortex (cyan, yellow) support the experience of fiction feelings.
The second, slow route is depicted in the lower panel of Figure  18.1. It is dominantly—but not
exclusively—activated during the reading of poetic texts characterized by a low BG/FG ratio. As hypoth-
esized in Jakobson’s (1960) famous “poetic function,” FG elements such as rhyme or metaphor attract read-
ers’ attention to formal text aspects and evoke a braking or even momentary stopping of the fluent reading

372
Literature and poetry

Figure 18.1 Extension of the neurocognitive poetics model sketching the likely main neural correlates of subprocesses
involved in implicit and explicit fiction processing, for example, situation model building, immersion, or
aesthetic appreciation. See text for details. Abbreviations: BG = Backgrounding elements, FG = Fore-
grounding elements.

process (often reflected by gaze refixations) due to difficult meaning making (open meaning gestalts) which
can start processes of self-reflection or a three-phase aesthetic trajectory characteristic for feelings of beauty
(Fitch et al., 2009): (1) familiar recognition; (2) surprise, ambiguity, and tension; and (3) closure of meaning
gestalts and tension. The third phase results from processes of integration and synthesis, occasionally sup-
plemented by an “aha” experience or feeling of good fit, “rightness,” or harmony motivating one to continue
to read even texts that are hard to comprehend. As shown in the left lower panel (C) the temporo-parietal
junction (cyan) and dorsolateral prefrontal cortex (magenta) support the effortful act of situation model
building and meaning computation during explicit figurative language processing, while—as shown in the
right lower panel (D)—the insula (red), OFC (blue) and caudate nucleus of the ventral striatum (green)
especially support aesthetic feelings.
The model has generated considerable research in neurocognitive poetics, and its main hypotheses—the
so-called Panksepp-Jakobson hypothesis and the fiction feeling hypothesis—have been tested and supported
by several studies (for detailed review, see Jacobs & Willems, 2019). The former claims that since evolution
did not have time to invent a proper neuronal system for art reception, and even less so for literary reading,
the affective and aesthetic processes we experience when reading must be linked to the ancient emotion
circuits we share with all mammals, as perhaps best described by Panksepp (1998), for example, the anterior
insula (Ziegler et al., 2018) or ventral striatum (Bohrn et al., 2013). Thus, aesthetic feelings during literature
reception are nothing special in the NCPM, much as for nonverbal art reception (Aleem et al., 2020; Skov &

373
Arthur M. Jacobs

Nadal, 2020a, 2020b). The latter hypothesis states that narratives with emotional content invite readers to
be empathic with the protagonists and immerse in the text world (e.g., by engaging the “affective empathy
network” of the brain) more than stories with neutral contents.
While the NCPM’s main predictions have been supported by the results of several computational and
experimental studies, there is evidence for crosstalk or partial overlap between the two routes in some con-
ditions (hence the dashed arrow in Figure 18.1). First, the results of Bohrn et al.’s (2013) study suggest a
potential contribution of the upper route to the aesthetic trajectory via feelings of familiarity and processing
fluency, as already hypothesized in Fitch et al.’s (2009) model. Second, there is some qualitative evidence
(i.e., obtained via interviews) for the possibility that foregrounding (as operationalized qualitatively by read-
ers reporting “a narrative element as being deviating”) is related to narrative absorption through its effects
on emotional engagement (Balint et  al., 2016). However, as the authors acknowledge themselves, their
qualitative data do not allow for either causal conclusions or generalizing statements and thus represent no
real constraint for a revision of the NCPM.
Still, much remains to be done, especially regarding the transformation of parts of the NCPM’s boxes
and arrows into computational sub-models allowing quantitative predictions (cf. Jacobs, 2021), but promis-
ing work is in progress. For example, reader-group–specific computational models simulating the seman-
tic memory or apperceptive mass (Kintsch, 1980) of, for example, young unexperienced vs. highly skilled
experienced readers are now in development ( Jacobs & Kinder, 2021). By modelling the associative net-
works underlying affective and aesthetic semantic judgments, a core part of the apperceptive mass crucial for
understanding and appreciating literature can be simulated and detailed predictions about which back- and
foreground elements of a text elicit which cognitive and emotional reader responses become possible, such
as which words or lines of a poem are perceived as the most beautiful ones (2018b; Jacobs & Kinder, 2019).
However, the scarcity of quantitative empirical data on poetry reception and the lack of standard databases
like the Gutenberg English Poetry Corpus ( Jacobs, 2018a) in languages other than English remain major
obstacles for the development of such computational models, which need both for proper cross-validation,
revision, and generalization.

Future challenges and goals


If poetry is the most challenging kind of fiction, potentially revealing new layers of meaning and opportuni-
ties for a rich variety of emotional responses such as sadness, joy, or aesthetic feelings at each and every (re)
reading act (Schrott & Jacobs, 2011), a major challenge lies in specifying and predicting the conditions under
which this meaning making (1) is easy or difficult, (2) is pleasant or not, and (3) elicits which emotional
responses. That is, we need models that offer quantitative predictions of both the comprehensibility and
enjoyment of verbal art. The lack of such models—in sharp contrast to the many psychological models of
nonverbal art reception (for review, see Pelowski et al., 2016)—is the number-one major challenge nowa-
days. Models require methods able to test their predictions. Challenge number two consists in convincing
researchers of scientific studies of literature to use more sophisticated methods combining direct with indi-
rect and online with offline measures to address at least two of the three levels of inquiry covered by the
NCPM (i.e., neural, experiential, and behavioural), as proposed in the debate between Dixon and Bortolussi
(2015), Kuiken (2015), and Jacobs (2015c, 2016). Thus, as stated initially, the integration of qualitative and
quantitative models and methods in a spiral approach with both contributing to hypothesis testing and model
development in their own right is what this field of research needs most.

374
Literature and poetry

References
Aleem, H., Correa-Herran, I., & Grzywacz, N. M. (2020). A theoretical framework for how we learn aesthetic values.
Frontiers in Human Neuroscience, 14, 345. https://doi.org/10.3389/fnhum.2020.00345
Aryani, A., Conrad, M., Schmidtke, D., & Jacobs, A. (2018a). Why “piss” is ruder than “pee”? The role of sound in
affective meaning making. PLoS One, 13(6), e0198430. https://doi.org/10.1371/journal.pone.0198430
Aryani, A., Hsu, C. T., & Jacobs, A. M. (2018b). The sound of words evokes affective brain responses. Brain Sciences, 8(6),
94. https://doi.org/10.3390/brainsci8060094
Aryani, A., Hsu, C.-T., & Jacobs, A. M. (2019). Affective iconic words benefit from additional sound—Meaning integra-
tion in the left amygdala. Human Brain Mapping, 15, 1–12.
Aryani, A., & Jacobs, A. M. (2018). Affective congruence between sound and meaning of words facilitates semantic deci-
sion. Behavioral Sciences, 8(6). https://doi.org/10.3390/bs8060056
Aryani, A., Kraxenberger, M., Ullrich, S., Jacobs, A. M., & Conrad, M. (2016). Measuring the basic affective tone of
poems via phonological saliency and iconicity. Psychology of Aesthetics, Creativity, and the Arts, 10(2), 191–204. https://
doi.org/10.1037/aca0000033
Bálint, K., Hakemulder, F., Kuijpers, M. M., Doicaru, M. M., & Tan, E. S. (2016). Reconceptualizing foregrounding:
Identifying response strategies to deviation in absorbing narratives. Scientific Study of Literature, 6(2), 176–207. https://
doi.org/10.1075/ssol.6.2.02bal
Bambini, V., Canal, P., Resta, D., & Grimaldi, M. (2018). Time course and neurophysiological underpinnings of meta-
phor in literary context. Discourse Processes, 56(1), 77–97. https://doi.org/10.1080/0163853X.2017.1401876
Belfi, A. M., Vessel, E. A., & Starr, G. G. (2018). Individual ratings of vividness predict aesthetic appeal in poetry. Psychol-
ogy of Aesthetics, Creativity, and the Arts, 12(3), 341–350. https://doi.org/10.1037/aca0000153
Bohrn, I. C., Altmann, U., Lubrich, O., Menninghaus, W., & Jacobs, A. M. (2013). When we like what we know—
A parametric fMRI analysis of beauty and familiarity. Brain and Language, 124(1), 1–8. http://doi.org/10.1016/j.
bandl.2012.10.003
Bortolussi, M.,  & Dixon, P. (2003). Psychonarratology: Foundations for the empirical study of literary response. Cambridge
University Press.
Brewer, W. F., & Lichtenstein, E. H. (1982). Stories are to entertain: A structural-affect theory of stories. Journal of Prag-
matics, 6(5–6), 473–486. https://doi.org/10.1016/0378-2166(82)90021-2
Carminati, M. N., Stabler, J., Roberts, A. M., & Fischer, M. H. (2006). Readers’ responses to sub-genre and rhyme
scheme in poetry. Poetics, 34(3), 204–218. http://doi.org/10.1016/j.poetic.2006.05.001
Chatterjee, A., & Vartanian, O. (2016). Neuroscience of aesthetics. Annals of the New York Academy of Sciences, 1369(1),
172–194. https://doi.org/10.1111/nyas.13035
Chen, Q., Zhang, J., Xu, X., Scheepers, C., Yang, Y., & Tanenhaus, M. K. (2016). Prosodic expectations in silent read-
ing: ERP evidence from rhyme scheme and semantic congruence in classic Chinese poems. Cognition, 154, 11–21.
http://doi.org/10.1016/j.cognition.2016.05.007
Delmonte, R. (2016). Exploring Shakespeare’s sonnets with SPARSAR. Linguistics and Literature Studies, 4(1), 61–95.
https://doi.org/10.13189/lls.2016.040110
Dixon, P., Bortolussi, M., Twilley, L. C., & Leung, A. (1993). Literary processing and interpretation: Towards empirical
foundations. Poetics, 22(1–2), 5–33. https://doi.org/10.1016/0304-422X(93)90018-C
Dixon, P., & Bortolussi, M. (2016). Measuring literary experience: Comment on Jacobs (2016). Scientific Study of Litera-
ture, 5(2), 178–182. https://doi.org/10.1075/ssol.5.2.03dix
Fitch, W. T., Graevenitz, A. V., & Nicolas, E. (2009). Bio-aesthetics and the aesthetic trajectory: A dynamic cognitive
and cultural perspective. In M. Skov & O. Vartanian (Eds.), Neuroaesthetics. Baywood.
Forgács, B., Bohrn, I. C., Baudewig, J., Hofmann, M. J., Pléh, C., & Jacobs, A. M. (2012). Neural correlates of com-
binatorial semantic processing of literal and figurative noun-noun compound words. Neuroimage, 63(3), 1432–1442.
https://doi.org/10.1016/j.neuroimage.2012.07.029
Fucks, W. (1952). On the mathematical analysis of style. Biometrika, 39(1–2), 122–129. https://doi.org/10.1093/
biomet/39.1-2.122
Gambino, R., Pulvirenti, G., Sylvester, T., Jacobs, A. M.,  & Lüdtke, J. (2020). The foregrounding assessment
matrix: An interface for qualitative-quantitative interdisciplinary research. Enthymema, 26, 254–277. http://doi.
org/10.13130/2037–2426/14387
Graesser, A. C., Person, N., & Johnston, G. S. (1996). Three obstacles in empirical research on aesthetic and literary
comprehension. In R. J. Kreuz & M. S. MacNealy (Eds.), Empirical approaches to literature and aesthetics. Ablex.
Grünbein, D. (1996). Katze und Mond. In D. Grünbein (Ed.), Galilei Vermisst Dantes Hölle und Bleibt an den Maßen Hän-
gen (pp. 116–128). Suhrkamp.

375
Arthur M. Jacobs

Hanauer, D. (1998). Reading poetry: An empirical investigation of formalist, stylistic, and conventionalist claims. Poetics
Today, 19(4), 565. https://doi.org/10.2307/1773260
Herrmann, J. B., Jacobs, A. M., & Piper, A. (2021). Computational stylistics. In D. Kuiken & A. M. Jacobs (Eds.), Hand-
book of empirical literary studies. De Gruyter.
Hoffstaedter, P. (1987). Poetic text processing and its empirical investigation. Poetics, 16(1), 75–91. https://doi.
org/10.1016/0304-422X(87)90037-4
Hoorn, J. F. (1996). Psychophysiology and literary processing: ERPs to semantic and phonological deviations in reading
small verses. In R. J. Kreuz & M. S. MacNealy (Eds.), Empirical approaches to literature and aesthetics (pp. 339–358).
Ablex.
Jacobs, A. M. (2011). Neurokognitive Poetik: Elemente eines Modells des literarischen Lesens [Neurocognitive poetics:
Elements of a model of literary reading]. In R. Schrott & A. M. Jacobs (Eds.), Gehirn und Gedicht: Wie wir unsere
Wirklichkeiten konstruieren [Brain and poetry: How we construct our realities] (pp. 492–520). Carl Hanser Verlag.
Jacobs, A. M. (2015a). Towards a neurocognitive poetics model of literary reading. In R. Willems (Ed.), Towards a cognitive
neuroscience of natural language use (pp. 135–159). Cambridge University Press.
Jacobs, A. M. (2015b). Neurocognitive poetics: Methods and models for investigating the neuronal and cognitive-
affective bases of literature reception. Frontiers in Human Neuroscience, 9, 186. https://doi.org/10.3389/fnhum.2015.
00186
Jacobs, A. M. (2015c). The scientific study of literary experience: Sampling the state of the art. Scientific Study of Literature,
5(2), 139–170. https://doi.org/10.1075/ssol.5.2.01jac
Jacobs, A. M. (2016). The scientific study of literary experience and neuro-behavioral responses to literature:
Reply to commentaries. Scientific Study of Literature, 6(1), 164–174. https://doi.org/10.1075/ssol.6.1.08jac ISSN
2210-4372.
Jacobs, A. M. (2018a). The Gutenberg English poetry corpus: Exemplary quantitative narrative analyses. Frontiers in
Digital Humanities, 5, 5. https://doi.org/10.3389/fdigh.2018.00005
Jacobs, A. M. (2018b). (Neuro-)cognitive poetics and computational stylistics. Scientific Study of Literature, 8(1), 165–208.
https://doi.org/10.1075/ssol.18002.jac
Jacobs, A. M. (2021). The neurocognitive poetics model of literary reading 10 years after. In A. Chatterjee & E. Cardillo
(Eds.), Neuroaesthetics in focus. Oxford University Press.
Jacobs, A. M., & Kinder, A. (2017). The brain is the prisoner of thought: A machine-learning assisted quantitative nar-
rative analysis of literary metaphors for use in neurocognitive poetics. Metaphor and Symbol, 32(3), 139–160. https://
doi.org/10.1080/10926488.2017.1338015
Jacobs, A. M., & Kinder, A. (2018). What makes a metaphor literary? Answers from two computational studies. Metaphor
and Symbol, 33(2), 85–100. https://doi.org/10.1080/10926488.2018.1434943
Jacobs, A. M., & Kinder, A. (2019). Computing the affective-aesthetic potential of literary texts. Artificial Intelligence, 1(1),
11–27. https://doi.org/10.3390/ai1010002
Jacobs, A. M., & Kinder, A. (2021). Computational models of readers’ apperceptive mass. Frontiers in Artificial Intelligence,
5, 718690. https://doi.org/10.3389/frai.2022.718690
Jacobs, A. M., & Lüdtke, J. (2017). Immersion into narrative and poetic worlds: A neurocognitive poetics perspective. In
F. Hakemulder, M. M. Kuijpers, E. S. Tan, K. Bálint & M. M. Doicaru (Eds.), Narrative absorption (pp. 69–96). John
Benjamins Publishing Company.
Jacobs, A. M., Lüdtke, J., Aryani, A., Meyer-Sickendieck, B.,  & Conrad, M. (2016). Mood-empathic and aesthetic
responses in poetry reception: A model-guided, multilevel, multimethod approach. Scientific Study of Literature, 6(1),
87–130. https://doi.org/10.1075/ssol.6.1.06jac
Jacobs, A. M., & Willems, R. M. (2019). The fictive brain: Neurocognitive correlates of engagement in literature. Review
of General Psychology, 22(2), 147–160. http://doi.org/10.1037/gpr0000106
Jakobson, R. (1960). Closing statement: Linguistics and poetics. In T. A. Sebeok (Ed.), Style in language (pp. 350–377).
MIT Press.
Jakobson, R., & Lévi-Strauss, C. (1962). Les Chats. Homme, 5–21. http://www.jstor.org/stable/25131017
Jockers, M. L., & Witten, D. M. (2010). A comparative study of machine learning methods for authorship attribution.
Literary and Linguistic Computing, 25(2), 215–223. https://doi.org/10.1093/llc/fqq001
Keidel, J. L., Davis, P. M., Gonzalez-Diaz, V., Martin, C. D., & Thierry, G. (2013). How Shakespeare tempests the brain:
Neuroimaging insights. Cortex, 49(4), 913–919. http://doi.org/10.1016/j.cortex.2012.03.011
Kim, E., Padó, S., & Klinger, R. (2017). Investigating the relationship between literary genres and emotional plot development
(pp.  17–26). Proceedings of the Joint SIGHUM Workshop on Computational Linguistics for Cultural Heritage,
Social Sciences, Humanities and Literature.
Kintsch, W. (1980). Learning from text, levels of comprehension, or: Why anyone would read a story anyway. Poetics,
9(1–3), 87–98. https://doi.org/10.1016/0304-422X(80)90013-3

376
Literature and poetry

Kuiken, D. (2015). The implicit erasure of “literary experience” in empirical studies of literature: Comment on “the
scientific study of literary experience: Sampling the state of the art” by Arthur Jacobs. Scientific Study of Literature, 5(2),
171–177. http://doi.org/10.1075/ssol.5.2.02kui
Liu, S., Erkkinen, M. G., Healey, M. L., Xu, Y., Swett, K. E., Chow, H. M., & Braun, A. R. (2015). Brain activity and
connectivity during poetry composition: Toward a multidimensional model of the creative process. Human Brain
Mapping, 36(9), 3351–3372. http://doi.org/10.1002/hbm.22849
Lüdtke, J., Meyer-Sickendieck, B., & Jacobs, A. M. (2014). Immersing in the stillness of an early morning: Testing the
mood empathy hypothesis of poetry reception. Psychology of Aesthetics, Creativity, and the Arts, 8(3), 363–377. https://
doi.org/10.1037/a0036826
Martindale, C. (1973). COUNT: A  PL/I program for content analysis of natural language. Behavioral Science, 18(2),
148–155. https://doi.org/10.1002/bs.3830180211
Martindale, C. (1984). Evolutionary trends in poetic style: The case of English metaphysical poetry. Computers and the
Humanities, 18(1), 3–21. https://doi.org/10.1007/BF02259803
Martindale, C., & Dailey, A. (1995). I.A. Richards revisited: Do people agree in their interpretations of literature? Poetics,
23, 299–314. https://doi.org/10.1016/0304422x(94)00025-2
Menninghaus, W., Bohrn, I. C., Altmann, U., Lubrich, O., & Jacobs, A. M. (2014). Sounds funny? Humor effects of
phonological and prosodic figures of speech. Psychology of Aesthetics, Creativity, and the Arts, 8(1), 71–76. https://doi.
org/10.1037/a0035309
Miall, D. S.,  & Kuiken, D. (1994). Foregrounding, defamiliarization, and affect: Response to literary stories. Poetics,
22(5), 389–407. https://doi.org/10.1016/0304-422X(94)00011-5
Moretti, F. (2000). Conjectures on world literature. New Left Review, 1, 54–68. https://newleftreview.org/issues/II1/
articles/franco-moretti-conjectures-on-world-literature
Nadal, M., & Chatterjee, A. (2019). Neuroaesthetics and art’s diversity and universality. Wiley Interdisciplinary Reviews:
Cognitive Science, 10(3), e1487. https://doi.org/10.1002/wcs.1487
Nicklas, P., & Jacobs, A. M. (2017). Rhetorics, neurocognitive poetics and the aesthetics of adaptation. Poetics Today,
38(2), 393–412. https://doi.org/10.1215/03335372-3869311
O’Sullivan, N., Davis, P., Billington, J., Gonzalez-Diaz, V., & Corcoran, R. (2015). “Shall I compare thee”: The neural
basis of literary awareness, and its benefits to cognition. Cortex; a Journal Devoted to the Study of the Nervous System and
Behavior, 73, 144–157. http://doi.org/10.1016/j.cortex.2015.08.014
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. Oxford University Press.
Pelowski, M., Markey, P. S., Lauring, J. O., & Leder, H. (2016). Visualizing the impact of art: An update and comparison
of current psychological models of art experience. Frontiers in Human Neuroscience, 10, 160. https://doi.org/10.3389/
fnhum.2016.00160
Pynte, J., Besson, M., Robichon, F. H., & Poli, J. (1996). The time course of metaphor comprehension: An event-related
potential study. Brain and Language, 55(3), 293–316. https://doi.org/10.1006/brln.1996.0107
Richards, A. (1929). Practical criticism: A study of literary judgment. Harcourt Brace Jovanovich.
Schrott, R., & Jacobs, A. M. (2011). Gehirn und Gedicht: Wie wir unsere Wirklichkeiten konstruieren [Brain and poetry: How
we construct our realities]. Hanser.
Simonto, D. K. (1989). Shakespeare’s sonnets: A case of and for single-case historiometry. Journal of Personality, 57(3),
695–721. https://doi.org/10.1111/j.1467-6494.1989.tb00568.x
Simonton, D. K. (1990). Lexical choices and aesthetic success: A computer content analysis of 154 Shakespeare sonnets.
Computers and the Humanities, 24, 251–264. https://doi.org/10.1007/BF00123412
Skov, M., & Nadal, M. (2020a). There are no aesthetic emotions: Comment on Menninghaus et al. (2019). Psychological
Review, 127(4), 640–649. https://doi.org/10.1037/rev0000187
Skov, M., & Nadal, M. (2020b). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Stockwell, P. (2005). Cognitive poetics: An introduction. Routledge.
Stone, P. J., Dunphy, D. C., Smith, M. S., & Ogilvie, D. M. (1966). The general inquirer: A computer approach to content
analysis. MIT Press.
Thierry, G., Martin, C. D., Gonzalez-Diaz, V., Rezaie, R., Roberts, N., & Davis, P. M. (2008). Event-related potential
characterisation of the Shakespearean functional shift in narrative sentence structure. Neuroimage, 40(2), 923–931.
https://doi.org/10.1016/j.neuroimage.2007.12.006
Van Peer, W. (1983). Poetic style and reader response: An exercise in empirical semics. Journal of Literary Semantics, 12(2),
3–18. https://doi.org/10.1515/jlse.1983.12.2.3
Van Peer, W. (1986). Stylistics and psychology: Investigations of foregrounding. Croom-Helm.
Van Peer, W., Hakemulder, J., & Zyngier, S. (2007). Lines on feeling: Foregrounding, aesthetics and meaning. Language
and Literature, 16(2), 197–213. https://doi.org/10.1177/0963947007075985

377
Arthur M. Jacobs

Van Peer, W., Sopčák, P., Castiglione, D., Fialho, O., Jacobs, A. M.,  & Hakemulder, F. (2021). Foregrounding. In
D. Kuiken & A. M. Jacobs (Eds.), Handbook of empirical literary studies. De Gruyter.
Vaughan-Evans, A., Trefor, R., Jones, L., Lynch, P., Jones, M. W., & Thierry, G. (2016). Implicit detection of poetic
harmony by the naïve brain. Frontiers in Psychology, 7, 1859. https://doi.org/10.3389/fpsyg.2016.01859
Wassiliwizky, E., Koelsch, S., Wagner, V., Jacobsen, T., & Menninghaus, W. (2017). The emotional power of poetry:
Neural circuitry, psychophysiology and compositional principles. Social Cognitive and Affective Neuroscience, 12(8),
1229–1240. http://doi.org/10.1093/scan/nsx069
Willems, R. M.,  & Jacobs, A. M. (2016). Caring about Dostoyevsky: The untapped potential of studying literature.
Trends in Cognitive Sciences, 20(4), 243–245. https://doi.org/10.1016/j.tics.2015.12.009
Xue, S., Jacobs, A. M., & Lüdtke, J. (2020). What is the difference? Rereading Shakespeare’s sonnets—An eye tracking
study. Frontiers in Psychology, 11, 421. https://doi.org/10.3389/fpsyg.2020.00421
Xue, S., Lüdtke, J., Sylvester, T., & Jacobs, A. M. (2019). Reading Shakespeare sonnets: Combining quantitative nar-
rative analysis and predictive modeling—An eye tracking study. Journal of Eye Movement Research, 12(5). https://doi.
org/10.16910/jemr.12.5.2
Yule, G. U. (1938). On sentence length as a statistical characteristic of style in prose with application to two cases of
disputed authorship. Biometrika, 30, 363–390.
Zeman, A., Milton, F., Smith, A., & Rylance, R. (2013). By hear: An fMRI study of brain activation by poetry and
prose. Journal of Consciousness Studies, 20, 132–158.
Ziegler, J. C., Montant, M., Briesemeister, B. B., Brink, T. T., Wicker, B., Ponz, A., Bonnard, M., Jacobs, A. M., &
Braun, M. (2018). Do words stink? Neural re-use as a principle for understanding emotions in reading. Journal of
Cognitive Neuroscience, 30(7), 1023–1032. https://doi.org/10.1162/jocn_a_01268

378
19
NARRATIVE
Franziska Hartung

Narratives are fundamental to human behavior. Humans spend a substantial part of their awake time being
engaged with narratives. Whether we read books, comics, or the newspaper; watch TV; talk to other people;
listen to podcasts; go to the movies; or play video games—much of our daily information input has narra-
tive structure. Even daydreaming or mind wandering often takes the shape of a narrative. Stories are used to
inform, entertain, teach, create shared perspectives, explore possibilities, and communicate cultural values.
As individuals and as communities, we construct narratives to make meaning of past events and create alter-
native realities and hypothetical or future worlds (Lee et al., 2020).
A story consists of causally and temporally related events that happen to a goal-based agent (protagonist)
and are confined within a context. Most narratives follow a basic plot structure that has already been high-
lighted by Aristotle: complication (i.e., the onset of conflict, an obstacle or problem) and unraveling (i.e., its
eventual resolution) (Butcher, 1907). In a meta-analysis of 40,000 popular narratives from around the world,
Boyd and colleagues (2020) found consistent underlying story structure that revealed three primary pro-
cesses: staging, plot progression, and cognitive tension. These substructures seem to hold across a wide range
of narrative types, including literary fiction, news stories, and TED talks. However, popularity of a narratives
is not predicted by adherence to normative story structure (Boyd et al., 2020).
Some psychologists even go as far to suggest that the very experience of our own life is narrative for
humans. For instance, Bruner (2004) argues that life consists of a stream of events involving characters,
actions, and objects in spatiotemporal contexts, with the main protagonist being oneself. Indeed, people
commonly talk about their own life and experiences in narrative form to themselves and others. A meaning-
ful and coherent “self-narrative” seems important for self-identity and coherence (see overview in Lee et al.,
2020).
Stories impact our beliefs and regulate our complex social behaviors from an early age on. Some argue
that storytelling is an evolutionary behavior for community regulation (see review in Jacobs  & Willems,
2018) and that the neurodynamics of engaging with or telling narratives are central to the understanding of
human neurobiology (Lee et al., 2020). Indeed, there is evidence that narrative structure is the optimal form
of information processing in humans. Information embedded in narratives is remembered longer and better
(see review in Lee et al., 2020), and the neuro-signature of processing information in narratives has a signal-
to-noise ratio superior to information presented without narrative organization (Kurby & Zacks, 2013).
Experimental neuroscience of language has long ignored narratives because they were thought to be too
complex and hence too difficult to control experimentally. However, narratives have long been used to study

DOI: 10.4324/9781003008675-21 379


Franziska Hartung

the underlying neural processes of other cognitive functions and in neuropsychological approaches with
patients. In the past 15 years, neuroimaging research with narratives significantly advanced our understanding
of (episodic) memory, event processing, attention, language, emotions, and social cognition (see Jääskeläinen
et al., 2021; Sonkusare et al., 2019 for review).
In this chapter, I will review what we currently know about the neuroarchitecture of narrative process-
ing and aesthetic responses to narratives. I will further discuss research at the interface of narrative and social
cognition—a topic of ongoing controversy. Two final sections are dedicated to the discussion of the impor-
tance and informativeness of differences between individuals, stories, genres, and situational motivation or
reading goals. The chapter will conclude with highlighting current challenges, future directions, and possible
applications.

The neuroarchitecture of narrative processing


Research with narratives is one of the few flourishing disciplines where language is studied in naturalistic
contexts, which stands in contrast to the often isolated and self-referential psycholinguistic approaches to
explain the neurobiology of language (see review in Hasson et al., 2018). In addition to auditory, word, and
sentence processing, understanding a narrative requires integrating cumulative information into a situation
model that represents the events in the narrative and their connection to each other (see review in Jacobs &
Willems, 2018). Constituents of such a situation model typically are agents and their motivations, shared
context, a defined time frame, and causal relations of events to each other. Throughout a narrative, a situation
model is constantly updated and reshaped and often retrospectively reinterpreted (e.g., plot twists). Research
on narratives is often better understood with cognitive models explaining event segmentation, mentalizing,
causal inference, and episodic thinking (Hasson et al., 2018) rather than models of language processing.
The neural network that is activated when people engage with narratives is different from what psy-
cholinguists typically call language processing areas. In a seminal study, Ferstl and colleagues (2008) showed
that the neuroarchitecture supporting narrative processing is drastically expanded from the areas associated
with phonological, lexical, and syntactic processing in the left perisylvian cortex, reminiscent of the first
neuropsychological work in aphasic patients that are still carrying the names of Broca, Wernicke, and Lich-
theim. Instead, connected discourse induces bilateral activation of a wide range of brain regions, including
the inferior frontal gyri, medial prefrontal cortex, posterior cingulate cortex, angular gyri, precunei, anterior
temporal lobes, and temporo-parietal junctions, compared to non-discourse baselines (AbdulSabur et  al.,
2014; Chow et al., 2015; Ferstl et al., 2008; Mar, 2011; Xu et al., 2005).
While this so-called “extended language network” was a surprising finding for language researchers, it
was not for researchers from other neuroscientific fields who used narratives to study cognitive functions
such as social cognition or memory. This finding highlights that our understanding of the neurobiological
underpinnings of phonological, word, and sentence processing does not scale up to contextualized language
and underscores the importance of research on discourse (e.g., with narratives) to understand the function of
our language faculty. More recent models of the neuroarchitecture supporting language have acknowledged
a wider distribution of labor beyond the perisylvian cortex, especially regarding the interface of language and
social cognition and the involvement of the right hemisphere in discourse processing (e.g., Hagoort, 2017;
see also Hagoort & Indefrey, 2014).
There is evidence for functional subdivisions in the extended language network. Aboud and colleagues
(2019) showed that the subnetwork that is engaged in narrative comprehension is qualitatively different from
the subnetwork that is engaged in the processing of non-narrative texts. In a large meta-analysis, Yang and
colleagues showed that the narrative subnetwork (2019) most heavily relies on the default mode network
(DMN), particularly on its dorsal and medial subsystem, with the left inferior frontal gyrus and left middle
temporal gyrus serving as core regions across all sub-processes of narrative processing (see also Tylén et al.,

380
Narrative

2015). The specialization and manifestation of the brain’s narrative network is documented in children as
young as 5 and seems not yet fully completed at age 18 (Szaflarski et al., 2012).
Experimental research shows that the information that accumulates based on narrative progression is
dependent on DMN structures that are involved in encoding, recall, and reconstruction of narrative events
(see review in Jääskeläinen et al., 2021). These structures have been shown to be hierarchically linked to tem-
poral receptive windows (Hasson et al., 2015; see Figure 19.1; Lerner et al., 2011). Plot formation modulates

Figure 19.1 Hierarchical processing memory for narratives by Hasson et al. (2015, Trends in Cognitive Sciences). Brain
areas process language input in hierarchical processing windows linked to hierarchically organized informa-
tion processing units in the brain that range from perceptual processes (recognizing speech or writing) to
words, sentences, and meaningful discourse. Processing windows with different time sensitivities are spe-
cialized in processing phonemes, words, sentences, paragraphs, or entire discourses, integrate information
with increasing semanticity and integration of meaning, and are modulated by the attention and control
network as well as episodic and relational memory processes. These processes are organized in the brain
along pathways ranging from perceptual areas to higher-order meaning integration and representation (e.g.,
primary visual cortex to episodic memory and default mode network for written language). Reproduced
from Hasson et al. (2015) with permission from Elsevier.

381
Franziska Hartung

DNM activity (Tylén et al., 2015): especially in right temporo-parietal regions, plot development has been
shown to be linearly modulated by cumulative integration of narrative information as a story progresses. This
linear modulation is dependent on semantic coherence of events rather than mere quantity of information
that has to be integrated into the situation model of the narrative (Tylén et al., 2015; see also abstraction in
Figure 19.1).
Synchronicity of neurodynamics related to narrative engagement and event segmentation have proven to be
highly robust across participants and seem to be driven by high-level narrative structure and coherence rather
than low-level stimulus or language properties (Simony et al., 2016; Speer et al., 2009; Yarkoni et al., 2008).
Boundaries of events and narrative structure result in different phase synchronization across the brain network
involved in processing narratives that can be used to detect event boundaries in stories (Baldassano et al., 2017;
see also Zacks et al., 2007). Coupling strength between brain areas within an individual’s DMN has been fur-
ther shown to predict memory performance (Simony et al., 2016). Shared interpretation of narrative meanings
also predicts similarity of brain activity and network coupling in the DMN (Nguyen et al., 2019).
Studies on storytelling in interactive settings show that speaker-listener neural synchrony predicts com-
prehension: there are short and consistent delays (several seconds) in this coupling wherein the speaker brain
signal precedes that of the listeners or vice versa (interpreted as predictive listener behavior) (Kuhlen et al.,
2012; Zadbood et al., 2017). These findings show that, quite literally, we transfer brain states from one brain
to another when we engage with other people’s stories. From an evolutionary viewpoint, this is the strongest
argument for an adaptive and regulatory function of the language faculty for human social behavior and the
rise of culture.

Aesthetic responses to narratives


People can experience a wide range of feelings when engaging with narratives, such as suspense, joy, iden-
tification with characters, awe, and beauty, with beauty and suspense seemingly being the most important
aspects of aesthetic responses to narratives (Knoop et al., 2016). For the most popular narrative genres—
novels and short stories—suspenseful, interesting, and (to a lesser degree) romantic seem to be central concepts
people use to describe aesthetic experiences (Knoop et al., 2016).
In Knoop and colleagues’ (2016) cross-genre study, the term beautiful stood out for describing aesthetic
emotions to literature as being listed most frequently across genres, including genres that are not primarily
narrative. The authors argue that the attribution of beauty to literary texts is independent of plot features but
rather seems to relate to the writing style and the expression of emotions, moods, and atmospheres (Knoop
et al., 2016). Experiencing beauty links aesthetic responses to narratives with aesthetic responses in other
domains such as visual art, human faces, music, or landscapes, where experiences of beauty are also promi-
nently featured among aesthetic emotions.
The second concept that was central to describe aesthetic responses to narrative fiction in Knoop et al.’s
(2016) study was suspenseful. Suspense in narrative experiences is linked to higher levels of transportation into
a narrative world, which in turn is linked to enjoyment (see also next section). Storytellers often use suspense
to capture the attention of an audience over longer stretches of time, as it contributes to anticipating certain
plot developments and the need for (sub)plot resolution that will fluctuate in likelihood and desirability
throughout the trajectory of a narrative (see review in Knoop et al., 2016). Research on visual narratives has
shown that suspense also affects the susceptibility of the audience and can regulate audience attention, which
is reflected in activation dynamics in brain networks linked to attention as well as the DMN (see review in
Bezdek et al., 2015).
The naturally fluctuating emotional intensity profile in narratives has been related to dynamics in heart
rate variation, as well as in activity in brain areas linked to attention, emotion, event, and language processing
(Wallentin et al., 2011a). In Wallentin and colleagues’ (2011a) study, they found increased brain activations

382
Narrative

linked to emotionally intense parts of a narrative have been interpreted as reflective of saliency or attentional
modulation. In a recent study that also modeled brain activations linked to emotional intensity in narratives,
emotionally intense segments of a novel literary story, however, were negatively correlated with activation
of the attention network, seemingly reflecting a decrease in top-down attention during emotionally intense
parts of a story (Hartung et al., 2021). While these findings seem to contradict the earlier study by Wallentin
and colleagues, the difference in the direction of activation in the attention network can be attributed to
novelty: Wallentin and colleagues’ study (2011a) used an extremely well-known folk fairytale with which all
participants were familiar since childhood, whereas Hartung and colleagues’ study used novel literary nar-
ratives. Hence, the differences in these two studies can be attributed to anticipation of story events in the
well-known story and letting the story guide information seeking in the study with the novel stories where
listeners have less accurate predictions for what will happen next and also fewer emotional attachments com-
pared to a story that is highly culturalized such as folklore and fairytales (see Hartung et al., 2021). Behavioral
evidence shows that rereading narratives can lead to increased appreciation and enjoyment. This effect is
independent of narrative or linguistic complexity and seems to be linked to better perceived comprehension
(Kuijpers & Hakemulder, 2018).
While aesthetics is often associated with positive emotions, the content of many narratives deals with
negatively valanced aspects of life such as conflict, despair, suffering, and pain. Still, unpleasant narratives can
be liked and enjoyed and in fact rank high among the most popular pieces of literature (see review in Maslej
et al., 2020). In one functional MRI study, Altmann and colleagues (2012) showed that activity in brain
areas related to social cognition and empathy linearly increased with negative story valence, with the medial
prefrontal cortex being particularly engaged if stories were liked despite being unpleasant. Being the only
neuroimaging study to my knowledge on valence and aesthetics, it is hard to say what this activation pattern
means. More research is needed to understand the neurodynamics of emotional and aesthetic responses to
different types of literary fiction.

Enjoyment and immersion into narratives


Most people experience engaging with narratives as being extremely rewarding. We often get completely
absorbed in the story and forget the world around us, temporarily escaping our real lives by diving into a
fictional world. Narrative aesthetic absorption describes a state in which we focus on the story world of
a narrative while becoming less aware of our surroundings and ourselves. It is characterized by focused
attention, vivid mental imagery, and emotional engagement. This process is often referred to as narrative
absorption, transportation, or immersion (Gerrig, 1993, 1999; Green & Brock, 2000; Kuijpers et al., 2014;
Sestir & Green, 2010). While these terms all highlight different aspects of narrative experiences, they are
closely related concepts and are often used interchangeably. Narrative immersion is linked to flow—a state
of absorption marked by “deep concentration, losing awareness of one’s self, one’s surroundings and track
of time” (Csikszentmihalyi, 1990) that people can experience in a multitude of activities, such as running,
writing, playing music, and dancing.
Immersion into narratives has been shown to correlate with increased activation in the dorsal anterior
cingulate cortex (Hsu et al., 2014). For visual narratives, immersion has also been shown to increase activity
in the left inferior frontal gyrus, the right lateral temporal cortex, and bilaterally in the posterior superior
temporal sulcus (see review in Jääskeläinen et al., 2021). Electrophysiology measures show that immersion
into visual narratives is also linked to high synchronicity of EEG signals between participants as measured by
intersubject correlations, which has been proposed to reflect heightened attention and emotional immersion
into the narrative (see review in Jääskeläinen et al., 2021). High synchronicity in brain activity between par-
ticipants has even been shown to predict movie box office success with 20 times higher accuracy than more
traditional box office estimates (Christoforou et al., 2017).

383
Franziska Hartung

Immersion is a multidimensional experience based on different components that can contribute dif-
ferently depending on the situation (Kuijpers et al., 2014). For instance, one component of immersion is
feeling transported into a narrative world. Narrative transportation is “a feeling of entering a story world,
without completely losing contact with the actual world,” thus the feeling of being part of a fictional world
during reading (Kuijpers et al., 2014). Narrative transportation correlates with story enjoyment and can be
facilitated by high levels of suspense.
Transportation into a fictional world is also linked to increased affective responses to fictional characters,
which is another important component of being immersed in a narrative (Sestir & Green, 2010). Emotional
engagement with fictional characters includes feelings of sympathy, empathy, and identification (Kuijpers
et al., 2014). Whether we get captured by a story is often affected by whether we care about the characters
in the story. Tal-Or and Cohen (2010) showed that identification with protagonists and transportation into a
narrative are separate processes, even though they often co-occur and both correlate with enjoyment. While
transportation is linked to suspense, identification is linked to having an emotional connection with the main
character(s).
Another important component of immersion is attention. A high level of attention towards the story is
often marked by a subjective experience of losing self-awareness, awareness of the surroundings, and track of
time (Kuijpers et al, 2014). This component more than other components of immersion is affected by the
context (e.g., location, mood, environmental noise, fatigue) in which people engage with narratives.
Immersion into narratives is often linked to experiencing mental imagery of surroundings, characters,
and situations (Green  & Brock, 2002; Kuijpers et  al., 2014). This component of narrative immersion is
associated with mental simulation ( Jacobs, 2015a, 2015b; Schrott & Jacobs, 2011; Zwaan, 2004), a cogni-
tive mechanism that supports semantic processing by activating experience-based knowledge from different
cognitive modalities. For instance, motor and sensory cortical activation is associated with a covert mental
“reenactment” of actions in a story. During engagement with narratives, mental simulation facilitates build-
ing multimodal situation models of the semantic contents of a narrative as the basis of narrative comprehen-
sion (Gernsbacher, 1997; Zwaan & van Oostendorp, 1993).
Factors inhibiting immersion are related to the processing of stylistic devices that are often referred to as
foregrounding (Hakemulder, 2004; Jacobs & Willems, 2018; van Peer et al., 2007). Foregrounding of stylistic
devices, such as unusual or poetic expressions, redirects the attention of the audience from the content to
the form of the writing and stands in contrast to backgrounding, which is linked to higher attention to story
contents and higher immersion ( Jacobs & Willems, 2018, see also Hartung et al., 2021). Foregrounding has
been linked to increased activity in brain areas linked to language processing, particularly semantic integra-
tion (Hartung et al., 2021), which can be interpreted as higher cognitive load on language comprehension
systems during foregrounded segments compared to backgrounded segments of a story. While foregrounding
and immersion are often understood as mutually exclusive, there is some evidence showing that foreground-
ing can actually facilitate absorption into a story (Bálint et al., 2016).
There is a consistent link between enjoyment and immersion (Busselle & Bilandzic, 2009; Green, 2004;
Hartung, Withers et al., 2017; Sestir & Green, 2010; Tal-Or & Cohen, 2010). Since immersion is based on
different subcomponents, individual and situational differences in the ability to engage in these subcompo-
nents will contribute differentially to experiences of immersion into narratives. For example, there is some
evidence that people with higher trait empathy can immerse more easily (e.g., Hartung et al., 2016), which
might be based on their ability to empathize or emotionally engage more easily with fictional characters,
which in turn contributes to immersion. Similarly, people who can control their attention better might have
an easier time blending out distracting environmental information and hence might immerse more easily
than people who get easily distracted. The same is true for mental simulation: people who are particularly
good at creating mental worlds from narrative input likely have richer and more in-depth experiences with
narrative worlds.

384
Narrative

Enjoyment of reading also might be linked to training effects, meaning that engaging with narratives
more frequently might improve the quality of narrative experiences. This aspect will be discussed in detail
in the following section on individual differences. The research on immersion and enjoyment of narratives
is young, and particularly the underlying neurodynamics are poorly understood. Cognitive models such as
mental simulation help explain and hypothesize about underlying dynamics, but the actual empirical research
so far is scarce. Understanding the neurodynamics of aesthetic engagement with narratives is crucial to relate
aesthetic experiences across different domains and narrative genres.

The social aspect of narratives


Narrative fiction is social by nature (Bruner, 1986). Fiction communicates and reinforces moral principles
and cultural values and is thought to train social and emotional competences both explicitly and implicitly
(see review in Wimmer et al., 2021). There is evidence that we treat fictional characters like real people and
that their role model function is important for moral development during childhood and adolescence (see
review in Maslej et al., 2020). Contrary to stereotypes of bookworms being antisocial, avid readers have been
shown to have higher social life satisfaction than infrequent readers and outperform them on measures assess-
ing social cognition abilities (Mar et al., 2006). Immersion into narratives also has been shown to improve
well-being in older adults (Poerio & Totterdell, 2020) and has been linked to longevity (Bavishi et al., 2016).
The idea that engaging with fiction is linked to better social skills is discussed at least since antiquity,
when it was proposed by Aristotle (Aristotle, 2013; Nussbaum, 1997; see recent review in Carlarco et al.,
2017). The rationale behind this hypothesis is that narrative engagement (1) enhances our abilities of moral
reflection and hence self-reflection and (2) trains empathy and taking perspectives of other people by means
of stepping into the shoes of fictional characters (Koopman & Hakemulder, 2015).
Engaging with fictional characters and their goals, beliefs, feelings, and actions is assumed to simulate
interacting with real-world agents, which results in both the positive effects of social interaction as well as
the training effect of social cognition skills by practicing understanding the mental world of others (Bruner,
1986; Gerrig, 1999; Jacobs & Willems, 2018; Mar & Oatley, 2008; Mar, 2018; Oatley, 2016, 1999a, 1999b;
Willems & Jacobs, 2016). Similar to how flight simulators train the skills of pilots, engaging with narratives
enables the individual to simulate interacting with social agents and thereby train the capacity to understand
the intentions, feelings, desires, and beliefs of other people. This has been shown to improve empathy and
relatability to outgroups: engaging with fiction and hence fictional characters has been linked to improved
attitudes towards stigmatized groups ( Johnson, 2013; Vezzali et  al., 2014; Kaufman  & Libby, 2012) and
enhanced understanding of one’s own and others’ ethnic identity (Vasquez, 2005). Engaging with narratives
can also increase prosociality ( Johnson, 2012; Koopman, 2015).
While the idea that engaging with fiction is linked to social abilities is a popular theory, it is difficult to
prove experimentally. In a seminal study, Kidd and Castano (2013) reported that after reading excerpts of
literature, people outperformed control groups that read scientific excerpts or non-literary stories on social
cognition and social ability measures (see also Bal & Veltkamp, 2013; Johnson, 2012; Kuijk et al., 2018).
Similar findings have been reported for watching visual narratives (Black & Barnes, 2015). However, mul-
tiple efforts to reproduce the findings of Kidd and Castano’s study (2013) failed to produce similar results
(e.g., Samur et al., 2018; Panero et al., 2016; Kidd & Castano, 2019). A single event of engaging with a
literary narrative does not seem to be sufficient to measurably increase social abilities. Consistently reported
instead is that people with higher lifetime exposure to literary fiction tend to perform better on social cog-
nition measures than people who read less (Samur et al., 2018; Mar et al., 2006; Mumper & Gerrig, 2017;
Tamir et al., 2016). One neuroimaging study further showed that parts of the social cognition brain network
were more strongly activated in frequent readers during comprehension of social content in brief excerpts
of fiction (Tamir et al., 2016). Yet the relation between engaging with fiction and social abilities remains

385
Franziska Hartung

correlational, and the question of whether people who are good at social cognition tend to like reading more
or whether lifetime exposure to fiction improves social abilities remains unanswered.

Narrative perspective and perspective taking


One aspect that is often talked about in reference to the social aspects of narratives is perspective taking.
Perspective taking is considered important in the construction and comprehension of fiction (Bal, 2009;
Dancygier, 2014; Genette, 1980; Rimmon-Kennan, 2002; Sanford  & Emmott, 2012), as well as in the
generation of situation models (Bower & Morrow, 1990; Johnson-Laird, 1983). Consequently, it also plays a
crucial role in immersing into narratives and mentalizing with fictional characters (Mano et al., 2009).
Readers often take the perspective of the protagonist and simulate their mental states as the source of
construal when constructing a situation model (Albrecht et al., 1995; Horton & Rapp, 2003). Taking the
viewpoint of characters is linked to identification: it is believed that the reader is more engaged when taking
a character’s viewpoint and adopting the character’s goals and intentions, which can result in experiencing
emotions of empathy (Oatley, 1999a). Indeed, adopting a protagonist’s perspective can cause changes in the
mental and emotional states of the reader (Gerrig, 1993; Green & Brock, 2000; Komeda et al., 2013), par-
ticularly when the reader experiences high levels of story immersion (Sestir & Green, 2010).
Alternatively to adopting a character’s viewpoint, readers can also become immersed in a story by tak-
ing the perspective of an eyewitness (Boyd, 2005; Oatley, 1999b, 1999a). While the narrative perspective
of a story (focalization and voice) does not seem to affect perspective taking in readers (Hartung, Hagoort
et al., 2017; Hartung, Withers et al., 2017; Wimmer et al., 2021), individuals seem to have preferences
for a default perspective with which they approach fiction. Taking the perspective of an eyewitness or a
character in a story are not mutually exclusive, as a significant number of people tend to do both (Hartung,
Hagoort et al., 2017; Hartung, Withers et al., 2017). Neuroimaging evidence shows that eyewitness and
character perspective taking are associated with the activation of different brain areas (Hartung, Hagoort
et al., 2017).
There is some evidence that first-person narration can facilitate transportation into story worlds and
enjoyment of reading experiences compared to third-person narration (Hartung et  al., 2016; Hartung,
Withers et al., 2017; Samur et al., 2021). However, this effect seems to be driven by people who prefer
character perspective taking: presumably, the match between their preference for character perspective taking
and the first-person narration makes the experience more rewarding, whereas narrative perspective seems to
have little to no effect on people who prefer eyewitness perspective or have no preference (Hartung, Withers
et al., 2017).
While perspective taking is a crucial part of engaging with fictional characters and immersing into sto-
ries, the factors that drive perspective taking and individual preferences are poorly understood. The social
impact of engaging with narratives is significant, yet few research endeavors are looking into the interaction
of cognitive and affective variables that affect story immersion and how stories influence our experiences,
knowledge, and beliefs. Neuroimaging can help us uncover the underlying mechanisms linked to the social
aspects of narrative engagement by offering a window into labor division and interaction of different brain
networks during story immersion. In conjunction with careful behavioral designs testing narrative experi-
ences and changes in emotions and beliefs, neuroimaging can help us understand how and why stories enrich
our social lives.

Individual differences
Perspective taking is of course not the only aspect of narrative engagement where people show drastic
individual variation and preferences in how they engage with narratives. Neuroimaging findings show that

386
Narrative

people differ in their sensitivity to different types of semantic content: Some people show higher sensitivity
in brain areas linked to social cognition during emotional content in narratives, while others seem to be more
sensitive to what is happening, which is reflected in brain activity in areas linked to action observation and
execution (Nijhof & Willems, 2015; see also Tamir et al., 2016). Frequent readers read stylistically marked
language differently than people who read less frequently (Hoven et al., 2016). Individual differences have
also been reported for mental imagery and reading motivations (Mak et al., 2020). All of these studies have
in common that style and content of a narrative do not seem to matter as much as individual biases and
preferences of the people engaging with them (Eekhof et al., 2021; Hartung, Hagoort et al., 2017; Hartung,
Withers et al., 2017). Each person brings their own preferences, skills, and knowledge, which in interaction
with the narrative itself create a unique narrative experience. Assessing individual differences between people
to test the effects of personality and lifestyle on reading experiences has a long tradition, but it is often dif-
ficult to make an informed decision on what measures are best suited.
A typical measure that is used to assess how well-read people are is the Author Recognition Test (ART),
as well as self-report measures of reading frequency and genre preferences (Kuijpers et al., 2020). People
who like reading fiction tend to read faster and are more likely to engage in protagonist perspective taking
(Hartung, Hagoort et al., 2017). How much people read also correlates with their openness of experience
as well as how easily they get transported into story worlds (Kuijpers et al., 2020). How frequently people
engage with narratives is also correlated with emotion processing and need for cognition (Cacioppo et al.,
1996)—a measure assessing how much reward people experience for solving cognitive problems. It seems
to be the case that open-mindedness, intellectual curiosity, and emotional responsiveness are directly linked
to trait absorption, which in turn affects how enjoyable people find engaging with narratives, particularly
fiction, which in turn affects habitual engagement with fiction.

Fiction and nonfiction stories


So far, I focused on research linked to narratives in fiction and literature, but of course the form “narrative”
has a much broader application. For instance, narration has also become increasingly important for journal-
ism (see review in van Krieken & Sanders, 2019). Narrative features in journalistic articles have been shown
to engage readers more deeply. Readers identify more with eyewitnesses and feel more present when read-
ing a narrative eyewitness report than when reading a non-narrative report of the same event (van Krieken,
Hoeken et al., 2015). While narrative journalism has a positive effect on audience engagement, it has been
heavily criticized for “the tensions it creates between ethics and aesthetics, fact and fiction, and objectivity
and subjectivity” (van Krieken & Sanders, 2019).
The available empirical evidence for differences in behavior and engagement between fiction and non-
fiction is mixed. Several studies have shown that the format of presentation (e.g., newspaper vs. book layout)
and expectations towards a format can affect reading behavior. For instance, Zwaan (1994) showed that
believing that a text was taken from a newspaper (factual) or from a novel (fictional) influences behavior and
memory. In two studies, he reported that texts believed to be factual were read faster compared to fictional
texts (see also Altmann et al., 2014; Wolfe, 2005). Moreover, readers showed a better performance on situ-
ational memory for factual texts but better memory for the text’s surface structure for fictional texts (Zwaan,
1994).
However, participants’ belief a story is factual or fictional does not necessarily affect emotional and expe-
riential aspects of engaging with narratives. In a large-scale study, Hartung, Withers and colleagues (2017)
found no evidence that believing a story is based on factual or fictional events plays a role in a multitude
of measures assessing aesthetic, emotional, imaginative, and experiential aspects of reading. Moreover, they
found no evidence that the fact or fiction manipulation affected reading behavior or situational memory, not
replicating previous findings such as Zwaan’s (1994) study.

387
Franziska Hartung

There is some evidence that subjective intensity of negative valence might be weaker for fiction than for
factual narratives, arousal being equal (Sperduti et al., 2016). Other studies suggest that there is no difference
in emotional response to fact and fiction (Goldstein, 2009) and instead that observed differences in emotional
response to factual or fictional information are mediated by individual variation in how much participants
scrutinize the information they are presented with (Green et al., 2006; Wolfe & Mienko, 2007). This again
highlights the importance of individual differences we discussed previously.
Processing of fictional and factual texts also seems to be supported by different neural networks (Abra-
ham et al., 2008; Altmann et al., 2014; Han et al., 2005; Mar et al., 2007; Metz-Lutz et al., 2010). For
instance, Altmann and colleagues (2014) found evidence for different neural network activation depending
on whether the text was believed to be factual or fictional. The activation pattern while reading factual
texts was associated with motor cortex activation suggesting “an action-based . . . reconstruction of what
happened” (Altmann et al., 2014, p. 26). Reading fictional stories, on the other hand, was associated with
networks linked to social cognition and imagining possible events suggesting “a constructive simulation of
what might have happened” (ibid, p. 27).
These findings from Altmann and colleagues’ (2014) functional MRI study support the idea of genre-
primed reading strategies: Some accounts argue that the expectation of reading fiction triggers a genre-
specific reading strategy which allows us to get immersed and experience strong emotions while engaging
with fiction (Mar & Oatley, 2008; Oatley, 1999b). Other accounts argue, however, that it is not the knowl-
edge about the factuality of a narrative but rather an engaging narrative style that causes readers to get more
immersed and to experience stronger emotions, regardless of whether the information is believed to be true
(van Krieken, Hoeken et al., 2015; van Krieken, Sanders et al., 2015).
Empirical evidence supports both accounts. Readers seem to use prior knowledge about the genre to sys-
tematically select criteria and strategies for comprehension linked to different reading goals (van den Broek
et al., 2001, 2005; Zwaan, 1994). Reading goals for factual texts are to obtain information about reality (e.g.,
reading for study purposes or reading the news) and are thought to prompt reading strategies which empha-
size connections in the text in order to reconstruct the contents (van den Broek et al., 2001). Reading for
enjoyment, on the other hand, is associated with a stronger motivation for subjective experience and is linked
to reduced scrutiny and attention to detail (Mar & Oatley, 2008; van den Broek et al., 2001; Zwaan et al.,
1995). At the same time, narrative features seem to facilitate attentional and emotional engagement regardless
of whether a text is factual or fictional. While most agree that there are different strategies associated with
different types of texts—whether narrative or not, it is unclear based on which criteria these strategies are
selected. Reading narratives activates different reading goals than non-narrative texts, but whether narratives
are factual or fictional does not seem to affect reading strategies as much. Systematically investigating reading
goals and the associated cognitive and neural mechanisms is a promising avenue for future research with a lot
of potential for clinical and social applications.
It is undeniable that the line between fiction and non-fiction is fuzzy. Many fiction narratives are “based
on true events,” whereas there is also plenty of fiction involving real historical persons, places, and events. At
the same time, media coverage of recent events has incentive to sell stories. In the age of fake news, question-
able reporting practices in the media, and lack of responsibility for spreading false information, this is a par-
ticularly important area of research that does not yet get the attention from empirical sciences that it deserves.

Major challenges, goals, and suggestions


Until recently, most experimental research on language barely looked beyond the sentence level, let alone
narratives. The two main reasons for this were (1) the common belief that effects from the laboratory easily
scale up to natural situations and (2) the fear of compromising experimental control when choosing a more
natural experimental setting. Especially in neuroimaging research, continuous stimulus presentation times

388
Narrative

have been avoided due to technical limitations (see review in Jacobs & Willems, 2018). However, in recent
years, methodological advancements in data analysis have allowed for a growing body of new research with
narratives (e.g., Chow et al., 2013; Lerner et al., 2011; Yarkoni et al., 2008, Hartung et al. 2021), movies
(Hasson et al., 2004; Kauttonen et al., 2015), theatre plays (Metz-Lutz et al., 2010), and even dance perfor-
mances (Bachrach et al., 2016).
Experimental research with narratives rightfully gained popularity in experimental psychology and cogni-
tive neuroscience over the last two decades. There is a growing body of research on narrative comprehension,
especially on sensorimotor simulation and mentalizing (e.g., Hartung, Withers et al., 2017; Kurby & Zacks,
2013; Nijhof & Willems, 2015; Speer et al., 2009; Wallentin et al., 2011b), proving how methodological
challenges can be overcome. More importantly, maybe, they show the value of experimental research with
narratives: Stories have a substantial advantage for experimental research by reducing inter-individual vari-
ability in semantics. The embedding of words in narratives reduces ambiguity by keeping components like
time, agent, and causality constant for multiple events in a design. As a result, such experimental designs have
better robustness and reliability of responses across subjects, increase motivation and relevance, and decrease
fatigue for participants in empirical studies. Moreover, these studies have superior external validity compared
to less naturalistic experimental materials. Methodological developments, especially in neuroimaging data
acquisition and analysis, but also linguistic corpus tools, have solved many previously existing issues. Yet
there are still challenges. One of them is the need for tools that can help us better estimate the linguistic and
semantic value of text elements in a given context, as well as understanding semantic connections within a
narrative.
The stories we read can provide us with quite different reading experiences. There are stories which make
us think or experience strong emotions or stories which create colorful pictures in our minds and take us on
wild adventures. Sometimes we are mere spectators of the scenes unfolding on the pages; sometimes we are
companions of our heroes. And sometimes we even merge with fiction and experience the events through the
eyes of a character. These experiences are thought to be a result of simulating what is written on the pages.
Our brains create mental models of what is going on in the stories and by doing so provide us with the illusion
of experience. Some even argue that engaging with narratives can prepare us for situations in the real world
(van Krieken, 2018; Mar & Oatley, 2008; Oatley, 1999b). Narratives offer us a playground for adventures
and scenarios that we would like to avoid in the real world (e.g., horror) and teach us about possible futures
and past events (war, famines). Engaging with stories helps us simulate real-world encounters and therefore
experience situations “vicariously” (Zwaan, 2004). This explains why we have developed such a rich tradition
of storytelling: We transmit knowledge and experience. Immersion into story worlds could therefore serve as
a type of learning to gain experience and knowledge, which in turn can lead to benefits in real life (Bruner,
1990; Mar & Oatley, 2008). By simulating information during reading or listening about the experiences of
others, we can learn new things about the world (Mar & Oatley, 2008; Taylor & Zwaan, 2009).
Following this argumentation, engaging with narratives is an effective method of learning, particularly
about cultural, social, and emotional aspects of life. This idea goes back to Aristotle (Aristotle, 2013; Bruner,
1986; Gerrig, 1993; Oatley, 1999b) and has been the subject of much debate. The rationale behind the con-
nection of social cognition and narratives is that by interacting and mentalizing with fictional characters, we
train our abilities to socialize with real people. While difficult to prove empirically, the link between social
abilities and engaging with narratives is undeniable. However, the question of whether this is because people
with good social abilities tend to like engaging with fiction more or whether lifetime exposure to fiction
improves social abilities is still unanswered and needs to be addressed in future research. Future research
should aim at understanding the cognitive mechanisms and neurodynamics of simulation and its contribu-
tions to knowledge development and social cognition, and well-being.
Understanding the causal contributions of narrative engagement and social abilities is particularly impor-
tant because of its possible applications for improving social abilities in typical and clinical populations. For

389
Franziska Hartung

instance, bibliotherapy seems to be a promising treatment for emotional disorders (den Boer et al., 2004).
Earlier in this chapter, I also reviewed the evidence regarding the benefits of narrative engagement for inter-
action between different social groups, improving cultural competence, and empathy towards and better
understanding of marginalized groups. We should systematically explore the clinical and social applications
of narratives to address current societal challenges. A future direction that is particularly promising is to inves-
tigate the interface of reading goals and strategies with how they interact with different narratives, learning,
persuasion, and enjoyment. How do we become the hero of a story? Does the way we immerse ourselves in a
story affect our enjoyment and susceptibility during reading? Are people who identify more easily with char-
acters more likely to be avid readers? What is it about a story that pulls us in and keeps us glued to the pages?
Another important goal for future research on narratives must be to increase diversity. A lot of the experi-
ments done in the past are based on “canonical” literature. While this is a practical solution to avoid having to
define and operationalize what “literature” or “high quality” means, canons are inherently biased in terms of
authors’ social identity, themes, topics, and esteem and ultimately the characters, topics, and narrative worlds
the stories are about. Being mindful of diversity also means to test representative samples of participants with
a range of group identities and variation in age, background, and education. Diversifying and decolonizing
our research is critical for advancing our understanding of individual differences and the complexity of inter-
action of individuals and narratives. Additionally, a wider selection of measures that assess narrative experi-
ences, as well as individual differences both between people and narratives, is needed.
Individual differences have not been taken seriously enough in previous research in empirical research
with narratives. To understand the unique interaction of an individual’s experiences with a narrative, as well
as shared cultural experiences with narratives, we need to rethink the ways we design studies. Future research
should explore the interface of individual unique experiences and shared cultural experiences with narratives
that contribute to building communities.
As many positive aspects that are associated with narrative engagement, there are also obvious dangers.
Narratives are a powerful tool in communicating ideas and manipulating others. For instance, in the adver-
tising industry, narratives are used to activate shared moral values of customers and brands and to support
brand-consumer connections, ultimately generating brand value and manipulating potential customers to
identify with brands (Sanders & van Krieken, 2018). Narratives are also used and abused in creating danger-
ous propaganda and stereotypes that dehumanize groups of people and create and perpetrate societal and
political tension.
Engaging in stories is special to humans. Narratives are a powerful tool for social and cultural regulation.
Through stories, we communicate, we transmit and preserve knowledge, and we create culture. If narra-
tives have the potential to create and preserve culture, they also have the potential to change it. We need to
explore the impact of narratives for culture and social regulation and be mindful of the narratives we create,
teach, and perpetrate. Like with every powerful tool, the power of narratives also invites abuse. As researchers
and as members of society, we have a responsibility to advance understanding as well to educate. Research on
language has to acknowledge that language is more than words and sentences and should continue to support
the growing body of systematic research on language in context.
I hope that I showed that using naturalistic and complex stimuli for experiments is viable and has relevant
insights into cognitive processes which otherwise remain unexplored. Understanding the neural dynamics
of narrative processing and its interface to other cognitive and neural systems is crucial. But we must also
ask ourselves if what we investigate is specific to narrative or if what we are looking at is actually just human
information processing in its most optimized form. As discussed earlier, narratives might just be an optimi-
zation of episodic processing. We have to ask ourselves how useful it is to make theoretical and cognitive
distinctions where there might be no difference in natural kinds.
The neural basis of aesthetic responses to narratives is still poorly understood. As a result, it is difficult
to relate aesthetic responses to narratives to aesthetic experiences from other domains. Not even within

390
Narrative

aesthetic domains that have narrative elements, such as video games, plays, and movies, do we have suffi-
cient understanding how they relate to each other either in behavioral or cognitive terms. How important
is the imaginative aspect of fiction reading that makes it different from engaging with movies? Is simulation
or imagination what facilitates learning? How does the agency aspect of video games relate to more passive
narrative experiences like reading a book or watching a movie? The big challenge for future research in
neuroaesthetics of narrative is to find meaningful categories that can be compared across narratives, different
types of narratives, and other aesthetic domains.
We are a growing international community of researchers who investigate the neurological and behav-
ioral correlates of narratives. With the help of continuously developing methodology and the crystallization
of core goals, we hopefully soon will be a flourishing field within neuroaesthetics and social and affective
neuroscience.

References
AbdulSabur, N. Y., Xu, Y., Liu, S., Chow, H. M., Baxter, M., Carson, J., & Braun, A. R. (2014). Neural correlates and
network connectivity underlying narrative production and comprehension: A combined fMRI and PET study. Cor-
tex, 57, 107–127. https://doi.org/10.1016/j.cortex.2014.01.017
Aboud, K. S., Bailey, S. K., Del Tufo, S. N., Barquero, L. A., & Cutting, L. E. (2019). Fairy tales versus facts: Genre mat-
ters to the developing brain. Cerebral Cortex, 29(11), 4877–4888. https://doi.org/10.1093/cercor/bhz025
Abraham, A., von Cramon, D. Y., & Schubotz, R. I. (2008). Meeting George Bush versus meeting Cinderella: The
neural response when telling apart what is real from what is fictional in the context of our reality. Journal of Cognitive
Neuroscience, 20(6), 965–976. https://doi.org/10.1162/jocn.2008.20059
Albrecht, J. E., O’Brien, E. J., Mason, R. A.,  & Myers, J. L. (1995). The role of perspective in the accessibility of
goals during reading. Journal of Experimental Psychology. Learning, Memory, and Cognition, 21(2), 364–372. https://doi.
org/10.1037//0278-7393.21.2.364
Altmann, U., Bohrn, I. C., Lubrich, O., Menninghaus, W.,  & Jacobs, A. M. (2012, June). The power of emotional
valence-from cognitive to affective processes in reading. Frontiers in Human Neuroscience, 6, 192. https://doi.org/
10.3389/fnhum.2012.00192
Altmann, U., Bohrn, I. C., Lubrich, O., Menninghaus, W.,  & Jacobs, A. M. (2014). Fact vs fiction—How paratex-
tual information shapes our reading processes. Social Cognitive and Affective Neuroscience, 9(1), 22–29. https://doi.
org/10.1093/scan/nss098
Aristotle. (2013). Poetics (reprint ed.). Oxford University Press.
Bachrach, A., Jola, C., & Pallier, C. (2016). Neuronal bases of structural coherence in contemporary dance observation.
NeuroImage, 124, 464–472. https://doi.org/10.1016/j.neuroimage.2015.08.072
Bal, M. (2009). Narratology introduction to the theory of narrative (2nd ed.). University of Toronto Press.
Bal, P. M. M., & Veltkamp, M. (2013). How does fiction reading influence empathy? An experimental investigation on
the role of emotional transportation. PLoS One, 8(1), e55341. https://doi.org/10.1371/journal.pone.0055341
Baldassano, C., Chen, J., Zadbood, A., Pillow, J. W., Hasson, U., & Norman, K. A. (2017). Discovering event struc-
ture in continuous narrative perception and memory. Neuron, 95(3), 709–721.e5. https://doi.org/10.1016/j.
neuron.2017.06.041
Bálint, K., Hakemulder, F., Kuijpers, M. M., Doicaru, M. M., & Tan, E. S. (2016). Reconceptualizing foregrounding.
Scientific Study of Literature, 6(2), 176–207. https://doi.org/10.1075/ssol.6.2.02bal
Bavishi, A., Slade, M. D., & Levy, B. R. (2016). A chapter a day: Association of book reading with longevity. Social Science
and Medicine, 164, 44–48. https://doi.org/10.1016/j.socscimed.2016.07.014
Bezdek, M. A., Gerrig, R. J., Wenzel, W. G., Shin, J., Pirog Revill, K., & Schumacher, E. H. (2015). Neural evidence
that suspense narrows attentional focus. Neuroscience, 303, 338–345. https://doi.org/10.1016/j.neuroscience.2015
.06.055
Black, J., & Barnes, J. L. (2015). Fiction and social cognition: The effect of viewing award-winning television dramas
on theory of mind. Psychology of Aesthetics, Creativity, and the Arts, 9, 423–429. https://doi.org/10.1037/aca0000031
Bower, G. H., & Morrow, D. G. (1990). Mental models in narrative comprehension. Science (New York, NY), 247(4938),
44–48. https://doi.org/10.1126/science.2403694
Boyd, B. (2005). Literature and evolution: A bio-cultural approach. Philosophy and Literature, 29(1), 1–23. https://doi.
org/10.1353/phl.2005.0002

391
Franziska Hartung

Boyd, R. L., Blackburn, K. G.,  & Pennebaker, J. W. (2020). The narrative arc: Revealing core narrative structures
through text analysis. Science Advances, 6(32), eaba2196. https://doi.org/10.1126/sciadv.aba2196
Bruner, J. (1990). Acts of meaning. Harvard University Press.
Bruner, J. (2004). Life as a narrative. Social Research, 71(3), 691–710. https://doi.org/10.1353/sor.2004.0045
Bruner, J. (1986). Actual minds, possible worlds. Harvard University Press.
Busselle, R., & Bilandzic, H. (2009). Measuring narrative engagement. Media Psychology, 12(4), 321–347. https://doi.
org/10.1080/15213260903287259
Butcher, S. H. (1907). The poetics of Ariostotle (4th ed.). Macmillan, and Co.
Cacioppo, J. T., Petty, R. E., Feinstein, J. A., & Jarvis, W. B. G. (1996). Dispositional differences in cognitive motivation:
The life and times of individuals varying in need for cognition. Psychological Bulletin, 119(2), 197–253. https://doi.
org/10.1037/0033-2909.119.2.197
Carlaco, N., Fong, K., Rain, M.,  & Mar, R. M. (2017). Absorption in narrative fiction and its possible impact on
social abilities. In F. Hakemulder, M. M. Kuijpers, E. S. Tan, K. Bálint, & M. M. Doicaru (Eds.), Narrative absorption
(pp. 293–313). Linguistic Approaches to Literature 27.
Chow, H. M., Mar, R. A., Xu, Y., Liu, S., Wagage, S., & Braun, A. R. (2013). Embodied comprehension of stories:
Interactions between language regions and modality-specific neural systems. Journal of Cognitive Neuroscience, 26(2),
279–295. https://doi.org/10.1162/jocn
Chow, H. M., Mar, R. A., Xu, Y., Liu, S., Wagage, S., & Braun, A. R. (2015). Personal experience with narrated events
modulates functional connectivity within visual and motor systems during story comprehension. Human Brain Map-
ping, 36(4), 1494–1505. https://doi.org/10.1002/hbm.22718
Christoforou, C., Papadopoulos, T. C., Constantinidou, F., & Theodorou, M. (2017, December). Your brain on the
movies: A  computational approach for predicting box-office performance from viewer’s brain responses to movie
trailers. Frontiers in Neuroinformatics, 11, 1–13. https://doi.org/10.3389/fninf.2017.00072
Csikszentmihalyi, M. (1990). Flow: The psychology of optimal experience. Harper & Row.
Dancygier, B. (2014). Conclusion: Multiple viewpoints, multiple spaces. In Viewpoint in language: A multimodal perspective
(pp. 219–231). Cambridge University Press.
den Boer, P. C. A. M., Wiersma, D., & Van den Bosch, R. J. (2004). Why is self-help neglected in the treatment of emotional
disorders? A meta analysis. Psychological Medicine, 34(6), 959–971. https://doi.org/10.1017/S003329170300179X
Eekhof, L. S., Kuijpers, M. M., Faber, M., Gao, X., Mak, M., van den Hoven, E., & Willems, R. M. (2021). Lost in a
story, detached from the words. Discourse Processes, 1–22. https://doi.org/10.1080/0163853X.2020.1857619
Ferstl, E. C., Neumann, J., Bogler, C., & von Cramon, D. Y. (2008). The extended language network: A meta-analysis
of neuroimaging studies on text comprehension. Human Brain Mapping, 29(5), 581–593. https://doi.org/10.1002/
hbm.20422
Genette, G. (1980). Narrative discourse. Cornell University Press.
Gernsbacher, M. A. (1997). Two decades of structure building. Discourse Processes, 23(3), 265–304. https://doi.
org/10.1080/01638539709544994
Gerrig, R. J. (1993). Experiencing narrative worlds. Yale University Press.
Gerrig, R. J. (1999). Experiencing narrative worlds: On the psychological activities of reading. Westview Press.
Goldstein, T. R. (2009). The pleasure of unadulterated sadness: Experiencing sorrow in fiction, nonfiction, and “in per-
son.” Psychology of Aesthetics, Creativity, and the Arts, 3(4), 232–237. https://doi.org/10.1037/a0015343
Green, M. C. (2004). Transportation into narrative worlds: The role of prior knowledge and perceived realism. Discourse
Processes, 38(2), 247–266. https://doi.org/10.1207/s15326950dp3802_5
Green, M. C., & Brock, T. C. (2000). The role of transportation in the persuasiveness of public narratives. Journal of
Personality and Social Psychology, 79(5), 701–721. https://doi.org/10.1037/0022-3514.79.5.701
Green, M. C.,  & Brock, T. C. (2002). In the mind’s eye: Transportation-imagery model of narrative persuasion. In
M. C. Green, J. J. Strange & T. C. Brock (Eds.), Narrative impact: Social and cognitive foundations (pp. 315–341). Law-
rence Erlbaum.
Green, M. C., Garst, J., Brock, T. C., & Chung, S. (2006). Fact versus fiction labeling: Persuasion parity despite height-
ened scrutiny of fact. Media Psychology, 8(3), 267–285. https://doi.org/10.1207/s1532785xmep0803_4
Hagoort, P. (2017). The core and beyond in the language-ready brain. Neuroscience and Biobehavioral Reviews, 81, 194–
204. https://doi.org/10.1016/j.neubiorev.2017.01.048
Hagoort, P., & Indefrey, P. (2014). The neurobiology of language beyond single words. Annual Review of Neuroscience, 37,
347–362. https://doi.org/10.1146/annurev-neuro-071013-013847
Hakemulder, J. F. (2004). Foregrounding and its effect on readers’ perception: The effects of personal involvement in nar-
rative discourse. A Special Issue of Discourse Processes, 6950, 193–218. https://doi.org/10.1207/s15326950dp3802_3

392
Narrative

Han, S., Jiang, Y., Humphreys, G. W., Zhou, T., & Cai, P. (2005). Distinct neural substrates for the perception of real and
virtual visual worlds. NeuroImage, 24(3), 928–935. https://doi.org/10.1016/j.neuroimage.2004.09.046
Hartung, F., Burke, M., Hagoort, P., & Willems, R. M. (2016). Taking perspective: Personal pronouns affect experiential
aspects of literary reading. PLoS One, 11(5), e0154732. https://doi.org/10.1371/journal.pone.0154732
Hartung, F., Hagoort, P., & Willems, R. M. (2017). Readers select a comprehension mode independent of pronoun:
Evidence from fMRI during narrative comprehension. Brain and Language, 170, 29–38. https://doi.org/10.1016/j.
bandl.2017.03.007
Hartung, F., Wang, Y., Mak, M.,  Willems, R. M.,  & Chatterjee, A. (2021). Aesthetic appraisals of literary style and
emotional intensity in narrative engagement are neurally dissociable. Communications Biology, 4(1), 1401. https://doi.
org/10.1038/s42003-021-02926-0
Hartung, F., Withers, P., Hagoort, P., & Willems, R. M. (2017). When fiction is just as real as fact: No differences in
reading behavior between stories believed to be based on true or fictional events. Frontiers in Psychology, 8, 1618.
https://doi.org/10.3389/fpsyg.2017.01618
Hasson, U., Chen, J., & Honey, C. J. (2015). Hierarchical process memory: Memory as an integral component of infor-
mation processing. Trends in Cognitive Sciences, 19(6), 304–313. https://doi.org/10.1016/j.tics.2015.04.006
Hasson, U., Egidi, G., Marelli, M., & Willems, R. M. (2018, June). Grounding the neurobiology of language in first
principles: The necessity of non-language-centric explanations for language comprehension. Cognition, 180(2017),
135–157. https://doi.org/10.1016/j.cognition.2018.06.018
Hasson, U., Nir, Y., Levy, I., Fuhrmann, G., & Malach, R. (2004). Intersubject synchronization of cortical activity during
natural vision. Science (New York, NY), 303(5664), 1634–1640. https://doi.org/10.1126/science.1089506
Horton, W. S., & Rapp, D. N. (2003). Out of sight, out of mind: Occlusion and the accessibility of information in nar-
rative comprehension. Psychonomic Bulletin and Review, 10(1), 104–110. https://doi.org/10.3758/bf03196473
Hsu, C. T., Conrad, M.,  & Jacobs, A. M. (2014). Fiction feelings in Harry Potter: Haemodynamic response in the
mid-cingulate cortex correlates with immersive reading experience. NeuroReport, 25(17), 1356–1361. https://doi.
org/10.1097/WNR.0000000000000272
Jääskeläinen, I. P., Sams, M., Glerean, E., & Ahveninen, J. (2021). Movies and narratives as naturalistic stimuli in neuro-
imaging. NeuroImage, 224, 117445. https://doi.org/10.1016/j.neuroimage.2020.117445
Jacobs, A. M. (2015a). Neurocognitive poetics: Methods and models for investigating the neuronal and cognitive-affective
bases of literature reception. Frontiers in Human Neuroscience, 9, 186. https://doi.org/10.3389/fnhum.2015.00186
Jacobs, A. M. (2015b). Towards a neurocognitive poetics model of literary reading. In R. M. Willems (Ed.), Cognitive
neuroscience of natural language use (pp. 135–159). Cambridge University Press.
Jacobs, A. M., & Willems, R. M. (2018). The fictive brain: Neurocognitive correlates of engagement in literature. Review
of General Psychology, 22(2), 147–160. https://doi.org/10.1037/gpr0000106
Johnson, D. R. (2012). Transportation into a story increases empathy, prosocial behavior, and perceptual bias toward
fearful expressions. Personality and Individual Differences, 52(2), 150–155. https://doi.org/10.1016/j.paid.2011.10.005
Johnson, D. R. (2013). Transportation into literary fiction reduces prejudice against and increases empathy for Arab-
Muslims. Scientific Study of Literature, 3(1), 77–92. https://doi.org/10.1075/ssol.3.1.08joh
Johnson-Laird, P. N. (1983). Mental models: Towards a cognitive science of language, inference and consciousness. Harvard Uni-
versity Press.
Kaufman, G. F., & Libby, L. K. (2012). Changing beliefs and behavior through experience-taking. Journal of Personality
and Social Psychology, 103(1), 1–19. https://doi.org/10.1037/a0027525
Kauttonen, J., Hlushchuk, Y., & Tikka, P. (2015). Optimizing methods for linking cinematic features to fMRI data.
NeuroImage, 110, 136–148. https://doi.org/10.1016/j.neuroimage.2015.01.063
Kidd, D. C., & Castano, E. (2013). Reading literary fiction improves theory of mind. Science (New York, NY), 342(6156),
377–380. https://doi.org/10.1126/science.1239918
Kidd, D. C., & Castano, E. (2019). Reading literary fiction and theory of mind: Three preregistered replications and
extensions of Kidd and Castano (2013).  Social Psychological and Personality Science,  10(4), 522–531.  https://doi.
org/10.1177/1948550618775410
Knoop, C. A., Wagner, V., Jacobsen, T., & Menninghaus, W. (2016). Mapping the aesthetic space of literature “from
below”. Poetics, 56, 35–49. https://doi.org/10.1016/j.poetic.2016.02.001
Komeda, H., Tsunemi, K., Inohara, K., Kusumi, T., & Rapp, D. N. (2013). Beyond disposition: The processing conse-
quences of explicit and implicit invocations of empathy. Acta Psychologica, 142(3), 349–355. https://doi.org/10.1016/j.
actpsy.2013.01.002
Koopman, E. M. E. (2015). Empathic reactions after reading: The role of genre, personal factors and affective responses.
Poetics, 50, 62–79. https://doi.org/10.1016/j.poetic.2015.02.008

393
Franziska Hartung

Koopman, E. M. E.,  & Hakemulder, F. (2015). Effects of literature on empathy and self-reflection: A  theoretical-
empirical framework. Journal of Literary Theory, 9(1), 79–111. https://doi.org/10.1515/jlt-2015-0005
Kuhlen, A. K., Allefeld, C., & Haynes, J. D. (2012, September). Content-specific coordination of listeners’ to speakers’
EEG during communication. Frontiers in Human Neuroscience, 6, 266. https://doi.org/10.3389/fnhum.2012.00266
Kuijpers, M. M., Douglas, S., & Kuiken, D. (2020). Capturing the ways we read. Anglistik, 31(1), 53–69. https://doi.
org/10.33675/ANGL/2020/1/6
Kuijpers, M. M., & Hakemulder, F. (2018). Understanding and appreciating literary texts through rereading. Discourse
Processes, 55(7), 619–641. https://doi.org/10.1080/0163853X.2017.1390352
Kuijpers, M. M., Hakemulder, F., Tan, E. S., & Doicaru, M. M. (2014). Exploring absorbing reading experiences: Devel-
oping and validating a self-report scale to measure story world absorption. Scientific Study of Literature, 4(1), 89–122.
https://doi.org/10.1075/ssol.4.1.05kui
Kurby, C. A., & Zacks, J. M. (2013). The activation of modality-specific representations during discourse processing.
Brain and Language, 126(3), 338–349. https://doi.org/10.1016/j.bandl.2013.07.003
Lee, H., Bellana, B., & Chen, J. (2020). What can narratives tell us about the neural bases of human memory? Current
Opinion in Behavioral Sciences, 32, 111–119. https://doi.org/10.1016/j.cobeha.2020.02.007
Lerner, Y., Honey, C. J., Silbert, L. J., & Hasson, U. (2011). Topographic mapping of a hierarchy of temporal receptive
windows using a narrated story. Journal of Neuroscience: the Official Journal of the Society for Neuroscience, 31(8), 2906–
2915. https://doi.org/10.1523/JNEUROSCI.3684-10.2011
Mak, M., de Vries, C., & Willems, R. M. (2020). The influence of mental imagery instructions and personality charac-
teristics on reading experiences. Collabra: Psychology, 6(1), 1–24. https://doi.org/10.1525/collabra.281
Mano, Y., Harada, T., Sugiura, M., Saito, D. N., & Sadato, N. (2009). Perspective-taking as part of narrative comprehension:
A functional MRI study. Neuropsychologia, 47(3), 813–824. https://doi.org/10.1016/j.neuropsychologia.2008.12.011
Mar, R. A. (2011). The neural bases of social cognition and story comprehension. Annual Review of Psychology, 62,
103–134. https://doi.org/10.1146/annurev-psych-120709-145406
Mar, R. A. (2018). Stories and the promotion of social cognition. Current Directions in Psyhological Science, 27, 257–262.
Mar, R. A., Kelley, W. M., Heatherton, T. F., & Macrae, C. N. (2007). Detecting agency from the biological motion of
veridical vs animated agents. Social Cognitive and Affective Neuroscience, 2(3), 199–205. https://doi.org/10.1093/scan/
nsm011
Mar, R. A., & Oatley, K. (2008). The function of fiction is the abstraction and simulation of social experience. Perspectives
on Psychological Science, 3(3), 173–192. https://doi.org/10.1111/j.1745-6924.2008.00073.x
Mar, R. A., Oatley, K., Hirsh, J., dela Paz, J., & Peterson, J. B. (2006). Bookworms versus nerds: Exposure to fiction
versus non-fiction, divergent associations with social ability, and the simulation of fictional social worlds. Journal of
Research in Personality, 40(5), 694–712. https://doi.org/10.1016/j.jrp.2005.08.002
Maslej, M. M., Quinlan, J. A., & Mar, R. M. (2020). Aesthetic responses to the characters, plots, worlds, and style of
stories. In M. Nadal & O. Vartanian (Eds.), Oxford handbook of empirical aesthetics. Oxford University Press. https://doi.
org/10.1093/oxfordhb/9780198824350.001.0001, ISBN: 9780198824350.
Metz-Lutz, M. N., Bressan, Y., Heider, N., & Otzenberger, H. (2010, August). What physiological changes and cerebral
traces tell us about adhesion to fiction during theater-watching? Frontiers in Human Neuroscience, 4, 1–10. https://doi.
org/10.3389/fnhum.2010.00059
Mumper, M. L., & Gerrig, R. J. (2017). Leisure reading and social cognition: A meta-analysis. Psychology of Aesthetics,
Creativity, and the Arts, 11(1), 109–120.  https://doi.org/10.1037/aca0000089
Nguyen, M., Vanderwal, T., & Hasson, U. (2019). Shared understanding of narratives is correlated with shared neural
responses. NeuroImage, 184, 161–170. https://doi.org/10.1016/j.neuroimage.2018.09.010
Nijhof, A. D., & Willems, R. M. (2015). Simulating fiction: Individual differences in literature comprehension revealed
with FMRI. PLoS One, 10(2), e0116492. https://doi.org/10.1371/journal.pone.0116492
Nussbaum, M. (1997). Poetic justice: The literary imagination and public life. Beacon Press.
Oatley, K. (1999a). Meeting of minds: Dialogue, sympathy, and identification in reading fiction. Poetics, 26(5–6), 439–
454. https://doi.org/10.1016/S0304-422X(99)00011-X
Oatley, K. (1999b). Why fiction may be twice as true as fact: Fiction as cognitive and emotional simulation. Review of
General Psychology, 3(2), 101–117. https://doi.org/10.1037/1089-2680.3.2.101
Oatley, K. (2016). Fiction: Simulation of social worlds. Trends in Cognitive Sciences, 20, 618–628.
Panero, M. E., Weisberg, D. S., Black, J., Goldstein, T. R., Barnes, J. L., Brownell, H., & Winner, E. (2016). Does read-
ing a single passage of literary fiction really improve theory of mind? An attempt at replication. Journal of Personality
and Social Psychology, 111(5), e46–e54. https://doi.org/10.1037/pspa0000064
Poerio, G., & Totterdell, P. (2020). The effect of fiction on the well-being of older adults: A longitudinal RCT interven-
tion study using audiobooks. Psychosocial Intervention, 29(1), 29–38. https://doi.org/10.5093/PI2019A16

394
Narrative

Rimmon-Kennan, S. (2002). Narrative fiction (2nd ed.). Routledge.


Samur, D., Tops, M., & Koole, S. L. (2018). Does a single session of reading literary fiction prime enhanced mentalis-
ing performance? Four replication experiments of Kidd and Castano (2013). Cognition and Emotion, 32(1), 130–144.
https://doi.org/10.1080/02699931.2017.1279591
Samur, D., Tops, M., Slapšinskaitė, R., & Koole, S. L. (2021). Getting lost in a story: How narrative engagement emerges
from narrative perspective and individual differences in alexithymia. Cognition and Emotion, 35(3), 576–588. https://
doi.org/10.1080/02699931.2020.1732876
Sanders, J., & van Krieken, K. (2018, September). Exploring narrative structure and hero enactment in brand stories.
Frontiers in Psychology, 9, 1–17. https://doi.org/10.3389/fpsyg.2018.01645
Sanford, A. J., & Emmott, C. (2012). Mind, brain, narrative. Cambridge University Press.
Schrott, R., & Jacobs, A. M. (2011). Gehirn und Gedicht: Wie wir unsere Wirklichkeiten konstruieren (Brain and poetry: How
we construct our realities). Hanser.
Sestir, M., & Green, M. C. (2010). You are who you watch: Identification and transportation effects on temporary self-
concept. Social Influence, 5(4), 272–288. https://doi.org/10.1080/15534510.2010.490672
Simony, E., Honey, C. J., Chen, J., Lositsky, O., Yeshurun, Y., Wiesel, A., & Hasson, U. (2016, May). Dynamical recon-
figuration of the default mode network during narrative comprehension. Nature Communications, 7(2015), 1–13.
https://doi.org/10.1038/ncomms12141
Speer, N. K., Reynolds, J. R., Swallow, K. M., & Zacks, J. M. (2009). Reading stories activates neural representations of vis-
ual and motor experiences. Psychological Science, 20(8), 989–999. https://doi.org/10.1111/j.1467-9280.2009.02397.x
Sperduti, M., Arcangeli, M., Makowski, D., Wantzen, P., Zalla, T., Lemaire, S., Dokic, J., Pelletier, J.,  & Piolino, P.
(2016). The paradox of fiction: Emotional response toward fiction and the modulatory role of self-relevance. Acta
Psychologica, 165, 53–59. https://doi.org/10.1016/j.actpsy.2016.02.003
Sonkusare, S., Breakspear, M., & Guo, C. (2019). Naturalistic stimuli in neuroscience: Critically acclaimed. Trends in
Cognitive Sciences, 23(8), 699–714. https://doi.org/10.1016/j.tics.2019.05.004
Szaflarski, J. P., Altaye, M., Rajagopal, A., Eaton, K., Meng, X., Plante, E., & Holland, S. K. (2012). A 10-year longi-
tudinal fMRI study of narrative comprehension in children and adolescents. NeuroImage, 63(3), 1188–1195. https://
doi.org/10.1016/j.neuroimage.2012.08.049
Tamir, D. I., Bricker, A. B., Dodell-Feder, D., & Mitchell, J. P. (2016). Reading fiction and reading minds: The role of
simulation in the default network. Social Cognitive and Affective Neuroscience, 11(2), 215–224. https://doi.org/10.1093/
scan/nsv114
Tal-Or, N., & Cohen, J. (2010). Understanding audience involvement: Conceptualizing and manipulating identification
and transportation. Poetics, 38(4), 402–418. https://doi.org/10.1016/j.poetic.2010.05.004
Taylor, L. J., & Zwaan, R. A. (2009). Action in cognition: The case of language. Language and Cognition, 1(1), 45–58.
https://doi.org/10.1515/LANGCOG.2009.003
Tylén, K., Christensen, P., Roepstorff, A., Lund, T., Østergaard, S., & Donald, M. (2015). Brains striving for coher-
ence: Long-term cumulative plot formation in the default mode network. NeuroImage, 121, 106–114. https://doi.
org/10.1016/j.neuroimage.2015.07.047
van den Broek, P., Lorch, R. F., Linderholm, T., & Gustafson, M. (2001). The effects of readers’ goals on inference
generation and memory for texts. Memory and Cognition, 29(8), 1081–1087. https://doi.org/10.3758/BF03206376
van den Broek, P., Rapp, D. N., & Kendeou, P. (2005). Integrating memory-based and constructionist processes in accounts
of reading comprehension. Discourse Processes, 39(2), 299–316. https://doi.org/10.1207/s15326950dp3902&3_11
van den Hoven, E., Hartung, F., Burke, M., & Willems, R. M. (2016). Individual differences in sensitivity to style during
literary reading. Insights from Eye-Tracking, 2(1), 1–16.
van Krieken, K., Hoeken, H.,  & Sanders, J. (2015). From reader to mediated witness: The engaging effects
of journalistic crime narratives. Journalism and Mass Communication Quarterly, 92(3), 580–596. https://doi.
org/10.1177/1077699015586546
van Krieken, K., & Sanders, J. (2019). What is narrative journalism? A systematic review and an empirical agenda. Jour-
nalism, 1–20. https://doi.org/10.1177/1464884919862056
van Krieken, K., Sanders, J., & Hoeken, H. (2015). Viewpoint representation in journalistic crime narratives: An analy-
sis of grammatical roles and referential expressions. Journal of Pragmatics, 88, 220–230. https://doi.org/10.1016/j.
pragma.2014.07.012
van Krieken, K. (2018). How reading narratives can improve our fitness to survive. Narrative Inquiry, 28(1), 139–160.
https://doi.org/10.1075/ni.17049.kri
van Kuijk, I., Verkoeijen, P., Dijkstra, K., & Zwaan, R. A. (2018, January 1). The effect of reading a short passage of
literary fiction on theory of mind: A replication of Kidd and Castano (2013). Collabra: Psychology, 4(1), 7. https://
doi.org/10.1525/collabra.117

395
Franziska Hartung

van Peer, W., Hakemulder, J., & Zyngier, S. (2007). Lines on feeling: Foregrounding, aesthetics and meaning. Language
and Literature, 16(2), 197–213. https://doi.org/10.1177/0963947007075985
Vasquez, J. M. (2005). Ethnic identity and Chicano literature: How ethnicity affects reading and reading affects ethnic
consciousness. Ethnic and Racial Studies, 28(5), 903–924. https://doi.org/10.1080/01419870500158927
Vezzali, L., Hewstone, M., Capozza, D., Giovannini, D.,  & Wölfer, R. (2014). Improving intergroup relations with
extended and vicarious forms of indirect contact. European Review of Social Psychology, 25(1), 314–389. https://doi.
org/10.1080/10463283.2014.982948
Wallentin, M., Nielsen, A. H., Vuust, P., Dohn, A., Roepstorff, A., & Lund, T. E. (2011a). Amygdala and heart rate
variability responses from listening to emotionally intense parts of a story. NeuroImage, 58(3), 963–973. https://doi.
org/10.1016/j.neuroimage.2011.06.077
Wallentin, M., Nielsen, A. H., Vuust, P., Dohn, A., Roepstorff, A., & Lund, T. E. (2011b). BOLD response to motion
verbs in left posterior middle temporal gyrus during story comprehension. Brain and Language, 119(3), 221–225.
https://doi.org/10.1016/j.bandl.2011.04.006
Willems, R. M.,  & Jacobs, A. M. (2016). Caring about Dostoyevsky: The untapped potential of studying literature.
Trends in Cognitive Sciences, 20, 243–245.
Wimmer, L., Friend, S., Currie, G., & Ferguson, H. J. (2021, February). Reading fictional narratives to improve social
and moral cognition: The influence of narrative perspective, transportation, and identification. Frontiers in Communica-
tion, 5, 1–19. https://doi.org/10.3389/fcomm.2020.611935
Wolfe, M. B. W. (2005). Memory for narrative and expository text: Independent influences of semantic associations and
text organization. Journal of Experimental Psychology. Learning, Memory, and Cognition, 31(2), 359–364. https://doi.
org/10.1037/0278-7393.31.2.359
Wolfe, M. B. W., & Mienko, J. A. (2007). Learning and memory of factual content from narrative and expository text.
British Journal of Educational Psychology, 77(3), 541–564. https://doi.org/10.1348/000709906X143902
Xu, J., Kemeny, S., Park, G., Frattali, C., & Braun, A. (2005). Language in context: Emergent features of word, sentence,
and narrative comprehension. NeuroImage, 25(3), 1002–1015. https://doi.org/10.1016/j.neuroimage.2004.12.013
Yang, X. H., Li, H. J., Lin, N., Zhang, X. P., Wang, Y. S., Zhang, Y., Zhang, Q., Zuo, X. N., & Yang, Y. F. (2019,
June). Uncovering cortical activations of discourse comprehension and their overlaps with common large-scale neural
networks. NeuroImage, 203, 116200. https://doi.org/10.1016/j.neuroimage.2019.116200
Yarkoni, T., Speer, N. K., & Zacks, J. M. (2008). Neural substrates of narrative comprehension and memory. NeuroImage,
41(4), 1408–1425. https://doi.org/10.1016/j.neuroimage.2008.03.062
Zacks, J. M., Speer, N. K., Swallow, K. M., Braver, T. S., & Reynolds, J. R. (2007). Event perception: A mind-brain
perspective. Psychological Bulletin, 133(2), 273–293. https://doi.org/10.1037/0033-2909.133.2.273
Zadbood, A., Chen, J., Leong, Y. C., Norman, K. A., & Hasson, U. (2017). How we transmit memories to other brains:
Constructing shared neural representations via communication. Cerebral Cortex, 27(10), 4988–5000. https://doi.
org/10.1093/cercor/bhx202
Zwaan, R. A. (1994). Effect of genre expectations on text comprehension. Journal of Experimental Psychology: Learning,
Memory, and Cognition, 20(4), 920–933. https://doi.org/10.1037/0278-7393.20.4.920
Zwaan, R. A. (2004). The immersed experiencer: Toward an embodied theory of language comprehension. Psychology
of Learning and Motivation—Advances in Research and Theory, 44, 35–62.
Zwaan, R. A., Magliano, J. P.,  & Graesser, A. C. (1995). Dimensions of situation model construction in narrative
comprehension. Journal of Educational Psychology: Learning, Memory, and Cognition, 21(2), 386–397. https://doi.org/
10.1037/0278-7393.21.2.386
Zwaan, R. A., & van Oostendorp, H. (1993). Do readers construct spatial representations in naturalistic story compre-
hension? Discourse Processes, 16(1–2), 125–143. https://doi.org/10.1080/01638539309544832

396
20
MUSIC-EVOKED EMOTIONS
Their contribution to aesthetic experiences,
health, and well-being

Liila Taruffi and Stefan Koelsch

Emotions are usually a pivotal component of the aesthetic experience originated by music and other art-
forms. At times, emotional reactions to music can become so profound to leave a trace in our memory, to
add a sense of meaning to our lives, to absorb completely our attention from the surroundings, or to lead to
transcendent experiences (Gabrielsson, 2010). Furthermore, the idea that aesthetic experiences with music
can make a powerful contribution to the healing process has been embraced by various cultures, and public
health across the world now recognizes the importance of art-based interventions, including music listening
and making. However, only in recent years have rigorous and systematic studies started to examine the ben-
efits of music; therefore, the extent to which music enhances people’s health is still largely a work in progress.
It is, nevertheless, undeniable that one main route through which music can lead to positive effects on health
and well-being is the experience and modulation of affect.
Over the last two decades, music psychologists have devoted a large and growing amount of attention
to the empirical study of music-evoked emotions (Anglada-Tort & Sanfilippo, 2019). We now know that
the experience and regulation of emotions are among the core reasons people seek music in their daily lives
( Juslin & Laukka, 2004) and that music evokes emotions most of the time we spend listening to it ( Juslin
et al., 2008). Music performers can reliably convey target emotions to listeners (to some extend even cross-
culturally; Fritz et al., 2009; Thompson & Balkwill, 2010) by relying on acoustic features and their percepts1
as well as expressive cues,2 which are also characteristics of the human vocal expression of emotion ( Juslin &
Laukka, 2003). While more cross-cultural research is needed to extend the importance of a number of self-
regulatory functions3 of music to non-WEIRD (Western, educated, industrialized, rich, and democratic)
individuals, emotions evoked by music can be organized across discrete categories or continuous dimensions
with consistent overlap between different cultures (Cowen et al., 2020). Importantly, music has the capability
to modulate all the core components of emotion, such as subjective feeling (i.e., the subjective experience of
an emotional state like joy, sadness, etc.), physiological arousal (e.g., body temperature, heart and respiration
rate, skin conductance), expressive behaviour (e.g., facial muscle activity like smiling or frowning and vocal
expressions like crying), action tendency (e.g., approaching others, dancing), and appraisal (e.g., the cognitive
evaluation of the music). Furthermore, music engages all the core structures of the “limbic” brain, which is
the neural system implicated in emotion-related processing (Koelsch, 2014, 2020).
The fields of affective neuroscience and neuroaesthetics overlap considerably. For example, a rock song
or an orchestral piece conveying and evoking emotions can be linked to a hedonic value and give rise
to aesthetic pleasure (Salimpoor et al., 2011). Vice versa, a number of neuroimaging studies investigating

DOI: 10.4324/9781003008675-22 397


Liila Taruffi and Stefan Koelsch

aesthetic appreciation experienced via different art forms have shown activity of emotion-related brain
regions (Cupchik et al., 2009; Di Dio et al., 2007) or regions involved in the coding of reward (Kawabata &
Zeki, 2004; Vartanian & Goel, 2004). Along these lines, a few music-specific aesthetic models include affec-
tive experiences (i.e., pleasure or enjoyment, discrete and aesthetic emotions) as their core components along
with cognitive components such as aesthetic judgment and liking or preference (Brattico, 2019; Brattico &
Pearce, 2013), although consistent work needs to be carried out to shed light on how the complex interac-
tions between cognitive and affective processes may give rise to a genuine aesthetic experience. Furthermore,
music research has rarely investigated aesthetic-related phenomena with “explicit reference to overarching
aesthetic frameworks, differently from what happens in visual research” (Brattico, 2019, p. 10), leading to
methodological flaws such as assuming that the participants have undergone an aesthetic experience simply
by listening to music (see also the final section of this chapter).
In this chapter, we provide an overview of neuroscientific and behavioural research on music and emo-
tion with an eye towards evaluating its unique role in aesthetic experiences. In the first part of the chapter,
we illustrate the neural underpinning of music-evoked affective experiences such as enjoyment or pleasure,
as well as discrete and aesthetic emotions. We also discuss research investigating self-controlled and uninten-
tional regulation of affective states via music and its relationship to health. In the second part, we deepen the
discussion on how music-evoked affect can significantly contribute to maintaining and enhancing individu-
als’ well-being and health and to its potential use in therapy by highlighting relevant health-domains and
psychological mechanisms through which music can act. In the final part, we argue that ecological validity
is the most pressing pitfall the neuroaesthetics and empirical aesthetics of music have come across, and we
highlight current research trends that can overcome this issue.

Emotions in musical experiences: types, neural correlates,


and regulatory processes
Aesthetic experiences of music are often characterized by the evocation of pleasure or enjoyment as well as
a wide range of emotions. Thus, the current state-of-the-art knowledge regarding the neural correlates of
music-evoked emotions is highly helpful to the scope of the neuroaesthetics of music.

Music-evoked pleasure
Pleasure or enjoyment is an emotional response that often accompanies aesthetic experiences of music and is
associated with liking and preference. At the bodily level, music-evoked pleasure can sometimes give rise to
chills or goosebumps. Chills can be characterized as a combination of a subjective feeling (generally a positive
feeling, although it can also be a negative one; Goldstein, 1980) and physiological arousal (e.g., changes in
heart rate, skin conductance, tingling sensations, feelings of warmth or coldness, shivering down the spine,
and lump-in-the-throat sensation). They are typically experienced in response to favourite self-selected
music (Salimpoor et al., 2009), are associated with structural and acoustic features of the music (Bannister &
Eerola, 2018), and are more common among individuals who score high on the personality trait of openness
to experience (McCrae, 2007).
At the brain level, music-evoked pleasure and chills are associated with the activation of the dopaminergic
mesolimbic system (in particular the nucleus accumbens and the dorsal striatum), commonly referred to as
the “reward system” of the brain (Blood & Zatorre, 2001; Salimpoor et al., 2011), which reinforces behav-
iours to approach stimuli (e.g., sex, food, drinks) that increase fitness and ensure survival of the individual
and the species. Notably, the reward system interacts with the auditory cortex in modulating music-evoked
reward and aesthetic pleasure. For example, in an auction paradigm in which listeners had to provide bids of
how much they were willing to spend to listen again to a piece of music, Salimpoor and colleagues (2013)

398
Music-evoked emotions

observed increased functional connectivity of the auditory cortices with the nucleus accumbens as the reward
value of music increased. Such brain network dynamics are reduced in individuals with music-specific anhe-
donia (i.e., a reduced capacity to experience pleasure from music; Mas-Herrero et al., 2014), indicating a
pivotal role of this specific neural interaction for the enjoyment of music as well as the existence of a wide
range of individual differences in how people take pleasure from music-related activities (Martínez-Molina
et al., 2016; Mas-Herrero et al., 2012).

Discrete emotions
One of music’s strongest appeals consists in its capability to represent and evoke emotions, and in contempo-
rary industrialized societies, music has become one of the most common tools for creating, enhancing, and
regulating affects (van Goethem & Sloboda, 2011), including emotions (i.e., short-lived feelings that come
from a known cause) and moods (i.e., long-lasting feelings that often have no clear cause nor starting point
of formation). In affective psychology and neuroscience, emotions are often conceptualized (and thus opera-
tionalized) as a set of “basic” emotions4 or as blends of different emotions or simply as blends of continuous
emotion dimensions such as valence and arousal. Regardless of the framework adopted by music research-
ers, everyday studies of music listening show that positive emotions and moods (e.g., happy, relaxed, being
moved) dominate among the most commonly felt affects ( Juslin & Laukka, 2004; Sloboda & O’Neill, 2001).
Two recent meta-analyses of neuroimaging studies on music and emotion (Koelsch, 2014; 2020) reported
activity changes in brain structures known to be crucially involved in emotion, including the amygdala, the
(anterior) hippocampal formation, a number of structures from the reward circuitry (ventral and dorsal stria-
tum, the orbitofrontal cortex, and the anterior cingulate cortex), and the auditory cortex. All together, these
results stress the rewarding and pleasurable component of music listening, the importance of attachment-
related emotions and social bonding promoted by music (through emotional and motor synchronization),
and the role of the auditory cortex in processing affect besides sensory perception.
Specifically, BOLD signal intensity in the superficial amygdala (a key brain region traditionally involved
in the processing of basic socio-affective information) increases during music-evoked joy and decreases
during music-evoked fear (Koelsch et al., 2013). Furthermore, functional connections between the super-
ficial amygdala and the nucleus accumbens on the one hand and the mediodorsal thalamus on the other
are stronger during music-evoked joy than music-evoked fear, pointing to a network of brain regions that
may potentially regulate approach–withdrawal behaviour in response to music (Koelsch et al., 2013). The
laterobasal amygdala, which is generally implicated in the coding of reward value of stimuli with different
valence, shows signal changes in response to both joyful (Koelsch et al., 2013; Mueller et al., 2011) and sad
or dissonant, unpleasant music (Koelsch et al., 2006; Mitterschiffthaler et al., 2007). The role of the amygdala
in the processing of salient negative emotions induced by aversive stimuli is also confirmed by older lesion
studies showing that bilateral amygdala damage was specifically linked to impaired recognition of scary music
while perceptual skills remained intact (Gosselin et al., 2007, 2005).
Hippocampal activity has been reported in response to a wide variety of music-evoked emotions, includ-
ing tenderness and peacefulness (Trost et al., 2012), joy (Mueller et al., 2011), sadness (Mitterschiffthaler
et al., 2007), unpleasantness, and fear (Eldar et al., 2007; Koelsch et al., 2006). Because some of these studies
controlled for familiarity effects of the music on the participants (meaning that the participants were unfa-
miliar with the music), the involvement of the hippocampus could be specific to emotion rather than the
evocation of autobiographical memories—the usual function of this brain area along with spatial orientation
and learning as established by previous work. The engagement of the hippocampal formation could in fact be
explained by its role in generating attachment-related, positive affects (Koelsch et al., 2015, Koelsch, 2020).
Such attachment-related emotions could be triggered by music owing to its capability to promote entrain-
ment or synchronization (Clayton, 2012), which can increase mutual trust and cooperation and is associated

399
Liila Taruffi and Stefan Koelsch

with positive emotional effects and possibly with the release of endogenous opioids, including endorphins
(Tarr et al., 2014).
Finally, the auditory cortex is also implicated in music-evoked emotions. Besides its role in processing
sensory information, there is strong evidence to suggest that the auditory cortex directly modulates affec-
tive processes. For example, Liu et al. (2017) reported auditory-limbic connectivity during non-evaluative
attentive listening of unfamiliar music. In addition, anterior and posterior regions of the auditory association
cortex exhibit emotion-characteristic functional connectivity with limbic/paralimbic, somatosensory, visual,
motor-related, and attentional structures (Koelsch et al., 2018).

Aesthetic emotions
Aesthetic emotions are a special subset of discrete emotions that always (1) feature an aesthetic evaluation
or appraisal of a target event and/or object; (2) are differentially tuned to, and predictive of, a specific type
of aesthetic appeal; (3) are associated with a subjective feeling of pleasure or displeasure; and (4) thus can
predict liking or disliking (Menninghaus et al., 2019). Some typical examples are being moved, amazement,
awe, feeling of beauty, feeling of the sublime, feeling bored. Aesthetic emotions can be experienced within
artistic domains but also occur beyond them (e.g., being awestruck by a natural landscape) and are often
characterized by a mixed affective nature including both positive and negative facets (see, for example, nos-
talgia, which is often labelled “bittersweet” because it involves a mixture of sadness and wistful joy; Batcho,
2007; another example is fear, a negative emotion, which can be experienced as pleasurable in the context
of a positively evaluated horror movie). Aesthetic emotions are typically “savoured” for their own sake, with
their subjective feeling being experienced as rewarding per se. Because of this lack of goals or implications in
“real life,” they have often been defined as “non-utilitarian” (Scherer, 2004). However, aesthetic emotions
can actually inform decision making and promote well-being. For example, nostalgia plays an important role
in maintaining personal well-being by increasing self-positivity (Hart et al., 2011), self-esteem (Wildschut
et al., 2006), and accessibility of positive self-attributes (Vess et al., 2012). Furthermore, positive aesthetic
emotions, if experienced as highly pleasurable, are associated with the activation of the reward circuitry and
the experience of chills or tears, along with their physiological arousal (see previous section on music-evoked
pleasure; see Menninghaus et al., 2019 for detailed discussion).
With regard to music, the Geneva Emotional Music Scale (GEMS; Zentner et al., 2008) was introduced
to specifically assess music-evoked aesthetic emotions. The GEMS is the result of a series of field experiments
conducted at a large music festival employing more than 1,000 participants, who were asked to rate the emo-
tions evoked in response to different styles of music on an extensive list of adjectives. The GEMS organizes
the affective experience underlying music listening around nine dimensions (wonder, transcendence, tender-
ness, nostalgia, sadness, peacefulness, power, joyful activation, and tension), which condense into three main
factors named sublimity, vitality, and unease. Trost and colleagues (2012) investigated the neural substrates of
aesthetic emotions evoked by music, as measured by the GEMS. By using parametric regression analyses
based on the intensity of the experienced emotions, they found a differentiated recruitment across emotions
of brain networks linked to reward, memory, self-reflection, and sensorimotor processes. These findings are
of particular importance because they illustrate the impact of musicevoked emotions on brain systems that
are not primarily emotion areas. However, as a note of caution, we suggest that for a proper assessment of
aesthetic emotions, it would have been ideal to collect emotion ratings (e.g., ratings for feeling wonder), in
addition to ratings of a positive or a negative appreciation of the music as artwork and ratings of overall liking.
The aesthetic emotion of nostalgia has received significant attention from researchers. We know that nos-
talgia often accompanies the occurrence of music-evoked autobiographical memories and that its intensity
can be predicted by both contextual (e.g., autobiographical saliency and familiarity with the music) and per-
sonal factors (e.g., nostalgia proneness, trait neuroticism) (Barrett et al., 2010) and their interplay. In a recent

400
Music-evoked emotions

fMRI experiment (Barrett & Janata, 2016), researchers evoked nostalgic experiences by exposing participants
to familiar music that was popular during their adolescent and teenage years. In this way, they were able to
identify a set of brain regions (involved in memory, emotion, and social cognition) whose activity was pre-
dicted by affective personality measures (e.g., nostalgia proneness) that are known to modulate the strength
of nostalgic experiences. These findings are important not only because they provide evidence for the neural
underpinning of music-evoked nostalgia but also because they stress how individual differences highly mod-
ulate people’s neural responses to aesthetic emotions, which are often idiosyncratic in nature (i.e., the same
piece of music can evoke different emotions in different listeners or even in the same listener in different situ-
ations). Along these lines, a recent study showed that a unique network of brain regions (including areas sub-
serving the coding of compassion, mentalizing, and visual mental imagery) related to individual differences
in trait empathy (measured by the Interpersonal Reactivity Index; Davis, 1980) is significantly more active
for instrumental sad music compared with instrumental happy music (Taruffi et al., 2021). These findings
corroborate previous behavioural evidence underscoring that empathic individuals experience sad music in
a compelling way, experiencing overall more pleasurable and intense emotions (e.g., Vuoskoski et al., 2012).

Regulation of emotion via music


Emotion regulation (ER) is an internal process that can be intentional or unintentional, through which a
person maintains a comfortable state of arousal by modulating one or more aspects of emotion (Gross &
Thompson, 2007). Effortful ER involves various strategies to regulate emotions, including altering atten-
tion, challenging interpretations, and reframing the meaning of situations. Impairments on ER can have a
lifelong impact on an individual’s mental health and well-being (Saxena et al., 2011) and are typically found
in a range of affective disorders such as anxiety and depression.5 The neural correlates underlying ER suggest
an interplay between cognitive, executive areas (including frontal regions) and areas from the limbic system
involved in emotion processing (Gyurak et al., 2011), with the former ones providing top-down inhibition
of the latter ones.
Most of the music and emotion studies described in the previous sections focus on evocation rather than
effortful, self-controlled ER. Although there is a substantial body of research investigating the neural cor-
relates and behaviour underlying ER, surprisingly there is a paucity of studies addressing the links between
music and ER.6 However, music can evoke neural responses in the same regions that are implicated in ER
(for a review, see Moore, 2013), thus pointing to the potential implications of the use of music in clinical
interventions. For most individuals, music serves a number of practical ER purposes, including relieving
boredom, alleviating tension, increasing energy, and providing a distraction from concerns (Hallam, 2010).
Besides positive outcomes of music-based ER, this can also take some maladaptive characteristics such as
enhancing negative emotions by listening to unpleasant music; therefore, more studies are needed to create
a clearer picture of both adaptive and maladaptive ER with music. At the behavioural level, music as a tool
for ER has been explored, in particular, in young people, in depressed individuals, and in relationship to
a number of personality traits. For instance, in young people (i.e., 10–24 years of age), the high use of the
ER strategy of discharge (i.e., venting of negative emotion through music) predicts high levels of depression,
anxiety, and stress, while the strategy of entertainment (i.e., maintenance and enhancement of positive emo-
tion) predicts low levels of depression (Thomson et al., 2014). Furthermore, individuals with tendencies to
depression are more likely to enjoy listening to self-selected sad music, even if this has a negative impact on
their moods (Garrido & Schubert, 2015), and depressed individuals listen to music to express their emotions
more often than non-depressed individuals (Wilhelm et al., 2013). Similarly, people higher in trait neuroticism
are more prone to use music as an accessible tool to regulate their negative emotions, although it is not clear
whether the outcome of this process is successful (Miranda & Blais-Rochette, 2020). Nevertheless, more
work is needed to provide comprehensive guidelines to therapists that could consider musical characteristics,

401
Liila Taruffi and Stefan Koelsch

personality traits, and contextual features that play a role in music-based ER. Beyond therapeutic contexts,
a comprehensive understanding of self-regulatory processes with music can enable a strategic, goal-oriented
use of music in daily life settings available to all listeners. Although people report often seeking music for ER,
it is not clear whether such behaviours are intentional or unintentional, and the basis on which people select
music pieces to listen to needs to be examined. Given the importance and omnipresence of music across the
globe, it is crucial to enhance our understanding of which factors play a role in music-based ER processes to
become more intentional about its use.

Music’s effects on health and well-being


According to the World Health Organisation, “health is a state of complete physical, mental, social well-
being and not merely and absence of disease or infirmity.”7 Over the last decades, healthcare has shifted from
a model centred around coping with disease and symptoms to encompass a broader framework into which
prevention and maintenance of well-being are crucial. This change has led to the promotion of good lifestyle
habits (e.g., fitness, meditation, mindfulness, etc.) where music can play a pivotal role as self-management
tool, in particular due to its capability to alleviate stress and arousal, modulate pain levels, and boost immunity
(Chanda & Levitin, 2013; see Chapter 21). As we will highlight in the following, many of these beneficial
outcomes of music are due to its power to evoke and modulate emotions as well as to its nature of “social”
stimulus. This capability of music to improve people’s health and well-being could ultimately contribute to
shedding novel light on aesthetic experiences, stressing their function and importance at both individual and
societal levels in contrast to the “non-utilitarian” perspective on them.

Stress
Stress is a neurochemical response to the loss of the homeostatic equilibrium, which involves multiple
feedback loops at the level of the central and peripheral nervous system. The beneficial effects of music
listening on stress are largely due to its ability to modulate arousal levels and mood and to distract from a
current activity such as a stressful medical procedure. Music can both relax and activate, and there is good
evidence for the effects of relaxing and activating music on heart rate and blood pressure (see, for example,
a meta-analysis by Koelsch & Jäncke, 2015). A meta-analysis of 42 randomized controlled trials8 shows that
music interventions had positive effects on reducing patients’ anxiety and stress in approximately half of the
reviewed studies (Nilsson, 2008). Another meta-analysis of 22 studies using recorded music combined with
a relaxation technique applied to patients before or during medical operations found that this treatment had
a significant effect on decreasing stress and that musicians, females, and participants under 18 years of age in
particular respond better to the music intervention, pointing to the importance of considering individual
variables (Pelletier, 2004). Taken together, these meta-analyses highlight that slow tempo can effectively
induce relaxation, thereby diminishing stress levels.

Pain
Pain is an individual perception consisting of an “unpleasant sensory and emotional experience associated
with actual or potential tissue damage.”9 The impact of pain on quality of life is significant because it can not
only affect physical levels but can also lead to psychological, social, and spiritual concerns. Cognitive pro-
cesses, such as distraction, re-appraisal, and perceived control over pain, have the potential to change the way
we perceive pain, and undoubtedly, the most impressive and extensively examined example is the placebo
analgesic response (Wiech, 2016). Using music as a pain reduction method in hospitals is still rare; however,
there are several potential advantages of music-based treatments such as being cost-effective, non-invasive,

402
Music-evoked emotions

and having no side effects. While evidence on the efficacy of music in reducing pain has to be substanti-
ated yet (given considerable inconsistencies that have been identified across studies), it is plausible that the
potential analgesic effect of music is likely to be due to its capability to modulate attention (specifically acting
as a distractor), to induce positive emotions and pleasure, and also to increase perceived control over pain
(Skevington, 1995; see Chapter 21).
A meta-analysis of 51 randomized controlled trials using music for short-term pain relief in a variety of
contexts (including pain during medical procedures, postoperative pain, labour pain, and cancer pain) dem-
onstrated that listening to music reduces pain intensity levels and opioid requirements by about 15% (Cepeda
et al., 2006). Thus, the size of this effect is relatively small but might be larger when patients have control
over the selection of the music. Another more recent systematic review (Howlin & Rooney, 2020) empha-
sized even more the role of the patients’ sense of agency in mediating the beneficial effects of music listening
through the processes of meaning-making, enjoyment, and musical integration. Furthermore, these findings
point to the necessity of re-conceptualizing music listening interventions as an active interaction between the
patient’s mental and emotional state and the music rather than a unidirectional process where certain types
of music evoke specific physiological or psychological responses (Howlin & Rooney, 2020). In other words,
people undergoing pain may need different types of music at different times: In some situations, music could
act as a pleasant distractor; in others, it may be used to motivate and energize them.

Immune system
There is a clear physiological association between stress and the immune system; for example, chronic stress
has negative effects on the immune system and health in general (Glaser & Kiecolt-Glaser, 2009). Therefore,
from a physiological perspective, the positive effects of music on stress reduction and mood (as reviewed
previously) can in turn also have a positive impact on the immune system and general health status. How-
ever, studies on this topic are challenging (e.g., due to relatively large numbers of participants needed to
achieve sufficient statistical power), and there is only limited evidence available on the effects of music on the
immune system (as reviewed in Koelsch et al., 2016). This might also be the reason, to our knowledge, no
meta-analysis on this topic is available yet.

Influence on mental states


Dysfunctional thoughts and negatively hued mental images are typical symptoms of a wide range of mental
disorders. For example, rumination (a repetitive style of thinking focused on worries and distress; Nolen-
Hoeksema, 2000) and intrusive memories of negative past events are found in depression, intrusive imagery is
a distinctive feature of post-traumatic stress disorder, and hallucinatory imagery is observed in schizophrenia
(Pearson et al., 2013). Also, in healthy individuals, mind-wandering can lead to detrimental effects on mood
(although these effects are mediated by the content of people’s thoughts; Killingsworth & Gilbert, 2010;
Ruby et al., 2013). During the past years, music psychologists have begun to investigate the effects of music-
evoked emotions on internally oriented mental states such as mind-wandering, daydreaming, and visual
mental imagery (for a review, see Taruffi & Küssner, 2019). A number of studies showed that the emotional
tone of the music reflects on the content of thought or images (i.e., heroic music evoked heroic thoughts;
Koelsch et al., 2019; Taruffi et al., 2017), as well as the frequency of mind-wandering (i.e., sad music is
associated with higher mind-wandering levels compared with happy music) and its neural correlates, specifi-
cally the brain’s default mode network (DMN; Taruffi et al., 2017). Furthermore, music-evoked emotions
predict thought valence of mind-wandering episodes during personal music listening in daily life (Taruffi,
2021), and music can systematically influence the vividness, the imagined time passed and distance travelled,
and the emotional tone of people’s imagined journeys in a directed visual imagery task (Herff et al., 2021).

403
Liila Taruffi and Stefan Koelsch

Interestingly, recent work has pointed to the involvement of the DMN in aesthetic experiences of visual art
(Belfi et al., 2019; Vessel et al., 2019); however, it should be noted that there is no evidence available that
links the DMN uniquely to the visual modality. Importantly, clinical studies revealed that a compromised
integrity of the DMN and altered functional connectivity patterns with other large-scale brain networks are
manifested in a wide range of clinical conditions, including depression and anxiety, schizophrenia, autism
spectrum disorder, Alzheimer’s disease, and attention deficit hyperactivity disorder (Andrews-Hanna et al.,
2014). Given the importance of mind-wandering for mental health and the putative role of the DMN in
mental disorders, the use of music to evoke emotions and in turn to regulate mental states has a huge, yet
unexplored, potential for health and well-being and for music-based interventions in clinical populations.

Major challenges and suggestions

Ecological validity
Although the use of artificial and highly controlled music stimuli in a laboratory environment is critical for
unveiling the functional architecture underlying aesthetic processes and its constituent parts, it would be
wrong to assume that brain functions, as observed in such an environment and in response to such stimuli,
would also apply to an aesthetic experience in real-life settings. A major challenge hampering the neuroaes-
thetics and empirical aesthetics of music has been the lack of ecological validity characterizing most of the
published research: Are the experiments conducted in the lab or inside an fMRI scanner, using standardized
music stimuli, really capable of capturing a genuine aesthetic experience? And, related to our previous sec-
tion: Do the findings on the effects of music on health apply to live music contexts?
Specifically, the role of the listening context and the music stimuli seem to be crucial factors underlying
the authenticity and the intensity of aesthetic experiences. For instance, Coutinho and Scherer (2017) com-
pared emotions evoked by the same performance of a Schubert lieder during a live concert and in a labora-
tory setting and found that participants achieved statistically higher levels of emotional arousal in the live
performance than in the laboratory context and that the experience of particular emotions was determined
by complex interactions between auditory and visual cues during the live performance. Moreover, young
adults reported strong positive experiences of music listening in a live listening situation, such as a gig or
festival, with other people and with a large audience (Lamont, 2011). In fact, the social component of music
listening seems to lead to beneficial outcomes for audience members, as witnessed by a couple of studies
showing that live music festivals are associated with positive effects on psychological and social well-being as
well as mood (Little et al., 2017; Packer & Ballantyne, 2010).
To circumvent the issue of ecological validity, researchers should embrace—more often and where
possible—naturalistic approaches with high ecological validity, such as conducting data collection in the
context of a live concert, gig, or music festival or by using real music as stimulus material (see Chap-
ter 24). While the latter suggestion is easier to implement and has been already adopted by several studies
(e.g., Toiviainen et al., 2014), the first one requires more attention and strategic preparation, especially
with regard to the choice of suitable “non-invasive” methods. However, some examples of research on
aesthetic and affective experiences during live concerts or performances are already available (e.g., Eger-
mann & Reuben, 2020; Deil et al., under review; Merrill et al., 2020; see also the virtual workshop “The
Psychology of Live Music Performance” that took place in June 2020). In particular, the implementation
of physiological measures to be combined with self-reports is promising (Egermann et al., 2013) and does
not seem to alter substantially the audience’s experience of music (although extensive preparation time is
needed and there are severe limitations on the participants’ movements). This naturalistic approach should
hopefully allow researchers to capture genuine aesthetic experiences with music. However, research in loco
should not replace simplified investigations in controlled environments, which are a necessary step of the

404
Music-evoked emotions

scientific method and allow us to understand single processes in isolation (on the other hand, naturalistic
paradigms are particularly complex given the dynamic interactions between different social, contextual,
and sensorimotor cues). The combination of such naturalistic approaches with an increased sophistication
of empirical designs in controlled settings should guarantee a rapid and smooth progress of the neuroaes-
thetics of music’s agenda.

Notes
1 Such as the sound’s frequency, which contributes to the percept of pitch.
2 Such as a legato articulation.
3 For example, to relax, to vent negative moods, to get comfort when feeling lonely, and so on.
4 A set of separate and fundamentally different constructs that are panculturally recognized, including fear, anger, sad-
ness, happiness, surprise, and disgust.
5 For example, depressed individuals exhibit a tendency to ruminate, which is focused attention on the symptoms of
one’s distress.
6 ER through music occurs when people engage in music listening to create, change, or maintain positive and negative
emotions (Baltazar & Saarikallio, 2016).
7 Preamble to the Constitution of WHO as adopted by the International Health Conference, New York, 19 June–22
July 1999; signed on 22 July 1999 by the representatives of 61 states (Official Records of WHO, no. 2, p. 100).
8 A randomized controlled trial is a type of scientific (often medical) design aiming at reducing certain sources of sys-
tematic bias when testing the effectiveness of new treatments. This is accomplished by randomly allocating subjects to
two or more groups, treating them differently, and then comparing them with respect to a measured response. One
group—the experimental group—has the intervention being assessed, while the other—usually called the control
group—has an alternative condition, such as a placebo or no intervention.
9 Merskey, H., & Bogduk, N. (Eds.). (1994). Classification of chronic pain: Descriptions of chronic pain syndromes and
definitions of pain terms. In Task force on taxonomy of the IASP. IASP Press.

References
Andrews-Hanna, J. R., Smallwood, J., & Spreng, R. N. (2014). The default network and self-generated thought: Com-
ponent processes, dynamic control, and clinical relevance. Annals of the New York Academy of Sciences, 1316, 29–52.
https://doi.org/10.1111/nyas.12360
Anglada-Tort, M., & Sanfilippo, K. R. M. (2019). Visualizing music psychology: A bibliometric analysis of psychol-
ogy of music, music perception, and musicae scientiae from 1973 to 2017. Music and Science, 2, 1–18. https://doi.
org/10.1177/2059204318811786
Baltazar, M.,  & Saarikallio, S. (2016). Toward a better understanding and conceptualization of affect self-regulation
through music: A  critical, integrative literature review. Psychology of Music, 44(6), 1500–1521. https://doi.org/
10.1177/0305735616663313
Bannister, S., & Eerola, T. (2018). Suppressing the chills: Effects of musical manipulation on the chills response. Frontiers
in Psychology, 9, 2046. https://doi.org/10.3389/fpsyg.2018.02046
Barrett, F. S., Grimm, K. J., Robins, R. W., Wildschut, T., Sedikides, C., & Janata, P. (2010). Music-evoked nostalgia:
Affect, memory, and personality. Emotion, 10(3), 390–403. https://doi.org/10.1037/a0019006
Barrett, F. S., & Janata, P. (2016). Neural responses to nostalgia-evoking music modeled by elements of dynamic musi-
cal structure and individual differences in affective traits. Neuropsychologia, 91, 234–246. https://doi.org/10.1016/j.
neuropsychologia.2016.08.012
Batcho, K. I. (2007). Nostalgia and the emotional tone of song lyrics. American Journal of Psychology, 120(3), 361–381.
Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., Pelli, D. G., & Starr, G. G. (2019).
Dynamics of aesthetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.
org/10.1016/j.neuroimage.2018.12.017
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Brattico, E. (2019). The neuroaesthetics of music: A research agenda coming of age. In M. H. Thaut & D. A. Hodges
(Eds.), The Oxford handbook of music and the brain. Oxford University Press.

405
Liila Taruffi and Stefan Koelsch

Brattico, E., & Pearce, M. (2013). The neuroaesthetics of music. Psychology of Aesthetics, Creativity, and the Arts, 7(1),
48–61. https://doi.org/10.1037/a0031624
Cepeda, M. S., Carr, D. B., Lau, J., & Alvarez, H. (2006). Music for pain relief. Cochrane Database of Systematic Reviews,
2(2), CD004843. https://doi.org/10.1002/14651858.CD004843.pub2
Chanda, M. L.,  & Levitin, D. J. (2013). The neurochemistry of music. Trends in Cognitive Sciences, 17(4), 179–193.
https://doi.org/10.1016/j.tics.2013.02.007
Clayton, M. (2012). What is entrainment? Definition and applications in musical research. Empirical Musicology Review,
7(1–2), 49–56. https://doi.org/10.18061/1811/52979
Coutinho, E., & Scherer, K. R. (2017). The effect of context and audio-visual modality on emotions elicited by a musical
performance. Psychology of Music, 45(4), 550–569. https://doi.org/10.1177/0305735616670496
Cowen, A. S., Fang, X., Sauter, D., & Keltner, D. (2020). What music makes us feel: At least 13 dimensions organize
subjective experiences associated with music across different cultures. Proceedings of the National Academy of Sciences of
the United States of America, 117(4), 1924–1934. https://doi.org/10.1073/pnas.1910704117
Cupchik, G. C., Vartanian, O., Crawley, A., & Mikulis, D. J. (2009). Viewing artworks: Contributions of cognitive con-
trol and perceptual facilitation to aesthetic experience. Brain and Cognition, 70(1), 84–91. https://doi.org/10.1016/j.
bandc.2009.01.003
Davis, M. H. (1980). A multidimensional approach to individual differences in empathy. JSAS Catalog of Selected Docu-
ments in Psychology, 10, 85.
Deil, J., Markert, N., Normand, P., Kammen, P., Küssner, M. B., & Taruffi, L. (Under review). Mind-wandering during
contemporary live music: An exploratory study.
Di Dio, C., Macaluso, E., & Rizzolatti, G. (2007). The golden beauty: Brain response to classical and renaissance sculp-
tures. PLoS One, 2(11), e1201. https://doi.org/10.1371/journal.pone.0001201
Egermann, H., Pearce, M. T., Wiggins, G. A., & McAdams, S. (2013). Probabilistic models of expectation violation
predict psychophysiological emotional responses to live concert music. Cognitive, Affective and Behavioral Neuroscience,
13(3), 533–553. https://doi.org/10.3758/s13415-013-0161-y
Egermann, H.,  & Reuben, F. (2020). “Beauty is how you feel inside”: Esthetic judgments are related to emotional
responses to contemporary music. Frontiers in Psychology, 11, 510029. https://doi.org/10.3389/fpsyg.2020.510029
Eldar, E., Ganor, O., Admon, R., Bleich, A., & Hendler, T. (2007). Feeling the real world: Limbic response to music
depends on related content. Cerebral Cortex, 17(12), 2828–2840. https://doi.org/10.1093/cercor/bhm011
Fritz, T., Jentschke, S., Gosselin, N., Sammler, D., Peretz, I., Turner, R., Friederici, A. D., & Koelsch, S. (2009). Uni-
versal recognition of three basic emotions in music. Current Biology, 19(7), 573–576. https://doi.org/10.1016/j.
cub.2009.02.058
Gabrielsson, A. (2010). Strong experiences with music. In P. N. Juslin & J. A. Sloboda (Eds.), Handbook of music and emo-
tion: Theory, research, applications (pp. 547–574). Oxford University Press.
Garrido, S., & Schubert, E. (2015). Music and people with tendencies to depression. Music Perception, 32(4), 313–321.
https://doi.org/10.1525/mp.2015.32.4.313
Glaser, R.,  & Kiecolt-Glaser, J. (2009). How stress damages immune system and health. Discovery Medicine, 5(26),
165–169.
Goldstein, A. (1980). Thrills in response to music and other stimuli. Physiological Psychology, 8(1), 126–129. https://doi.
org/10.3758/BF03326460
Gosselin, N., Peretz, I., Johnsen, E., & Adolphs, R. (2007). Amygdala damage impairs emotion recognition from music.
Neuropsychologia, 45(2), 236–244. https://doi.org/10.1016/j.neuropsychologia.2006.07.012
Gosselin, N., Peretz, I., Noulhiane, M., Hasboun, D., Beckett, C., Baulac, M., & Samson, S. (2005). Impaired recogni-
tion of scary music following unilateral temporal lobe excision. Brain, 128(3), 628–640. https://doi.org/10.1093/
brain/awh420
Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. In J. J. Gross (Ed.), Handbook of
emotion regulation (pp. 3–24). Guilford Press.
Gyurak, A., Gross, J. J., & Etkin, A. (2011). Explicit and implicit emotion regulation: A dual-process framework. Cogni-
tion and Emotion, 25(3), 400–412. https://doi.org/10.1080/02699931.2010.544160
Hallam, S. (2010). Music education: The role of affect. In P. N. Juslin & J. A. Sloboda (Eds.), Handbook of music and emo-
tion: Theory, research, applications (pp. 719–817). Oxford University Press.
Hart, C. M., Sedikides, C., Wildschut, T., Arndt, J., Routledge, C., & Vingerhoets, A. J. J. M. (2011). Nostalgic rec-
ollections of high and low narcissists. Journal of Research in Personality, 45(2), 238–242. https://doi.org/10.1016/j.
jrp.2011.01.002
Herff, S. A., Cecchetti, G., Taruffi, L., & Déguernel, K. (2021). Music influences vividness and content of imagined jour-
neys in a directed visual imagery task. Scientific Reports, 11(1), 15990. https://doi.org/10.1038/s41598-021-95260-8

406
Music-evoked emotions

Howlin, C., & Rooney, B. (2020). The cognitive mechanisms in music listening interventions for pain: A scoping review.
Journal of Music Therapy, 57(2), 127–167. https://doi.org/10.1093/jmt/thaa003
Juslin, P. N., & Laukka, P. (2003). Communication of emotions in vocal expression and music performance: Different
channels, same code? Psychological Bulletin, 129(5), 770–814. https://doi.org/10.1037/0033-2909.129.5.770
Juslin, P. N., & Laukka, P. (2004). Expression, perception, and induction of musical emotions: A review and a ques-
tionnaire study of everyday listening. Journal of New Music Research, 33(3), 217–238. https://doi.org/10.1080/
0929821042000317813
Juslin, P. N., Liljeström, S., Västfjäll, D., Barradas, G., & Silva, A. (2008). An experience sampling study of emotional
reactions to music: Listener, music, and situation. Emotion, 8(5), 668–683. https://doi.org/10.1037/a0013505
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Killingsworth, M. A., & Gilbert, D. T. (2010). A wandering mind is an unhappy mind. Science, 330(6006), 932. https://
doi.org/10.1126/science.1192439
Koelsch, S. (2014). Brain correlates of music-evoked emotions. Nature Reviews Neuroscience, 15(3), 170–180. https://doi.
org/10.1038/nrn3666
Koelsch, S. (2020). A coordinate-based meta-analysis of music-evoked emotions. Neuroimage, 223, 117350. https://doi.
org/10.1016/j.neuroimage.2020.117350
Koelsch, S., Bashevkin, T., Kristensen, J., Tvedt, J., & Jentschke, S. (2019). Heroic music stimulates empowering thoughts
during mind-wandering. Scientific Reports, 9(1), 10317. https://doi.org/10.1038/s41598-019-46266-w
Koelsch, S., Boehlig, A., Hohenadel, M., Nitsche, I., Bauer, K., & Sack, U. (2016). The impact of acute stress on hor-
mones and cytokines and how their recovery is affected by music-evoked positive mood. Scientific Reports, 6, 23008.
https://doi.org/10.1038/srep23008
Koelsch, S., Fritz, T., V Cramon, D. Y., Müller, K., & Friederici, A. D. (2006). Investigating emotion with music: An
fMRI study. Human Brain Mapping, 27(3), 239–250. https://doi.org/10.1002/hbm.20180
Koelsch, S., Jacobs, A. M., Menninghaus, W., Liebal, K., Klann-Delius, G., von Scheve, C.,  & Gebauer, G. (2015).
The quartet theory of human emotions: An integrative and neurofunctional model. Physics of Life Reviews, 13, 1–27.
https://doi.org/10.1016/j.plrev.2015.03.001
Koelsch, S.,  & Jäncke, L. (2015). Music and the heart. European Heart Journal, 36(44), 3043–3049. https://doi.
org/10.1093/eurheartj/ehv430
Koelsch, S., Skouras, S., Fritz, T., Herrera, P., Bonhage, C., Küssner, M. B., & Jacobs, A. M. (2013). The roles of super-
ficial amygdala and auditory cortex in music-evoked fear and joy. Neuroimage, 81, 49–60. https://doi.org/10.1016/j.
neuroimage.2013.05.008
Koelsch, S., Skouras, S., & Lohmann, G. (2018). The auditory cortex hosts network nodes influential for emotion pro-
cessing: An fMRI study on music-evoked fear and joy. PLoS One, 13(1), e0190057. https://doi.org/10.1371/journal.
pone.0190057
Lamont, A. M. (2011). University students’ strong experiences of music: Pleasure, engagement, and meaning. Musicae
Scientiae, 15, 229–249.
Little, N., Burger, B., & Croucher, S. M. (2017). EDM and ecstasy. The lived experiences of electronic dance music
festivals attendees. Journal of New Music Research, 47(1), 78–95. https://doi.org/10.1080/09298215.2017.1358286
Liu, C., Brattico, E., Abu-Jamous, B., Pereira, C. S., Jacobsen, T., & Nandi, A. K. (2017). Effect of explicit evaluation
on neural connectivity related to listening to unfamiliar music. Frontiers in Human Neuroscience, 11, 611. https://doi.
org/10.3389/fnhum.2017.00611
Martínez-Molina, N., Mas-Herrero, E., Rodríguez-Fornells, A., Zatorre, R. J.,  & Marco-Pallarés, J. (2016). Neural
correlates of specific musical anhedonia. Proceedings of the National Academy of Sciences of the United States of America,
113(46), E7337–E7345. https://doi.org/10.1073/pnas.1611211113
Mas-Herrero, E., Marco-Pallares, J., Lorenzo-Seva, U., Zatorre, R. J., & Rodriguez-Fornells, A. (2012). Individual dif-
ferences in music reward experiences. Music Perception, 31(2), 118–138. https://doi.org/10.1525/mp.2013.31.2.118
Mas-Herrero, E., Zatorre, R. J., Rodriguez-Fornells, A., & Marco-Pallarés, J. (2014). Dissociation between musical and
monetary reward responses in specific musical anhedonia. Current Biology, 24(6), 699–704. https://doi.org/10.1016/j.
cub.2014.01.068
McCrae, R. R. (2007). Aesthetic chills as a universal marker of openness to experience. Motivation and Emotion, 31(1),
5–11. https://doi.org/10.1007/s11031-007-9053-1
Menninghaus, W., Wagner, V., Wassiliwizky, E., Schindler, I., Hanich, J., Jacobsen, T., & Koelsch, S. (2019). What are
aesthetic emotions? Psychological Review, 126(2), 171–195. https://doi.org/10.1037/rev0000135
Merrill, J., Czepiel, A., Fink, L., Toelle, J., & Wald-Fuhrmann, M. (2020). The aesthetic experience of live concerts:
Self-reports and psychophysiology. Psychiatry, https://doi.org/10.31234/osf.io/g829v

407
Liila Taruffi and Stefan Koelsch

Miranda, D.,  & Blais-Rochette, C. (2020). Neuroticism and emotion regulation through music listening: A  meta-
analysis. Musicae Scientiae, 24(3), 342–355. https://doi.org/10.1177/1029864918806341
Mitterschiffthaler, M. T., Fu, C. H., Dalton, J. A., Andrew, C. M., & Williams, S. C. (2007). A functional MRI study
of happy and sad affective states evoked by classical music. Human Brain Mapping, 28(11), 1150–1162. https://doi.
org/10.1002/hbm.20337
Moore, K. S. (2013). A systematic review on the neural effects of music on emotion regulation: Implications for music
therapy practice. Journal of Music Therapy, 50(3), 198–242. https://doi.org/10.1093/jmt/50.3.198
Mueller, K., Mildner, T., Fritz, T., Lepsien, J., Schwarzbauer, C., Schroeter, M. L., & Möller, H. E. (2011). Investigating
brain response to music: A comparison of different fMRI acquisition schemes. Neuroimage, 54(1), 337–343. https://
doi.org/10.1016/j.neuroimage.2010.08.029
Nilsson, U. (2008). The anxiety- and pain-reducing effects of music interventions: A systematic review. AORN Journal,
87(4), 780–807. https://doi.org/10.1016/j.aorn.2007.09.013
Nolen-Hoeksema, S. (2000). The role of rumination in depressive disorders and mixed anxiety/depressive symptoms.
Journal of Abnormal Psychology, 109(3), 504–511. https://doi.org/10.1037/0021-843X.109.3.504
Packer, J., & Ballantyne, J. (2010). The impact of music festival attendance on young people’s psychological and social
well-being. Psychology of Music, 39(2), 164–181. https://doi.org/10.1177/0305735610372611
Pearson, D. G., Deeprose, C., Wallace-Hadrill, S. M., Burnett Heyes, S. B., & Holmes, E. A. (2013). Assessing mental
imagery in clinical psychology: A review of imagery measures and a guiding framework. Clinical Psychology Review,
33(1), 1–23. https://doi.org/10.1016/j.cpr.2012.09.001
Pelletier, C. L. (2004). The effect of music on decreasing arousal due to stress: A meta-analysis. Journal of Music Therapy,
41(3), 192–214. https://doi.org/10.1093/jmt/41.3.192
Ruby, F. J., Smallwood, J., Engen, H., & Singer, T. (2013). How self-generated thought shapes mood—The relation
between mind-wandering and mood depends on the socio-temporal content of thoughts. PLoS One, 8(10), e77554.
https://doi.org/10.1371/journal.pone.0077554
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music.  Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Salimpoor, V. N., Benovoy, M., Longo, G., Cooperstock, J. R.,  & Zatorre, R. J. (2009). The rewarding aspects of
music listening are related to degree of emotional arousal. PLoS One, 4(10), e7487. https://doi.org/10.1371/journal.
pone.0007487
Salimpoor, V. N., van den Bosch, I., Kovacevic, N., McIntosh, A. R., Dagher, A., & Zatorre, R. J. (2013). Interactions
between the nucleus accumbens and auditory cortices predict music reward value. Science, 340(6129), 216–219.
https://doi.org/10.1126/science.1231059
Saxena, P., Dubey, A., & Pandey, R. (2011). Role of emotion regulation difficulties in predicting mental health and well-
being. SIS Journal of Projective Psychology and Mental Health, 18, 147–155.
Scherer, K. R. (2004). Which emotions can be induced by music? What are the underlying mechanisms? And how can
we measure them? Journal of New Music Research, 33, 239–251.
Skevington, S. M. (1995). Psychology of pain. John Wiley & Sons.
Sloboda, J. A., & O’Neill, S. A. (2001). Emotions in everyday listening to music. In P. Juslin & J. Sloboda (Eds.), Music
and emotion: Theory and research (pp. 415–429). Oxford University Press.
Tarr, B., Launay, J., & Dunbar, R. I. (2014). Music and social bonding: “self-other” merging and neurohormonal mecha-
nisms. Frontiers in Psychology, 5, 1096. https://doi.org/10.3389/fpsyg.2014.01096
Taruffi, L. (2021). Mind-wandering during personal music listening in everyday life: Music-evoked emotions pre-
dict thought valence. International Journal of Environmental Research and Public Health, 18(23), 12321. https://doi.
org/10.3390/ijerph182312321
Taruffi, L., & Küssner, M. B. (2019). A review of music-evoked visual mental imagery: Conceptual issues, relation to
emotion, and functional outcome. Psychomusicology: Music, Mind, and Brain, 29(2–3), 62–74. https://doi.org/10.1037/
pmu0000226
Taruffi, L., Pehrs, C., Skouras, S., & Koelsch, S. (2017). Effects of sad and happy music on mind-wandering and the
default mode network. Scientific Reports, 7(1), 14396. https://doi.org/10.1038/s41598-017-14849-0
Taruffi, L., Skouras, S., Pehrs, C., & Koelsch, S. (2021). Trait empathy shapes neural responses toward sad music. Cogni-
tive, Affective and Behavioral Neuroscience, 21(1), 231–241. https://doi.org/10.3758/s13415-020-00861-x
Thompson, W. F., & Balkwill, L.-L. (2010). Cross-cultural similarities and differences. In P. N. Juslin & J. A. Sloboda
(Eds.), Handbook of music and emotion: Theory, research, applications (pp. 755–788). Oxford University Press.
Thomson, C. J., Reece, J. E., & Di Benedetto, M. (2014). The relationship between music-related mood regulation and
psychopathology in young people. Musicae Scientiae, 18(2), 150–165. https://doi.org/10.1177/1029864914521422

408
Music-evoked emotions

Toiviainen, P., Alluri, V., Brattico, E., Wallentin, M.,  & Vuust, P. (2014). Capturing the musical brain with Lasso:
Dynamic decoding of musical features from fMRI data. Neuroimage, 88, 170–180. https://doi.org/10.1016/j.
neuroimage.2013.11.017
Trost, W., Ethofer, T., Zentner, M., & Vuilleumier, P. (2012). Mapping aesthetic musical emotions in the brain. Cerebral
Cortex, 22(12), 2769–2783. https://doi.org/10.1093/cercor/bhr353
van Goethem, A., & Sloboda, J. (2011). The functions of music for affect regulation. Musicae Scientiae, 15, 208–228.
Vartanian, O., & Goel, V. (2004). Neuroanatomical correlates of aesthetic preference for paintings. NeuroReport, 15(5),
893–897. https://doi.org/10.1097/00001756-200404090-00032
Vess, M., Arndt, J., Routledge, C., Sedikides, C., & Wildschut, T. (2012). Nostalgia as a resource for the self. Self and
Identity, 11(3), 273–284. https://doi.org/10.1080/15298868.2010.521452
Vessel, E. A., Isik, A. I., Belfi, A. M., Stahl, J. L., & Starr, G. G. (2019). The default-mode network represents aesthetic
appeal that generalizes across visual domains. Proceedings of the National Academy of Sciences of the United States of America,
116(38), 19155–19164. https://doi.org/10.1073/pnas.1902650116
Vuoskoski, J. K., Thompson, W. F., McIlwain, D., & Eerola, T. (2012). Who enjoys listening to sad music and why? Music
Perception, 29(3), 311–317. https://doi.org/10.1525/mp.2012.29.3.311
Wiech, K. (2016). Deconstructing the sensation of pain: The influence of cognitive processes on pain perception. Science,
354(6312), 584–587. https://doi.org/10.1126/science.aaf8934
Wildschut, T., Sedikides, C., Arndt, J., & Routledge, C. (2006). Nostalgia: Content, triggers, functions. Journal of Person-
ality and Social Psychology, 91(5), 975–993. https://doi.org/10.1037/0022-3514.91.5.975
Wilhelm, K., Gillis, I., Schubert, E., & Whittle, E. L. (2013). On a blue note: Depressed peoples’ reasons for listening to
music. Music and Medicine, 5(2), 76–83. https://doi.org/10.1177/1943862113482143
Zentner, M., Grandjean, D., & Scherer, K. R. (2008). Emotions evoked by the sound of music: Characterization, clas-
sification, and measurement. Emotion, 8(4), 494–521. https://doi.org/10.1037/1528-3542.8.4.494

409
21
THE HEALTH BENEFITS OF
ART EXPERIENCE
Claire Howlin

Art experiences can have direct and measurable health benefits. Since 2019, the World Health Organization
has been advocating for the inclusion of arts-based interventions in routine clinical practice (WHO, 2019).
This recommendation is based on a high-quality evidence base that demonstrates the effectiveness of arts-
based interventions for a range of different outcomes. There is strong evidence that demonstrates the clinical
effectiveness of arts-based interventions for some health outcomes such as pain (Lu et al., 2020; Lee, 2016),
stress (de Witte et al., 2020), and depression (Tang et al., 2020). However, the evidence remains less clear
in terms of the impact of dance and drama interventions in the context of schizophrenia (Ruddy & Milnes,
2005; Ruddy & Dent-Brown, 2007) and dementia (Deshmukh et al., 2018). This may indicate that specific
types of art experiences are only good for specific types of health outcomes. Alternatively, the lack of strong
evidence for some art forms or for some health conditions may simply reflect that this is an emerging area
of research and that answers to these questions will take time to establish. It may also be the case that differ-
ent aspects of art experiences may benefit different health outcomes, for different reasons. For example, in
Parkinson’s, the tempo of the music is important in developing a stronger gait, but for pain management,
the tempo is less important. Either way, it is clear that to understand how art experiences benefit health
outcomes, we need to pay attention to the precise form of art experience that we introduce into specific
health contexts.
This chapter is centred around two examples of art experiences, music listening and art making, for two
different health outcomes: pain management and stress management, respectively. These art experiences have
not been chosen because they are considered superior or more effective than other types of art experience.
Instead, these art experiences have been chosen because they appear to have the most neurobiological evi-
dence to support how they work. Several neuroimaging studies have been conducted to identify the neural
changes in related to reported pain decreases in the presence of music, including fMRI studies and EEG stud-
ies (Coen et al., 2009; Wagner et al., 2009; Dobek et al., 2014; Guo et al., 2020; Garza-Villarreal et al., 2015;
Pando-Naude et al., 2019). Emerging evidence has also started to identify neurobiological changes associ-
ated with mindfulness-based art therapy in terms of decreases in blood cortisol (Walsh et al., 2007; Lawson
et al., 2012; Kaimal et al., 2016; Beerse et al., 2019) and changes in neural activity (Monti et al., 2012; Kruk
et al., 2014; Kaimal et al., 2017). Additionally, these two art experiences provide a nice juxtaposition, in that
music listening is predominantly an activity of appreciation, while clay sculpting is predominantly an activity
of expression. By examining the health benefits of both appreciation and expression, it should be possible

410 DOI: 10.4324/9781003008675-23


The health benefits of art experience

to use the insights gained from examining these contexts in other forms of appreciation and expression or
performance.

Arts, aesthetics, creativity, and health

Arts experience or arts intervention?


Both arts experiences and arts interventions can lead to health and well-being benefits. Everyday arts experi-
ences include a range of activities such as attending galleries, museums, or concerts or engaging with the arts
at home through making, painting, or drawing. Arts interventions encompass a broad class of activities across
all art forms and modalities, including dance, drama, poetry, music, film, sculpture, and painting. Indeed, a
great reason that arts interventions are so appealing is that they can be adapted and implemented according to
the needs of individuals and groups. Arts activities can lead to a wide range of emotional and social benefits
and have been identified to contribute to health and well-being outcomes across the life span (World Health
Organization, 2019). It is being increasingly recognized that cognitive and emotional responses to music
contribute to the health and well-being benefits of art experiences (Howlin & Rooney, 2020).
The key difference between everyday arts experiences and arts interventions is that arts interventions
tend to involve a skilled arts intervention practitioner (e.g., a music therapist or art therapist) and focus on a
specific health issue or problem. Regular art experiences may be more beneficial for general well-being and
stress management, whereas specific arts interventions may help to target more complex issues like childhood
trauma, PTSD, or Parkinson’s disease. It is important to recognize that everyday art experiences, like attend-
ing a gallery or a film with friends, should not be medicalized by referring to them as arts interventions.
At the same time, arts interventions should be duly recognized for their ability to make art more accessible
to different groups and bring art into new contexts and spaces. For example, music interventions are easily
integrated into hospital environments, and community arts programmes help to make everyday art experi-
ences more accessible to groups that may otherwise feel marginalized from the typical gallery or museum
experience. One of the overarching challenges in this area is that although many studies have identified that
arts-based interventions are very effective, there is still much work to be done in terms of identifying how
arts interventions work.

Aesthetic engagement in arts interventions


Aesthetic engagement is too often overlooked in the context of arts interventions. For example, in music
interventions, much attention is often placed on specific acoustic features such as tempo, valence, or depth
without consideration of how people enjoy or interpret these features. This is problematic because there
is very little evidence to support this idea, and several decades of research are quite definitive that specific
acoustic features cannot account for the well-being effects attributed to music interventions (Bradt et al.,
2015; Lee, 2016; Tsai et al., 2014). Instead a person’s own choice of music is a stronger predictor of a success-
ful outcome. On examination of a database looking at the impact of music on pain from more than 90,000
patients, the person’s own choice of music was seen as fundamental to the success of the music intervention.
Additionally, the importance of personal choice in music interventions has been replicated in several system-
atic meta-analyses of the literature (Bradt et al., 2015; Lee, 2016; Tsai et al., 2014). One meta-analysis with
3731 cancer patients demonstrated that personally chosen music had a much larger impact on pain reduction
compared to researcher-chosen music (Bradt et al., 2016). Subsequently, an even larger meta-analysis with
6,430 patients also demonstrated the importance of self-chosen music over experimenter-chosen music (Lee,
2016). However, despite the huge amount of data demonstrating the importance of personal choice in music

411
Claire Howlin

interventions, some studies still search for specific acoustic features related to well-being benefits (Basiński
et al., 2021) but without much success. Perhaps unsurprisingly, big data models are also coming to the same
conclusion. Acoustic features are less important than personal preferences (Basiński et al., 2021).
Personal choice in art increases the chances that people will have a meaningful art experience. On the
surface, asking someone to choose their favourite piece of music or art may seem straightforward, but in
doing so, you are increasing the chance that the person will have a meaningful art experience. It may be that
people experience greater levels of enjoyment, or greater levels of beauty, but regardless of how the aesthetic
experience is conceptualized or described, it can lead to a different pattern of neurophysiological changes
compared to a boring or confusing art experience. Neuroaesthetic approaches highlight that processes of
decision-making and cognitive appraisal underpin neural reward from arts engagement rather than simply
the sensory components of the stimulus (Brattico & Pearce, 2013). Functional magnetic resonance imaging
(fMRI) studies have found differential activation in the right striatum when consonant chords are perceived
as being beautiful compared to consonant cords that are not perceived as beautiful (Nemoto et al., 2010).
This distinction dissociates the experience of pleasure from mere pleasantness and would explain how people
may feel the benefits of a music listening intervention, even if sad or angry music were chosen. Although
dopaminergic neural reward can occur in response to a range of aesthetic stimuli (Brattico, 2015), it is much
more likely to occur in response to an individual’s self-chosen favourite piece of music (Salimpoor et al.,
2011). This indicates that dopaminergic neural reward is not simply an automated physiological response
because it requires a personally meaningful interaction.
The impact of aesthetic engagement is also apparent when we make direct comparisons between arts
interventions and activities that rely on other forms of cognitive engagement. The analgesic benefits related
to music listening are not observed in relation to other cognitive tasks such as listening to a lecture (Zhao &
Chen, 2009) or performing mental arithmetic (Mitchell et al., 2006). Conversely, listening to poetry pro-
vides similar analgesic benefits compared to music listening (Arruda et al., 2016). Taken together, these find-
ings imply that there is something different about using arts interventions as a distraction compared to other
cognitive tasks. It may be that music and poetry present an added opportunity compared to other cognitive
tasks because they can be pleasurable, which intrinsically motivates sustained engagement with the activity.
Aesthetic engagement may also be useful because it provides people with opportunities for self-reflection
and challenges people to search for meaning in potentially abstract stimuli. Of course, searching for meaning
can be a rewarding endeavour in and of itself, but it may also help people to integrate potentially challeng-
ing or abstract aspects of their daily experience, which could also lead to indirect benefits through cognitive
reappraisal.

Using the cognitive vitality model to understand cognitive processes


There are many different processes that mediate health benefits from art engagement. Recently a theoretical
framework called the cognitive vitality model (Howlin & Rooney, 2020) was developed to characterize the
different cognitive mechanisms involved in mediating the analgesic benefits of music. A systematic review
of 75 research papers was used to identify five key cognitive mechanisms: automated attention, cognitive agency,
meaning-making, musical integration, and cognitive vitality. The automated attention mechanism reflects how the
presence of any music will initiate the orienting response and attract lower levels of attention. Based on cur-
rent theories of emotional engagement with music ( Juslin & Västfjäll, 2008; Juslin, 2013; Juslin et al., 2013;
Koelsch, 2012), the cognitive vitality model does not propose that specific music features will affect attention
or emotion in a universal way. However, some level of attention to the music is necessary as a foundation for
deeper levels of music engagement. Once music has activated our attention, cognitive agency is required to
direct attention to different streams within the music and relies on the individual’s specific reasons for listen-
ing and a conscious awareness of the music. Cognitive agency reflects that music listening is an interactive

412
The health benefits of art experience

process between the listener and the characteristics of a piece of music, where the listener can monitor and
influence their response to the music in a process of emotion regulation, similar to Gross’ model (1998). As
people continue to actively listen for a specific reason, they begin to engage in meaning-making processes,
initiated either by spontaneous emotional engagement with the music or by cognitive appraisals of the lyrical
meaning or aesthetic features of the music. Meaning-making processes involve both emotional and cogni-
tive components which may occur simultaneously, reciprocally, or in isolation. Ultimately, this emotional
and cognitive engagement will lead to a rewarding experience for the listener and act as an incentive for the
person to effortlessly maintain active engagement, leading to absorption in the musical experience. Musi-
cal integration brings the music into the forefront of their consciousness not just as an additional sensory
component but as an entirely absorbing phenomenological experience on a cognitive and emotional level.
Subsequent studies with clinical populations with chronic pain have identified that patients are consciously
aware of the role of musical integration and cognitive absorption in facilitating analgesic benefits (Howlin
et al., in review). As a result of this meaningful, rewarding, absorbing music listening experience, people
develop a strengthened sense of self, which characterizes enhanced cognitive vitality. Cognitive vitality pro-
vides the basis for vigilant coping in a flexible manner that is appropriate to specific contexts and the type
of pain being experienced.
The role of cognitive agency has been demonstrated in a recent experiment using the cold pressor task
with 52 participants (Howlin & Rooney, 2021). An experiment was devised to give participants different
degrees of perceived control over a selection of music, when the music was actually predetermined by the
experimenter. When participants had the perception that they were selecting the music, their pain tolerance
increased. This study demonstrates that part of the reason that arts interventions make people feel better is
because it gives them a greater sense of control of their immediate environment, which helps to increase their
locus of control and self-efficacy (Howlin & Rooney, 2021). The act of making a choice is not unique to
music and could explain how similar effects can be observed from quite different forms of art and art inter-
ventions. In this way, the cognitive vitality model could be used as a framework to understand the cognitive
processes involved in a range of arts and health contexts.

Neuroaesthetics of music for pain management

Music for pain management


Music listening has been used to successfully manage pain. A series of meta-analyses have consistently shown
beneficial effects (Bradt et al., 2016; Cepeda et al., 2006; Garza-Villarreal et al., 2017; Lee, 2016; Lu et al.,
2020). In clinical contexts, music has been shown to reduce the need for both opioid and non-opioid medi-
cation (Lee, 2016). Similarly, in experimental contexts using tasks to induce pain, music can help to increase
pain tolerance and threshold (Lee, 2016; Lu et al., 2020). These meta-analyses also suggest that the effect of
music on pain has increased in the last 10 to 15 years, as we learn more about music interventions and imple-
ment them more carefully. Early meta-analysis indicated a small effect size (Cepeda et al., 2006), whereas
more recent meta-analyses provide evidence of medium effect sizes (Lee, 2016; Lu et al., 2020).
However, despite the strength of these potential effects, these effect sizes are not always observed in indi-
vidual studies (Cepeda et al., 2006; Lunde et al., 2019). This is partly due to a lack of consideration given
to the prospective neural and cognitive mechanisms of action underlying music interventions. In fact, most
music intervention studies in healthcare contexts fail to provide an intervention theory or other rationale
for the delivery of music (Robb et al., 2018), which reduces the likelihood that the potential benefits of
the intervention will be observed. To address this shortfall, researchers and practitioners are encouraged to
embrace the principles of neuroaesthetics to help understand how music interventions can be used success-
fully. The current focus in music intervention research moves beyond the basic sensory properties of music

413
Claire Howlin

Figure 21.1 The cognitive mechanisms identified in the literature and how they fit together.

and considers the wider aspects of music engagement that facilitate sustained attention (Garza-Villarreal
et al., 2017) and absorption (Garcia & Hand, 2016). Increasingly, it is recognized that cognitive (Bradshaw
et al., 2012; Howlin & Rooney, 2021) and emotional (Garza-Villarreal et al., 2017) interactions mediate the
analgesic benefits of music engagement (Garcia & Hand, 2016; Gold & Clare, 2013; Guétin, 2012).

Neural changes related to pain tolerance in response to sad music


Neuro-imaging studies provide new insights to help us understand how music listening can impact the
experience of pain in more detail. Functional magnetic resonance imaging has been used to identify the
neural changes that correspond to the analgesic effects of music listening. It seems that music that makes
us sad also makes us feel more pain. Initial fMRI studies that focussed on the impact of sad music on pain
experience found that sad music can increase the experience of pain (Coen, 2009; Wagner et al., 2009).
Coen and colleagues (2009) used a balloon inflated in the oesophagus to induce either a painful sensation or
a non-painful sensation in 12 male healthy participants while completing an fMRI on a 1.5-Tesla scanner.
They then exposed people to the same levels of internal balloon inflation while listening to music that was
rated as either sad or neutral by the participants. They identified that listening to sad music in the presence
of the non-painful stimulus led to an increase in perceived pain, to the degree that there was no longer a
difference in pain rating between the painful stimulus and the non-painful stimulus. Coen and colleagues
(2009) also identified that there were striking similarities in neural activity between the two experimental
conditions. The insula, inferior insula, and anterior cingulate cortex demonstrated increased activity during
both the painful sensation and the non-painful sensation in the presence of sad music compared to baseline
activity (Coen et al., 2009). This indicates that perceived music valence can influence the perception of

414
The health benefits of art experience

pain severity and suggested a degree of functional overlap between the areas of the brain that interpret pain
and music.
A very similar result was also found Wagner and colleagues (2009), who also identified that music par-
ticipants perceive as sad can make physical sensations seem more painful. Wagner and colleagues (2009) used
a thermal stimulus to induce pain in 40 healthy female participants during an fMRI on a 1.5-T scanner.
During the fMRI, participants listened to samples of their own music that they had indicated would make
them feel sad and samples of neutral music. Wagner and colleagues (2009) also found that sad music made
the same degree of thermal stimulation seem more painful. However, contrary to Coen and colleagues
(2009), Wagner and colleagues identified an increased BOLD response bilaterally in a cluster of ventrolateral
nuclei of the thalamus, an area known to be involved in processing both sensations of temperature and pain
(Sheridan & Tadi, 2019).
However, a more recent EEG study with 40 participants in total identified a sub-group of participants
that experience increased pain tolerance from listening to sad music (Guo et al., 2020). Guo and colleagues
(2020) used a cold water paradigm to measure pain tolerance while asking participants to listen to music that
was determined to be sad by professional musicians. They identified that half of the participants experienced
increases in pain tolerance in response to sad music and that this was related brain oscillations in a higher
beta band and the gamma band at the O2 and P4 electrodes. Unfortunately Guo and colleagues (2020) did
not ask participants to rate the music as sad themselves, which makes it difficult to determine if the observed
responses can be attributed to feelings of sadness. It is possible that this sub-group of participants could have
interpreted the music to be beautiful or moving, which could implicate processes of reward that are not
unique to the specific experience of sadness.

Identifying the neural network responsible for music analgesia


A more important question than how we make people feel more pain is: How do we make people feel less
pain? Clinical studies indicate that listening to your favourite music is the strongest predictor of music anal-
gesia (Lee, 2016). The first fMRI study to examine impact of favourite music on pain used a thermal heat
paradigm to induce pain in 12 healthy female participants (Dobek et al., 2014). Participants were exposed to
the thermal stimuli while listening to their favourite music and in silence. Dobek (2014) identified a wide
network of neural changes involved in mediating pain decreases. Specifically, they identified that music can
activate an area that used to be known as the “pain matrix” but is also referred to as the “saliency matrix”.
The areas of the non-specific saliency matrix activated by both music and pain include the insula, thala-
mus, and secondary somatosensory cortex. It seems that different salient stimuli, whether thermal heat pain
stimuli or auditory, can similarly activate certain brain regions. These regions are described as multimodal
areas, and their degree of activation is related to the saliency of the stimulus. Crucially, when music activated
the saliency matrix, it led to a different pattern of neural activity. When people were exposed to music in
the presence of pain, they experienced greater levels of BOLD activation in brain regions associated with
emotion and reward and in the dorsolateral prefrontal cortex and decreased activity in the dorsal horn of the
spinal cord. The dorsolateral prefrontal cortex (DLPFC) is involved in attention processes and is important
because of its opioid sensitivity, which can influence supraspinal networks (Taylor et al., 2012) and inhibit the
orbitofrontal cortex (Lorenz et al., 2003). Repetitive stimulation of the DLPFC can increase pain tolerance in
participants (Taylor et al., 2012), and music listening appears to achieve this. Decreased activity in the dorsal
horn is also crucial in terms of pain modulation, given that this is the initial region of the spinothalamic tract
that modulates pain.
Collectively these patterns of findings suggest that music listening creates a shift in neural processing via
the saliency matrix, where processes related to emotion and reward are upregulated and processes related to
attention contribute to inhibition of the spinothalamic tract. This study provided important insight into the

415
Claire Howlin

specific neural mechanisms involved in mediating the analgesic benefits of music listening. However, BOLD
activation responses only provide limited information about the precise nature of how different brain regions
interact when people experience analgesic benefits from music engagement.
Functional connectivity studies identify how specific areas of the brain interact in response to music-
mediated analgesia. Garza-Villareal and colleagues (2015) asked 22 patients with fibromyalgia to listen to
either music or a pink noise control for 5 minutes. Resting-state fMRIs were taken before and after patients
listened to the music. All patients provided pain ratings before and after they listened to the music, and the
music was chosen by the experimenter based on the individual’s preferences (Garza-Villarreal et al., 2015).
Garza-Villareal and colleagues identified noticeably higher activity in the left angular gyrus after music lis-
tening and compared to the pink noise condition. Crucially, the increased signal in the left angular gyrus
was related to decreases in pain rating (Garza-Villarreal et al., 2015). Further post-hoc connectivity analysis
identified that music listening led to increased connectivity between the left angular gyrus with the right
dorsolateral prefrontal cortex and the left caudate and decreased connectivity with the right anterior cingu-
late cortex, precuneus, right supplementary motor area, and right precentral gyrus. Changes in connectivity
between the left angular gyrus and the right precentral gyrus were directly related to pain intensity ratings
(Garza-Villarreal et al., 2015).
A follow-up study by the same research team then repeated the study with a healthy group of age-
matched controls (Pando-Naude et al., 2019). Similar to earlier studies, they found that music listening in
healthy controls primarily corresponded with connectivity changes in a region known as the “pain matrix”
or “saliency matrix” (Dobek et al., 2014). Healthy controls demonstrated lower resting-state functional con-
nectivity between the pain matrix and left precuneus, left superior frontal gyrus, right midfrontal gyrus, right
Pa CiG, and left PCC and increased resting state connectivity between the right insula when compared with
patients with fibromyalgia. Although the degree of activity in the pain matrix was much lower in healthy
controls compared to people with fibromyalgia, healthy controls tended to show an increase in functional
connectivity between the right amygdala and angular gyrus with superior lateral occipital areas in both hemi-
spheres (Pando-Naude et al., 2019). At first this might seem strange, given that the healthy controls were not
experiencing pain, but the regions of the pain matrix are not exclusive for pain processing and are involved
in other cognitive processes. Specifically, it seems that the healthy controls were recruiting areas overlapping
with the pain matrix to engage with visual imagery. It seems that music listening can recruit neural process-
ing areas in the pain matrix and redirect them to music processing tasks involved in emotional engagement
and visual imagery.
More recently another study examined the impact of music listening on functional connectivity in patients
with fibromyalgia (Usui et al., 2020). Usui and colleagues (2020) asked 23 female patients with fibromyalgia
to listen to 17 minutes of music composed by Mozart in a quiet room. Functional state resting fMRI images
were taken on a 3-T scanner before and after music listening. Similar to Garza-Villareal and colleagues
(2015), they found significant differences in connectivity within the default mode network and decreased
activity between the default mode network and the supplementary motor area. However, the specific pattern
of changes in functional connectivity within the default mode network was notably different than the pattern
of changes found by Garza-Villareal and colleagues (2015). Usui and colleagues (2020) found a significant
increase in connectivity between the right insular cortex and the posterior cingulate cortex, the precuneus,
and the parahippocampus. Conversely, significant decreases in functional connectivity were found between
the insular cortex and the frontal orbital cortex and the supplementary motor area. Increases in functional
connectivity between the insular cortex and the precuneus were related to decreases in pain scores. However,
despite these differences in the specific changes in connectivity, collectively these studies identify that music
analgesia occurs as a result of top-down regulation of the pain modulatory network by the default mode
network (Garza-Villarreal et al., 2015; Usui et al., 2020). In line with the initial findings by Dobek and

416
The health benefits of art experience

colleagues (2014), the saliency of music competes with pain signals and disrupts further pain processing due
to inhibition of the spinothalamic tract.

Neurobiological changes associated with art making

Art making and art therapy decrease biomarkers associated with stress
Art making leads to physiological and neural changes related to health outcomes. Emerging evidence
has started to identify neurobiological changes associated with mindfulness-based art therapy in terms of
decreases in blood cortisol (Walsh et al., 2007; Lawson et al., 2012; Kaimal et al., 2016; Beerse et al., 2019)
and changes in neural activity (Monti et al., 2012; Kruk et al., 2014; Kaimal et al., 2017). Changes in blood
cortisol levels can help to quantify the degree to which people experience physiological benefits from art
experiences and art interventions. Cortisol is a hormone that is used as a biomarker because it can be used
as a biological index of the physiological stress response. It is quite normal to have increases and decreases in
cortisol production as a result of everyday events. Prolonged dysregulation of cortisol, where there is either
too much or too little, can occur as a result of ongoing stress and can lead to a number of health issues includ-
ing anxiety and depression, heart disease, and weight gain. For this reason, measures of cortisol are used to
understand and quantify the impact of arts interventions at the biological level.
Kaimal and colleagues (2016) examined the impact of art making using modelling clay, collage materials,
or markers on salivary cortisol levels. They used a before and after quasi-experimental research design with
39 participants. Salivary cortisol samples were taken before and after the art-making session, and salivary
cortisol samples were blinded for analysis, so that the analyst did not know whether each sample was taken
before or after the art-making session. Kaimal and colleagues (2016) observed significant decreases in salivary
cortisol levels after art mindfulness-based art therapy sessions for approximately 75% of participants. Indi-
viduals who already had quite low levels of salivary cortisol did not see further decreases.
It is important to consider the details of the specific art experience that led to this significant decrease
in cortisol. Very few restraints were placed on the participants’ creative participation: they could com-
bine all three media if preferred, create any type of imagery, and were told that there was no expectation
that this would result in a final artwork. There was an art therapist present, but they did not give spe-
cific directives; instead, the art therapist provided assistance and interpersonal interactions as directed by
participants. In this regard, Kaimal and colleagues (2016) placed a focus on the art-making process rather
than the outcome. This emphasizes the importance of creating an environment where people are more
likely to be creative and have a meaningful creative experience rather than simply encouraging people to
produce objects. This may be why art classes don’t tend to elicit the statistically significant decreases in
salivary cortisol in caregivers of people with cancer (Walsh et al., 2007). Art classes may focus more on
implementing specific techniques with precision, to the detriment of facilitating personal expression and
creative processes.
Art making has also been shown to reduce salivary cortisol levels in college students with anxiety (Beerse
et al., 2020) and patients receiving blood and bone marrow transplants (Lawson et al., 2012). In a study
that compared mindfulness-based art therapy with a neutral clay-making task, both types of art making
were found to lead to reductions in salivary cortisol (Beerse et al., 2020). Both groups reported significant
decreases in self-reported outcomes related to anxiety, but these outcomes were not strongly related to
cortisol changes. Although the mindfulness-based art therapy group experienced greater decrease on self-
reported outcomes related to anxiety, this was also not directly related to the physiological changes observed.
This study helps to identify that the actual art-making process with clay can help to reduce salivary cortisol
levels, which is indicative of a decreased stress response. It also identifies that having an art therapist present

417
Claire Howlin

may help to provide additional clinical benefits related to decreases in anxiety, which may account for more
holistic benefits alongside neurobiological changes.
Real-life art experiences can also reduce biomarkers related to stress in terms of salivary cortisol (Clow &
Fredhoi, 2006) and blood pressure (Mastandrea et al., 2019). In an ecologically valid study, 29 office workers
were invited to attend the Guildhall Art Gallery in London over their lunch break (Clow & Fredhoi, 2006).
Participants were free to explore the gallery in any way they pleased for a period of 35–40 minutes and
provided salivary cortisol samples along with self-report measures of stress at the beginning and end of their
gallery attendance. Cortisol levels dropped dramatically during the visit to the gallery, and this was directly
related to decreases in self-reported levels of stress. This study provides a strong example of the potential for
real-life art experiences to help reduce stress, without the need for any specific directives or instructions for
the individuals. This finding is supported by another study that examined the impact of attending an art gal-
lery on cardiovascular measures associated with stress (Mastandrea et al., 2019). Seventy-seven participants
were assigned to view figurative artworks or modern artworks or visit the office space in the gallery as a
control condition. Participants viewed the artworks for five minutes and had their blood pressure and heart
rate measured before and after viewing the artworks. Mastandrea and colleagues (2019) found that the figu-
rative art condition was the only condition that led to a significant decrease in blood pressure. There was no
decrease in blood pressure observed for the modern art condition. This was despite the fact that there was no
difference in liking ratings between the modern art condition and the figurative art condition. Additionally,
there was no correlations identified between aesthetic appreciation and blood pressure for any of the three
groups. This indicates that although attendance at an art gallery can lead to decreases in blood pressure, this
is not necessarily guaranteed and may in fact be related to the type of art that you view. Different pieces of
art and art styles are likely to evoke very different types of responses, and it’s possible that not every type of
response will be directly conducive to well-being.

Neural changes associated with art making


Neuroimaging research has begun to explore the neural changes elicited by art making and art therapies.
Preliminary evidence suggests that decreases in anxiety reported after mindfulness-based art therapy are
related to increased levels of blood flow in the caudate, insula, and amygdala (Monti et al., 2012). In a study
with 18 participants, 8 people who completed 8 weeks of mindfulness-based therapy were compared with
10 participants in a control group receiving educational support. Participants were imaged 2 weeks before
the programme and 2 weeks after the end of the 8-week programme. The programme itself consisted of
weekly group meetings of 2.5 hours where participants engaged in a range of expressive art tasks with differ-
ent instructions from the art therapist according to a pre-specified curriculum. Participants who completed
the programme reported significant decreases in anxiety scores that were strongly correlated with increases in
cerebral blood flow in the left caudate, whereas participants in the control condition did not show significant
changes in blood flow in the same region, despite the fact that they also reported significant decreases in
anxiety scores after the control intervention. Although this finding has yet to be replicated, it does provide
evidence of additional neural activity beyond what is accounted for using self-report methods.
To date very few studies have examined how neural changes related to art activities correspond with
specific clinical outcomes. However, several studies have identified specific neural changes associated in self-
expressive (Chamberlain et al., 2014; Kaimal et al., 2017) creative tasks similar to those used mindfulness-
based art therapy studies. The neural reward network has been shown to be more active in drawing, painting,
and doodling (Kaimal et al., 2017). Quantitative electroencephalograms have also started to characterize the
neural changes related to art making (Belkofer et al., 2014; Kruk et al., 2014). Clay making and drawing have
both shown to lead to increases in theta power in the right medial parietal lobe (Kruk et al., 2014). A slightly
different result was identified in a study that used oil pastels with 10 participants. Belkofer and colleagues

418
The health benefits of art experience

(2014) identified that that using oil pastels led to an increase in alpha band activity in the temporooccipital
region. However, given that both studies used relatively small sample sizes without any control groups with
non-art activities, it is difficult to be conclusive about these results at this stage. It also must be acknowledged
that neither study directly measured health or well-being outcomes. Although it seems plausible that neural
changes observed in relation to art activities may help to account for improved health outcomes, it is not
possible to be definite in this regard without more direct evidence of how neural changes relate to clinical
outcomes. As our understanding of neural changes in relation to art and creative activities increases overall,
perhaps more attention will be paid to how these changes in neural activity relate to changes in health and
well-being outcomes. The next logical step would appear to be to explore the degree to which neural activ-
ity associated with art experiences can account for the well-being effects identified in clinical studies.

Future directions

Suggestions for future research and practice


Previous attempts to synthesize the results of studies on arts experiences and interventions tend to be frus-
trated by a lack of scientific rigour. For example, although many studies have examined the role of dance
movement therapy (Karkou & Meekums, 2017) and music therapy on dementia (Fusar-Poli et al., 2018),
there is still no conclusive evidence that supports the introduction of these studies into clinical contexts.
Small sample sizes often lead to inconclusive results and frustrate attempts to make sub-group comparisons.
Similarly, suitable controls need to be introduced so that if health benefits are identified, they can be attrib-
uted to the intervention. Trials of high methodological quality, large sample sizes, and clarity in the way the
intervention is put together and delivered are needed to assess whether different types of arts experiences can
lead to improved health and well-being outcomes (Karkou & Meekums, 2017).
It seems plausible that individual differences in terms of art experience could contribute to or detract from
the effectiveness of arts interventions. Future studies should use specific instruments designed to account for
different types of art experience that demonstrate good statistical reliability and construct validity. This would
help to gain a more granular insight into how previous art experience may impact health and well-being out-
comes. It would also help to identify who will be more likely to benefit from different types of interventions
and may be used as a rationale to introduce or modify different interventions in different contexts.

Embrace interdisciplinary approaches


There is currently a dichotomy in the arts and health literature between arts interventions designed by arts
practitioners and arts interventions designed by medical professionals. This poses a challenge, since each
group prioritizes different aspects of the creative program, in line with their own practice, often at the
expense of other aspects of the intervention. It is often the case that individuals with a background in art
will provide high-quality art experiences that are likely integral to the overall success of the intervention
yet will have poorly defined health outcomes that make it difficult to quantify the overall effects. Similarly,
individuals from a health or medical background tend to excel at defining and measuring the health outcome
that they wish to improve but often fail to preserve a full art experience in the intervention, leading to a
reductionistic intervention with limited success. This is exemplified in a recent systematic review of 75 music
and pain interventions that identified that only 1 of the included 75 articles was actually designed to evaluate
or measure the role of aesthetic engagement in mediating the analgesic effect of music listening (Howlin &
Rooney, 2020). Furthermore, a number of music therapy studies report that the music experience may be
interrupted to allow for medical examinations, and perhaps unsurprisingly, the effectiveness of the interven-
tions is found to be reduced. This suggests that when the integrity of the art experience is disrupted, we do

419
Claire Howlin

not see the same degree of health and well-being benefits. In order to maximize the benefits of arts interven-
tions, it is important to acknowledge that a successful arts intervention will require a meaningful interaction
between the person and the art experience.
In order to help distinguish arts interventions from arts experiences, arts intervention practitioners should
provide clear descriptions about the nature of the arts activity involved, how it has been designed or curated,
and the specific health benefits expected. Making this distinction will help to create a common taxonomy
that can be used to describe and compare different activities and support greater specificity in how peo-
ple use the arts to improve health and well-being. This does not mean that arts intervention practitioners
should not encourage people to interact directly with everyday arts experiences. As with many other health
habits, a “little and often” approach can often lead to the greatest long-term benefits. For example, regular
exercise and fresh green vegetables will always be better than extreme fitness regimes or crash diets. In line
with this, arts intervention practitioners should consider their role as a source of encouragement to help
people to integrate arts experiences into their daily routines. Even when financial barriers are removed,
many people will still struggle to attend events, concerts, and galleries simply because they lack motivation
and encouragement.

References
Arruda, M. A. L. B., Garcia, M. A., & Garcia, J. B. S. (2016). Evaluation of the effects of music and poetry in oncologic
pain relief: A randomized clinical trial. Journal of Palliative Medicine, 19, 943–948.
Basiński, K., Zdun-Ryżewska, A., Greenberg, D. M., & Majkowicz, M. (2021). Preferred musical attribute dimensions
underlie individual differences in music-induced analgesia.  Scientific Reports,  11(1), 1–8. https://doi.org/10.1038/
s41598-021-87943-z
Beerse, M. E., Van Lith, T., & Stanwood, G. D. (2019). Is there a biofeedback response to art therapy? A technology-
assisted approach for reducing anxiety and stress in college students.  SAGE Open,  9(2). https://doi.org/10.1177/
2158244019854646
Beerse, M. E., Van Lith, T., & Stanwood, G. D. (2020). Therapeutic psychological and biological responses to mindful-
ness-based art therapy. Stress and Health, 36(4), 419–432. https://doi.org/10.1002/smi.2937
Belkofer, C. M., Van Hecke, A. V.,  & Konopka, L. M. (2014). Effects of drawing on alpha activity: A  quantitative
EEG study with implications for art therapy. Art Therapy, 31(2), 61–68. https://doi.org/10.1080/07421656.2014.
903821
Bradshaw, D. H., Chapman, C. R., Jacobson, R. C., & Donaldson, G. W. (2012). Effects of music engagement on responses
to painful stimulation. Clinical Journal of Pain, 28(5), 418–427. https://doi.org/10.1097/AJP.0b013e318236c8ca
Bradt, J., Dileo, C., Magill, L., & Teague, A. (2016). Music interventions for improving psychological and physical out-
comes in cancer patients. Cochrane Database of Systematic Reviews, 8(8), CD006911. https://doi.org/10.1002/14651858.
CD006911.pub3
Bradt, J., Potvin, N., Kesslick, A., Shim, M., Radl, D., Schriver, E., . . . & Komarnicky-Kocher, L. T. (2015). The impact
of music therapy versus music medicine on psychological outcomes and pain in cancer patients: A mixed methods
study. Supportive Care in Cancer, 23, 1261–1271.
Brattico, E. (2015). From pleasure to liking and back: Bottom-up and top-down neural routes to the aesthetic enjoyment
of music. In J. P. Huston, M. Nadal, F. Mora, L. F. Agnati, & C. J. Cela-Conde (Eds.), Art, aesthetics and the brain
(pp. 303–318). New York: Oxford University Press.
Brattico, E., & Pearce, M. (2013). The neuroaesthetics of music. Psychology of Aesthetics, Creativity, and the Arts, 7(1),
48–61. https://doi.org/10.1037/a0031624
Cepeda, M. S., Carr, D. B., Lau, J., & Alvarez, H. (2006). Music for pain relief. Cochrane Database of Systematic Reviews,
19(2), CD004843. https://doi.org/10.1002/14651858.CD004843.pub2
Chamberlain, R., McManus, I. C., Brunswick, N., Rankin, Q., Riley, H., & Kanai, R. (2014). Drawing on the right
side of the brain: A voxel-based morphometry analysis of observational drawing. NeuroImage, 96, 167–173. https://
doi.org/10.1016/j.neuroimage.2014.03.062
Coen, S. J., Yágüez, L., Aziz, Q., Mitterschiffthaler, M. T., Brammer, M., Williams, S. C., & Gregory, L. J. (2009). Nega-
tive mood affects brain processing of visceral sensation. Gastroenterology, 137, 253–261.
Clow, A., & Fredhoi, C. (2006). Normalisation of salivary cortisol levels and self-report stress by a brief lunchtime visit
to an art gallery by London City workers. Journal of Holistic Healthcare, 3(2), 29–32.

420
The health benefits of art experience

de Witte, M., Spruit, A., van Hooren, S., Moonen, X., & Stams, G. J. (2020). Effects of music interventions on stress-
related outcomes: A systematic review and two meta-analyses. Health Psychology Review, 14(2), 294–324. https://doi.
org/10.1080/17437199.2019.1627897
Deshmukh, S. R., Holmes, J., & Cardno, A. (2018). Art therapy for people with dementia. Cochrane Database of Systematic
Reviews (9), CD011073. https://doi.org/10.1002/14651858.CD011073.pub2
Dobek, C. E., Beynon, M. E., Bosma, R. L., & Stroman, P. W. (2014). Music modulation of pain perception and pain-
related activity in the brain, brain stem, and spinal cord: A functional magnetic resonance imaging study. The Journal
of Pain, 15, 1057–1068.
Fusar-Poli, L., Bieleninik, Ł., Brondino, N., Chen, X. J., & Gold, C. (2018). The effect of music therapy on cognitive
functions in patients with dementia: A systematic review and meta-analysis. Aging & Mental Health, 22, 1103–1112.
Garcia, R. L., & Hand, C. J. (2016). Analgesic effects of self-chosen music type on cold pressor-induced pain: Motivating
vs. relaxing music. Psychology of Music, 44(5), 967–983. https://doi.org/10.1177/0305735615602144
Garza-Villarreal, E. A., Jiang, Z., Vuust, P., Alcauter, S., Vase, L., Pasaye, E., . . . Barrios. (2015). Music reduces pain
and increases resting state fMRI BOLD signal amplitude in the left angular gyrus in fibromyalgia patients. Frontiers in
Psychology, 6, 1051. https://doi.org/10.3389/fpsyg.2015.01051
Garza-Villarreal, E. A., Pando, V., Vuust, P., & Parsons, C. (2017). Music-induced analgesia in chronic pain conditions:
A systematic review and meta-analysis. Pain Physician, 20, 597–610.
Gold, A., & Clare, A. (2013). An exploration of music listening in chronic pain. Psychology of Music, 41(5), 545–564.
https://doi.org/10.1177/0305735612440613
Gross, J. J. (1998). The emerging field of emotion regulation: An integrative review. Review of General Psychology, 2(3),
271–299. https://doi.org/10.1037/1089-2680.2.3.271
Guétin, S., Giniès, P., Siou, D. K. A., Picot, M. C., Pommié, C., Guldner, E., . . . Touchon. (2012). The effects of
music intervention in the management of chronic pain: A single-blind, randomized, controlled trial. Clinical Journal
of Pain, 28(4), 329–337. https://doi.org/10.1097/AJP.0b013e31822be973
Guo, S., Lu, J., Wang, Y., Li, Y., Huang, B., Zhang, Y., . . . Xia, Y. (2020). Sad music modulates pain perception: An
EEG study. Journal of Pain Research, 13, 2003. https://doi.org/10.2147/JPR.S264188
Howlin, C., & Rooney, B. (2020). The cognitive mechanisms in music listening interventions for pain. A scoping review.
Journal of Music Therapy, 57(2), 127–167. https://doi.org/10.1093/jmt/thaa003
Howlin, C., & Rooney, B. (2021). Cognitive agency in music interventions: Increased perceived control of music pre-
dicts increased pain tolerance. European Journal of Pain, 25(8), 1712–1722. https://doi.org/10.1002/ejp.1780
Howlin, C., Walsh, R., D’Alton, P., & Rooney, B. (in review). Validating the cognitive vitality model of music with a
chronic pain population.
Juslin, P. N. (2013). From everyday emotions to aesthetic emotions: Towards a unified theory of musical
emotions. Physics of Life Reviews, 10(3), 235–266. https://doi.org/10.1016/j.plrev.2013.05.008
Juslin, P. N., Harmat, L., & Eerola, T. (2013). What makes music emotionally significant? Exploring the underlying
mechanisms. Psychology of Music, 42(4), 599–623. https://doi.org/10.1177/0305735613484548
Juslin, P. N., & Västfjäll, D. (2008). Emotional responses to music: The need to consider underlying mechanisms. Behav-
ioral and Brain Sciences, 31, 559–575. https://doi.org/10.1017/S0140525X08005293
Kaimal, G., Ayaz, H., Herres, J., Dieterich-Hartwell, R., Makwana, B., Kaiser, D. H., & Nasser, J. A. (2017). Functional
near-infrared spectroscopy assessment of reward perception based on visual self-expression: Coloring, doodling, and
free drawing. Arts in Psychotherapy, 55, 85–92. https://doi.org/10.1016/j.aip.2017.05.004
Kaimal, G., Ray, K., & Muniz, J. (2016). Reduction of cortisol levels and participants’ responses following art mak-
ing. Art Therapy, 33(2), 74–80. https://doi.org/10.1080/07421656.2016.1166832
Karkou, V., & Meekums, B. (2017). Dance movement therapy for dementia. Cochrane Database of Systematic Reviews, 2,
CD011022. https://doi.org/10.1002%2F14651858.CD011022.pub2
Koelsch, S. (2012). Brain and music. Wiley-Blackwell.
Kruk, K. A., Aravich, P. F., Deaver, S. P., & deBeus, R. (2014). Comparison of brain activity during drawing and clay
sculpting: A preliminary qEEG study. Art Therapy, 31(2), 52–60. https://doi.org/10.1080/07421656.2014.903826
Lawson, L. M., Williams, P., Glennon, C., Carithers, K., Schnabel, E., Andrejack, A., & Wright, N. (2012). Effect of art
making on cancer-related symptoms of blood and marrow transplantation recipients. Oncology Nursing Forum, 39(4),
E353–E360. https://doi.org/10.1188/12.ONF.E353-E360
Lee, J. H. (2016). The effects of music on pain. Journal of Music Therapy Advance Access, 1–48.
Lorenz, J., Minoshima, S., & Casey, K. L. (2003). Keeping pain out of mind: The role of the dorsolateral prefrontal cortex
in pain modulation. Brain, 126(5), 1079–1091. https://doi.org/10.1093/brain/awg102
Lu, X., Yi, F.,  & Hu, L. (2020). Music-induced analgesia: An adjunct to pain management.  Psychology of Music.
0305735620928585.

421
Claire Howlin

Lunde, S. J., Vuust, P., Garza-Villarreal, E. A., & Vase, L. (2019). Music-induced analgesia: How does music relieve pain?
Pain, 160, 989–993.
Mastandrea, S., Maricchiolo, F., Carrus, G., Giovannelli, I., Giuliani, V., & Berardi, D. (2019). Visits to figurative art
museums may lower blood pressure and stress. Arts and Health, 11(2), 123–132. https://doi.org/10.1080/17533015
.2018.1443953
Mitchell, L. A., MacDonald, R. A., & Brodie, E. E. (2006). A comparison of the effects of preferred music, arithmetic and
humour on cold pressor pain. European Journal of Pain, 10(4), 343–351. https://doi.org/10.1016/j.ejpain.2005.03.005
Monti, D. A., Kash, K. M., Kunkel, E. J., Brainard, G., Wintering, N., Moss, A. S., . . . Newberg, A. B. (2012). Changes
in cerebral blood flow and anxiety associated with an 8-week mindfulness programme in women with breast can-
cer. Stress and Health, 28(5), 397–407. https://doi.org/10.1002/smi.2470
Nemoto, I., Fujimaki, T., & Wang, L. Q. (2010). fMRI measurement of brain activities to major and minor chords and
cadence sequences. In Annual international conference of the IEEE engineering in medicine and biology, 5640–5643. https://
doi.org/10.1109/IEMBS.2010.5628044
Pando-Naude, V., Barrios, F. A., Alcauter, S., Pasaye, E. H., Vase, L., Brattico, E., . . . Garza-Villarreal, E. A. (2019). Func-
tional connectivity of music-induced analgesia in fibromyalgia. Scientific Reports, 9(1), 1–17. https://doi.org/10.1038/
s41598-018-37186-2
Robb, S. L., Hanson-Abromeit, D., May, L., Hernandez-Ruiz, E., Allison, M., Beloat, A., .  .  .  & Wolf, E. (2018).
Reporting quality of music intervention research in healthcare: A  systematic review. Complementary Therapies in
Medicine, 38, 24–41.
Ruddy, R., & Dent-Brown, K. (2007). Drama therapy for schizophrenia or schizophrenia-like illnesses. Cochrane Database
of Systematic Reviews (1). Art. no.: CD005378, September 3, 2021. https://doi.org/10.1002/14651858.CD005378.
pub2
Ruddy, R., & Milnes, D. (2005). Art therapy for schizophrenia or schizophrenia-like illnesses. Cochrane Database of Sys-
tematic Reviews, 2005(4), Art. no.: CD003728, September 3, 2021. https://doi.org/10.1002/14651858.CD003728.
pub2
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Sheridan, N., & Tadi, P. (2019). Neuroanatomy. Thalamic Nuclei.
Tang, Q., Huang, Z., Zhou, H., & Ye, P. (2020). Effects of music therapy on depression: A meta-analysis of randomized
controlled trials. PLoS One, 15(11), e0240862. https://doi.org/10.1371/journal.pone.0240862
Taylor, J. J., Borckardt, J. J., & George, M. S. (2012). Endogenous opioids mediate left dorsolateral prefrontal cortex
rTMS-induced analgesia. Pain, 153(6), 1219–1225. https://doi.org/10.1016/j.pain.2012.02.030
Tsai, H. F., Chen, Y. R., Chung, M. H., Liao, Y. M., Chi, M. J., Chang, C. C., & Chou, K. R. (2014). Effectiveness
of music intervention in ameliorating cancer patients’ anxiety, depression, pain, and fatigue: A meta-analysis. Cancer
Nursing, 37, E35–E50.
Usui, C., Kirino, E., Tanaka, S., Inami, R., Nishioka, K., Hatta, K., . . . Inoue. (2020). Music intervention reduces
persistent fibromyalgia pain and alters functional connectivity between the insula and default mode network. Pain
Medicine, 21(8), 1546–1552. https://doi.org/10.1093/pm/pnaa071
Wagner, G., Koschke, M., Leuf, T., Schlösser, R., & Bär, K. J. (2009). Reduced heat pain thresholds after sad-mood
induction are associated with changes in thalamic activity. Neuropsychologia, 47, 980–987.
Walsh, S. M., Radcliffe, R. S., Castillo, L. C., Kumar, A. M., & Broschard, D. M. (2007). Oncology Nursing Forum, 34(1),
E10, 38. https://doi.org/10.1188/07.ONF.E9-E16
World Health Organization. (2019). What is the evidence on the role of the arts in improving health and well-being? A scoping
review. World Health Organization. Regional Office for Europe.
Zhao, H., & Chen, A. C. (2009). Both happy and sad melodies modulate tonic human heat pain. Journal of Pain, 10(9),
953–960. https://doi.org/10.1016/j.jpain.2009.03.006

422
22
EXPERIENCING ART IN MUSEUMS
Aniko Illes and Pablo P. L. Tinio

More than a century’s worth of psychological aesthetics research has led to a great body of knowledge about
how people experience art. However, many unanswered questions remain. For example, what makes the act
of looking at an object an art experience, and, especially relevant to this chapter, what distinguishes experi-
encing art in a museum from other settings? What is it about the museum space and the manner in which
artworks are presented there that leads to intense, meaningful, and often pleasurable interactions with the
artworks? This chapter focuses on questions like these and describes research that has tried to answer them.
The focus of our discussions will be primarily on art museums and on visual art, which has received the most
attention from aesthetics researchers during its long history as a distinct field. We will draw on research that
has used behavioral, physiological, and neuroscientific methodologies to help increase our understanding of
what happens during the experience of art in museums.
Even with the omnipresence of technology in our society and the ever-increasing digitization of texts,
sounds, images, and other human artifacts, museums—as a space for collecting, preserving, interpreting, and
presenting works of art and other objects—remain relevant. In fact, relevant is a conservative term, given that
the last decade has seen a rise in museum attendance all over the world (Brieber, Nadal & Leder, 2015).
Museums, along with performance venues, libraries, and other cultural institutions, serve as centers for
the preservation and communication of culture and spaces for social, cultural, and intellectual growth and
transformation (Carr, 2003; Smith, 2014). In the following, we discuss the psychology of art and aesthetics
within the context of museums and show why museums are fertile ground for scientific research on human
behavior, especially those related to the experience of art.

The museum as a context for aesthetics research


For many people, there is a traditional view of what characterizes a typical museum. In this view, the
museum is a dedicated space for art, often housed in a grand, ornate building that is located in the busiest
and often most central location in a city or town. The museum has a collection of “high art”—paintings,
sculptures, and drawings by famous artists or up-and-coming and recently “discovered” talent. Inside the
museum, there is an air of reverence of the space, the art, and the activity of looking at art. Visitors give the
artworks in the museum’s galleries the space that they deserve, and any proximity to the art that is physi-
cally “too close” is immediately flagged by museum guards or loud alarms triggered by hidden sensors. As
visitors stroll around the museum halls, they speak in hushed voices, and their encounters with the art are

DOI: 10.4324/9781003008675-24 423


Aniko Illes and Pablo P. L. Tinio

marked by self-referential, reflective, and even transformative experiences (Carr, 2003). In this traditional
view, museum visitors themselves also fit a somewhat similar profile. They tend to be educated, and many are
repeat museum visitors who are comfortable with how museums operate, who know what the expectations
are as far as their behavior in the museum and who have some background knowledge about art that helps
them to make sense of the various artworks they encounter.
The traditional view of the museum is arguably outdated (Gürel & Nielsen, 2019). Museums have under-
gone, and continue to undergo, significant changes regarding their day-to-day operations as well as how they
engage with their audience and the programs that they offer to the public (Kotler et al., 2008). Examples of
these changes include presenting artworks from much wider groups of artists beyond the mainstream and
from a much broader range of types of art (e.g., performance, installation, and conceptual art), opening up
the museum to a wider audience beyond the highly educated and financially well off and arranging the pres-
entation of art in a way that not only engages and challenges but also educates their audience (e.g., Zhang
et al., 2018). For example, the Museum of Modern Art’s renewed mission states that

The Museum is dedicated to its role as an educational institution and provides a complete program
of activities intended to assist both the general public and special segments of the community in
approaching and understanding the world of modern and contemporary art.
(The Museum of Modern Art, n.d.)

Perhaps the most pronounced change that museums around the world have undergone during the last
decade is their role in educating their audience and engaging with the local community of which they are
a part (Kotler et al., 2008). Almost all museums nowadays have education departments that not only have a
say in how the museum operates but also how these institutions’ missions, strategic development goals, and
funding are administered. The focus of museums today has therefore started to shift from being socially and
economically privileged places to more open institutions that welcome patrons who seek aesthetic experi-
ences and are eager to learn about their cultural heritage and the cultural artifacts that reflect the world and
society in which they live.
The changes in museum practices as well as the issues associated with them are highly significant to the
work of empirical researchers trying to understand how people experience art in museum contexts (Tinio
et al., 2014). The key research questions are no longer just what types of art people like, prefer, or find
attractive but also what motivates people to visit museums; what goals and expectations they might have
for their visit; what the effects of the physical characteristics of the museum itself are on their visit; how
information impacts their visit; and what emotional, personally transformative, and educational outcomes
result from not only looking at specific works of art but also experiencing the totality of the museum
experience.
These key questions underscore the shift in the focus of museums from a relatively institution- and
collection-focused approach to a more open and visitor-centered approach and the significance of this shift
on aesthetics research. The shift has led to greater interest in how people experience art, in general and in
the museum experience in particular. Cognitive, emotional, behavioral, and social outcomes are now at the
forefront when considering the mission’s place in public life and discourse. Falk (2009) described this shift in
the museum world’s perspective:

For many in the museum community this is all we need to know—visitors come to see the exhibi-
tions, they see the exhibitions, and they leave knowing about the exhibitions. However, years of
visitor research has shown that there’s more subtlety to the issue; we need to know something about
the visitor and the conditions of visit as well.
(pp. 18–19)

424
Experiencing art in museums

For aesthetics researchers, this means that, more than ever, there are opportunities to venture out of the
lab and to conduct research in museums—the context that is most often associated with the art experience
(Pelowski et al., 2017; Tinio & Specker, 2020). In this chapter, we examine these issues from the perspective
of what Tinio (2019b, 2020) has termed as the aesthetics triumvirate, or the interrelated effects of three com-
ponents of the aesthetic experience of art: the art object, the perceiver, and the context in which the experi-
ence takes place, with a particular emphasis on subtopics most relevant to experiencing art in museums. We
then relate these discussions to what we believe neuroaesthetics can contribute to our understanding of the
experience of art in museums.

The museum visitor and the museum visit


The current trend towards engaging a wider audience means that museums must now consider all types of
visitors, at least much more so than previously. Research has consistently demonstrated that the nature of the
museum experience is greatly impacted by personal characteristics such as how much knowledge or interest
a visitor has in the art that they are viewing (Specker, Stamkou et al., 2020), the frequency of museum visita-
tion (Smith & Smith, 2001), and the motives and expectations that visitors bring with them when entering
museums (Pekarik et al., 1999; Pekarik & Schreiber, 2012). Of course, the extent to which these factors
influence the museum experience may depend on whether the museum has a traditional and mainstream
collection of works, such as Renaissance and Impressionist paintings, or contemporary or more obscure
works, such as multimedia installation, conceptual, or interactive performance art. However, there are certain
factors that are known to be highly influential to the experience of art, in general.

Art knowledge
Art knowledge is one factor known to highly influence the experience of art. In empirical aesthetics research,
art knowledge is often determined by the extent to which a person has received formal training or school-
ing in topics related to art (art history or studio classes, professional work in an art-related field) or informal
exposure to art (frequency of museum visitation, being raised in a home that emphasized art experiences).
The amount of experience of, and knowledge about, art that people bring with them when they enter a
museum will affect their level of engagement with art and if they decide to come back to the museum or
recommend a particular exhibition to others (e.g., Daskalaki et al., 2020).
Art knowledge has clear effects on the aesthetic experience, and this applies to the most fundamental
behavior involved in such experience: the movement of one’s eyes around a picture frame. How one goes
about looking at an artwork is affected by how much knowledge one has about the work. For example, those
with a lot of knowledge about a painting will tend to look more globally at the canvas, taking in not only the
elements being depicted in the image but also how they are being depicted—how do individual elements in
the image relate to each other visually, where are the elements placed in terms of the entire canvas, and how is
the use of the elements consistent or inconsistent with artistic conventions of the time, tradition, or art style to
which the painting belongs? Therefore, for the highly knowledgeable, content and composition are observed,
analyzed, and often judged (e.g., Hekkert & Wieringen, 1996; Kozbelt & Seeley, 2007). A different picture
emerges when someone with less art knowledge looks at an artwork, because for this person, looking behav-
ior becomes focused on what objects are being depicted and on the central region of the canvas (e.g., Francuz
et al., 2018; Massaro et al., 2012). In other words, content and center are prioritized over convention and
composition (e.g., Pihko et al., 2011; Vogt & Magnussen, 2007). Because what is depicted is the primary
focus for a person with less art knowledge, artworks that do not depict anything recognizable, or that depict
things in unusual ways—such as purely abstract works like Cy Twombly paintings and Surrealist art—become
more difficult to look at or less preferred than works in which the content is obvious (e.g., Rawlings, 2003).

425
Aniko Illes and Pablo P. L. Tinio

The previous is just one of many examples of how art knowledge directly affects looking behavior.
From the perspective of the museum, taking into account the knowledge levels of their visitors is of utmost
importance because knowledge facilitates engagement with art. Regardless of the level of art knowledge,
the fundamental activities that museum visitors are preoccupied with involve meaning making, understand-
ing, problem solving, and evaluating artworks. If these cognitive processes are to transpire, visitors must stop
at, attend to, and think about the works on display. It is therefore imperative that museums know, or suf-
ficiently estimate, just what their visitors know and do not know so that they may develop effective means
for supporting the information needs of the visitors and in turn promoting engagement with the artworks.
During the past decade, museums have placed ever-increasing emphasis on ascertaining the amount and
types of knowledge their visitors bring with them when they enter the museum (Pekarik & Schreiber, 2012;
Cairns, 2013).
Asking someone questions about their formal or informal experiences with art is one way to ascertain
visitors’ knowledge. However, what such an approach does not take into account very well is the level of
knowledge that these experiences have actually imparted in the individual. Smith and Smith (2006), work-
ing out of the Metropolitan Museum of Art’s Office of Research and Evaluation, suggested a more direct
means for determining level of art knowledge through their Aesthetic Fluency Scale (AFS). The AFS is easy
to administer and score and involves presenting the following art concepts or artists to visitors or research
participants: Impressionism, Abstract Expressionism, John Singer Sargent, Mary Cassatt, Alessandro Boticelli,
Fauvism, Egyptian Funerary Stelae, Isamu Noguchi, Gian Lorenzo Bernini, and Chinese Scrolls. Participants
then use a 5-point scale to rate how well they know each art concept or artist. The ratings are summed to
derive an overall aesthetic fluency score. Smith and Smith found that the AFS correlated highly with fre-
quency of museum visitation and training in art or art history. Yet unlike asking people how many art classes
they have taken, the AFS ascertains specific knowledge that presumably should be gained from experiences
with art. Smith and Smith liken the development of art knowledge to vocabulary development, which takes
place incrementally and supported by linguistic experiences such as reading. The more one reads, the big-
ger one’s vocabulary; the more one experiences art, the higher one’s aesthetic fluency. The AFS is arguably
the most commonly used measure of art-related knowledge in the field of empirical aesthetics. However, it
has become clear in recent years that factors closely related to art knowledge, like art interest, should also be
taken into account. Researchers such as Specker, Forster, and colleagues (2020) have recently developed the
Vienna Art Interest & Art Knowledge Questionnaire, which seems promising and should prove a useful tool
for museum researchers.

Motivations and goals


Art knowledge could be thought of as the foundation for how one experiences art in museums. Its devel-
opmental nature (Smith & Smith, 2006) means that it takes time for this knowledge to accumulate—likely
weeks and months, but more likely many years. As with other types of knowledge, art knowledge is stable
over time: you might forget some details, but you tend to remember larger topics and themes (e.g., art
movements, famous artworks and artists, and categories of art media). As visitors enter the museum knowing
certain things, they also enter with specific motivations for why they came to the museum to begin with and
expectations regarding their museum visit (Zhang et al., 2018). Just as general art knowledge serves as the
foundation for how visitors engage with art, motivation serves as the foundation for how visitors structure
their museum visit, what artworks they focus on, and how they go about interacting with the artworks and
other aspects of the museum (Tinio, 2017).
Motivations are critical because they underlie specific goals that visitors might have for their visit to the
museum. Examples of goals include to socialize with companions and other visitors, to learn something new
or exciting, or simply for leisure or to pass time. Goals drive behavior. Art enthusiasts who visit a museum to

426
Experiencing art in museums

see a new exhibit on Abstract Expressionism, a style of art that they love and have been studying their entire
adult lives, will have a strong desire to learn about the works on display and about the artists that created
them. The physical path that they take in the museum will be goal-directed and specific. However, goals for
a visit could also be non-specific or implicit in that visitors are not aware of exactly what they would like to
do during a visit (Illes & Bodor, 2016, 2018): A tourist walking by a museum that they have read about in
the past might enter the museum just to see what it is like and what artworks are on display. The goal for
this person’s visit is less specific and not as pre-determined as the previous visitor who entered the museum
to learn about Abstract Expressionist art. One would expect the nature of the visits of these two visitors to
be very different because their goals are very different.
Motivations and goals drive visitor behavior and lead to expectations for how the museum visit will tran-
spire. This is why it is critical for the museum to determine the goals and motivations of their visitors so that
they can create a context in which their expectations are met or purposefully challenged. Expectations have
a deep impact on the aesthetic experience and might be as important to the museum experience as the types
of artworks shown and how they are shown (e.g., Pekarik et al., 1999; Pekarik & Schreiber, 2012).

The museum and authenticity of art objects


The characteristics of the museum visitor are a critical component of the aesthetics triumvirate, as are the
artworks that are on display. However, the effects of each of these components on the aesthetic experience
are moderated by the context in which artworks are being viewed. But context does not just have a moder-
ating effect. It presents specific conditions that directly impact the aesthetic experience, a point that will be
addressed in the following section.
Although the museum context is one of the first things that might come to mind when thinking about
the experience of art, it is but one of many contexts for engaging with art—art books, art in public spaces,
art displayed at home and in workplaces, and of course, as has become clear in recent years, art that is viewed
online and in virtual spaces. The question that arises is: What makes the viewing of art in a museum or
gallery particularly distinct from these other contexts and formats? What makes looking at a painting or a
sculpture or a video projection in a formal, dedicated space special (and is it really, truly special)? Finally, what
features of the museum context are most impactful to the experiences of museum visitors?
Let us consider what a typical museum visit is like. With the exception of the museum visit that occurs
by happenstance, as in the previous example of a tourist visiting a new city who happens to just wander
into a museum, museum visits are usually occasions that are planned (at times highly so). Museumgoers are
committed the minute they begin their commute to their destination. The trip to and from the museum
as well as time spent there may require an entire morning or afternoon to be set aside. Once they enter the
museum, visitors are subjected to an inspection of their bags and other belongings, a practice that in recent
years has become more common than not. Most museums require an entrance fee, and those that do not
often ask for donations. At this point of the visit, money and time have already been spent even before a
single artwork has been examined. Before further entering the museum, visitors might pick up a brochure
or similar materials, browse the information that they contain, and decide what galleries, exhibitions, and
artworks they want to look at. If visiting the museum with others, these decisions must first be discussed.
Visitors then spend the next hour or so browsing the museum’s collection of original artworks, permanent
and special exhibitions, and notable features of the museum space. They may even take some time to browse
the museum shop or to sit down for a break in the museum café or restaurant. After having experienced all
of this, the visitors leave the museum with new knowledge and experiences and having committed time,
money, and mental resources.
The museum visitors are deliberate and often strategic in their decision to visit a museum and in their
conduct when they are there. As Carr (2003) wrote regarding museums, “Nothing is there by accident, not

427
Aniko Illes and Pablo P. L. Tinio

even its users. At its best, a museum offers a constructed situation, a place we seek out purposefully” (p. 1).
The museum experience is a far cry from just mere exposure to a piece of art. For one, an object displayed
in a museum has a distinctiveness not necessarily due to unique physical properties of the object but instead
to the object being associated with a certain artist or artistic process and being displayed in the museum.
Moreover, authenticity and originality matter. This was demonstrated in a study by Huang et al. (2011), who
showed participants a set of paintings that were attributed to Rembrandt and a set of paintings that were not
attributed to Rembrandt but that were painted in his style by another artist. Participants viewed the paintings
that were randomly labeled as either authentic or a copy while fMRI data were collected. Results showed
that there were no significant and meaningful differences in cortical areas implicated in the visual perception
of objects and faces in response to authenticated Rembrandt paintings or copies of his paintings. There was
therefore no evidence that participants were able to identify differences in the visual characteristics of the two
sets of paintings. However, data showed different responses in the frontopolar cortex and right precuneus as a
result of being told that a painting was authentic or a copy, a finding that suggests that assignment of authen-
ticity and copy influenced the way that participants were thinking about the paintings.
It is also the case that original artworks hold more monetary value than their exact copies, which New-
man and Bloom (2012) stated were due to two types of beliefs that people have about art. First, artworks are
the result of a unique creative performance by an artist. Such a performance requires a particular set of skills
and the devotion of time, effort, and resources to the creative process. Second, original artworks are deemed
more valuable because their creators have physically touched them, therefore imparting these objects with
a special aura or quality, a phenomenon referred to as contagion (Frazer, 1890/1959; Newman & Bloom).
Because an artwork was hand made by a famous artist, it is seen as more special and valuable than the same
artwork reproduced by a lesser known person or by a machine. Contagion helps to explain why people are
willing to pay large sums of money to obtain the belongings of famous people—such as pens of famous
writers—when these same objects would be very inexpensive to purchase in a store.
Museums display original artworks that are considered one of a kind, valuable, and possessing special qual-
ities. But the museum space is, in itself, seen as special. In an fMRI study, Silveira et al. (2015) presented par-
ticipants with images of paintings that were labeled as belonging to the Museum of Modern Art (MOMA)
or an adult education center. Each painting was presented with one of the two types of contexts, randomized
across the participants. Both fMRI data and aesthetic appreciation ratings were collected. Results showed
that although there was no significant difference in the behavioral ratings of the paintings as a function of
context, the MOMA condition was associated with more intense neural activations in brain areas associated
with aesthetic value than the adult education center condition.
The finding that the museum context modulates aesthetic experiences of art was also demonstrated by
Specker et al. (2017). Their study directly compared the effects of presenting art in the laboratory and in the
museum. The same painting by the artist William Gropper was presented as the actual, original version in
the museum and as its digital reproduction in the laboratory. While viewing the painting in either context,
participants rated the intensity of their aesthetic experience of the work and answered questions that tested
the extent to which they learned information about the work that was presented to them. Results showed
that experiencing the Gropper painting in the museum led to both more intense aesthetic experiences and
better learning than experiencing the reproduction in the laboratory.
Another study that highlighted the effects of the museum environment on the aesthetic experience of
art was conducted by Brieber, Nadal and Leder (2015), who stated as a primary goal of their research to
determine “the psychological value of experiencing art in a museum” (p. 37). An important feature of their
study was the use of a repeated-measures design with participants able to view the same set of artworks in
the museum and in the laboratory (the same exhibition, but in a virtual format). They were assigned to
three groups that determined the order of the context: museum first—laboratory second, laboratory first—
museum second, and laboratory first and second, as a control group. Results showed the extent to which the

428
Experiencing art in museums

museum environment enhanced the aesthetic experience of the artworks. Participants found the artworks
in the museum more positive, likeable, interesting, understandable, and arousing. Interestingly, the order
of the context of viewing also had an effect, with significant differences in ratings of positivity, likeability,
interestingness, understandability, and arousal between the museum and laboratory viewing sessions for the
participants who viewed the artworks in the museum first; no such differences were found in the two other
groups. As a secondary analysis, Brieber, Nadal and Leder (2015) found that participants who viewed the
artworks in the museum were better able to remember the artworks. The authors’ conclusions were very
telling of what the museum brings to the aesthetic experience:

It enriches the experience and memory of art. By allowing people to encounter authentic artworks
in a special context that enables actual physical exploration, art is experienced as more arousing,
positive, and interesting, it is liked more and remembered better. This added value explains, in part
at least, why people are willing to invest time and resources to visit museums, instead of taking
inexpensive virtual tours.
(p. 42)

Research has therefore consistently shown key differences between the experience of art in museums and
other contexts (see also Locher et al., 1999, 2001). However, such studies raise an important issue: original
artworks such as paintings are physically different from their photographic reproductions, and because of this,
stimulus and context are confounded. A study by Brieber et al. (2014) addressed this issue in their study,
which used as stimuli photographs that were being exhibited in a museum. The use of art photographs had
the advantage of closing the gap between the physical characteristics of the originals displayed in the museum
and their corresponding images displayed in the lab. Put simply, a photograph and its photographic repro-
duction are more similar than a painting or sculpture and its photographic reproduction. Participants in the
study viewed the artworks either in the museum or in the lab while eye-tracking data were collected as a
means to objectively measure how long participants looked at the artworks. Participants in the museum spent
more time viewing the artworks than those in the lab. In addition, the participants in the museum found the
artworks more likeable and interesting. As with Specker et al. (2017), Brieber et al. showed that the museum
context enhanced the experience of the artworks.
Brieber et al.’s (2014) use of photographs as experimental stimuli helped to bridge the physical and visual
gap between real artworks and their reproductions. However, the approach still has its limitations, as a pho-
tograph and its reproduction are still fundamentally different stimuli, even if the differences between them
are minimal, such as regarding their image resolution, color characteristics, and size (e.g., Estrada-Gonzalez
et al., 2020; Tinio & Leder, 2009; Tinio & Leder, 2013; Tinio et al., 2010).
The optimal research design to examine the added value that the museum adds to the aesthetic experi-
ence would involve people viewing the same original artworks in different spaces. This, of course, is not a
straightforward and simple matter, as it involves significant logistical problems that a typical psychology study
would find difficult, if not impossible, to address. Consider a study with a similar research design as was used
by Brieber et al. (2014) or Specker et al. (2017). The study will require the use of the exact originals to be
presented in both the museum and laboratory. In other words, the original works would have to be physically
moved from one place to another. Such works would likely be highly valuable, if not priceless, and would be
considered significant and important works not only within a museum’s collection but the art world at large.
To use the works as experimental stimuli, the researchers would have to secure the approval of the museum
and address issues surrounding the safety of the works (proper handling during transport and secure display
in the mock gallery) and their insurance policies. To date, only one study has conducted such research with
highly valuable original works belonging to a real art collection (Brieber, Leder & Nadal, 2015). However,
the results did not show the independent contributing effects on the aesthetic experience of physical context

429
Aniko Illes and Pablo P. L. Tinio

and genuineness of the artworks. The researchers concluded that the artworks’ physical removal from the gal-
leries and relocation to the laboratory likely disrupted the ability of the artworks to elicit the strong aesthetic
responses that they would have in the museum. In this sense, artwork and art context are intertwined, an
issue we will delve into more in the following section.

Curatorial features of the museum


The previous findings serve as clear evidence for the museum’s added value to the aesthetic encounter and
speak to the concept of situated cognition—that human behavior and experiences are a result of interactions
with specific events, processes, and physical aspects of the environment. In other words, the situation and
the experience are intertwined, and one cannot be considered independent of the other (e.g., Barsalou,
2008; Clark, 1998). According to Carr (2003), museums “[capture and hold] far more than the sum of
their visible contents, museums enfold the multiple, infinite, and simultaneous constructions of their users”
(p. 2). Artworks in museums therefore become more than just the qualities of the art objects they contain
because the museum contexts themselves define the nature and interpretation of the artworks presented
within their walls.
Museums, as cultural institutions, hold, restore, and protect cultural artifacts, and for art museums, these
include objects of high monetary value and fame and often priceless iconic works that represent what people
consider typical art. Included in this group of artworks are van Gogh’s Starry Night, Picasso’s Les Demoiselles
d’Avignon, Michelangelo’s David, Munch’s The Scream, and perhaps most world renowned, da Vinci’s Mona
Lisa. These are highly recognizable objects that for many non-art experts might come to mind if asked to
think about artworks they know. There is no doubt that museums hold treasures and masterpieces. The
physical buildings in which they are housed are often carefully designed and decorated, with some having
rich histories themselves. As such, museums radiate a certain aura that makes them highly distinctive as spaces
and as cultural institutions (Bourdieu, 1984; Carr, 1991, 1992; Tröndle, 2014), and those who visit them
hold the experience with a certain sense of reverence (Smith & Wolf, 1996).
Museums also have an air of authority about them that colors people’s reactions to the artworks they hold.
Even the mere suggestion that an object is an artwork could have a big impact on how that object is evalu-
ated. For example, Van Dongen et al. (2016) showed that labeling photographs of various events as artworks
were liked more than when they were not labeled as artworks. They also showed ERP evidence for implicit
emotional regulation in response to the artworks. Moreover, Kirk et al. (2009) showed greater activity in
pleasure areas of the brain when people were told that the artworks they were viewing came from an art gal-
lery as opposed to being created using a computer.
One of the more significant, illuminating, and fairly consistent findings from psychology of art research
over the last two decades is related to the amount of time visitors spend looking at art in museums. These
findings suggest the impact of the museum context on the experience of art. One might think that a person
standing in front of a masterpiece would look at the object for at least several minutes to examine, ponder,
and savor the work. It is a masterpiece, after all, and it should get the attention that it deserves. Further-
more, when asked about their visit to museums, visitors often report that the experience was transformative,
emotional, self-referential, awe-inspiring, and even life-changing (e.g., Pelowski, 2015; Pelowski & Akiba,
2011; Smith, 2014; Tinio & Gartus, 2018; Tinio et al., 2010). Aesthetic experiences of art in museums could
also be considered events that are highly memorable. What is remarkable is that when people are observed
interacting with artworks in museums, they do not spend much time at all looking at individual artworks.
Smith and Smith (2001), in their landmark study on the amount of time visitors spent looking at artworks
at the Metropolitan Museum of Art, found that on average, visitors spent 27.2 seconds looking at individual
artworks. This duration included the time spent reading labels accompanying the works. In fact, a few outli-
ers drove up this duration, and a more representative time was the median at 17.0 seconds. Incredibly, the

430
Experiencing art in museums

artworks included in the study were considered masterpieces in the museum’s permanent collection and
included works such as Emmanuel Leutze’s Washington Crossing the Delaware, Rembrandt van Rijn’s Aristotle
with a Bust of Homer, Paul Cezanne’s The Card Players, and Raphael’s Madonna and Child Enthroned. Smith
et al. (2017) replicated this study at the Art Institute of Chicago, with a much larger sample, more diverse set
of artworks, and separate measurements of time spent looking at artworks and reading labels. They found a
mean looking time of 28.6 seconds and a median of 21.0 seconds, values that are consistent with those found
in the first study. In another replication study, Carbon (2017) examined the viewing times of six paintings by
the artist Gerhard Richter being exhibited at the State Museum for Art and Design in Nuremberg, Germany.
He also found viewing times that were similar to previous studies, with the mean at 33.9 seconds and the
median at 25.1 seconds. The fact that these studies found that interactions with individual artworks in muse-
ums occur in about 30 seconds raises an important question about the museum experience: How can such
brief experiences with individual artworks produce intense and meaningful experiences in their viewers?
The answer to this question lies in the very features that make museums distinct spaces for experienc-
ing art. Museums do not just present artworks; instead, they curate the presentation of artworks (Reitstätter
et al., 2020; Specker, Stamkou et al., 2020). The curation of artworks, usually done by a museum curator, is
one of the most important jobs in a museum. Museum curators are highly trained in art history, art theory,
and related fields and are experts in one or more types of art. It is not uncommon for a curator to spend an
entire career developing expertise in one very specific art style or artist. This curator would have in-depth
knowledge of the techniques and materials used to produce works within the area of expertise. The curator
would also know the historical context of when the works were produced, how the works might have been
transferred from one collection or owner to another, and how the works eventually became a part of the
museum’s collection. The curator would also know the influences that other artists and the art world have
had on the production of the works (Tinio, 2019a).
With this knowledge, curators lead the work of acquiring, protecting, representing, and, most relevant
to this chapter, exhibiting artworks in a coherent, curated manner. For example, a museum might devote an
entire exhibition to modern artworks by artists who were part of the New York art scene around the 1950s.
The focus of the exhibition could be narrowed even further to paintings by artists living around 9th Street
in New York around the 1950s, and even further to paintings by a select group of women artists working in
this same time and place, as Mary Gabriel (2018) described in her book, Ninth Street Women, which focused
on the lives, works, and relationships among the artists Lee Krasner, Elaine de Kooning, Grace Hartigan,
Joan Mitchell, and Helen Frankenthaler. An exhibition such as this would require in-depth knowledge of
each of these artists, including their life histories, development as artists, art education and training, theories
of and perspectives on art, creative breakthroughs, favored techniques and materials, successes and failures in
their career, personal relationships, influences from art history and contemporaries, and, of course, the art
that they produced. A curator heading such an exhibition would likely have knowledge about most if not
all of this and, when organizing the display of the works, would ensure that there is a shared narrative or a
conceptual thread connecting the selected works (Tröndle et al., 2014). Wall texts describing the exhibition
and labels accompanying each work would communicate this narrative and help to scaffold a museum visi-
tor’s experience of the art.
Therefore, the moment visitors step into an exhibition, such as in our previous example, they are pre-
sented with a carefully organized, coherent, and curated set of artworks and information that not only
describe each artwork’s creator, title, provenance, and materials but also information about the larger con-
texts in which the artwork should be considered: the artist’s personal context, such as the psychological,
socioemotional, and intellectual state of the artist during the time that a particular work was made, and larger
contexts, such as the historical, sociological, and economic factors that may have influenced the creation and
dissemination of the work. The visitor will also be guided by a theme or set of themes relating a work to
other artworks or an artist to other artists and by well-considered sequencing and placements of the works

431
Aniko Illes and Pablo P. L. Tinio

(e.g., Mullennix et al., 2020; Specker, Stamkou et al., 2020). The curation of artworks applies not only to
special exhibitions but also to a museum’s permanent collection, where artworks could vary tremendously in
terms of their artistic media (e.g., painting, sculpture, photography), styles, and time period. Even in these
cases, curators still attempt to find a common thread that connects one work to another. Museum curation
thus involves both the physical aspects of how works are exhibited and the conceptual aspects of how they
fit and relate together (Tröndle et al., 2014).
The curatorial work that happens in museums is the answer to the question of how brief experiences with
artworks are able to produce personal, intense, and meaningful experiences in their viewers. It is not the
mere collection of roughly 30-second experiences that produces these outcomes. Instead, it is the cumulative
effects of these individual experiences that matter (Smith & Smith, 2001; Smith et al., 2017). In other words,
the museum experience is greater than the sum of its parts. The work of presenting artworks in a museum is
focused less on brief looks at individual pieces of art than the totality of the museum experience.
Although you may enter a museum with a desire to learn about a specific painting or to see, or perhaps
revisit, a particular sculpture, you will still be subjected, if even just minimally, to what the museum intended
you to see, learn, and ponder. Whether implicitly or explicitly, the museum is entrusted with facilitating
a meaningful experience (Smith, 2014). The significance of what the actual museum brings to the aes-
thetic experience becomes even greater when we consider that museum visitors invested time and resources
towards their visit and when we factor in that they will be interacting with genuine works by accomplished,
if not famous, artists within a dedicated architectural space (e.g., Pelowski, Graser et al., 2019; Pelowski,
Leder et al., 2018).

The museum challenge for future research


As shown in the previous discussions, there is ample evidence showing how various factors associated with
the museum visitor, the artwork, and the viewing context can influence the experience of art (e.g., Chat-
terjee, 2003; Chatterjee & Vartanian, 2014; Leder & Nadal, 2014; Leder et al., 2004; Tinio, 2013). The psy-
chology of aesthetics and the arts is the second-oldest field in psychology, having been around for a century
and a half. During this time, most of the research on the experience of art has focused on how the character-
istics of the perceiver and of the artwork impact aesthetic experiences. We therefore have some fairly robust
findings that have been replicated using different samples, stimuli, and testing conditions.
Of the perceiver, the artwork, and the context, the latter has received the least attention. What we do
know is that there is a significant difference in people’s responses to art in the laboratory and in museum
contexts. The museum has a significant and positive impact on the experience of art, and experiencing art
in museums produce more meaningful, intense, and personally relevant outcomes (Mastandrea et al., 2019;
Pelowski et al., 2017; Specker et al., 2017; Tinio & Gartus, 2018). Given all of this, it is not surprising that
museums are considered the prototypical context for experiencing art, and even in our digital and online age,
museums are still tourist-attracting, revenue-generating, and highly respected cultural institutions (Styliani
et al., 2009). More empirical research in museums is needed.
Among the aspects of the experience of art in museums that are least understood and one of the most
deserving of attention from researchers is the social aspect of a museum visit. In addition to the authenticity
of artworks and the curatorial context, the social aspect is one of the factors that distinguishes art experiences
in museums from those in laboratories and other environments. One only needs to visit a museum to see that
the museum visit is often a social event (Kirchberg & Tröndle, 2015). In their study at the Art Institute of
Chicago, Smith et al. (2017) found that 39% of visitors were with one other person and 11% were with three
or more people. Visiting museums with others has been shown to impact outcomes such as time spent looking
at artworks, which increases when visitors are with other people (e.g., Smith & Smith, 2001). This increase
in viewing time is not surprising because being accompanied by others increases the likelihood of discussions

432
Experiencing art in museums

about the artworks (Tröndle et al., 2012). There are many important questions that remain to be asked about
how social interactions might impact key outcomes of the experience of art, including understanding and
meaning-making of, learning about, and judgments of art. It would also be important to find out what types
of social situations could positively or negatively impact aesthetic experiences (Pelowski et al., 2014). While
these questions are interesting from a scientific point of view, practically, from the museum’s perspective, they
are critical because they can directly inform practices, programming, and the ability to meet the needs of the
visitors.
Another aspect of the museum visit that has yet to be sufficiently examined is the impact that a museum
visit might have on a person’s health and well-being. In addition to the museum’s role in educating its audi-
ence, it is also considered a place for leisure, a place into which one could escape from the hustle and bustle
of daily life (Guintcheva & Passebois, 2009; Gürel & Nielsen, 2019).). This sort of restorative effect has been
shown in other contexts, such as natural environments (e.g., Kaplan, 1995; Ulrich et  al., 1991; Van den
Berg et al., 2007). The few studies that have examined the positive impact of experiencing art in museums
on health have been promising. For example, in a study at the National Gallery of Modern Art in Rome,
Mastandrea et al. (2019) measured visitors’ heart rate and blood pressure as they viewed modern and figura-
tive artworks. Results showed that viewing figurative artworks led to a significant decrease in blood pressure,
which is indicative of lower stress. Further research is needed to identify the exact conditions in which inter-
actions with artworks in museums could be beneficial to the health and well-being of visitors.
An additional factor of the museum visit is related to a visitor’s frame of mind regarding the decision to
visit a museum and how the visitor mentally prepares for the event (and whether this process is deliberate).
Moreover, as a visitor enters a museum, what cues suggest that their experience is now aesthetic in nature
versus just another daily life experience? Issues related to the visitor’s mindset or museum schema are relevant
to neuroscience, particularly in relation to an aesthetic mode of viewing that might be common to all people.
Nadal and Chatterjee (2018) have argued that the universality and diversity of art may be considered within
the same neuroaesthetic framework, and that

Art’s universality is the product of our common evolution and shared experiences. Art’s diversity
is the product of our complex, culturally permeable, and flexible brain responding to different
environments.
(p. 8, Nadal & Chattejee)

The nature of such an art response is beginning to be uncovered, with recent neuroaesthetics research
uncovering activities in the default-mode network and reward centers of the brain during aesthetic experi-
ences of visual art (Belfi et al., 2019). There are therefore no specific neural structures for aesthetic experi-
ences but rather patterns of neural activity.
Following Nadal and Chatterjee’s (2018) line of reasoning, questions related to what makes aesthetic expe-
riences unique become relevant to museums because they are social and cultural constructions—institutions
that were carefully developed for artistic encounters—to which humans have learned to respond in a specific
manner. An example of a specific art-related response is people’s experiences of installation art, where the
line between artwork and the context in which it is presented is often blurred (Pelowski et al., 2018). Perhaps
these types of art will be most illuminating and informative to study if we are to shed light on what makes
the experience of art truly special.

References
Barsalou, L. W. (2008). Grounded cognition. Annual Review of Psychology, 59, 617–645. https://doi.org/10.1146/
annurev.psych.59.103006.093639

433
Aniko Illes and Pablo P. L. Tinio

Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., Pelli, D. G., & Starr, G. G. (2019).
Dynamics of aesthetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.
org/10.1016/j.neuroimage.2018.12.017
Bourdieu, P. (1984). Distinction: A social critique of the judgement of taste. Harvard University Press.
Brieber, D., Leder, H., & Nadal, M. (2015). The experience of art in museums: An attempt to dissociate the role of physi-
cal context and genuineness. Empirical Studies of the Arts, 33(1), 95–105. https://doi.org/10.1177/0276237415570000
Brieber, D., Nadal, M., & Leder, H. (2015). In the white cube: Museum context enhances the valuation and memory of
art. Acta Psychologica, 154, 36–42. https://doi.org/10.1016/j.actpsy.2014.11.004
Brieber, D., Nadal, M., Leder, H., & Rosenberg, R. (2014). Art in time and space: Context modulates the relation
between art experience and viewing time. PLoS One, 9(6), e99019. https://doi.org/10.1371/journal.pone.0099019
Cairns, S. (2013). Mutualizing museum knowledge: Folksonomies and the changing shape of expertise. Curator: The
Museum Journal, 56(1), 107–119. https://doi.org/10.1111/cura.12011
Carbon, C. C. (2017). Art perception in the museum: How we spend time and space in art exhibitions. i-Perception, 8(1),
2041669517694184. https://doi.org/10.1177/2041669517694184
Carr, D. W. (1991). Minds in museums and libraries: The cognitive management of cultural institutions [Teach-
ers College record]. Teachers College Record: The Voice of Scholarship in Education, 93(1), 6–27. https://doi.
org/10.1177/016146819109300105
Carr, D. W. (1992). Cultural institutions as structures for cognitive change. New Directions for Adult and Continuing Educa-
tion, 1992(53), 21–35. https://doi.org/10.1002/ace.36719925305
Carr, D. W. (2003). The promise of cultural institutions. Alta Mira Press.
Chatterjee, A. (2003). Prospects for a cognitive neuroscience of visual aesthetics. Bulletin of Psychology and the Arts, 4,
55–60.
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Clark, A. (1998). Being there: Putting brain, body, and world together again. MIT Press.
Daskalaki, V. V., Voutsa, M. C., Boutsouki, C., & Hatzithomas, L. (2020). Service quality, visitor satisfaction and future
behavior in the museum sector. Journal of Tourism, Heritage & Services Marketing, 6(1), 3–8.
Estrada-Gonzalez, V., East, S., Garbutt, M., & Spehar, B. (2020). Viewing art in different contexts. Frontiers in Psychology,
11, 569. https://doi.org/10.3389/fpsyg.2020.00569
Falk, J. H. (2009). Identity and the museum visitor experience. Lef Coast Press.
Francuz, P., Zaniewski, I., Augustynowicz, P., Kopiś, N., & Jankowski, T. (2018). Eye movement correlates of expertise
in visual arts. Frontiers in Human Neuroscience, 12, 87. https://doi.org/10.3389/fnhum.2018.00087
Frazer, J. G. (1959). The new golden bough: A new abridgment of the classic work. Macmillan (original work published 1890).
Gabriel, M. (2018). Ninth Street women: Krasner, L., de Kooning, E., Hartigan, G., Mitchell, J., & Frankenthaler, H. Five
painters and the movement that changed modern art. Little, Brown, and Co.
Guintcheva, G., & Passebois, J. (2009). Exploring the place of museums in European leisure markets: An approach based
on consumer values. International Journal of Arts Management, 11, 4–19.
Gürel, E., & Nielsen, A. (2019). Art museum visitor segments: Evidence from Italy on omnivores and highbrow univores.
International Journal of Arts Management, 21(2), 55–69.
Hekkert, P.,  & Van Wieringen, P. C. (1996). Beauty in the eye of expert and nonexpert beholders: A  study in the
appraisal of art. The American Journal of Psychology, 109(3), 389–407. https://doi.org/10.2307/1423013
Huang, M., Bridge, H., Kemp, M. J., & Parker, A. J. (2011). Human cortical activity evoked by the assignment of authen-
ticity when viewing works of art. Frontiers in Human Neuroscience, 5, 134. https://doi.org/10.3389/fnhum.2011.00134
Illes, A., & Bodor, P. (2016). Patterns of museum perception in Hungarian university students. In S. Mastandrea & F.
Maricchiolo (Eds.), The role of the museum in the education of young adults. Motivation, emotion and learning (pp. 185–201).
Roma Tre-Press.
Illes, A., & Bodor, P. (2018). Cultural heritage and the meaning of museums for young Hungarians. In M. Banks (Ed.),
IECON18: Interpret Europe conference—conference proceedings (pp. 81–82). Kőszeg.
Kaplan, S. (1995). The restorative benefits of nature: Towards an integrative framework. Journal of Environmental Psychol-
ogy, 15(3), 169–182. https://doi.org/10.1016/0272-4944(95)90001-2
Kirchberg, V., & Tröndle, M. (2015). The museum experience: Mapping the experience of fine art. Curator: The Museum
Journal, 58(2), 169–193. https://doi.org/10.1111/cura.12106
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009). Modulation of aesthetic value by semantic con-
text: An fMRI study. Neuroimage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009
Kotler, N. G., Kotler, P., & Kotler, W. I. (2008). Museum marketing and strategy: Designing missions, building audiences, generat-
ing revenue and resources. John Wiley & Sons.

434
Experiencing art in museums

Kozbelt, A.,  & Seeley, W. P. (2007). Integrating art historical, psychological, and neuroscientific explanations of art-
ists’ advantages in drawing and perception. Psychology of Aesthetics, Creativity, and the Arts, 1(2), 80–90. https://doi.
org/10.1037/1931-3896.1.2.80
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Locher, P. J., Smith, J. K.,  & Smith, L. F. (2001). The influence of presentation format and viewer training in the
visual arts on the perception of pictorial and aesthetic qualities of paintings. Perception, 30(4), 449–465. https://doi.
org/10.1068/p3008
Locher, P. J., Smith, L. F., & Smith, J. K. (1999). Original paintings versus slide and computer reproductions: A com-
parison of viewer responses. Empirical Studies of the Arts, 17(2), 121–129. https://doi.org/10.2190/R1WN-TAF2-
376D-EFUH
Massaro, D., Savazzi, F., di Dio, C., Freedberg, D., Gallese, V., Gilli, G.,  & Marchetti, A. (2012). When art moves
the eyes: A  behavioral and eye-tracking study. PLoS One, 7(5), e37285. https://doi.org/10.1371/journal.pone.
0037285
Mastandrea, S., Maricchiolo, F., Carrus, G., Giovannelli, I., Giuliani, V., & Berardi, D. (2019). Visits to figurative art
museums may lower blood pressure and stress. Arts & Health, 11(2), 123–132.
Mullennix, J. W., Kristo, G. M., & Robinet, J. (2020). Effects of preceding context on aesthetic preference. Empirical
Studies of the Arts, 38(2), 149–171. https://doi.org/10.1177/0276237418805687
The Museum of Modern Art history: MoMA. (n.d.). Retrieved March  10, 2021. http://www.moma.org/about/
who-we-are/moma-history
Nadal, M., & Chatterjee, A. (2018). Neuroaesthetics and art’s diversity and universality. WIREs Cognitive Science, 10(3),
e1487.
Newman, G. E., & Bloom, P. (2012). Art and authenticity: The importance of originals in judgments of value. Journal of
Experimental Psychology. General, 141(3), 558–569. https://doi.org/10.1037/a0026035
Pekarik, A. J., Doering, Z. D., & Karns, D. A. (1999). Exploring satisfying experiences in museums. Curator: The Museum
Journal, 42(2), 152–173. https://doi.org/10.1111/j.2151-6952.1999.tb01137.x
Pekarik, A. J., & Schreiber, J. B. (2012). The power of expectation. Curator: The Museum Journal, 55(4), 487–496. https://
doi.org/10.1111/j.2151-6952.2012.00171.x
Pelowski, M. (2015). Tears and transformation: Feeling like crying as an indicator of insightful or “aesthetic” experience
with art. Frontiers in Psychology, 6, 1006. https://doi.org/10.3389/fpsyg.2015.01006
Pelowski, M., & Akiba, F. (2011). A model of art perception, evaluation and emotion in transformative aesthetic experi-
ence. New Ideas in Psychology, 29(2), 80–97. https://doi.org/10.1016/j.newideapsych.2010.04.001
Pelowski, M., Forster, M., Tinio, P. P. L., Scholl, M., & Leder, H. (2017). Beyond the lab: An examination of key fac-
tors influencing interaction with “real” and museum-based art. Psychology of Aesthetics, Creativity, and the Arts, 11(3),
245–264. https://doi.org/10.1037/aca0000141
Pelowski, M., Graser, A., Specker, E., Forster, M., von Hinüber, J., & Leder, H. (2019). Does gallery lighting really have
an impact on appreciation of art? An ecologically valid study of lighting changes and the assessment and emotional
experience with representational and abstract paintings. Frontiers in Psychology, 10, 2148. https://doi.org/10.3389/
fpsyg.2019.02148
Pelowski, M., Leder, H., Mitschke, V., Specker, E., Gerger, G., Tinio, P. P. L., Vaporova, E., Bieg, T.,  & Husslein-
Arco, A. (2018). Capturing aesthetic experiences with installation art: An empirical assessment of emotion, valua-
tions, and mobile eye tracking in Olafur Eliasson’s “Baroque, Baroque!” Frontiers in Psychology, 9, 1255. https://doi.
org/10.3389/fpsyg.2018.01255
Pelowski, M., Liu, T., Palacios, V., & Akiba, F. (2014). When a body meets a body: An exploration of the negative impact
of social interactions on museum experiences of art. International Journal of Education and the Arts, 15(14), 1–48.
Pihko, E., Virtanen, A., Saarinen, V. M., Pannasch, S., Hirvenkari, L., Tossavainen, T., Haapala, A., & Hari, R. (2011).
Experiencing art: The influence of expertise and painting abstraction level. Frontiers in Human Neuroscience, 5, 94.
https://doi.org/10.3389/fnhum.2011.00094
Rawlings, D. (2003). Personality correlates of liking for’unpleasant’paintings and photographs. Personality and Individual
Differences, 34(3), 395–410. https://doi.org/10.1016/S0191-8869(02)00062-4
Reitstätter, L., Brinkmann, H., Santini, T., Specker, E., Dare, Z., Bakondi, F., Miscená, A., Kasneci, E., Leder, H., &
Rosenberg, R. (2020). The display makes a difference: A mobile eye tracking study on the perception of art before
and after a museum’s rearrangement. Journal of Eye Movement Research, 13(2). https://doi.org/10.16910/jemr.13.2.6

435
Aniko Illes and Pablo P. L. Tinio

Silveira, S., Fehse, K., Vedder, A., Elvers, K., & Hennig-Fast, K. (2015). Is it the picture or is it the frame? An fMRI
study on the neurobiology of framing effects. Frontiers in Human Neuroscience, 9, 528. https://doi.org/10.3389/
fnhum.2015.00528
Smith, J. K. (2014). The museum effect: How museums, libraries, and cultural institutions educate and civilize society. Rowman &
Littlefield Publishing Group.
Smith, J. K.,  & Smith, L. F. (2001). Spending time on art. Empirical Studies of the Arts, 19(2), 229–236. https://doi.
org/10.2190/5MQM-59JH-X21R-JN5J
Smith, L. F., & Smith, J. K. (2006). The nature and growth of aesthetic fluency. In P. Locher, C. Martindale & L. Dorf-
man (Eds.), Foundations and frontiers in aesthetics. New directions in aesthetics, creativity and the arts (pp. 47–58). Baywood
Publishing.
Smith, L. F., Smith, J. K., & Tinio, P. P. L. (2017). Time spent viewing art and reading labels. Psychology of Aesthetics,
Creativity, and the Arts, 11(1), 77–85. https://doi.org/10.1037/aca0000049
Smith, J. K., & Wolf, L. F. (1996). Museum visitor preferences and intentions in constructing aesthetic experience. Poetics,
24(2–4), 219–238. https://doi.org/10.1016/0304-422X(95)00006-6
Specker, E., Forster, M., Brinkmann, H., Boddy, J., Pelowski, M., Rosenberg, R., & Leder, H. (2020). The Vienna Art
Interest and Art Knowledge Questionnaire (VAIAK): A unified and validated measure of art interest and art knowl-
edge. Psychology of Aesthetics, Creativity, and the Arts, 14(2), 172–185. https://doi.org/10.1037/aca0000205
Specker, E., Stamkou, E., Pelowski, M., & Leder, H. (2020). Radically revolutionary or pretty flowers? The impact of
curatorial narrative of artistic deviance on perceived artist influence. Psychology of Aesthetics, Creativity, and the Arts
[Advance online publication]. https://doi.org/10.1037/aca0000320
Specker, E., Tinio, P. P., & van Elk, M. (2017). Do you see what I see? An investigation of the aesthetic experience in
the laboratory and museum. Psychology of Aesthetics, Creativity, and the Arts, 11, 265–275.
Styliani, S., Fotis, L., Kostas, K., & Petros, P. (2009). Virtual museums, a survey and some issues for consideration. Journal
of Cultural Heritage, 10(4), 520–528. https://doi.org/10.1016/j.culher.2009.03.003
Tinio, P. P. L. (2013). From artistic creation to aesthetic reception: The mirror model of art. Psychology of Aesthetics, Crea-
tivity, and the Arts, 7(3), 265–275. https://doi.org/10.1037/a0030872
Tinio, P. P. L. (2017). On viewer motivation, unit of analysis, and the VIMAP: Comment on “Move me, astonish me . . .
delight my eyes and brain: The Vienna Integrated Model of top-down and bottom-up processes in Art Perception
(VIMAP) and corresponding affective, evaluative, and neurophysiological correlates” by Matthew Pelowski et al. Phys-
ics of Life Reviews, 21, 152–154. https://doi.org/10.1016/j.plrev.2017.05.003
Tinio, P. P. L. (2019a). Creative mutual facilitation in the arts and sciences. International Journal of Creativity and Problem
Solving, 29, 65–76.
Tinio, P. P. L. (2019b). Creativity and aesthetics. In J. C. Kaufman & R. J. Sternberg (Eds.), Cambridge handbook of creativ-
ity (pp. 691–708). Cambridge University Press.
Tinio, P. P. L. (2020). Aesthetics. In R. Runco & S. Pritzker (Eds.), Encyclopedia of creativity (3rd ed). Elsevier.
Tinio, P. P. L., & Gartus, A. (2018). Characterizing the emotional response to art beyond pleasure: Correspondence
between the emotional characteristics of artworks and viewers’ emotional responses. Progress in Brain Research, 237,
319–342. https://doi.org/10.1016/bs.pbr.2018.03.005
Tinio, P. P. L., & Leder, H. (2009). Natural scenes are indeed preferred, but image quality might have the last word.
Psychology of Aesthetics, Creativity, and the Arts, 3(1), 52–56. https://doi.org/10.1037/a0014835
Tinio, P. P. L., & Leder, H. (2013). The means to art’s end: Styles, creative devices, and the challenge of art. In O. Varta-
nian, A. S. Bristol, & J. C. Kaufman (Eds.), The neuroscience of creativity (pp. 273–298). MIT Press.
Tinio, P. P. L., Leder, H., & Strasser, M. (2010). Image quality and the aesthetic judgment of photographs: Contrast,
sharpness, and grain teased apart and put together. Psychology of Aesthetics, Creativity, and the Arts, 5(2), 165–176.
https://doi.org/10.1037/a0019542
Tinio, P. P. L., Smith, J. K., & Potts, K. (2010). The object and the mirror: The nature and dynamics of museum tours.
International Journal of Creativity and Problem Solving, 20, 37–52.
Tinio, P. P. L., Smith, J. K., & Smith, L. F. (2014). The walls do speak: Psychological aesthetics and the museum experi-
ence. In P. P. L. Tinio & J. K. Smith (Eds.), Cambridge handbook of the psychology of aesthetics and the arts (pp. 195–218).
Cambridge University Press.
Tinio, P. P. L., & Specker, E. (2020). Observation as a method. In M. Nadal & O. Vartanian (Eds.), Cambridge handbook
of empirical aesthetics (pp. 1–17). Cambridge University Press.
Tröndle, M. (2014). Space, movement and attention: Affordances of the museum environment. International Journal of
Arts Management, 17(1), 4–17.
Tröndle, M., Greenwood, S., Bitterli, K., & van den Berg, K. (2014). The effects of curatorial arrangements. Museum
Management and Curatorship, 29(2), 140–173. https://doi.org/10.1080/09647775.2014.888820

436
Experiencing art in museums

Tröndle, M., Greenwood, S., Kirchberg, V., & Tschacher, W. (2014). An integrative and comprehensive methodology for
studying aesthetic experience in the field: Merging movement tracking, physiology, and psychological data. Environ-
ment and Behavior, 46(1), 102–135. https://doi.org/10.1177/0013916512453839
Tröndle, M., Wintzerith, S., Wäspe, R., & Tschacher, W. (2012). A museum for the twenty-first century: The influence
of ’sociality’on art reception in museum space. Museum Management and Curatorship, 27(5), 461–486. https://doi.org/
10.1080/09647775.2012.737615
Ulrich, R. S., Simons, R. F., Losito, B. D., Fiorito, E., Miles, M. A., & Zelson, M. (1991). Stress recovery during expo-
sure to natural and urban environments. Journal of Environmental Psychology, 11(3), 201–230. https://doi.org/10.1016/
S0272-4944(05)80184-7
Van den Berg, A. E., Hartig, T., & Staats, H. (2007). Preference for nature in urbanized societies: Stress, restoration, and
the pursuit of sustainability. Journal of Social Issues, 63(1), 79–96. https://doi.org/10.1111/j.1540-4560.2007.00497.x
Van Dongen, N. N., Van Strien, J. W., & Dijkstra, K. (2016). Implicit emotion regulation in the context of viewing
artworks: ERP evidence in response to pleasant and unpleasant pictures. Brain and Cognition, 107, 48–54. https://doi.
org/10.1016/j.bandc.2016.06.003
Vogt, S., & Magnussen, S. (2007). Expertise in pictorial perception: Eye-movement patterns and visual memory in artists
and laymen. Perception, 36(1), 91–100. https://doi.org/10.1068/p5262
Zhang, H., Chang, P. C., & Tsai, M. F. (2018). How physical environment impacts visitors’ behavior in learning-based
tourism—The example of technology museum. Sustainability, 10(11), 3880. https://doi.org/10.3390/su10113880

437
23
CONTEXT AND COMPLEXITY OF
AESTHETIC EXPERIENCES
A neuroscientific view

Julia Crone and Helmut Leder

In this chapter, we discuss the effect of context on the aesthetic appreciation of art from the perspective of
cognitive psychology and neuroscience. Our main goal is to show how many of the distinctive features of the
aesthetic appreciation owe to its embeddedness in a context. This goal stems from the fact that, by its very
nature, art is always part of and encountered in complex physical, social, cultural, and historical contexts.
Engaging with art in these contexts requires processing and integrating different sorts of information and
resolving ambiguity and elicits a variety of emotional responses which may modify and intensify aesthetic
experiences. Our discussion of the role of context in the appreciation of art bridges the cognitive and neu-
roscientific levels of explanation, too often dealt with separately.

The aesthetic experience


The capacity to appreciate art seems to be unique to the human species (Dissanayake, 2008; see also Chap-
ter 29). For this and other reasons, the aesthetic appreciation of art has attracted philosophers, humanists, and
psychologists for centuries. More recently, it has attracted scholars from other fields, such as neuroscience and
biology. From the perspective of psychology, the fundamental goal of empirical aesthetics is to understand
the aesthetic experience (Leder et al., 2004; Leder & Nadal, 2014). However, there has been little consensus
among these approaches and disciplines about what constitutes an aesthetic experience. There is disagree-
ment both about the most fruitful perspective and about how to define even the most fundamental terminol-
ogy. Some authors have argued that a domain-general neural system processes the affective values underlying
the appreciation of art (e.g., Skov & Nadal, 2020). Others see the experience of art as an exceptional state of
mind, a transcending experience of aesthetically pleasing stimuli that is detached from simple reward mecha-
nisms, and have criticized the reductionism and simplification of neuroscience (e.g., Marković, 2012; Noë,
2009). One of the reasons that make it difficult to find common ground between these opposing views is the
lack of a clear definition of the aesthetic experience and its specific components.
Nonetheless, there is agreement on some points. For example, most would accept that aesthetic experi-
ences can be elicited by natural and artificial sources of sensory stimulation that are not directly related to
survival (Zeki, 2013). According to Leder and colleagues (2004), aesthetic episodes consist of several cogni-
tive, evaluative, and affective processing stages. Chatterjee and Vartanian (2014) have proposed that these
can be organized as an “aesthetic triad” comprising sensory–motor, emotional–valuation, and meaning–
knowledge systems. Aesthetic experiences consist of many different emotional states and subjective feelings

438 DOI: 10.4324/9781003008675-25


Complexity of aesthetic experiences

(Anderson & Adolphs, 2014) which are described with different terms and concepts, including but not lim-
ited to hedonic response, emotional response, emotional intensity, enjoyment, pleasure, arousal, liking, interest, beauty,
moving experience, emotional valence, hedonic response, and aesthetic appreciation. Common to these descriptors is
the idea that an aesthetic experience always induces a certain level of pleasure (Chatterjee & Vartanian, 2014;
Nadal & Skov, 2018; Skov, 2020). In addition to mere feelings of pleasure, aesthetic stimuli can elicit intense
hedonic and moving experiences associated with so-called aesthetic chills, an emotional state that also mani-
fests itself in physiological responses such as goosebumps or a cold shiver running down the spine (Bannister,
2019; Blood & Zatorre, 2001; Grewe, 2007; Pelowski, 2015; Pelowski et al., 2018; Schoeller & Perlovsky,
2016; Schoeller et al., 2018). Aesthetic stimuli may even cause a transformative experience in which the
perspective of an individual is changed, resulting in new opinions and views about many different issues,
including themselves (Pelowski, 2015; Piff et al., 2015). Thus, any discussion of the aesthetic experience
should consider its multidimensionality and the fact that research is still making progress by investigating
parts of the whole.

The problem of isolating processes: pleasure and its underlying neurobiology


It has been a long-standing tradition in psychology, neurobiology, and neuroscience to break down psycho-
logical as well as neurobiological phenomena into singular processes, which are then investigated in isolation.
Although this analytical approach is unable to capture the inherent complexity of mental phenomena, it
reflects a desire for scientific accuracy in the way cognitive processes are measured in controlled condi-
tions. It also reflects the simple necessity to reduce complexity: Studying the neurobiological mechanisms
underlying cognitive processes necessarily entails isolating the process of interest. For example, recording all
neural responses of someone observing a van Gogh painting in a museum would provide little insight into
that experience due to the extreme complexity of the data. This is because the brain is not only processing
the aesthetic experience itself. It is also processing motion, actions, thoughts, and attentional shifts. The
brain activity measured in this specific situation could be associated with a wide range of different processes,
including the sensory stimulation of the painting itself, of the surroundings (e.g., conversations of other
people in the room), olfactory input (such as the scent of paintings or other people), personal associations
evoked by the painting or the artist, the retrieval of stored knowledge or memories, or processes related to
motion while moving through the museum. It would be impossible to know which of these neurobiological
mechanisms corresponds to the aesthetic experience.
A more fruitful approach begins by dividing the complex aesthetic experience into its constituent parts
to study them in isolation. Psychology and neuroscience, thus, begin by identifying the fundamental com-
ponents of the aesthetic experience. A typical approach is to break down the experience into psychological
concepts that are familiar. A very basic starting point is the psychological concept of the hedonic response. The
hedonic response comprises the experience of pleasure (i.e., liking—the consummatory aspect) and the moti-
vation to seek out a pleasurable experience (i.e., wanting—the motivational aspect) (Berridge et al., 2009;
see also Chapter 3). As Berridge and Kringelbach (2015) put it, “In a sense, pleasure can be thought of as
evolution’s boldest trick, serving to motivate an individual to pursue rewards necessary for fitness” (p. 647).
Thus, the most fundamental and prevalent aspect of an aesthetic experience is pleasure. However, the
affective valence and thus the induced feeling of pleasure is not an inherent quality of the stimulus. It is,
rather, the result of the way the stimulus is processed. The consummatory and the motivational aspects can
occur consciously or not (Childress et al., 2008; Winkielman et al., 2005), and the conscious and uncon-
scious aspects of liking can even be dissociated (Gilbert & Wilson, 2009): The amount of pleasure induced
is measured by subjective liking scales in humans (which represent only the conscious, subjective feeling of
pleasure) or objective hedonic responses, such as facial expression or consumption movements in both ani-
mals and humans (capturing conscious and unconscious correlates of pleasure).

439
Julia Crone and Helmut Leder

These different aspects of pleasure are mediated by activity in the brain’s mesolimbic reward system. The
reward system comprises a network of brain regions located primarily in the cortico-basal ganglia-thalamo-
cortical circuit (Yager et al., 2015) that interact through several glutamatergic, GABAergic, dopaminergic,
and opioid pathways (Berridge & Kringelbach, 2015; Dunlop & Nemeroff, 2007; Yager et al., 2015). One
essential component of the reward circuit is the dopamine pathway, which originates in the ventral tegmental
area and extends to the nucleus accumbens (located in the ventral striatum), the amygdala, the hippocam-
pus, and the prefrontal cortex (Dunlop & Nemeroff, 2007). Other components include the dorsal striatum
(i.e., the caudate nucleus and putamen), substantia nigra, thalamus, globus pallidus, ventral pallidum, parabra-
chial nucleus, anterior cingulate cortex, insular cortex, and orbitofrontal cortex.
The reward system is involved in both the consummatory and the motivational aspects of hedonic
responses. Both aspects can be dissociated in respect to specific mechanisms within this circuit. The consum-
matory aspect is primarily associated with the opioid system and the motivational aspect mainly with mes-
olimbic dopamine pathways. Animal studies demonstrate that dopaminergic suppression reduces anticipated
pleasure (Galistu & D’Aquila, 2012) but not expressions of pleasure (Peciña et al., 1997) and that dopamin-
ergic stimulation of these specific sites leads to enhanced wanting but not to enhanced liking (Smith et al.,
2011). Evidence from human studies support this dissociation, demonstrating that dopamine is not causally
involved in the experience of pleasure but instead induces the motivation to seek out a pleasurable experi-
ence (Brauer & De Wit, 1997; Leyton et al., 2007; see Chapters 3 and 7). In line with this, the anticipation of
aesthetic experiences is related to activity in the caudate nucleus, whereas the feeling of pleasure is related to
activity in the nucleus accumbens (Salimpoor et al., 2011). The nucleus accumbens is one of several regions
in the reward system that contain hedonic hotspots, neural sites that cause the experience of pleasure. Other
hedonic hotspots are found in other subcortical areas, such as the ventral pallidum and the pontine parabrachial
nucleus, and also in cortical areas such as the orbitofrontal cortex and insula (Castro & Berridge, 2017).
Only a small portion of these structures have the capacity to enhance liking reactions when stimulated
(Castro & Berridge, 2014; Smith & Berridge, 2005), and only the posterior ventral pallidum appears to be
a necessary hotspot for liking (Ho & Berridge, 2014; Khan et al., 2020). To induce a liking response, these
regions need to work together. If opioid activation is blocked in one site, stimulation of the other hotspot
does not affect the liking response (Smith & Berridge, 2007; Smith et al., 2011). A particular interesting
aspect is the neuronal plasticity of subcortical hotspots of pleasure; they expand and contract depending
on the environment (Reynolds & Berridge, 2008; Richard & Berridge, 2011). This rewiring seems to be
orchestrated by cortical hotspots, such as the orbitofrontal cortex, which regulate the intensity and valence
of motivations and pleasure (Richard & Berridge, 2013). There is much evidence showing that aesthetic
experiences involve activity throughout this network of reward-related brain regions.
The processing of pleasure seems to be independent of the source of sensory stimulation. Different stimuli
such as food, sex, drugs, music, and art activate a common pattern of reward activity throughout the main
pathways of the reward system (Georgiadis & Kringelbach, 2012; Kringelbach et al., 2012; Salimpoor et al.,
2011; Vartanian & Skov, 2014), a “common neural currency” (Berridge & Kringelbach, 2015). However, the
studies providing these results have used coarse methods to identify neuronal correlates such as fMRI blood-
level oxygen-dependent (BOLD) activity or connectivity. It may be that finer-grained methods would reveal
significant differences between stimulation types. Moreover, according to Skov (2019)

neuroscientific evidence suggests that aesthetic appreciation is not a distinct neurobiological pro-
cess assessing certain objects, but a general system, centered on the mesolimbic reward circuit, for
assessing the hedonic value of any sensory object. Furthermore, neuroscientific research also makes
it clear that hedonic values are not determined solely by object properties, but subject to a range of
object-extrinsic modulatory factors.
(p. 220)

440
Complexity of aesthetic experiences

Context and its role in the aesthetic experience


As noted previously, it can be argued that aesthetic experiences encompass other forms of hedonic responses
apart from the feeling of pleasure and the motivation to seek reward. Especially in art, the intensity of an
aesthetic experience (such as the feeling of beauty and experiencing chills) goes beyond the mere sensation
of pleasure. For example, chills involve not just states of intense pleasure but also a distinct engagement of
arousal, respiration, and heart rate that is specific to the feeling of “frisson” (Salimpoor et al., 2009). Impor-
tantly, chills are rare and appear to be elicited in different ways by different kinds of stimuli (Grewe et al.,
2009). Moreover, the different terms used to assess the aesthetic response to art (such as pleasure, liking, aesthetic
appreciation, beauty, etc.) do not differ merely in the choice of the precise wording which otherwise describes
the same phenomenon. They actually emphasize a distinction in the quality of the experience itself. There
is also evidence that different evaluative judgments that humans can use to appraise art (e.g., liking vs. beauty)
involve executive functions to different degrees. For instance, Brielmann and Pelli (2017) found beauty judg-
ments to be more affected by a concurrent working memory task than liking judgments. Similarly, Che and
colleagues (2021) demonstrated that high loading of visual working memory interfered more with judgments
of beauty than with judgments of liking. Recent imaging research suggests that aesthetic experiences engage
higher cortical areas, including medial prefrontal regions and the default mode network (Belfi et al., 2019;
Vessel et al., 2012, 2013; Williams et al., 2018), suggesting that some forms of aesthetic appreciation rely on
higher-order cognition.
It has long been realized that semantic meaning can influence aesthetic responses to art (Leder & Nadal,
2014; Pelowski, 2015). Known as cognitive penetration, this phenomenon shows that object-external infor-
mation (see Chapter 25) and conceptual expectations affect aesthetic appraisals. For example, if an object is
believed to be art, then negative emotional responses to it are attenuated (Frijda, 1988; Gross, 1998; Mocai-
ber et al., 2010), possibly because this expectation engenders a psychological distance toward the object (Cup-
chik, 2002). In contrast, positive emotions can be enhanced, as demonstrated by studies that have compared
responses to visual objects experienced in a museum and a laboratory setting (Brieber et al., 2015; Brieber
et al., 2014; Grüner et al., 2019; Specker et al., 2017; Szubielska et al., 2019).
The experience of art involves a social component as well. This social aspect is an important part of
the contextual setting and needs to be considered when discussing the role of context in the emergence
of an aesthetic response. However, the presence of other people in natural settings (such as museums or
art galleries), the social interactions, and their implications for the quality of an aesthetic episode have
only been investigated in very few studies so far. Pelowski et al. (2014) discussed the interaction between
museum space, the mere presence of others, and art exploration as well as possible effects of social
exchange, discourse, and learning. Steier et al. (2015) provided an interesting analysis of people encoun-
tering art together. A neuroscientific approach linking context as social interaction and brain network
interaction has not been implemented yet. A  promising direction to gather valuable insights into this
topic in the future are hyper-scanning studies employing portable functional near-infrared spectroscopy
(fNIRS).
Appraisals of art can also be affected by accompanying semantic information, for example, about the
work (Leder et al., 2006; Belke et al., 2010) or about the creator (Gerger & Leder, 2015; Huang et al., 2011;
Kirk et al., 2009; Millis, 2001; Szubielska et al., 2019). Intriguingly, different forms of semantic informa-
tion appear to have different effects. For instance, explanatory descriptions (compared to only a title and
the artist’s name or no information at all) increase the pleasantness the art induces but not necessarily the
meaningfulness of the art (Russell, 2003). Gerger and Leder (2015) studied context in form of a descriptive,
semantic-matching title compared to a random, non-matching or no title at all and found these to increase
liking ratings but not interest ratings. This finding was supported by objective measures of facial expression as
indexed by electromyography (facial EMG). When titles matched the content of the artwork, participants

441
Julia Crone and Helmut Leder

engaged in more smiling responses. When the titles did not match the content, more frowning responses
were observed.
There is evidence to suggest that the relevance of the semantic context influences the aesthetic experi-
ence by increasing understanding. Content-specific information compared to less relevant information
(such as genre-specific information, the title, or no information at all) increases the subjective feeling of
understanding of the artwork, and this increase in understanding has a mediating effect on aesthetic appre-
ciation (Swami, 2013). Park et al. (2015) found that increases in aesthetic judgments were a function of
the relevance of the information presented as part of the task instructions. Similarly, Leder et al. (2006)
found aesthetic valuations were higher when the provided information made the painting more meaning-
ful to the viewer, with the amount of modulation dependent on the level of the participant’s subjective
agreement with the context provided. In contrast, elaborative titles (such as “Hibernation”), compared to
plainly descriptive titles (such as “Houses in Snow”), did not affect judgments of liking or personal meaning,
interest, emotional affection, and thoughts evoked by the artwork but only the feeling of understanding the artist’s
intention.
The ambiguity of the information plays also a role in shaping aesthetic experiences. In a study investigat-
ing the effects of ambiguity, the experimenter provided multiple different descriptive words that were pre-
sented together with pieces of art, and the participants were asked to rate whether each statement matched
the painting, as well as providing ratings of liking and interestingness ( Jakesch & Leder, 2009). Results showed
that a certain amount of ambiguity affected liking scores. If the ambiguity became too high, however,
the effect reversed, suggesting an inverted U-shaped effect. Furthermore, different vehicles of context also
affect the aesthetic experience. When comparing contexts provided by background information to contexts
through real-world experiences in a study by Szubielska and colleagues (2019), specific aspects of aesthetic
ratings (emotional valence, arousal, dominance, origin heart vs mind, significance) were found to be dependent on
the type of context provided.
Finally, it is well known that framing can also modulate aesthetic appraisals (see also Chapter 2 and Chap-
ter  25). Frames are object-external descriptions of the object being assessed which can take the form of
labels, price information, and so on. For instance, Kirk and colleagues (2009) showed that framing abstract
art as either examples of “prestigious gallery work” or “work generated by a computer” influenced aesthetic
ratings: participants reporting higher liking for images when they were labelled with the former than for the
latter. Kirk et al. (2009) also found that when images were presented with the “gallery” frame, activity in
the orbitofrontal cortex was higher than when they were presented with the “computer” frame. Similarly,
framing photography as “art” enhanced liking compared to framing it as “journalism” (Gerger et al., 2014).
Interestingly, the framing used in this study (“art” vs. “journalism”) increased the positive aspects of the
aesthetic experience (making the photos more joyful and more liked) for the negative images and decreased
them (less joyful and less liked) for the positive images, but did not affect negative emotions (anger, disgust, fear,
sadness, and shame). This finding has been ecologically validated in a quasi-experimental field study (Wagner
et al., 2016) and provides evidence for the effect of psychological distancing on aesthetic experiences (see
also Leder et al., 2014).
In sum, both the complexity of the context and the content of the information can have significant effects
on certain aspects of an aesthetic experience. Context has the potential to intensify the aesthetic experience,
and information has the potential to regulate it more specifically, depending on the specific cognitive pro-
cesses the content induces. It seems as if an inherent feature of art is its embeddedness in complex context
that provokes integration of information, resolution of ambiguity, and multiple different emotional responses
which may, in cumulation, lead to more intense aesthetic experiences. The lack of systematic investigation
of the complexity of contexts as well as the ambiguity of self-rating scales used to assess aesthetic experiences
makes it difficult, at this time, to present a clearer picture of the effects of complex contexts and information
on the aesthetic experience of art.

442
Complexity of aesthetic experiences

Conclusions and outlook


The evidence we have reviewed here suggests that aesthetic experiences are generated by a system that inte-
grates information delivered by the specific object; the motivation and inner emotional state of the beholder;
and the embeddedness in a complex physical, cultural, and social context. There are several theoretical
approaches that describe the processes involved in aesthetic experiences of art (Leder et al., 2004; Jacobsen,
2006; Pelowski et al., 2017). The potential configurations and capacity of this system to process information
and respond are constrained by its structure but also depend on the total external input it receives. Beyond
the nature of the aesthetic object, such as an artwork, all other surrounding features can elicit some effects.
Therefore, further studies of artworks in systematically varying contexts (laboratories, homes, museums, gal-
leries) are required.
This system responds to all input and is thus a domain-general system including the hedonic hotspots
discussed previously. Nevertheless, art objects are always embedded within complex contexts that contribute
to the way they are processed by also modulating expectations and semantic information. Consequently,
“exceptional” states can emergence as a function of domain-general systems, giving rise to intense, complex,
and sometimes even transformative aesthetic experiences. Also, recent approaches study psychological and
neuronal phenomena in a more complex setting (Pelowski et al., 2018, 2020), and some address the question
of whether an aesthetic experience in response to art is induced within a domain-general system or is actually
an exceptional state of a very domain-specific system (see Nadal & Skov, 2018).
Therefore, due to its increasing pool of knowledge and interdisciplinary research strategies, psychology
and neuroscience have begun to overcome the initial phase of investigating single aspects of the aesthetic
experience in isolation. In this vein, there is a steady increase in the number of studies in which art is studied
in realistic contexts to see how art is differently processed in the museum when directly comparing museum
settings with more controlled laboratory conditions (Brieber et al., 2015; Grüner et al., 2019). Other stud-
ies exclusively focused on the museum context, for example, to study conceptual art, such as Olafur Elias-
son’s room-filling installations, that can only be fully experienced in natural, usually in museum contexts
(Pelowski et al., 2018). This study is also an example of systematic combinations of evaluations using newly
developed questionnaires and sets of scales, with on-site measures of physiology (eye-tracking). Progress in
lighter, cheaper, and more efficient portable measures of physiological and neurophysiological tools enable
more and more testing in natural field settings.
In this chapter, we have discussed the effects of context on aesthetic art appreciation jointly from the
perspectives of cognitive psychology and neuroscience. We have shown how context influences behavioral
and neural responses and how these findings have spurred a trend towards more field studies in empiri-
cal aesthetics. The ultimate goal of this trend is to provide a thorough explanation of the exciting inter-
play between objects and environment in general and, in empirical aesthetics, the context-dependence of
dynamically emerging hedonic states. Empirical aesthetics is also a fascinating field because art appreciation
ideally represents examples of individual variation in the way in which specific objects—artworks—resonate
with individual states of previous experiences, individual taste, current emotional states, or knowledge about
the artwork (Leder et al., 2004). Despite these recent developments, many questions remain unanswered
and require a wholistic approach accounting for the complexity of the environment, the brain, and their
interaction.

References
Anderson, D. J.,  & Adolphs, R. (2014). A  framework for studying emotions across species. Cell, 157(1), 187–200.
https://doi.org/10.1016/j.cell.2014.03.003
Bannister, S. (2019). Distinct varieties of aesthetic chills in response to multimedia. PLoS One, 14(11), e0224974. https://
doi.org/10.1371/journal.pone.0224974

443
Julia Crone and Helmut Leder

Belke, B., Leder, H., Strobach, T., & Carbon, C. C. (2010). Cognitive fluency: High-level processing dynamics in art
appreciation. Psychology of Aesthetics, Creativity, and the Arts, 4(4), 214–222. https://doi.org/10.1037/a0019648
Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., . . . Starr. (2019). Dynamics of aes-
thetic experience are reflected in the default-mode network. NeuroImage, 188, 584–597. https://doi.org/10.1016/j.
neuroimage.2018.12.017
Berridge, K. C.,  & Kringelbach, M. L. (2015). Pleasure systems in the brain. Neuron, 86(3), 646–664. https://doi.
org/10.1016/j.neuron.2015.02.018
Berridge, K. C., Robinson, T. E., & Aldridge, J. W. (2009). Dissecting components of reward: “liking”, “wanting”, and
learning. Current Opinion in Pharmacology, 9(1), 65–73. https://doi.org/10.1016/j.coph.2008.12.014
Blood, A. J., & Zatorre, R. J. (2001). Intensely pleasurable responses to music correlate with activity in brain regions
implicated in reward and emotion. Proceedings of the National Academy of Sciences of the United States of America, 98(20),
11818–11823. https://doi.org/10.1073/pnas.191355898
Brauer, L. H., & De Wit, H. (1997). High dose pimozide does not block amphetamine-induced euphoria in normal
volunteers. Pharmacology Biochemistry and Behavior, 56(2), 265–272. https://doi.org/10.1016/S0091-3057(96)00240-7
Brieber, D., Nadal, M., & Leder, H. (2015). In the white cube: Museum context enhances the valuation and memory of
art. Acta Psychologica (Amst), 154, 36–42. https://doi.org/10.1016/j.actpsy.2014.11.004
Brieber, D., Nadal, M., Leder, H., & Rosenberg, R. (2014). Art in time and space: Context modulates the relation
between art experience and viewing time. PLoS One, 9(6), e99019. https://doi.org/10.1371/journal.pone.0099019
Brielmann, A. A., & Pelli, D. G. (2017). Beauty requires thought. Current Biology, 27(11), 1706. https://doi.org/10.1016/j.
cub.2017.05.045
Castro, D. C., & Berridge, K. C. (2014). Opioid hedonic hotspot in nucleus accumbens shell: Mu, delta, and kappa
maps for enhancement of sweetness “liking” and “wanting”. Journal of Neuroscience, 34(12), 4239–4250. https://doi.
org/10.1523/JNEUROSCI.4458-13.2014
Castro, D. C., & Berridge, K. C. (2017). Opioid and orexin hedonic hotspots in rat orbitofrontal cortex and insula.
Proceedings of the National Academy of Sciences of the United States of America, 114(43), E9125–E9134. https://doi.
org/10.1073/pnas.1705753114
Chatterjee, A., & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18, 370–375.
Che, J., Sun, X., Skov, M., Vartanian, O., Rosselló, J., & Nadal, M. (2021). The role of working memory capacity in
evaluative judgments of liking and beauty. Cognition and Emotion, 35(7), 1407–1415. https://doi.org/10.1080/0269
9931.2021.1947781
Childress, A. R., Ehrman, R. N., Wang, Z., Li, Y., Sciortino, N., Hakun, J., .  .  . O’Brien. (2008). Prelude to pas-
sion: Limbic activation by “unseen” drug and sexual cues. PLoS One, 3(1), e1506. https://doi.org/10.1371/journal.
pone.0001506
Cupchik, G. C. (2002). The evolution of psychical distance as an aesthetic concept. Culture and Psychology, 8(2), 155–187.
https://doi.org/10.1177/1354067X02008002437
Dissanayake, E. (2008). The arts after Darwin: Does art have an origin and adaptive function. I K. Zijlmans and W. van Damme
(Red.), World Art Studies: Exploring Concepts and approaches (s. 241–263). Valiz.
Dunlop, B. W., & Nemeroff, C. B. (2007). The role of dopamine in the pathophysiology of depression. Archives of General
Psychiatry, 64(3), 327–337. https://doi.org/10.1001/archpsyc.64.3.327
Frijda, N. H. (1988). The laws of emotion. American Psychologist, 43(5), 349–358. https://doi.
org/10.1037/0003-066x.43.5.349
Galistu, A., & D’Aquila, P. S. (2012). Effect of the dopamine D1-like receptor antagonist SCH 23390 on the micro-
structure of ingestive behaviour in water-deprived rats licking for water and NaCl solutions. Physiology and Behavior,
105(2), 230–233. https://doi.org/10.1016/j.physbeh.2011.08.006
Georgiadis, J. R., & Kringelbach, M. L. (2012). The human sexual response cycle: Brain imaging evidence linking sex to
other pleasures. Progress in Neurobiology, 98(1), 49–81. https://doi.org/10.1016/j.pneurobio.2012.05.004
Gerger, G., & Leder, H. (2015). Titles change the esthetic appreciations of paintings. Frontiers in Human Neuroscience, 9,
464. https://doi.org/10.3389/fnhum.2015.00464
Gerger, G., Leder, H., & Kremer, A. (2014). Context effects on emotional and aesthetic evaluations of artworks and IAPS
pictures. Acta Psychologica (Amst), 151, 174–183. https://doi.org/10.1016/j.actpsy.2014.06.008
Gilbert, D. T., & Wilson, T. D. (2009). Why the brain talks to itself: Sources of error in emotional prediction. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 364(1521), 1335–1341. https://doi.org/10.1098/
rstb.2008.0305
Grewe, O. (2007). Listening to music as A re-creative process: Physiological, psychological, and psychoacoustical cor-
relates of chills and strong emotions. Music Perception, 24(3), 297–314. https://doi.org/10.1525/mp.2007.24.3.297

444
Complexity of aesthetic experiences

Grewe, O., Kopiez, R., & Altenmüller, E. (2009). Chills as an indicator of individual emotional peaks. Annals of the New
York Academy of Sciences, 1169, 351–354. https://doi.org/10.1111/j.1749-6632.2009.04783.x
Gross, J. J. (1998). Antecedent- and response-focused emotion regulation: Divergent consequences for expe-
rience, expression, and physiology. Journal of Personality and Social Psychology, 74(1), 224–237. https://doi.
org/10.1037/0022-3514.74.1.224
Grüner, S., Specker, E., & Leder, H. (2019). Effects of context and genuineness in the experience of art. Empirical Studies
of the Arts, 37(2), 138–152. https://doi.org/10.1177/0276237418822896
Ho, C. Y.,  & Berridge, K. C. (2014). Excessive disgust caused by brain lesions or temporary inactivations: Mapping
hotspots of the nucleus accumbens and ventral pallidum. European Journal of Neuroscience, 40(10), 3556–3572. https://
doi.org/10.1111/ejn.12720
Huang, M., Bridge, H., Kemp, M. J., & Parker, A. J. (2011). Human cortical activity evoked by the assignment of authen-
ticity when viewing works of art. Frontiers in Human Neuroscience, 5, 134. https://doi.org/10.3389/fnhum.2011.00134
Jacobsen, T. (2006). Bridging the arts and sciences: A  framework for the psychology of aesthetics. Leonardo, 39(2),
155–162. https://doi.org/10.1162/leon.2006.39.2.155
Jakesch, M., & Leder, H. (2009). Finding meaning in art: Preferred levels of ambiguity in art appreciation. Quarterly
Journal of Experimental Psychology, 62(11), 2105–2112. https://doi.org/10.1080/17470210903038974
Khan, H. A., Urstadt, K. R., Mostovoi, N. A., & Berridge, K. C. (2020). Mapping excessive “disgust” in the brain:
Ventral pallidum inactivation recruits distributed circuitry to make sweetness “disgusting”. Cognitive, Affective and
Behavioral Neuroscience, 20(1), 141–159. https://doi.org/10.3758/s13415-019-00758-4
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009). Modulation of aesthetic value by semantic con-
text: An fMRI study. Neuroimage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009
Kringelbach, M. L., Stein, A.,  & van Hartevelt, T. J. (2012). The functional human neuroanatomy of food pleasure
cycles. Physiology and Behavior, 106(3), 307–316. https://doi.org/10.1016/j.physbeh.2012.03.023
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811
Leder, H., Carbon, C. C., & Ripsas, A. L. (2006). Entitling art: Influence of title information on understanding and
appreciation of paintings. Acta Psychologica (Amst), 121(2), 176–198. https://doi.org/10.1016/j.actpsy.2005.08.005
Leder, H., Gerger, G., Brieber, D., & Schwarz, N. (2014). What makes an art expert? Emotion and evaluation in art
appreciation. Cognition and Emotion, 28(6), 1137–1147. https://doi.org/10.1080/02699931.2013.870132
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Leyton, M., aan het Rot, M., Booij, L., Baker, G. B., Young, S. N., & Benkelfat, C. (2007). Mood-elevating effects of
d-amphetamine and incentive salience: The effect of acute dopamine precursor depletion. Journal of Psychiatry and
Neuroscience, 32(2), 129–136.
Marković, S. (2012). Components of aesthetic experience: Aesthetic fascination, aesthetic appraisal, and aesthetic emo-
tion. i-Perception, 3(1), 1–17. https://doi.org/10.1068/i0450aap
Millis, K. (2001). Making meaning brings pleasure: The influence of titles on aesthetic experiences. Emotion, 1(3),
320–329. https://doi.org/10.1037/1528-3542.1.3.320
Mocaiber, I., Pereira, M. G., Erthal, F. S., Machado-Pinheiro, W., David, I. A., Cagy, M., . . . de Oliveira. (2010). Fact or
fiction? An event-related potential study of implicit emotion regulation. Neuroscience Letters, 476(2), 84–88. https://
doi.org/10.1016/j.neulet.2010.04.008
Nadal, M., & Skov, M. (2018). The pleasure of art as a matter of fact. Proceedings. Biological Sciences, 285(1875), 20172252.
http://doi.org/10.1098/rspb.2017.2252
Noë, A. (2009). Out of our heads: Why you are not your brain, and other lessons from the biology of consciousness. Hill and Wang.
Park, S. A., Yun, K., & Jeong, J. (2015). Reappraising abstract paintings after exposure to background information. PLoS
One, 10(5), e0124159. https://doi.org/10.1371/journal.pone.0124159
Peciña, S., Berridge, K. C., & Parker, L. A. (1997). Pimozide does not shift palatability: Separation of anhedonia from
sensorimotor suppression by taste reactivity. Pharmacology Biochemistry and Behavior, 58(3), 801–811. https://doi.
org/10.1016/S0091-3057(97)00044-0
Pelowski, M. (2015). Tears and transformation: Feeling like crying as an indicator of insightful or “aesthetic” experience
with art. Frontiers in Psychology, 6, 1006. https://doi.org/10.3389/fpsyg.2015.01006
Pelowski, M., Leder, H., Mitschke, V., Specker, E., G., Tinio, P., Vaporova, E., Bieg, T., & Husslein-Arco, A. (2018).
Capturing aesthetic experiences with installation art: An empirical assessment of emotion, evaluations, and mobile
eye tracking in Olafur Eliasson’s “Baroque, Baroque!” Frontiers in Psychology, 9.1255.10.3389/fpsyg.2018.01255

445
Julia Crone and Helmut Leder

Pelowski, M., Liu, T., Palacios, V., & Akiba, F. (2014). When a body meets a body: An exploration of the negative impact
of social interactions on museum experiences of art. International Journal of Education and the Arts, 15(14). http://www.
ijea.org/v15n14/.
Pelowski, M., Markey, P., Forster, M., Gerger, G., & Leder, H. (2017). Move me, astonish me . . . delight my eyes and
brain: The Vienna Integrated Model of top-down and bottom-up processes in Art Perception (VIMAP) and cor-
responding affective, evaluative, and neurophysiological correlates. Physics of Life Reviews, 21, 80–125. https://doi.
org/10.1016/j.plrev.2017.02.003
Pelowski, M., Markey, P., & Leder, H. (2018). Chills, aesthetic experience, and new versus old knowledge—What do
chills actually portend? Comment on “Physics of mind: experimental confirmations of theoretical predictions” by.
Physics of Life Reviews, 25, 83–87. https://doi.org/10.1016/j.plrev.2018.03.014
Pelowski, M., Specker, E., Gerger, G., Leder, H., & Weingarden, L. S. (2020). Do you feel like I do? A study of sponta-
neous and deliberate emotion sharing and understanding between artists and perceivers of installation art. Psychology
of Aesthetics, Creativity, and the Arts, 14(3), 276–293.  https://doi.org/10.1037/aca0000201
Piff, P. K., Dietze, P., Feinberg, M., Stancato, D. M., & Keltner, D. (2015). Awe, the small self, and prosocial behavior.
Journal of Personality and Social Psychology, 108(6), 883–899. https://doi.org/10.1037/pspi0000018
Reynolds, S. M., & Berridge, K. C. (2008). Emotional environments retune the valence of appetitive versus fearful func-
tions in nucleus accumbens. Nature Neuroscience, 11(4), 423–425. https://doi.org/10.1038/nn2061
Richard, J. M., & Berridge, K. C. (2011). Nucleus accumbens dopamine/glutamate interaction switches modes to gen-
erate desire versus dread: D(1) alone for appetitive eating but D(1) and D(2) together for fear. Journal of Neuroscience,
31(36), 12866–12879. https://doi.org/10.1523/JNEUROSCI.1339-11.2011
Richard, J. M., & Berridge, K. C. (2013). Prefrontal cortex modulates desire and dread generated by nucleus accumbens
glutamate disruption. Biological Psychiatry, 73(4), 360–370. https://doi.org/10.1016/j.biopsych.2012.08.009
Russell, P. A. (2003). Effort after meaning and the hedonic value of paintings. British Journal of Psychology, 94(1), 99–110.
https://doi.org/10.1348/000712603762842138
Salimpoor, V. N., Benovoy, M., Larcher, K., Dagher, A., & Zatorre, R. J. (2011). Anatomically distinct dopamine release
during anticipation and experience of peak emotion to music. Nature Neuroscience, 14(2), 257–262. https://doi.
org/10.1038/nn.2726
Salimpoor, V. N., Benovoy, M., Longo, G., Cooperstock, J. R.,  & Zatorre, R. J. (2009). The rewarding aspects of
music listening are related to degree of emotional arousal. PLoS One, 4(10), e7487. https://doi.org/10.1371/journal.
pone.0007487
Schoeller, F., & Perlovsky, L. (2016). Aesthetic chills: Knowledge-acquisition, meaning-making, and aesthetic emotions.
Frontiers in Psychology, 7, 1093. https://doi.org/10.3389/fpsyg.2016.01093
Schoeller, F., Perlovsky, L., & Arseniev, D. (2018). Physics of mind: Experimental confirmations of theoretical predic-
tions. Physics of Life Reviews, 25, 45–68. https://doi.org/10.1016/j.plrev.2017.11.021
Skov, M. (2019). Aesthetic appreciation: The view from neuroimaging. Empirical Studies of the Arts, 37, 220–248. https://
doi.org/https://doi.org/10.1177/0276237419839257
Skov, M. (2020). The neurobiology of sensory valuation. In M. Nadal & O. Vartanian (Eds.), The Oxford handbook of
empirical aesthetics. Oxford University Press.
Skov, M., & Nadal, M. (2020). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Smith, K. S., & Berridge, K. C. (2005). The ventral pallidum and hedonic reward: Neurochemical maps of sucrose “liking”
and food intake. Journal of Neuroscience, 25(38), 8637–8649. https://doi.org/10.1523/JNEUROSCI.1902-05.2005
Smith, K. S.,  & Berridge, K. C. (2007). Opioid limbic circuit for reward: Interaction between hedonic hotspots of
nucleus accumbens and ventral pallidum. Journal of Neuroscience, 27(7), 1594–1605. https://doi.org/10.1523/
JNEUROSCI.4205-06.2007
Smith, K. S., Berridge, K. C., & Aldridge, J. W. (2011). Disentangling pleasure from incentive salience and learning
signals in brain reward circuitry. Proceedings of the National Academy of Sciences of the United States of America, 108(27),
E255–E264. https://doi.org/10.1073/pnas.1101920108
Specker, E., Tinio, P. P. L., & van Elk, M. (2017). Do you see what I see? An investigation of the aesthetic experience
in the laboratory and museum see what I see? Psychology of Aesthetics, Creativity, and the Arts, 11(3), 265–275. https://
doi.org/10.1037/aca0000107
Steier, R., Pierroux, P., & Krange, I. (2015). Embodied interpretation: Gesture, social interaction, and meaning makingin
a nationalart museum. Learning, Culture and Social Interaction, 7, 28–42.  http://doi.org/10.1016/j.lcsi.2015.05.002
Swami, V. (2013). Context matters: Investigating the impact of contextual information on aesthetic appreciation of
paintings by Max Ernst and Pablo Picasso. Psychology of Aesthetics, Creativity, and the Arts, 7(3), 285–295. https://doi.
org/10.1037/a0030965

446
Complexity of aesthetic experiences

Szubielska, M., Imbir, K., & Szymańska, A. (2019). The influence of the physical context and knowledge of artworks on
the aesthetic experience of interactive installations. Current Psychology, 40(8), 3702–3715. https://doi.org/10.1007/
s12144-019-00322-w
Vartanian, O., & Skov, M. (2014). Neural correlates of viewing paintings: Evidence from a quantitative meta-analysis
of functional magnetic resonance imaging data. Brain and Cognition, 87, 52–56. https://doi.org/10.1016/j.bandc.
2014.03.004
Vessel, E. A., Starr, G. G., & Rubin, N. (2012). The brain on art: Intense aesthetic experience activates the default mode
network. Frontiers in Human Neuroscience, 6, 66. https://doi.org/10.3389/fnhum.2012.00066
Vessel, E. A., Starr, G. G., & Rubin, N. (2013). Art reaches within: Aesthetic experience, the self and the default mode
network. Frontiers in Neuroscience, 7, 258. https://doi.org/10.3389/fnins.2013.00258
Wagner, V., Klein, J., Hanich, J., Shah, M., Menninghaus, W., & Jacobsen, T. (2016). Anger framed: A field study on
emotion, pleasure, and art. Psychology of Aesthetics, Creativity, and the Arts, 10(2), 134–146. https://doi.org/10.1037/
aca0000029
Williams, P. G., Johnson, K. T., Curtis, B. J., King, J. B., & Anderson, J. S. (2018). Individual differences in aesthetic
engagement are reflected in resting-state fMRI connectivity: Implications for stress resilience. Neuroimage, 179, 156–
165. https://doi.org/10.1016/j.neuroimage.2018.06.042
Winkielman, P., Berridge, K. C., & Wilbarger, J. L. (2005). Unconscious affective reactions to masked happy versus
angry faces influence consumption behavior and judgments of value. Personality and Social Psychology Bulletin, 31(1),
121–135. https://doi.org/10.1177/0146167204271309
Yager, L. M., Garcia, A. F., Wunsch, A. M., & Ferguson, S. M. (2015). The ins and outs of the striatum: Role in drug
addiction. Neuroscience, 301, 529–541. https://doi.org/10.1016/j.neuroscience.2015.06.033
Zeki, S. (2013). Clive Bell’s “Significant Form” and the neurobiology of aesthetics. Frontiers in Human Neuroscience, 7,
730. https://doi.org/10.3389/fnhum.2013.00730

447
24
EXPERIENCING ART IN SOCIAL
SETTINGS
Haeeun Lee and Guido Orgs

Whether in galleries, theatres, concert halls or out on the street, art is often experienced together, not alone.
The performing arts, including dance, theatre and music, have likely evolved in social settings (Brown et al.,
2005; Merker et al., 2015), and practicing them facilitates social bonding (Dunbar, 2012; Savage et al., 2015;
Von Zimmermann et al., 2018). However, while research in aesthetics has thrived in the last two decades
(Chatterjee & Vartanian, 2014; Kawabata & Zeki, 2004), the sociality of aesthetic experience remains under-
investigated. Most existing studies (and theories) exclusively focus on aesthetic experiences of individual
people in isolation. Only in recent years has the social context of art experience come into focus, when
people collectively watch a film in the cinema or attend a live concert, a dance performance or a play in the
theatre (Hanich, 2014). In this chapter, we will review behavioural and neuroimaging studies that explore
art experiences in these social settings.
This social gap in the empirical aesthetics literature may be due to the methodological constraints of neu-
roimaging methods and laboratory settings, which typically can only record measurements from one person
at a time. Aptly, Schilbach et  al. (2013) labelled measuring the collective experience as the “dark matter”
of social neuroscience. Recently, hyperscanning, that is, simultaneously recording the brain activities from
more than one subject interacting (for review, see Babiloni & Astolfi, 2014; Hasson et al., 2012), and mobile
neuroimaging approaches such as functional near-infrared spectroscopy (fNIRS, e.g., Pinti et al., 2020) and
mobile electroencephalography (EEG, e.g., Mustile et al., 2021) have allowed researchers to conduct stud-
ies outside of the conventional laboratory environment, moving into socially dynamic, real-world settings.
These technologies can thus be applied to the collective aesthetic experience and record the brain activity of
multiple audience members in live or virtual performance contexts (Hamilton et al., 2018).
Existing research on the collective aesthetic experience often manipulates the number of people experi-
encing art together (e.g., attending to artwork alone or in a group, Kaltwasser et al., 2019) or the social con-
text in which art appreciation occurs, for example, the degree of shared experience (Dunand et al., 1984) or
the layout of a gallery space (Pelowski et al., 2014). Measurement of the collective experience often takes the
form of quantifying interpersonal synchrony, that is, the temporal and spatial coordination of behavioural,
cognitive or neurophysiological processes between two or more people (Czeszumski et al., 2020; Delaherche
et al., 2012; Mayo & Gordon, 2020). Interpersonal synchrony can emerge automatically in social situations,
as when large crowds clap in synchrony (Néda et al., 2000). Notably, coordinating actions with others is
rewarding in itself (Kokal et al., 2011), an experience that has been termed “collective effervescence,” or the
joy of belonging (Durkheim, 1912; Rennung & Göritz, 2016).

448 DOI: 10.4324/9781003008675-26


Experiencing art in social settings

In this chapter, we will review social aesthetic appreciation across different art domains. We will discuss
social influences on the aesthetics of dance and theatre, live and recorded music and watching film. We
will focus specifically on studies that have either: (1) investigated co-presence between a performer and a
spectator; (2) directly compared individual and social settings of aesthetic appreciation; (3) investigated inter-
personal synchrony between people attending to the same artwork, recorded separately; or (4) investigated
interpersonal synchrony among audience members experiencing art together.

Social influences on the aesthetics of the live performing arts


(dance and theatre)
Before the arrival of recording technologies, the performing arts could only be experienced “live” (Auslander,
1999), in a defined space and time where and when performers and spectators share their experience (Orgs
et  al., 2016). In live performances, performers and the audience are typically co-present (Fischer-Lichte,
2008). Indeed, one study suggests that co-presence of performer and spectator can produce greater corti-
cal excitability than the experience of a recorded performance which does not involve co-presence ( Jola &
Grosbras, 2013). In recent years, there has been a growing number of studies conducted during live per-
formances, in which physiological and behavioural responses of the audience and performers are recorded.
Live performance research has studied the role of audience size (Lemasson et al., 2018; Rennung & Göritz,
2016), the spatial layout of the audience seats (Lemasson et al., 2019), the social closeness between the per-
former and the spectator, active engagement of the participant (Konvalinka et al., 2011) or attentional focus
(Bachrach et al., 2015; Himberg et al., 2018).
Another social aspect of the performing arts is watching a group of performers act or dance together.
Vicary et al. (2017) studied the influence of synchronous movement among a group of dancers on physi-
ological and continuous behavioural responses of spectators in contemporary dance choreography which
continuously manipulated synchrony. Rather than a fixed series of steps, dancers in this study worked with a
defined movement vocabulary that emerged from a set of task instructions that relied on the social interac-
tions and decisions made among the performers. Across four live dance performances, 101 participants were
equipped with wrist sensors that recorded cardiac activity while they attended to the performance, as an
implicit index of engagement and arousal. A subset of participants also evaluated synchrony and enjoyment
during the performance using a tablet. For the dancers, their movement acceleration in three-dimensional
space was recorded during the performances. Performed group synchrony was computed from the dancers’
acceleration data with cross recurrence quantification analysis (CRQA, Fusaroli et al., 2014). It was found
that, although audience members gave homogenous subjective ratings within their groups (65–70%), four
performances elicited different levels of subjective enjoyment, as one performance was most favoured (third),
and another was least favoured (first) by the audiences. Granger causality analysis (an analysis identifying
directional correlation between two time series data; Dean & Bailes, 2010) revealed that synchrony among
dancers predicted the spectators’ heart rate and enjoyment, but only for the most and the least favoured per-
formances, whereas for performances with intermediate levels of enjoyment, the relationships were absent or
even reversed, suggesting that forming a strong and stable evaluation of a live performance is associated with
sustained temporal coupling of the spectators’ reactions to the performers’ actions. In sum, the findings show
that the aesthetic appreciation of live dance indeed depends on the co-presence of performers and spectators
in a defined here and now and that movement synchrony is an important feature of the aesthetics of group
dancing.
The communicative effect between the performer and the audience in live performing arts settings is not
necessarily a one-way interaction from the performer to the spectator. The performers can also be influ-
enced, for example, by the size or composition of the audience and their reactions. Lemasson and research-
ers (2019) focused on studying spatial positioning of live audiences and its impact on both the actors and

449
Haeeun Lee and Guido Orgs

spectators’ implicit and explicit reactions. In theatres, proscenium arch or end-on stage seating arrangements
are common, where the audience faces the stage only frontally. Alternative traverse (bi-frontal), in-the-round
(quadri-frontal) or other configurations that blur the distinction between stage and the audience are often
associated with a more immersive experience or audience participation. Lemasson et al. investigated three
different types of staging on emotional responses of the actors and spectators. Fifteen acting school students
performed a combination of monologues and dialogues in a fixed order, repeated in three performances,
each time in a different staging condition. Different audience members were recruited for each day (for
frontal = 35, bi-frontal = 28, quadri-frontal = 43 participants). For both actors and spectators, galvanic skin
responses (GSRs, Bodie, 2010) were recorded during the performance as a measure of arousal. At the end of
the performances, spectators and actors were asked to rate their aesthetic experience.
The results showed that the actors preferred to perform while being surrounded by the audience. In this
condition, performers felt that they interacted the most with the audience and that the audience emotions
were most heightened, although they also reported that they were more anxious in the bi-frontal and quadri-
frontal layout. Reciprocally, actors’ GSRs were highest in the quadri-frontal layout. Similarly, spectators
reported highest attention in the quadri-frontal staging and the lowest attention in the bi-frontal staging.
However, spectators reported to have interacted the most with other spectators in the bi-frontal staging, and
their GSRs were significantly higher in this configuration compared to the other two. Overall, the findings
indicate that there is a disparity of felt and speculated affect, interaction and attention between the actors and
the spectators. However, the sample size in this study was small; more research is needed to investigate the
effect of spatial layout of the performance space on the sociality of aesthetic experience.
Ardizzi et al. (2020) investigated cardiac synchrony between spectators during the live performance of
a monologue, in groups of four people, and studied how implicit physiological synchronization could be
related to subjective reports of emotional intensity. The electrocardiogram (ECG) was recorded from 12
professional actors and 12 quartets of spectators before and during the monologue performances. After the
monologue, ratings for emotional intensity and quality of the monologues were collected. Cardiac synchrony
was significantly higher for groups of spectators who saw the performance together (in-group) than for
those who were in different quartets (out-group), both during the performance and rest periods. Contrary
to their hypothesis, the authors did not find a positive correlation between cardiac synchrony and emotional
intensity ratings. However, further analysis revealed that the convergence score, that is, the similarity between
emotional ratings of each participant and the other in-group spectators, correlated with in-group heart rate
synchrony. Thus, the more similarly people had rated their experience, the more synchronous their heart
rates. These findings suggest that the mere presence of others in attending theatrical performance does syn-
chronize behaviour and psychophysiology, but this synchronization does not necessarily lead to the experi-
ence of more intense emotions.
Other studies have directly compared art appreciation between individuals and groups. Nomura and col-
leagues (2015a, 2015b) investigated unconscious eyeblink synchronization among spectators as an index of
cognitive entrainment while watching traditional Japanese Rakugo theatre. In Rakugo, a story is narrated by a
single actor, with a specific focus on the performer’s voice and facial expressions. In a first study, Nomura et al.
(2015a) found that expert viewers of Rakugo showed significantly higher eyeblink synchrony at the beginning
of the performance compared to novice spectators. Both groups watched a 50-minute recorded performance
in the lab. However, novice participants’ eyeblinks became synchronized over the course of the performance,
and the difference between the two groups disappeared in the last 15 minutes. In a follow-up study, Nomura
et al. (2015b) compared eye motion synchrony during a live and a recorded Rakugo performance. Eyeblink
synchronization was significantly stronger in the live social condition than the alone lab condition for both
expert and novice spectators. For expert spectators, eye blink synchronization was moderately stronger in the
live social condition compared to the alone recorded condition. For novice spectators, eye blink synchrony
during in the live social condition was significantly stronger from the beginning and throughout the entire

450
Experiencing art in social settings

performance compared to the alone recorded condition. Such results suggest that a live and social perfor-
mance context facilitates eye blink synchronization, especially if spectators have little or no prior experience
with a specific art form or style. Perhaps other spectators’ reactions to a performance (e.g., laughter) provide
a cue to engaging with art forms that are more difficult to understand if experienced alone.
Together, these studies show a clear influence of liveness and social context on the appreciation of dance
and theatre. Future studies should aim to further disentangle interpersonal synchrony between the perform-
ers and the audience at neural, physiological and behavioural levels.

Social influences on the aesthetics of live and recorded music


In contrast to dance and theatre, listening to recorded music in personal devices has become the dominant
format of music consumption (Bull, 2006; Wald-Fuhrmann et  al., 2021). But live music events remain
popular, particularly among people who seek novel experiences and social connectedness (Brown & Knox,
2017). Accordingly, many music listeners rate live concerts as their favourite way of engaging with music
(Krause et al., 2020).
In one of the first studies on social influences on music listening, Egermann et al. (2011) compared lis-
tening to music in solitary versus group conditions. The study investigated whether the social setting would
change subjective or physiological measures of emotion. Fourteen participants from an amateur orchestra
listened to 10 sets of one-minute classical music excerpts which were previously found to induce aesthetic
chills (Grewe et al., 2007). Participants listened to these excerpts once alone and once again in a group.
Subjective reports of emotional reaction were taken at the end of each musical excerpt using felt intensity
questionnaires (Krumhansl, 1997). Aesthetic chills, described to the participants as “strong emotions accom-
panied by shivers down the spine or goosebumps,” were recorded both explicitly—participants pressing a
button with one hand whenever they experienced a chill—and implicitly by measuring skin conductance
response (SCR). Interestingly, Egermann et al. (2011) found that there was no significant social influence on
subjective emotional ratings. However, participants exhibited greater skin conductance in the solitary listen-
ing condition, across all 10 musical pieces, and the difference between the alone and social conditions peaked
when listeners reported that they experienced aesthetic chills. In fact, 11 out of 14 participants reported that
they enjoyed music listening more when they were alone.
This finding is important, as it indicates that social settings may not necessarily heighten audience engage-
ment but instead can have negative effects on aesthetic appreciation (Sutherland et al., 2009). Perhaps, the
presence of other people can be distracting if the musical work requires more sustained and focused atten-
tion to be appreciated. However, this experimental paradigm involved a button-pressing task to be carried
out individually even in the group listening condition, which is only partially comparable to real-life social
music listening in a concert. Moreover, recorded classical music in particular may be typically experienced
alone rather than in groups.
Shoda and researchers (2016) investigated cardiovascular responses of audiences attending to live vs.
recorded piano performance, both times as a group. Out of 118 audience members, 37 participants’ continu-
ous ECG was measured, first during a pianist playing six solos, then once again 10 weeks later, while partici-
pants listened again to the recorded version. The six musical pieces were presented in blocks of two tempi
(fast and slow) by three composers (Bach, Schumann, Debussy). The audience’s average HR changed in line
with the tempo of the music, but only in the live condition. During live concerts, participants’ heartbeat
became faster with fast-tempo music and slower with slow-tempo music. Moreover, the vagal nerve activi-
ties indicating stress reduction (the amount of high-frequency ECG components relative to total-frequency
components, Eckberg, 1997) were found to be significantly higher in the live condition, while sympatho-
vagal balance indicating mental stress (calculated as the amount of low-frequency components relative to
high-frequency components; Nakahara et al., 2009) was moderately lower in the live condition. Overall, this

451
Haeeun Lee and Guido Orgs

study supports the idea that a pianist’s live performance may not only lead to stronger entrainment of heart
rate in a group of audiences, inducing audience attention, but also has a greater calming effect on the listener
than listening to a recorded performance.
Bernardi et al., 2017 assessed interpersonal synchrony among group of 14 participants with and 13 partici-
pants without musical training. Over four days, both groups of participants listened to seven religious musical
pieces repeated twice, played live by a professional organist in a cathedral. All participants gave subjective
ratings on pleasure, familiarity and loudness variability after the first performance. During both baseline and
live performance conditions, participants’ ECG, breathing rate, finger vasomotion and blood pressure were
recorded simultaneously. Interpersonal physiological synchrony was computed as generalized partial directed
coherence (Baccalá et  al., 2007) for each of the four physiological measures in low- and high-frequency
bands. The effect of live organ performance on group synchrony depended on the specific music composi-
tions: For simple harmonic progressions, all physiological signals showed stronger synchronization in the
live performance compared to the baseline condition, while during a hymn, no differences on physiologi-
cal group synchrony were found between live performance and baseline. Music training background had a
significant effect on promoting stronger group synchronization in breathing and blood flow distribution, but
not in cardiovascular and blood pressure coherence. While subjective ratings of pleasantness and familiarity
were found not to be good predictors of group synchrony, loudness variability alone accounted up to 80%
of variance in all measures of physiological group synchrony. Specifically, audiences listening to music with
simpler patterns showed greater physiological synchrony than when they listened to loud music with a com-
plex structure. This study shows that live performed music can lead to greater synchronization of various
autonomic signals, but this effect strongly depends on musical structure, with simpler structure generating
stronger group synchrony.
The study by Bernardi et al. (2017) involved co-presence of the musician and the listener, but it did not
actually manipulate it. More recently, Belfi and researchers (2021) studied music appreciation while one
group of participants attended a live concert in a social setting and another group of participants listened to
recordings of the concert in a laboratory setting. Twenty participants who were allocated for the live condi-
tion watched a joint concert of a military band and a university symphony orchestra together. Separately, 12
participants were invited to watch the video recording of the concert alone in a laboratory setting. For both
live and lab conditions, overall ratings were measured once before, once in the intermission and once after
the performance, while participants provided continuous aesthetic ratings on their smartphones (Brielmann
et al., 2017) as they listened to four pieces of music: two songs played by the military band and two by the
university orchestra. One of the compositions had a patriotic theme; the other piece was non-patriotic.
The overall ratings did not significantly differ between live social vs. recorded alone conditions, except
for liking for the pieces played by the military band. Positive feelings listening to the military band increased
between the intermission and after the concert, but only in the live social condition. With continuous
measures, researchers found that both the average and highest peak of continuous pleasure depended on the
congruency of musicians with their repertoire. People reported higher pleasure listening to the military band
playing patriotic music and the university orchestra playing non-patriotic music. There was no overall differ-
ence between the live and recorded conditions. Additionally, in a follow-up study where participants listened
to recorded music only, without visual information to acknowledge which musical piece was played by
which band, there was no effect of congruency between musicians and their repertoire. This study suggests
that continuous pleasure of music depends more on who’s playing what rather than the performance context.
However, in this study, physical liveness (live vs. recorded) is confounded with social liveness (listening alone
or in a group). Future studies should aim to disentangle these two distinct influences on music appreciation.
Apart from physiological recordings and subjective measures, body movement of the audience has also
been explored in music appreciation in live vs. recorded events (Swarbrick et al., 2019; Jensenius et al., 2017).
Swarbrick et al. (2019) captured the head movements of audience members during a live rock concert and

452
Experiencing art in social settings

in a control condition in which the audience collectively listened to recordings of the same songs. After the
performance, researchers not only collected linking ratings for the songs but also for the band. Audience
members were both either unfamiliar with the band or already fans of the band prior to the experiment. All
participants were randomly allocated to either attend the live event or an album release event in which only
recorded songs were played. Participants’ head movement data were collected with motion capture during
live and recorded concerts which were held in the same venue. All songs except the last one were unfamiliar
to the audience.
Participants showed more vigorous head movements during the live concert compared to the recorded
concert, and this effect was greater if listeners were fans of the band. Notably, liveness did not influence head
movements during the performance of unfamiliar songs. However, during the well-known song, fans nod-
ded their heads more vigorously to the beat in the live than the recorded condition. In both conditions, fans
showed more vigorous head movements than listeners who did know the band in advance. Together, these
findings suggest that social influences on music appreciation are more pronounced if the audience has a col-
lective positive attitude towards the musicians and their music. However, in this study, the recorded condition
did not provide any visual information, that is, a recorded video of the concert. Accordingly, it is not fully
clear whether more vigorous and synchronous head movements in the live condition were driven by co-
presence of audience and band or by the availability of visual information, that is, watching the movements
of the performers on stage.
The studies summarized previously suggest that the co-presence of musicians and listeners physiological
and behavioural synchrony among audience members. However, does synchrony also occur at the neural
level? Yingying Hou et al. (2020) investigated neural synchrony between the musician and the listener and
its relationship with listener’s subjective appreciation. Researchers used dual fNIRS to record cerebral blood
flow of a violinist playing 12 musical pieces. Sixteen listeners watched a video recording of this performance.
After each piece, participants rated their subjective fondness towards the music. Inter-brain coherence (IBC)
was computed for each violinist-participant dyad. IBCs between the violinist and the listeners were consist-
ently higher than at resting state across the left temporal cortex, the right post-central and the inferior frontal
cortices. Aesthetic ratings significantly correlated with the IBCs from all four brain regions, and this correla-
tion was most pronounced at the performance.
Such results imply that brain-to-brain synchrony between the musician and the listener is stronger when
the listener enjoys the experience more and that this synchronization occurs in brain areas associated with
interpersonal communication and musical structural comprehension (Dai et al., 2018; Abrams et al., 2010).
Arguably, music can function as a social signal, and stronger neural synchrony may be related to passing inten-
tions or emotions from the performer to the listener via music. In line with this idea, Leong et al. (2017)
had conducted a brain imaging study with infants listening to pre-recorded and lively performed nursery
rhymes while manipulating social gaze. The researchers found that direct gaze between the singer (adult)
and the listener (infant) significantly increased the inter-brain coupling between the singer-listener dyads.
When infants were watching a pre-recorded video of the adult singing, this effect was unidirectional; that is,
the singer’s EEG signal preceded the infant’s EEG signal. However, in the co-present hyperscanning setting,
that is, when infants were in the same room as the adult, inter-brain connectivity was bidirectional. This
suggests that social gaze and the co-presence of the listener also impact the singer’s brain activity. Together,
Hou et al. (2020) and Leong et al. (2017) show that synchronized neural activity between a musician and a
listener correlates with enjoyment of the audience and involves brain areas associated with communication
social interaction.
While the two previous experiments studied inter-brain coupling between the musician and the listener,
numerous other studies have investigated inter-brain synchrony among participants listening to music on
their own. In these studies, inter-brain synchrony typically increases in brain areas related to auditory per-
ception, verbal communication, body movement and emotion processing (Trost et al., 2015; Abrams et al.,

453
Haeeun Lee and Guido Orgs

2013; Alluri et al., 2012). For example, Trost et al. (2015) identified musical sections that elicited the high-
est inter-subjective correlations (ISCs) among 17 participants while they listened to three classical music
pieces. These segments were in then characterized along nine musical features, including rhythmic varia-
tions, timber and spectral entropy. In a follow-up session, 14 new participants provided continuous ratings
of arousal and valence, all while listening to the same musical pieces. Greater ISCs were observed in key
brain regions associated with reward and affect states such as the insula and the ventral striatum. The amyg-
dala was also found to be significantly activated during inter-brain synchrony, but only in the pieces which
elicited the strongest emotional reactions. ISCs in the left insula, the amygdala and the right caudate nucleus
were positively correlated with arousal, while the activity in the same brain areas was negatively correlated
with valence. Additionally, the authors discovered that BOLD activation during inter-personal synchrony
correlated with objective musical features, and the direction of the relationships varied depending on the
regions—for example, increases in synchronized amygdala activations were found to be accompanied by an
increase in energy-related music features such as event density or entropy. More research linking brain syn-
chrony to both subjective experience and objective musical features in this way is needed in order to better
understand to what extent brain synchrony is driven by external, stimulus-related or internal, listener-related
factors.
The aforementioned studies found synchronized brain activity among audience members or between
the musician and the listener, and the strength of these shared neural activities in auditory, sensorimo-
tor, attention, emotion, aesthetic appreciation and communication brain areas was indeed associated with
music appreciation. Overall, studies show positive and negative influences on the effect of social settings on
music appreciation. Notably, none of the studies described here recorded brain activity from multiple audi-
ence members simultaneously; ISC measures are always based on recordings from individual listening to
music alone and thus do not account for the effects of collective listening on music appreciation and brain
synchrony.

Social influences on the aesthetics of film watching


There are only very few studies on social influences on watching film, yet in real life, collective spectatorship
is the norm in cinemas and may even involve spectator participation, for example, during screenings of films
like The Rocky Horror Picture Show (Hanich, 2014) or Bollywood movies (Srinivas, 2002). Arguably, a col-
lective cinematic experience can elicit stronger emotional engagement and enjoyment than watching a film
alone at home (Fröber & Thomaschke, 2021).
Kaltwasser et al. (2019) studied the joint cinematic experience by measuring the spectators’ physiologi-
cal responses as well as behavioural self-report. Thirty-nine healthy participants watched emotional movie
clips in a cinema, once alone and once again in a gender-balanced group with three confederates. Ten two-
minute movie clips used were previously validated to be evoking five different emotions (amusement, anger,
fear, tenderness and neutral; Schaefer et  al., 2010). While film clips were shown in a randomized order,
the participants’ breathing (respiratory sinus arrhythmia; RSA), HR and GSR were recorded. After each
clip, participants rated their emotional experience of the film clips. In both conditions, participants’ resting
state tonic RSA was recorded for two minutes as baseline. RSA, that is, the rhythmical variation of HR in
synchrony with respiration, quantifies the vagal mediation on cardiac output (Porges, 2007). Low RSA is
associated with feeling safe and positive mood states. More importantly, resting state tonic RSA is known to
reflect individual differences in self-regulation abilities and pro-social traits (Muhtadie et al., 2015).
The findings indeed indicated that there were individual differences in one’s vagal mediation (baseline
tonic RSA) from the mere presence of others. However, the manipulation of collectiveness did not elicit sig-
nificant differences in either HR, GSR or behavioural responses. Instead, the emotional content of the clips
impacted physiological responses: Fearful films elicited higher HR, while angry films elicited higher GSR.

454
Experiencing art in social settings

For behavioural self-reports, only empathy ratings showed a main effect of baseline tonic RSA difference.
People who showed lower RSA in the social setting empathized more with the protagonists on screen. All
behavioural measures, apart from memory clip content, varied significantly according to the type of emotion
conveyed by the films. Such results suggest that the physiological and behavioural responses to film watching
primarily reflect responses to the content of the film clips rather than the social context. However, individual
differences in self-regulating abilities reflected in one’s physiological reaction to sharing a space with other
people may influence the way one empathizes with film characters. Further manipulation of group behav-
iour such as active responding (Dunand et al., 1984) and increased interaction between participants (Dikker
et al., 2017) could be explored in the future to understand the physiological correlates of the social apprecia-
tion of emotional films (Rennung & Göritz, 2016).
Continuing their research on affective processes of film watching (Kostoulas et al., 2015), Muszynski and
colleagues (2018) investigated social influences on film watching by testing whether strong aesthetic reac-
tions coincided with higher levels of viewer movement and skin conductance synchrony. The 13 participants
watched 30 films with scenes pre-classified by a group of film experts. The authors compared several group
and pairwise indices of synchrony computed from the EDA and motion acceleration data. Using a data-
driven approach, the study showed that EDA and motion can classify key moments in the film clips, but skin
conductance synchrony was found to be a better predictor than movement synchrony among the viewers.
In addition, pairwise measures performed more robustly than group measures, irrespective of film genre.
Watching film is known to elicit reliably similar brain activations not only within an individual in repeated
trials but also across people (Hasson et  al., 2004; Hasson et  al., 2008). Such co-varying brain activations
(ISCs) are typically computed as pairwise correlations of the individual viewers’ BOLD signals and then
averaging the results for the entire group. ISCs increase with stronger arousal, negative emotions and famili-
arity with the film stimulus (Hanson et al., 2009; Dmochowski et al., 2012). Hasson et al. (2008) proposed
that such shared brain responses among viewers should form the basis of a cognitive neuroscience of film or
“neurocinematics.” Yet, despite using activity from multiple subjects to study the brain mechanisms of film
watching, these studies do not inform about the influence of social context, because participants are always
tested individually and not in groups.
Dmochowski et al. (2014) adopted the ISC approach, recording EEG from a small group of participants
attending to a narrative TV series and TV commercials. Interestingly, neural reliability measures from only
12 participants were more predictive of the results of an online survey from 7000 people than the sample’s
own ratings. This suggests that shared temporal dynamics at cortical level among a small group of participants
may be a better predictor of large population behaviour than of their own behaviour. This study suggests
that collective brain activity among a small group of spectators can be used to predict popularity of films or
commercials in large samples.
Poulsen et al. (2017) recorded simultaneous EEG activity from two groups of nine participants watch-
ing excerpts from acclaimed short films while sitting together in a classroom. ISCs during group viewing
were compared to the ISCs from a previous study where 12 individuals had watched the same stimuli in an
isolated, controlled laboratory setting with high-density EEG (Dmochowski et al., 2012). The study reports
similar relationships between ISCs and attentional engagement with the short films for both social and
non-social data collection. The experiment included an additional control condition where the narrative of
the short film was disrupted with scrambled scenes, which resulted in significantly lower ISCs both within
and between viewing groups. Last, researchers computed average luminance difference (ALD) of the films
and found significant relationships between ISCs and ALD, which were reduced for the scrambled video.
Accordingly, similarity of neural responses among viewers can be partially explained by a combination of
low-level engagement with a coherent narrative structure.
Another factor found to impact collective experience reflected in brain-to-brain synchrony is self-
reported social dynamics (e.g., rapport, Bevilacqua et al., 2019; Dikker et al., 2017). In real life, people often

455
Haeeun Lee and Guido Orgs

visit the cinema with their friends, romantic partners or family, and the degree of social closeness can impact
the level of collective attention of spectatorship in film watching (Hanich, 2014). Recently, Parkinson and
colleagues (2018) studied whether social closeness is associated with how people attend and respond to films
using fMRI. The study investigated whether friendship could be predicted by BOLD signal similarities when
watching video clips.
Brain synchrony was clearly related to social closeness and was most pronounced in brain areas associated
with motivation, emotional processing and learning (e.g., nucleus accumbens, caudate nucleus, putamen
and amygdala; Ben-Yakov & Dudai, 2011), as well as language processing, attention, narrative and meaning-
making (Mar, 2011) in the parietal lobe. Finally, with a subsection of the collected data, researchers trained
a machine learning algorithm to classify social distance based on a dyad’s fMRI time series similarity. This
classifier reliably predicted friendship status. These findings suggest that social closeness between spectators is
an important predictor of the neural correlates of collective film viewing.
As is the case for the role of social influences on dance, theatre and music appreciation, existing studies
on film viewing raise more questions than they answer. Yet measures of brain synchrony emerge as a robust
measure of narrative engagement and appear to be modulated by a number social and individual factors
such as social facilitation, individual difference in physiological response to the mere presence of others and
friendship.

Conclusion
To date, there are only a handful of studies which have investigated collective aesthetic experiences. The
studies reviewed in this chapter reveal a heterogeneous picture of how social settings impact the experience
of various forms of art. Importantly, experiencing art together can both enhance and diminish the aesthetic
experience of the individual. Pelowski and others (2014) have proposed that encountering other people dur-
ing art appreciation may be detrimental to one’s aesthetic experience, because one’s attention enters a com-
petition between self-focused enjoyment and social awareness. On the other hand, merely being involved in
a live social especially when in synchrony can be rewarding (Kokal et al., 2011).
Moreover, many studies of aesthetic appreciation of live performance do not separate social influences
from the location in which the artwork is experienced. Films can be watched in the cinema or at home,
together or alone. In order to understand the role of social influences on aesthetic appreciation, future studies
should aim to carefully disentangle these social and physical components of the live experience.
Recently, Shamay-Tsoory et  al. (2019) proposed that interpersonal movement synchrony, emotional
contagion and social conformity may be interlinked processes relying on shared brain networks. Hyperscan-
ning paradigms investigating brain patterns among multiple spectators simultaneously in the museum or the
theatre provide an exciting new avenue of opportunity for research in empirical aesthetics. Other relevant
technologies like virtual reality can further help overcome the difficulties of studying social dynamics outside
the lab (Kourtesis et al., 2020). At the same time, theories of aesthetic appreciation have largely focused on
the individual and ignored social influences. New theoretical approaches to understanding aesthetic expe-
riences will need to incorporate the inherently social and situated nature of aesthetic experience and art
appreciation.

Acknowledgements
GO and HL are funded by a European Research Council (ERC) consolidator grant awarded to GO
under the European Union’s Horizon 2020 research and innovation programme (grant agreement No.
864420–Neurolive).

456
Experiencing art in social settings

References
Abrams, D. A., Bhatara, A., Ryali, S., Balaban, E., Levitin, D. J., & Menon, V. (2010). Decoding temporal structure in
music and speech relies on shared brain resources but elicits different fine-scale spatial patterns. Cerebral Cortex, 21(7),
1507–1518. https://doi.org/10.1093/cercor/bhq198
Abrams, D. A., Ryali, S., Chen, T., Chordia, P., Khouzam, A., Levitin, D. J., & Menon, V. (2013). Inter-subject synchro-
nization of brain responses during natural music listening. European Journal of Neuroscience, 37(9), 1458–1469. https://
doi.org/10.1111/ejn.12173
Alluri, V., Toiviainen, P., Jääskeläinen, I. P., Glerean, E., Sams, M., & Brattico, E. (2012). Large-scale brain networks
emerge from dynamic processing of musical timbre, key and rhythm. Neuroimage, 59(4), 3677–3689. https://doi.
org/10.1016/j.neuroimage.2011.11.019
Ardizzi, M., Calbi, M., Tavaglione, S., Umiltà, M. A., & Gallese, V. (2020). Audience spontaneous entrainment during
the collective enjoyment of live performances: Physiological and behavioural measurements. Scientific Reports, 10(1),
3813. https://doi.org/10.1038/s41598-020-60832-7
Auslander, P. (1999). Liveness: Performance in a mediatized culture. Routledge.
Babiloni, F., & Astolfi, L. (2014). Social neuroscience and hyperscanning techniques: Past, present and future. Neuroscience
and Biobehavioral Reviews, 44, 76–93. https://doi.org/10.1016/j.neubiorev.2012.07.006
Baccalá, L. A., Sameshima, K., & Takahashi, D. Y. (2007). Generalized partial directed coherence. In 15th international
conference on digital signal processing (pp. 163–166). IEEE Publications.
Bachrach, A., Fontbonne, Y., Joufflineau, C.,  & Ulloa, J. L. (2015). Audience entrainment during live contempo-
rary dance performance: Physiological and cognitive measures. Frontiers in Human Neuroscience, 9, 179. https://doi.
org/10.3389/fnhum.2015.00179
Belfi, A. M., Samson, D. W., Crane, J.,  & Schmidt, N. L. (2021). Aesthetic judgments of live and recorded music:
Effects of congruence between musical artist and piece. Frontiers in Psychology, 12, 618025. https://doi.org/10.3389/
fpsyg.2021.618025
Ben-Yakov, A.,  & Dudai, Y. (2011). Constructing realistic engrams: Poststimulus activity of hippocampus and dorsal
striatum predicts subsequent episodic memory. Journal of Neuroscience: the Official Journal of the Society for Neuroscience,
31(24), 9032–9042. https://doi.org/10.1523/JNEUROSCI.0702-11.2011
Bernardi, N. F., Codrons, E., di Leo, R., Vandoni, M., Cavallaro, F., Vita, G., & Bernardi, L. (2017). Increase in synchro-
nization of autonomic rhythms between individuals when listening to music. Frontiers in Physiology, 8, 785. https://
doi.org/10.3389/fphys.2017.00785
Bevilacqua, D., Davidesco, I., Wan, L., Chaloner, K., Rowland, J., Ding, M., Poeppel, D., & Dikker, S. (2019). Brain-to-
brain synchrony and learning outcomes vary by student–teacher dynamics: Evidence from a real-world classroom elec-
troencephalography study. Journal of Cognitive Neuroscience, 31(3), 401–411. https://doi.org/10.1162/jocn_a_01274
Bodie, G. D. (2010). A racing heart, rattling knees, and ruminative thoughts: Defining, explaining, and treating public
speaking anxiety. Communication Education, 59(1), 70–105. https://doi.org/10.1080/03634520903443849
Brielmann, A. A., Vale, L., & Pelli, D. G. (2017). Beauty at a glance. The feeling of beauty and the amplitude of pleasure
are independent of stimulus duration. Journal of Vision, 17(14), 9. https://doi.org/10.1167/17.14.9
Brown, S. C., & Knox, D. (2017). Why go to pop concerts? The motivations behind live music attendance. Musicae
Scientiae, 21(3), 233–249. https://doi.org/10.1177/1029864916650719
Brown, W. M., Cronk, L., Grochow, K., Jacobson, A., Liu, C. K., Popović, Z., & Trivers, R. (2005). Dance reveals sym-
metry especially in young men. Nature, 438(7071), 1148–1150. https://doi.org/10.1038/nature04344
Bull, M. (2006). Investigating the culture of mobile listening: From Walkman to iPod. In K. O’Hara & B. Brown (Eds.),
Consuming music together: Social and collaborative aspects of music consumption technologies (pp. 131–149). Springer. https://
doi.org/10.1007/1-4020-4097-0_7
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Czeszumski, A., Eustergerling, S., Lang, A., Menrath, D., Gerstenberger, M., Schuberth, S., . . . König. (2020). Hyper-
scanning: A valid method to study neural inter-brain underpinnings of social interaction. Frontiers in Human Neurosci-
ence, 14, 39. https://doi.org/10.3389/fnhum.2020.00039
Dai, B., Chen, C., Long, Y., Zheng, L., Zhao, H., Bai, X., . . . Lu, C. (2018). Neural mechanisms for selectively tuning
in to the target speaker in a naturalistic noisy situation. Nature Communications, 9(1), 2405. https://doi.org/10.1038/
s41467-018-04819-z
Dean, R. T., & Bailes, F. (2010). Time series analysis as a method to examine acoustical influences on real-time percep-
tion of music. Empirical Musicology Review, 5(4), 152–175. https://doi.org/10.18061/1811/48550

457
Haeeun Lee and Guido Orgs

Delaherche, E., Chetouani, M., Mahdhaoui, A., Saint-Georges, C., Viaux, S., & Cohen, D. (2012). Interpersonal syn-
chrony: A survey of evaluation methods across disciplines. IEEE Transactions on Affective Computing, 3(3), 349–365.
https://doi.org/10.1109/T-AFFC.2012.12
Dikker, S., Wan, L., Davidesco, I., Kaggen, L., Oostrik, M., McClintock, J., . . . Poeppel. (2017). Brain-to-brain syn-
chrony tracks real-world dynamic group interactions in the classroom. Current Biology, 27(9), 1375–1380. https://
doi.org/10.1016/j.cub.2017.04.002
Dmochowski, J. P., Bezdek, M. A., Abelson, B. P., Johnson, J. S., Schumacher, E. H., & Parra, L. C. (2014). Audience
preferences are predicted by temporal reliability of neural processing. Nature Communications, 5, 4567. https://doi.
org/10.1038/ncomms5567
Dmochowski, J. P., Sajda, P., Dias, J., & Parra, L. C. (2012). Components of ongoing EEG with high correlation point
to emotionally laden attention—A possible marker of engagement? Frontiers in Human Neuroscience, 6, 112. https://
doi.org/10.3389/fnhum.2012.00112
Dunand, M., Berkowitz, L., & Leyens, J. P. (1984). Audience effects when viewing aggressive movies. British Journal of
Social Psychology, 23(1), 69–76. https://doi.org/10.1111/j.2044-8309.1984.tb00610.x
Dunbar, R. I. (2012). Bridging the bonding gap: The transition from primates to humans. Philosophical Transactions of the
Royal Society of London. Series B, Biological Sciences, 367(1597), 1837–1846. https://doi.org/10.1098/rstb.2011.0217
Durkheim, É. (1912). Les formes élémentaires de la vie religieuse [The elementary forms of religious life]. Alcan.
Eckberg, D. L. (1997).  Sympathovagal balance: A  critical appraisal.  Circulation,  96(9), 3224–3232. https://doi.
org/10.1161/01.CIR.96.9.3224
Egermann, H., Sutherland, M. E., Grewe, O., Nagel, F., Kopiez, R., & Altenmüller, E. (2011). Does music listening in
a social context alter experience? A physiological and psychological perspective on emotion. Musicae Scientiae, 15(3),
307–323. https://doi.org/10.1177/1029864911399497
Fischer-Lichte, E., & Jain, S. (2008). The transformative power of performance a new aesthetics. Routledge.
Fröber, K., & Thomaschke, R. (2021). In the dark cube: Movie theater context enhances the valuation and aesthetic
experience of watching films. Psychology of Aesthetics, Creativity, and the Arts, 15(3), 528–544. https://doi.org/10.1037/
aca0000295
Fusaroli, R., Konvalinka, I., & Wallot, S. (2014). Analyzing social interactions: The promises and challenges of using cross
recurrence quantification analysis. In N. Marwan, M. Riley, A. Giuliani & C. L. Webber, Jr. (Eds.), Translational recur-
rence. Springer. Proceedings of the in Mathematics (pp. 137–155). Springer. https://doi.org/10.1007/978-3-319-09531-8_9
Grewe, O., Nagel, F., Kopiez, R., & Altenmüller, E. (2007). Listening to music as a re-creative process: Physiological,
psychological, and psychoacoustical correlates of chills and strong emotions. Music Perception, 24, 297–314.
Hamilton, A., Paola, P., Davide, P., & Jamie, A. D. (2018). Seeing into the brain of an actor with mocap and fNIRS. Inter-
national Symposium on Wearable Computers (ISWC), 216–217. https://doi.org/10.1145/3267242.3267284
Hanich, J. (2014). Watching a film with others: Towards a theory of collective spectatorship. Screen (London), 55(3),
338–359.
Hanson, S. J., Gagliardi, A. D., & Hanson, C. (2009). Solving the brain synchrony eigenvalue problem: Conservation of
temporal dynamics (fMRI) over subjects doing the same task. Journal of Computational Neuroscience, 27(1), 103–114.
https://doi.org/10.1007/s10827-008-0129-z
Hasson, U., Ghazanfar, A. A., Galantucci, B., Garrod, S.,  & Keysers, C. (2012). Brain-to-brain coupling: A  mecha-
nism for creating and sharing a social world. Trends in Cognitive Sciences, 16(2), 114–121. https://doi.org/10.1016/j.
tics.2011.12.007
Hasson, U., Landesman, O., Knappmeyer, B., Vallines, I., Rubin, N., & Heeger, D. J. (2008). Neurocinematics: The
neuroscience of film. Projections, 2(1), 1–26. https://doi.org/10.3167/proj.2008.020102
Hasson, U., Nir, Y., Levy, I., Fuhrmann, G.,  & Malach, R. (2004). Intersubject synchronization of cortical activity
during natural vision. Science. American Association for the Advancement of Science, 303(5664), 1634–1640. https://doi.
org/10.1126/science.1089506
Himberg, T., Laroche, J., Bigé, R., Buchkowski, M., & Bachrach, A. (2018). Coordinated interpersonal behaviour in
collective dance improvisation: The aesthetics of kinaesthetic togetherness. Behavioral Sciences, 8(2), 23. https://doi.
org/10.3390/bs8020023
Hou, Y., Song, B., Hu, Y., Pan, Y., & Hu, Y. (2020). The averaged inter-brain coherence between the audience and
a violinist predicts the popularity of violin performance. NeuroImage, 211, 116655. https://doi.org/10.1016/j.
neuroimage.2020.116655
Jensenius, A. R., Zelechowska, A., & Gonzalez Sanchez, V. (2017). The musical influence on people’s micromotion
when standing still in groups. Proceedings of the 14th sound and music computing conference. Aalto University.
Jola, C., & Grosbras, M. H. (2013). In the here and now: Enhanced motor corticospinal excitability in novices when
watching live compared to video recorded dance. Cognitive Neuroscience, 4(2), 90–98. https://doi.org/10.1080/1758
8928.2013.776035

458
Experiencing art in social settings

Kaltwasser, L., Rost, N., Ardizzi, M., Calbi, M., Settembrino, L., Fingerhut, J., . . . Gallese. (2019). Sharing the filmic
experience—The physiology of socio-emotional processes in the cinema. PLoS One, 14(10), e0223259. https://doi.
org/10.1371/journal.pone.0223259
Kawabata, H., & Zeki, S. (2004). Neural correlates of beauty. Journal of Neurophysiology, 91(4), 1699–1705. https://doi.
org/10.1152/jn.00696.2003
Kokal, I., Engel, A., Kirschner, S., & Keysers, C. (2011). Synchronized drumming enhances activity in the caudate and
facilitates prosocial commitment—If the rhythm comes easily. PLoS One, 6(11), e27272. https://doi.org/10.1371/
journal.pone.0027272
Konvalinka, I., Xygalatas, D., Bulbulia, J., Schjødt, U., Jegindø, E. M., Wallot, S., . . . Roepstorff. (2011). Synchronized
arousal between performers and related spectators in a fire-walking ritual. Proceedings of the National Academy of Sciences,
108(20), 8514–8519. https://doi.org/10.1073/pnas.1016955108
Kostoulas, T., Chanel, G., Muszynski, M., Lombardo, P., & Pun, T. (2015). Identifying aesthetic highlights in movies
from clustering of physiological and behavioural signals. Seventh International Workshop on Quality of Multimedia Experi-
ence (QoMEX), 1–6.
Kourtesis, P., Korre, D., Collina, S., Doumas, L. A. A., & Macpherson, S. E. (2020). Guidelines for the development of
immersive virtual reality software for cognitive neuroscience and neuropsychology: The development of virtual reality
everyday assessment lab (VR-EAL), a neuropsychological test battery in immersive virtual reality. Frontiers in Computer
Science, 1. https://doi.org/10.3389/fcomp.2019.00012
Krause, A. E., Maurer, S., & Davidson, J. W. (2020). Characteristics of self-reported favourite musical experiences. Music
and Science, 3. 205920432094132. https://doi.org/10.1177/2059204320941320
Krumhansl, C. L. (1997). An exploratory study of musical emotions and psychophysiology. Canadian Journal of Experi-
mental Psychology/Revue Canadienne de Psychologie Expérimentale, 51, 336–353.
Lemasson, A., André, V., Boudard, M., Lippi, D., Cousillas, H., & Hausberger, M. (2019). Influence of theatre hall layout
on actors’ and spectators’ emotions. Animal Cognition, 22(3), 365–372. https://doi.org/10.1007/s10071-019-01249-2
Lemasson, A., André, V., Boudard, M., Lippi, D.,  & Hausberger, M. (2018). Audience size influences actors’ anxi-
ety and associated postures on stage. Behavioural Processes, 157, 225–229. https://doi.org/10.1016/j.beproc.2018.
10.003
Leong, V., Byrne, E., Clackson, K., Georgieva, S., Lam, S., & Wass, S. (2017). Speaker gaze increases information cou-
pling between infant and adult brains. Proceedings of the National Academy of Sciences, 114(50), 13290–13295. https://
doi.org/10.1073/pnas.1702493114
Mar, R. A. (2011). The neural bases of social cognition and story comprehension. Annual Review of Psychology, 62(1),
103–134. https://doi.org/10.1146/annurev-psych-120709-145406
Mayo, O., & Gordon, I. (2020). In and out of synchrony—Behavioural and physiological dynamics of dyadic interper-
sonal coordination. Psychophysiology, 57(6), E13574-N/a.
Merker, B., Morley, I., & Zuidema, W. (2015). Five fundamental constraints on theories of the origins of music. Philo-
sophical Transactions of the Royal Society of London Series B, 370(1664), 1–11. https://doi.org/10.1098/rstb.2014.0095
Muhtadie, L., Koslov, K., Akinola, M., & Mendes, W. B. (2015). Vagal flexibility: A physiological predictor of social
sensitivity. Journal of Personality and Social Psychology, 109(1), 106–120. https://doi.org/10.1037/pspp0000016
Mustile, M., Kourtis, D., Ladouce, S., Learmonth, G., Edwards, M. G., Donaldson, D. I., & Ietswaart, M. (2021, Janu-
ary 19). Mobile EEG reveals functionally dissociable dynamic processes supporting real-world ambulatory obstacle
avoidance: Evidence for early proactive control. European Journal of Neuroscience, 54(12), 8106–8119. https://doi.
org/10.1111/ejn.15120
Muszynski, M., Kostoulas, T., Lombardo, P., Pun, T., & Chanel, G. (2018). Aesthetic highlight detection in movies based
on synchronization of spectators’ reactions. ACM Transactions on Multimedia Computing, Communications, and Applica-
tions, 14(3), 1–23. https://doi.org/10.1145/3175497
Nakahara, H., Furuya, S., Obata, S., Masuko, T., & Kinoshita, H. (2009). Emotion-related changes in heart rate and its
variability during performance and perception of music. Annals of the New York Academy of Sciences, 1169(1), 359–362.
https://doi.org/10.1111/j.1749-6632.2009.04788.x
Néda, Z., Ravasz, E., Brechet, Y., Vicsek, T., & Barabási, A. L. (2000). The sound of many hands clapping. Nature,
403(6772), 849–850. https://doi.org/10.1038/35002660
Nomura, R., Hino, K., Shimazu, M., Liang, Y., & Okada, T. (2015a). Emotionally excited eyeblink-rate variability pre-
dicts an experience of transportation into the narrative world. Frontiers in Psychology, 6, 447. https://doi.org/10.3389/
fpsyg.2015.00447
Nomura, R., Liang, Y., & Okada, T. (2015b). Interactions among collective spectators facilitate eyeblink synchronization.
PLoS One, 10(10), e0140774. https://doi.org/10.1371/journal.pone.0140774
Orgs, G., Caspersen, D., Haggard, P., Obhi, S., & Cross, E. (2016). You move, I watch, it matters: Aesthetic communication in
dance. Cambridge University Press.

459
Haeeun Lee and Guido Orgs

Parkinson, C., Kleinbaum, A. M., & Wheatley, T. (2018). Similar neural responses predict friendship. Nature Communica-
tions, 9(1), 332. https://doi.org/10.1038/s41467-017-02722-7
Pelowski, M., Liu, T., Palacios, V., & Akiba, F. (2014). When a body meets a body: An exploration of the negative impact
of social interactions on museum experiences of art. International Journal of Education and the Arts, 15(14).
Pinti, P., Tachtsidis, I., Hamilton, A. (2020, March). The present and future use of functional near-infrared spectros-
copy (fNIRS) for cognitive neuroscience. Annals of the New York Academy of Sciences, 1464(1), 5–29. https://doi.
org/10.1111/nyas.13948
Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143. https://doi.org/10.1016/j.
biopsycho.2006.06.009
Poulsen, A. T., Kamronn, S., Dmochowski, J., Parra, L. C., & Hansen, L. K. (2017). EEG in the classroom: Synchronised
neural recordings during video presentation. Scientific Reports, 7(1), 43916. https://doi.org/10.1038/srep43916
Rennung, M., & Göritz, A. S. (2016). Prosocial consequences of interpersonal synchrony: A Meta-Analysis. Zeitschrift
Für Psychologie, 224(3), 168–189. https://doi.org/10.1027/2151-2604/a000252
Savage, P. E., Brown, S., Sakai, E., & Currie, T. E. (2015). Statistical universals reveal the structures and functions of human
music. Proceedings of the National Academy of Sciences, 112(29), 8987–8992. https://doi.org/10.1073/pnas.1414495112
Schaefer, A., Nils, F., Sanchez, X., & Philippot, P. (2010). Assessing the effectiveness of a large database of emotion-
eliciting films: A  new tool for emotion researchers. Cognition and Emotion, 24(7), 1153–1172. https://doi.org/
10.1080/02699930903274322
Schilbach, L., Timmermans, B., Reddy, V., Costall, A., Bente, G., Schlicht, T., & Vogeley, K. (2013). Toward a second-
person neuroscience. Behavioral and Brain Sciences, 36(4), 393–414. https://doi.org/10.1017/S0140525X12000660
Shamay-Tsoory, S. G., Saporta, N., Marton-Alper, I. Z., & Gvirts, H. Z. (2019). Herding brains: A core neural mecha-
nism for social alignment. Trends in Cognitive Sciences, 23(3), 174–186. https://doi.org/10.1016/j.tics.2019.01.002
Shoda, H., Adachi, M., & Umeda, T. (2016). How live performance moves the human heart. PLoS One, 11(4), e0154322.
https://doi.org/10.1371/journal.pone.0154322
Srinivas, L. (2002). The active audience: Spectatorship, social relations and the experience of cinema in India. Media,
Culture and Society, 24(2), 155–173. https://doi.org/10.1177/016344370202400201
Sutherland, M. E., Grewe, O., Egermann, H., Nagel, F., Kopiez, R.,  & Altenmüller, E. (2009). The influence of
social situations on music listening.  Annals of the New York Academy of Sciences, 1169, 363–367. https://doi.org/
10.1111/j.1749-6632.2009.04764.x
Swarbrick, D., Bosnyak, D., Livingstone, S. R., Bansal, J., Marsh-Rollo, S., Woolhouse, M. H., & Trainor, L. J. (2019).
How live music moves us: Head movement differences in audiences to live versus recorded music. Frontiers in Psychol-
ogy, 9, 2682. https://doi.org/10.3389/fpsyg.2018.02682
Trost, W., Frühholz, S., Cochrane, T., Cojan, Y.,  & Vuilleumier, P. (2015). Temporal dynamics of musical emotions
examined through intersubject synchrony of brain activity. Social Cognitive and Affective Neuroscience, 10(12), 1705–
1721. https://doi.org/10.1093/scan/nsv060
Vicary, S., Sperling, M., von Zimmermann, J., Richardson, D. C., & Orgs, G. (2017). Joint action aesthetics. PLoS One,
12(7), e0180101. https://doi.org/10.1371/journal.pone.0180101
Von Zimmermann, J., Vicary, S., Sperling, M., Orgs, G., & Richardson, D. C. (2018). The choreography of group affili-
ation. Topics in Cognitive Science, 10(1), 80–94. https://doi.org/10.1111/tops.12320
Wald-Fuhrmann, M., Egermann, H., Czepiel, A., O’Neill, K., Weining, C., Meier, D., . . . Tröndle. (2021). Music
listening in classical concerts: Theory, literature review, and research program. Frontiers in Psychology, 12, 638783.
https://doi.org/10.3389/fpsyg.2021.638783

460
25
TOP-DOWN PROCESSES IN ART
EXPERIENCE
Aenne A. Brielmann

What is top-down processing? A quick literature search shows that the answer is far from obvious (see, e.g.,
Gaspelin & Luck, 2018; Rauss & Pourtois, 2013; Shea, 2014). Like many other terms in psychology and
neuroscience, such as “attention,” the phrases “top-down” and “bottom-up” are used with such ease that one
is rarely pressed to verbalize a concrete definition of either. We all feel like we know what we mean when we
use these terms. So why have we collectively been unable to translate this feeling into a concise definition?
The reason for this is, as we shall see, that there is no such thing as pure bottom-up or top-down processing.
So should you just skip this chapter because apparently there is no such thing as pure top-down pro-
cessing? I hope not. After all, there are some characteristics that do distinguish bottom-up from top-down
processes. Bottom-up processes are driven by the physical properties of the perceived object. They are rela-
tively fast and carried out as part of highly stereotyped processing routes. In contrast, top-down processes
are driven by features of the environment other than the object, such as the task at hand or the state and the
knowledge of the observer. Since they are less stereotyped, top-down processes are usually considered slow
and deliberate.
Take the example of looking at an orange. Your visual system will process its perceptual features like shape
and colour, no matter whether you are hungry or angry at the person next to you. The associations with
and evaluation of the very same orange will, however, radically differ depending on whether you intend to
eat the orange or throw it at a person who so annoyed you. Continuing the example, a person who knows
that she loves the taste of oranges and is a fan of the Oranje Dutch football team will have a very different
emotional response to looking at the colour orange than someone who had recently bitten into a half-rotten
orange and associates the colour orange with the traffic cones that mark the construction site in front of
their house. I could further belabour this example ad nauseam, but the bottom line remains this: There is no
doubt that parts of the perceptual processing of an object’s features are stable across situations and individuals.
We call those bottom-up. At the same time, the observer’s past experiences, current state, and future goals can
modify the way a stimulus is represented and the evaluations that are based on it. We call processes that take
these additional factors into account top-down. Most, if not all, of our experiences depend on both kinds of
processes.
For the sake of this chapter, I will refer to processes that show some form of involvement of informa-
tion not contained in the stimulus as “top-down” processes. These are contrasted to “bottom-up” processes,
which are driven by the information contained in the stimulus. In this chapter we will use this distinction to

DOI: 10.4324/9781003008675-27 461


Aenne A. Brielmann

Figure 25.1 A simplified schema illustrating the interplay of top-down and bottom-up processes involved in perception
and evaluation. The main text elaborates on the two stimulus examples depicted here. Note that the relative
weight of bottom-up and top-down processes may differ according to either the stimulus or the observer.

look at the way particular situations and past experiences of an observer (top-down) serve to influence her
experience of individual artworks (bottom-up).

What this chapter is (not) about


There is little doubt that bottom-up processes are involved in art appreciation—after all, we have to perceive
the artwork with our eyes, ears, or sometimes hands. However, the sensory processing of art falls short of
explaining the entire art experience. For one, art evaluations and preferences are notoriously subjective
(e.g., Vessel et al., 2018). Second, art expertise is known to change how we perceive and judge artworks
(e.g., Kirk, Skov, Christensen et al., 2009; Leder et al., 2014; Vartanian et al., 2019). Third, there is at least
some neurobiological evidence that art engagement depends on brain regions located outside the networks
associated with bottom-up stimulus processing (e.g., Cupchik et al., 2009; Jacobs et al., 2012; Vessel et al.,
2012). This chapter will review evidence for each of these three phenomena.
Even though this chapter mainly focuses on the experience of art, I will also draw on closely related
research on aesthetic experiences in general. Art is a complex stimulus class. It is, therefore, often imperative
to begin examination of prototypical art-associated experiences, such as beauty, by using simplified stimuli
like abstract patterns or shapes. Furthermore, I will follow the bias of the field in primarily focusing on work

462
Top-down processes in art experience

involving visual art, supplemented by discussions of selected studies on music. I will not discuss other forms
of art, such as dance, literature, or film.

An outline
To start, I will review how previous theories of art experiences have conceived the role of top-down process-
ing. I will start with the first writings on aesthetics from a psychological perspective—the work by Gustav
Fechner—and end with the most current and extensive theoretical framework for visual art experiences—
the Vienna integrated model of top-down and bottom-up processes in art perception (VIMAP; Pelowski
et al., 2017).
The second, main part of the chapter will review current empirical findings that cast light on top-down
processing in art experience. I will first discuss if top-down processes are necessary at all by reviewing work
that shows that people can evaluate art in a quick and seemingly automatic fashion. I will then present the
state of the art of what is currently known about top-down processing in art experiences. This review will
focus heavily on recent findings from the past decade.
Finally, we will I discuss the questions still left unanswered and outline future directions for the research
on top-down processing in art experiences.

Top-down processes in theory: a brief history and current models


The notion that there is more to the experience of art than perceptual responses to the artwork’s physical
properties is as old as the empirical study of aesthetics. The first one to acknowledge them was also the first
one to write about aesthetics from an empirical point of view: Gustav Fechner.
For the most part, Gustav Fechner’s approach to studying aesthetics was a bottom-up approach. Still, one
of the principles he stipulated in his Vorschule der Aesthetik (“Preschool of aesthetics,” 1876) underscored the
importance of top-down processes. He called it “Aesthetisches Assoziationsprinzip,” that is, the “principle
of aesthetic association.” In brief, this principle states that aesthetic experiences are shaped by the memories
and knowledge associated with the object. Thus, Fechner acknowledged the importance of factors other
than stimulus properties in aesthetic perception and judgments, even though they often interfered with his
experimental goals. A telling example is his perhaps best-known experiment. He asked observers to pick
their most and least favourite rectangle from a set of rectangles with different aspect ratios (Fechner, 1876,
chapter XIV). Fechner explicitly instructed his participants to look only at the shape and to disregard any
potential functionality or association with other objects such as envelopes or books. Still, in the end, he
expressed doubt that his participants were able to do so. Fechner admits that top-down factors almost inevi-
tably influence aesthetic evaluations.
Daniel Berlyne did not embrace the contribution of top-down factors in the way Fechner did. In Studies
in the New Experimental Aesthetics (1974), Berlyne clearly expressed that, in his view, stimulus properties are
the chief determinants of aesthetic responses. Even though he acknowledged that the variables he considered
most crucial to aesthetic evaluation, that is, complexity, novelty, and so on, are subjective, he strongly empha-
sized the existence of a stable link between subjective reports and objective measures. Differences between
individuals were, in his opinion, far less relevant than population averages. Effectively, Berlyne discouraged
his readers from attempting to study top-down processes and to instead focus on linking quantitative, meas-
urable stimulus properties to people’s responses.
Silvia’s position (2005) is a direct response to Berlyne’s neglect of top-down processes’ import for aes-
thetic experiences. His view borrows ideas from the appraisal theory of emotion and applies them to how
people experience and evaluate art. Put briefly, Silvia’s appraisal view does not see a direct link between
stimulus properties and observer responses. Instead, the observer’s current state and goals influence how

463
Aenne A. Brielmann

the object’s properties are subjectively perceived and, in turn, how the object is evaluated. As such, Silvia’s
theoretical approach clearly emphasizes the importance of top-down factors over bottom-up processing. In
Silvia’s model, the evaluation of an artwork can only take place with reference to the observer’s state and
goals; there is no aesthetic experience without a contribution from top-down processes.
Originally conceived to describe a user’s interaction with design products, Locher and colleagues (2010)
devised a model that makes an explicit distinction between top-down and bottom-up processes. In Locher’s
model, experiences are dependent upon two different stages. The first one is largely driven by the initial
(bottom-up) perceptual processing of the object. The second one is dominated by a “Central Executive.”
The Central Executive exerts top-down control over further object exploration and integration of short- and
long-term memories. Locher and colleagues’ model posits that there is a continuous interaction between
top-down and bottom-up processing. Here, a person’s past experiences and current state influence both how
the object is evaluated and how it is perceived.
Chatterjee and Vartanian’s (2014) model is heavily informed by our current understanding of the neural
processes involved in art experiences. Even though they do not explicitly denote it as such, the “meaning-
knowledge” system in their triad of neural circuits involved in aesthetic experiences most closely resembles a
representation of top-down processing. Chatterjee and Vartanian note that we know less about the meaning-
knowledge system than about the other parts of the aesthetic triad. They argue that this is likely due to its
widespread influences throughout the brain and its variability across individuals. It is worth noting that these
authors also acknowledge that the relative contribution of each system differs depending on the experience
and observer in question.
One of the best-known contemporary models of art appreciation was originally proposed by Leder
and colleagues (2004) and later updated by Leder and Nadal (2014). Both versions prominently feature
several top-down mechanisms. The original model explicitly declares that the influence of expertise and
knowledge stems from top-down cognitive processes. Within Leder and Nadal’s framework, evaluation
changes produced by added information about the artwork are also considered evidence for top-down
influences on aesthetic judgments. The revised version of the model does not make explicit statements
about “top-down” effects; however, the expertise and knowledge modules remain in place and feature more
explicit interactions with affective variables. In this, the Leder and Nadal model mirrors a general trend
in psychology: an increasing emphasis on bidirectional interactions between bottom-up and top-down
processes.
Leder and colleagues’ model (2004, 2014) has also served as a precursor to the most extensive current
model on art appreciation: the VIMAP, developed by Pelowski, Leder, and others (Pelowski et al., 2017).
As flagged by its name, the VIMAP model explicitly features both bottom-up and top-down processes. The
VIMAP model posits that top-down processing commences around 6–8 seconds after stimulus onset (in the
context of visual perception). However, a large number of processes that many would consider top-down
processes, such as the integration of knowledge, experience, taste, and the current state of the observer, are
included in the earlier, bottom-up phase of the model.
While most models of art appreciation have either implicitly or explicitly focused on visual experiences,
Brattico and colleagues (2013) proposed a model of aesthetic musical experiences. They explicitly denoted
the effects of what they call attitude as a top-down process. Overall, they listed a series of “modulating fac-
tors” (internal context, attitude, expertise, mood, attention, and intentionality) that should be considered
part of top-down processing in the experience of music. Notably, they placed the influence of these factors
outside the temporal sequence proposed for music experience and argued that they affect music experience
throughout, on a different time scale than the remaining processes.
In sum, almost every major model of aesthetic or art experience features some form of top-down pro-
cesses. Figure 25.2 illustrates this with a schematic overview of a selection across the history outlined here.
What differs between models is how much emphasis the individual model puts on top-down processes, to

464
Top-down processes in art experience

Figure 25.2 Schematic overview of the development of models of aesthetic experiences. Light arrows indicate pre-
dominantly bottom-up processing, darker arrows predominantly top-down processing. See main text for a
description of the depicted models, references, and further models.

465
Aenne A. Brielmann

what extent they are viewed as necessary or inevitable, their hypothesized onset time, and how much they
are thought to interact with the other components of the model.

Current empirical findings

Are top-down processes necessary?


Given the omnipresence of top-down influences in models of art appreciation, it seems self-evident that they
are an essential part of any art experience. Yet, there are studies that cast doubt on this intuition. These stud-
ies showcase how fast people make up their mind about art and can report aesthetic evaluations, especially
with respect to how much they like an artwork or not.
Belfi and colleagues (2018) showed that a person only needs to hear 750 ms of a piece of music to deter-
mine whether they like it. Their findings are in line with those of Istók and colleagues (2013): EEG signals
differentiate between preferred vs. non-preferred music genres as early as 230–370 ms. On a slightly longer
time scale, pleasure and beauty ratings for viewing images and tasting candy do not differ for presentation
durations between 1 and 30 seconds (Brielmann et al., 2017). Finally, both pleasantness ratings and the type
of verbal comments on paintings in Paul Locher and colleagues’ study show remarkable correspondence
between limited 100 ms  and unlimited viewing time (Locher et  al., 2008). Despite this early formation
of consistent aesthetic judgments, a distinct neural signature of making an aesthetic rather than perceptual
judgment only appears later (after about 500 ms; see “Neural Correlates of Top-Down Processing” in the
following).
The studies described previously show that at least the most intuitive form of top-down processing,
that is, response changes driven by processes that occur after full perceptual processing, may not always be
required for a full aesthetic appreciation to occur. However, top-down processes can not only occur after the
fact; they can also occur at the same time as bottom-up processes, and their contribution may still be crucial
for the overall aesthetic experience.
Indeed, we at least need to have the cognitive capacities for such processing available to experience
intense beauty. Brielmann and Pelli (2017) showed that felt pleasure and beauty of otherwise highly beautiful
images are diminished when participants have to engage with a two-back task while viewing these images.
Even though the images we used were not art, our results do speak to one of the most important aesthetic
experiences we can have in response to art: beauty. The cognitive processes involved in the two-back task are
essential for intense beauty experiences.

Do titles exert top-down effects on art experience?


The perhaps most frequently mentioned example of top-down influences on art appreciation is the way titles
and similar object-external information change the way we evaluate artworks (Anglada-Tort et al., 2018;
Belke et al., 2010; Gerger & Leder, 2015; Leder et al., 2006; Millis, 2001; Mullennix & Robinet, 2018;
Szubielska et al., 2019). I will not discuss these findings in depth, but I will discuss how they inform our
understanding of top-down processes in art experiences.
In laboratory settings, related titles elicit greater liking of artworks than unrelated ones (Belke et  al.,
2010; Gerger & Leder, 2015). Whether these titles are purely descriptive versus elaborative has varied effects,
depending on presentation duration (Leder et al., 2006), and sometimes affects only self-rated understanding
(Leder et al., 2006), sometimes only liking (Mullennix & Robinet, 2018), and sometimes aesthetic evalua-
tion overall (Millis, 2001).

466
Top-down processes in art experience

It seems that knowing the title of an artwork does not affect its evaluation when presented in the con-
text of a museum or gallery (Krauss et al., 2019; Szubielska et al., 2019). It is unclear what the cause of this
selective null effect is. On the one hand, it could be that a richer stimulus experience in real-life situations
(compared to computer screen displays used in lab experiments) overrides title effects and enhances the con-
tribution of bottom-up information. On the other hand, it could also be that people bring a different mindset
to lab versus gallery or museum, rendering the difference in findings itself a top-down–driven differentiation.
In sum, based on the evidence reviewed here, it is not clear what kind of top-down processing underlies
the effect of titles on art evaluation. It might even be questioned if title effects are evidence for top-down
processing at all. The addition of a title can be regarded as an addition to the stimulus, eliciting a change in
response solely based on bottom-up processing. Possibly, titles represent a change at the very bottom of the
hierarchy rather than an influence from the top.

Tagged as (original) art: framing effects


Title addition, as discussed previously, is closely related to a different experimental approach that more con-
vincingly induces top-down effects. That is: the manipulation of the framework in which objects are viewed
rather than the addition of information to any given object. I will refer to these studies as framing studies.
Framing manipulations include changes to the physical “framework” in which the objects are shown, like
Szubielska and colleagues (2019) have done by showing the same artworks in a gallery as well as videotaped
in a classroom.
Wagner and colleagues (2014) looked at the effect of framing on the rated positivity, negativity, intensity,
and beauty of pictures depicting typically disgust-eliciting objects like rotten food. Positivity ratings were
higher when photographs were presented as artworks rather than as hygiene education material. At the same
time, negativity and intensity ratings did not differ depending on how the images were framed. In addition,
the link between beauty and positivity ratings was stronger in the art compared to the hygiene framing.
In a similar setup, Gerger and colleagues (2014) framed pictures from the International Affective Picture
System (IAPS) as either artworks or press photographs. Like Wagner and colleagues, they found that images
with negative content received higher positivity ratings when framed as artworks rather than real-life images.
It is noteworthy, though, that the activity of the facial muscles was affected in a different way: the smile mus-
cle M. zygomaticus responded less when images were presented as art, while the frown muscle M. currogator
remained unaffected.
These studies are prototypical examples for one of the most puzzling characteristics of art. As a part of art,
the most negative emotions and experiences, like sadness, anger, and even fear, are sought out rather than
avoided. Think about encountering the bloodshed and monsters present in every other blockbuster movie:
What would your response be if you encountered them in real life? I would wager that one would have to
pay you significantly more than the price of a movie ticket to take just a glance at the gore that seems so
enjoyable from the theatre seat. It seems like we perceive things very differently depending on whether we
think of them as art or reality. In the following section on neural correlates of top-down processing, we will
see that similar framing effects do indeed change brain activity while processing the artwork in question.
Another line of studies has investigated the effect of framing objects as artist created and contrasted
this to framing as computer- or forger-generated objects. Newman and Bloom (2012) demonstrated in
a series of five experiments that people prefer and are willing to pay more for artworks created by the
original artist compared to those that were not. Kirk, Skov, Hulme and colleagues (2009) told their partici-
pants that the abstract art images they viewed were either from a museum or computer generated. In this
study, supposedly computer-generated art was rated less appealing than the same art framed as belonging

467
Aenne A. Brielmann

to a museum. At the same time, viewing art images labelled as museum pieces (in contrast to computer gen-
erated) was associated with greater activity in the entorhinal cortex and the temporal and medial temporal
lobe as well as a cluster in the visual cortex (independent of appeal).
In contrast to the previous studies on art images, Steinbeis and Koelsch (2009) did not observe a difference
in pleasantness ratings for musical pieces described either as “composed” or “computer generated.” They did,
however, find that intentionality ratings were higher for composed compared to computer-generated pieces
and that these intentionality ratings were positively linked to activity in the anterior medial frontal cortex
(aMFC).
In sum, information about the provenance of a work of art does seem to alter how we evaluate it. Yet how
evaluations are affected might depend on the kind of art being evaluated.

Inter-individual differences: expertise, familiarity, and associations


Top-down processes do not only become apparent when they are “forced” into existence by experimental
manipulations. Pre-existing differences between observers are also valuable demonstrations of how stimulus-
external, and thus likely top-down, changes affect art experiences. Aesthetic judgments, even of relatively
simple stock images, show greater divergence than agreement between observers (Brielmann & Pelli, 2019).
These differences are exaggerated when people judge art (Vessel et al., 2018; see also Wallisch & Whritner,
2017).
There is little doubt that these striking differences arise because people have different expectations of
and associations with a particular work of art (or any stimulus, for that matter). In research and practice, the
impact of these differences is referred to as effects of expertise, familiarity, or culture. While all three of these
sources of inter-individual differences can have distinct effects, it is worth keeping in mind that they all, in
essence, come down to differences in prior knowledge about the (art) object in question and how these shape
the experience of that object.
Effects of expertise (discussed in more depth in other chapters of this book; see Chapters 2 and 28) are
one example of how prior knowledge of the observer influences her art experience. These expertise effects
are, at the same time, one of many examples of inter-individual differences in art experience. In brief,
expertise effects showcase how prior knowledge can shape which aspect of the artwork is paid attention
to (e.g., Kirk, Skov, Christensen et  al., 2009; Vogt  & Magnussen, 2007) and how it is—subsequently—
evaluated (e.g., Leder et al., 2014; Leder et al., 2019; Pihko et al., 2011; van Paasschen et al., 2015; Vartanian
et al., 2019).
Expertise effects go hand in hand with another difference within individuals: familiarity. Experts are
much more likely to be familiar with stimuli that fall within the scope of their expertise. By itself, familiar-
ity is often studied via two routes. One is the direct route of experimentally manipulating how familiar an
observer is with a particular object through exposure during the experiment. The effects of such a manipu-
lation are referred to as “mere-exposure” effects, famously documented in 1968 by Zajonc. Mere-exposure
effects find frequent mention even though they are much less prominent than often claimed (for a meta-
analysis, see Montoya et al., 2017). When it comes to art in particular, long-term repeated exposure does
seem to be associated with higher preference ratings, as argued by Cutting (2003). He found that frequency
with which the painter Caillebot’s works of art were depicted in books was indeed predictive of adult’s—but
not children’s—preferences.
Whether due to a general presence of art or expertise, long-term rather than acute familiarity seems asso-
ciated with more favourable aesthetic evaluations. Taken to the extreme of life-long differences in familiar-
ity, they become a cultural phenomenon. In that sense, comparing responses of participants from different
cultures to objects of their own and another cultural background is a second way of assessing the influence
of familiarity on art experience.

468
Top-down processes in art experience

Most research on cultural differences in art experience and expression compares people from Western
(American and European) to people from East-Asian cultures (mostly Chinese and Japanese). It is evident
that artistic styles differ between these two broad cultural regions, as, for instance, documented by analyses
of landscape paintings (Masuda et al., 2008). Horizon lines are higher and central objects and people take up
relatively less space in East Asian art compared to Western art. Notably, Senzaki et al. (2014) found equivalent
differences today in landscape drawings and collages made by Japanese and Canadian children from grade two
onwards. These differences in production mirror cultural differences in preferences (e.g., Bao et al., 2016;
Masuda et al., 2008). Yet same-culture preferences are not always apparent. For instance, Yang and colleagues
(2019) found higher valence ratings for Western paintings in both cultures and no difference between Asian
and Western art preferences in their Chinese participants.
Findings on colour preferences add to the evidence that aesthetic experiences are shaped by one’s cultural
background (e.g., Taylor et al., 2013). These cultural idiosyncrasies are most likely shaped by differences in
the associations that accompany different hues (though one may argue that exposure to certain colours does
also vary by surroundings). A theory proposed by Schloss and Palmer, dubbed “Ecological Valence Theory”
(2010), is backed by some experimental evidence. For example, Strauss and colleagues (2013) documented
that colour preferences change when people see liked or disliked objects of a certain colouring. Schloss and
colleagues also observed that within individuals, colour preferences change according to the season (Schloss
et al., 2017). Thus, aesthetic preferences even for objects as simple as colour patches are not as stable as would
be expected from a purely bottom-up–driven process. Instead, they are malleable by recently triggered or
culturally acquired associations.
Outside the realm of visual art, effects of prior knowledge are studied under a different lens. A proposal
that is currently gaining popularity with music scholars links (musical) art experiences to reward learning,
especially the phenomenon of prediction errors (see, e.g., Belfi & Loui, 2019; Koelsch et al., 2019; see also
Chapters 7 and 15). The basic idea underlying this theory is that people like passages of music that differ from
their expectation to such an extent that they afford learning. This learning process then presumably triggers a
positive reward signal. Empirically, this notion has recently found support: Musical stimuli are more pleasant
or liked if they are characterized by low uncertainty and high surprise or high uncertainty and low surprise
(Cheung et al., 2019; Gold et al., 2019). We like to hear what we expect when we are not one hundred
percent sure what to expect. But when we already have strong expectations, we like a little surprise.
Prediction errors are inherently an example of top-down processing since they involve the comparison of
an expectation formed based on prior knowledge to the current stimulus. The appeal of the prediction-error
framework is that it allows one to accommodate both long-term implications of art and cultural expertise
and short-term effects of the associations triggered by the current context. After all, one’s current expecta-
tions should be based on both. The prediction error account also links the study of aesthetically rewarding
experiences to the broader field of reward learning. The main downside of the approach is that it has yet to
be applied to other kinds of art than music.

Neural correlates of top-down processing


The most definitive manipulation that showcases the effects of top-down processes on art appreciation is a
change in the task the observer has to do while looking at an artwork. Precisely because the stimulus itself
remains unchanged, any difference between tasks needs to be due to top-down modulations of stimulus
processing. In such a setting, we cannot compare the behavioural data between tasks precisely because tasks
differ. The changes of interest here occur on the level of brain activity.
Cupchik and colleagues (2009) carried out such a study with 32 representational paintings, viewed for
10 seconds. Their participants were either told to look at the paintings with an objective, detached stance
or with a subjective, engaged manner focused on evoked feelings. The direct contrast between those two

469
Aenne A. Brielmann

conditions showed greater activation in the left lateral prefrontal cortex (LPFC) when participants were
instructed to take a more subjective, aesthetic stance.
Even though they did not use art but simpler texture stimuli, Jacobs and colleagues (2012) documented
a task effect in an fMRI study. They contrasted an aesthetic judgment condition (rate beauty) with a percep-
tual one (rate roughness). Like Cupchik and colleagues, they found greater activation in the LPFC, albeit
bilateral, as well as in the posterior cingulate cortex and the amygdala during the aesthetic compared to the
perceptual judgments. Of these differentially activated areas, the frontomedian cortex and the amygdala also
exhibited a linear increase in activation for ugly to neutral to beautiful textures. Jacobs and colleagues inter-
pret this as evidence that these regions are specifically devoted to making beauty judgments, or at least a type
of evaluative rather than perceptual judgment.
In an EEG study using black-and-white patterns as stimuli, Höfel and Jacobsen (2007) found a globally
larger late positivity between 500 and 770 ms after stimulus onset for participants who were instructed to
contemplate the beauty of the patterns compared to those who only viewed them. A similar effect, not dis-
cussed by the authors, already emerged in the 130–200-ms time window. Both participant groups had the
same probe-detection task (unrelated to the pattern’s aesthetics). Höfel and Jacobsen thus argue that aesthetic
appreciation in the absence of an overt judgment already shows a relatively late but distinct neural activation
pattern.
Taken together, these differential activation patterns due to task demands speak for top-down modulation
of brain activity associated with the (visual) perception of stimuli with an aesthetic rather than perceptual
focus. However, the involvement of the LPFC in evaluative judgments is by no means special to aesthetic or
art experiences. In the general context of value-based decision making, the LPFC is thought of as a region
involved in exerting top-down attentional control (e.g., Grabenhorst & Rolls, 2011). Descriptions of the
object as well as task instructions are also known to influence the neural representation of an object’s value
in the orbitofrontal cortex (OFC) and anterior cingulate cortex (ACC) in a top-down fashion (e.g., Graben-
horst & Rolls, 2009; Grabenhorst et al., 2008; McCabe et al., 2008). Thus, whilst the previous studies show
evidence for top-down effects on the neural processing of art and that taking an aesthetic perspective does
alter brain responses to (art) stimuli, their results are not specific to artworks or aesthetic judgements. They
may simply reflect a general top-down modulation of valuation.
However, there is one line of research that documents a neural signature that might be special to intensely
moving art experiences. Vessel and his colleagues have repeatedly shown (2012, 2019; Belfi et al., 2019)
that when people view artworks that score highest on how “aesthetically moving” they are, the default
mode network (DMN) is more active than when viewing less moving artworks. Usually, activation in the
DMN is suppressed when actively engaging with a stimulus, and activation of the DMN is associated with
self-reflexion. Thus, the case of viewing aesthetically moving art may be a special one. Instead of a trade-off
between task- and self-relevant network activations, one observes co-activation.

Outstanding issues and future directions


The first obvious challenge we need to face in the discussion about top-down processing in art experience is
to define what we mean by “top-down.” Neuroaesthetics is not alone in this quest for achieving a uniform
understanding of the term, and I would encourage my colleagues to learn from the discussion that has been
ongoing in different subfields of psychology (e.g., Gaspelin & Luck, 2018; Rauss & Pourtois, 2013) and
philosophy (e.g., Shea, 2014). It is clear that in a complex system like the brain, there is no clear-cut differ-
entiation between processes that operate “bottom-up” versus “top-down,” and some will lie directly on the
blurry line in between. It is therefore all the more crucial that, until an explicit definition has been widely
accepted, every author make her definition of top-down (or bottom-up, for that matter) clear to her readers.

470
Top-down processes in art experience

At the same time, we will also need to acknowledge that bottom-up and top-down processing are not
entirely separable. Nadal and Skov (2017) have phrased it well:

a truly integrated model of top-down and bottom-up processes in art perception should explain
how viewers: (1) use their art knowledge, prior art experiences, their goals, and current situation to
predict and anticipate sensory features; (2) match the incoming signals with those predictions, and
(3) manage the prediction errors that propagate through the system.

So far, we have seen a fair amount of research on the separate aspects regarding (1), that is, the effects of
the observer on her art experience (e.g., Bao et al., 2016; Kirk, Skov, Christensen et al., 2009; Pihko et al.,
2011; Taylor et al., 2013). Studies on how we experience music have shed some light on (2), the importance of
matching expectations (e.g., Cheung et al., 2019; Gold et al., 2019). However, we know next to nothing about
(3): how both taken together influence the perception and evaluation of art. The construction of an integrated
model of top-down and bottom-up processing of art necessitates that we start to tackle this third point.
The road to such an ideal, integrated model may seem long, longer perhaps than the lifespan of a single
academic career. This should not prevent us from finding immediate value in working towards one. It may
serve as a valuable tool to evaluate existing models in the meantime. The models mentioned in the first
section of this chapter outline a series of possible top-down mechanisms. Few of these models have been
rigorously tested. Since bottom-up processing is a given, testing the different predictions that these models
make about which top-down processes influence art experiences could serve as an essential tool for testing
and eventually judging their usefulness.
One major difference between models that could be crucial in this process lies in the distinction between
perceiving and evaluating art. Top-down influences on the perception of art literally change how we see an
artwork, effectively altering bottom-up processing. Top-down influences on the evaluation of art change how
we judge an artwork, given the same percept, leaving bottom-up processing untouched. These are two funda-
mentally different routes of modifying an art experience. Both experiments and theories should make the role
of each of these two routes explicit. It may prove difficult, but certainly very fruitful, to dissociate the two.
And once their separate roles have been characterized, we can and need to also consider their interaction.
In my review of studies, I have intentionally included both studies that use artworks as stimuli and those
that used others, like simple patterns and photographs. At the moment, research on art experiences is too
scarce to allow a direct comparison between these two stimulus kinds. An open question is: How well do the
effects observed in simple stimuli translate to art? Are there top-down processes involved in art appreciation
that are unique to art? Given that our notion of what things are considered art and what things are not is in
constant flux, I doubt that there is a clear answer to this question. However, it may be that people do recruit
different top-down processes when experiencing things they consider art than those they do not consider art.
Given the fact that the neural signature of top-down processes involved in aesthetic appreciation (of art)
closely resembles a more general top-down modulation of value signals, I doubt that the search for a unique
art signature in the brain will be fruitful. That being said, it will be highly informative to document which
top-down modulations of the neural circuits implied in reward processing apply to the evaluation of art and
how. Like this, we can integrate our knowledge about art experience into the much broader framework of
how people perceive, value, and evaluate objects.

References
Anglada-Tort, M., Steffens, J., & Müllensiefen, D. (2018). Names and titles matter: The impact of linguistic fluency and
the affect heuristic on aesthetic and value judgements of music. Psychology of Aesthetics, Creativity, and the Arts, 13(3),
277–292. https://doi.org/10.1037/aca0000172

471
Aenne A. Brielmann

Bao, Y., Yang, T., Lin, X., Fang, Y., Wang, Y., Pöppel, E., & Lei, Q. (2016). Aesthetic preferences for eastern and west-
ern traditional visual art: Identity matters. Frontiers in Psychology, 7, 1596. https://doi.org/10.3389/fpsyg.2016.01596
Belfi, A. M., Kasdan, A., Rowland, J., Vessel, E. A., Starr, G. G., & Poeppel, D. (2018). Rapid timing of musical aesthetic
judgments. Journal of Experimental Psychology. General, 147(10), 1531–1543. https://doi.org/10.1037/xge0000474
Belfi, A. M., & Loui, P. (2019). Musical anhedonia and rewards of music listening: Current advances and a proposed
model. Annals of the New York Academy of Sciences, 1464(1), 99–114. https://doi.org/10.1111/nyas.14241
Belfi, A. M., Vessel, E. A., Brielmann, A., Isik, A. I., Chatterjee, A., Leder, H., . . . Starr, G. G. (2019). Dynamics of aes-
thetic experience are reflected in the default-mode network. Neuroimage, 188, 584–597. https://doi.org/10.1016/j.
neuroimage.2018.12.017
Belke, B., Leder, H., Strobach, T., & Carbon, C. C. (2010). Cognitive fluency: High-level processing dynamics in art
appreciation. Psychology of Aesthetics, Creativity, and the Arts, 4(4), 214–222. https://doi.org/10.1037/a0019648
Brattico, E., Bogert, B., & Jacobsen, T. (2013). Toward a neural chronometry for the aesthetic experience of music.
Frontiers in Psychology, 4, 206. https://doi.org/10.3389/fpsyg.2013.00206
Brielmann, A. A., & Pelli, D. G. (2017). Beauty requires thought. Current Biology, 27(10), 1506–1513.e3. https://doi.
org/10.1016/j.cub.2017.04.018
Brielmann, A. A., & Pelli, D. G. (2019). Intense beauty requires intense pleasure. Frontiers in Psychology, 10, 2420. https://
doi.org/10.3389/fpsyg.2019.02420
Brielmann, A. A., Vale, L., & Pelli, D. G. (2017). Beauty at a glance: The feeling of beauty and the amplitude of pleasure
are independent of stimulus duration. Journal of Vision, 17(14), 9–9. https://doi.org/10.1167/17.14.9
Chatterjee, A.,  & Vartanian, O. (2014). Neuroaesthetics. Trends in Cognitive Sciences, 18(7), 370–375. https://doi.
org/10.1016/j.tics.2014.03.003
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology,
29(23), 4084–4092.e4. https://doi.org/10.1016/j.cub.2019.09.067
Cupchik, G. C., Vartanian, O., Crawley, A., & Mikulis, D. J. (2009). Viewing artworks: Contributions of cognitive con-
trol and perceptual facilitation to aesthetic experience. Brain and Cognition, 70(1), 84–91. https://doi.org/10.1016/j.
bandc.2009.01.003
Cutting, J. E. (2003). Gustave Caillebotte, French impressionism, and mere exposure. Psychonomic Bulletin and Review,
10(2), 319–343. https://doi.org/10.3758/bf03196493
Fechner, G. T. (1876). Vorschule der aesthetik (Vol. 1). Breitkopf & Härtel.
Gaspelin, N., & Luck, S. J. (2018). Top-down. Journal of Cognition, 1(1). https://doi.org/10.5334/joc.28
Gerger, G., & Leder, H. (2015). Titles change the esthetic appreciations of paintings. Frontiers in Human Neuroscience, 9,
464. https://doi.org/10.3389/fnhum.2015.00464
Gerger, G., Leder, H., & Kremer, A. (2014). Context effects on emotional and aesthetic evaluations of artworks and IAPS
pictures. Acta Psychologica, 151, 174–183. https://doi.org/10.1016/j.actpsy.2014.06.008
Gold, B. P., Pearce, M. T., Mas-Herrero, E., Dagher, A.,  & Zatorre, R. J. (2019). Predictability and uncertainty in
the pleasure of music: A reward for learning? Journal of Neuroscience, 39(47), 9397–9409. https://doi.org/10.1523/
JNEUROSCI.0428-19.2019
Grabenhorst, F., & Rolls, E. T. (2009). Different representations of relative and absolute subjective value in the human
brain. Neuroimage, 48(1), 258–268. https://doi.org/10.1016/j.neuroimage.2009.06.045
Grabenhorst, F., & Rolls, E. T. (2011). Value, pleasure and choice in the ventral prefrontal cortex. Trends in Cognitive Sci-
ences, 15(2), 56–67. https://doi.org/10.1016/j.tics.2010.12.004
Grabenhorst, F., Rolls, E. T., & Bilderbeck, A. (2008). How cognition modulates affective responses to taste and fla-
vor: Top-down influences on the orbitofrontal and pregenual cingulate cortices. Cerebral Cortex, 18(7), 1549–1559.
https://doi.org/10.1093/cercor/bhm185
Höfel, L., & Jacobsen, T. (2007). Electrophysiological indices of processing aesthetics: Spontaneous or intentional pro-
cesses? International Journal of Psychophysiology, 65, 20–31.
Istók, E., Brattico, E., Jacobsen, T., Ritter, A.,  & Tervaniemi, M. (2013). ‘I love rock ‘n’ roll’—Music genre pref-
erence modulates brain responses to music. Biological Psychology, 92(2), 142–151. https://doi.org/10.1016/j.
biopsycho.2012.11.005
Jacobs, R. H., Renken, R., & Cornelissen, F. W. (2012). Neural correlates of visual aesthetics—Beauty as the coalescence
of stimulus and internal state. PLoS One, 7(2), e31248. https://doi.org/10.1371/journal.pone.0031248
Kirk, U., Skov, M., Christensen, M. S., & Nygaard, N. (2009). Brain correlates of aesthetic expertise: A parametric fMRI
study. Brain and Cognition, 69(2), 306–315. https://doi.org/10.1016/j.bandc.2008.08.004
Kirk, U., Skov, M., Hulme, O., Christensen, M. S., & Zeki, S. (2009). Modulation of aesthetic value by semantic con-
text: An fMRI study. Neuroimage, 44(3), 1125–1132. https://doi.org/10.1016/j.neuroimage.2008.10.009

472
Top-down processes in art experience

Koelsch, S., Vuust, P., & Friston, K. (2019). Predictive processes and the peculiar case of music. Trends in Cognitive Sciences,
23(1), 63–77. https://doi.org/10.1016/j.tics.2018.10.006
Krauss, L., Ott, C., Opwis, K., Meyer, A., & Gaab, J. (2019). Impact of contextualizing information on aesthetic experi-
ence and psychophysiological responses to art in a museum: A naturalistic randomized controlled trial. Psychology of
Aesthetics, Creativity, and the Arts, 15(3), 505–516. https://doi.org/10.1037/aca0000280
Leder, H., Belke, B., Oeberst, A., & Augustin, D. (2004). A model of aesthetic appreciation and aesthetic judgments.
British Journal of Psychology, 95(4), 489–508. https://doi.org/10.1348/0007126042369811
Leder, H., Carbon, C. C., & Ripsas, A. L. (2006). Entitling art: Influence of title information on understanding and
appreciation of paintings. Acta psychologica, 121(2), 176–198. https://doi.org/10.1016/j.actpsy.2005.08.005
Leder, H., Gerger, G., Brieber, D., & Schwarz, N. (2014). What makes an art expert? Emotion and evaluation in art
appreciation. Cognition and Emotion, 28(6), 1137–1147. https://doi.org/10.1080/02699931.2013.870132
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Leder, H., Tinio, P. P. L., Brieber, D., Kröner, T., Jacobsen, T., & Rosenberg, R. (2019). Symmetry is not a universal law
of beauty. Empirical Studies of the Arts, 37(1), 104–114. https://doi.org/10.1177/0276237418777941
Locher, P., Krupinski, E. A., Mello-Thoms, C., & Nodine, C. F. (2008). Visual interest in pictorial art during an aesthetic
experience. Spatial Vision, 21(1–2), 55–77. https://doi.org/10.1163/156856807782753868
Locher, P., Overbeeke, K.,  & Wensveen, S. (2010). Aesthetic interaction: A  framework. Design Issues, 26(2), 70–79.
https://doi.org/10.1162/DESI_a_00017
Masuda, T., Gonzalez, R., Kwan, L., & Nisbett, R. E. (2008). Culture and aesthetic preference: Comparing the atten-
tion to context of East Asians and Americans. Personality and Social Psychology Bulletin, 34(9), 1260–1275. https://doi.
org/10.1177/0146167208320555
McCabe, C., Rolls, E. T., Bilderbeck, A., & McGlone, F. (2008). Cognitive influences on the affective representation of
touch and the sight of touch in the human brain. Social Cognitive and Affective Neuroscience, 3(2), 97–108. https://doi.
org/10.1093/scan/nsn005
Millis, K. (2001). Making meaning brings pleasure: The influence of titles on aesthetic experiences. Emotion, 1(3),
320–329. https://doi.org/10.1037/1528-3542.1.3.320
Montoya, R. M., Horton, R. S., Vevea, J. L., Citkowicz, M., & Lauber, E. A. (2017). A re-examination of the mere
exposure effect: The influence of repeated exposure on recognition, familiarity, and liking. Psychological Bulletin,
143(5), 459–498. https://doi.org/10.1037/bul0000085
Mullennix, J. W., & Robinet, J. (2018). Art expertise and the processing of titled abstract art. Perception, 47(4), 359–378.
https://doi.org/10.1177/0301006617752314
Nadal, M., & Skov, M. (2017). Top-down and bottom-up: Front to back: Comment on ‘Move me, astonish me’. Physics
of Life Reviews, 21, 148–149. https://doi.org/10.1016/j.plrev.2017.06.013
Newman, G. E., & Bloom, P. (2012). Art and authenticity: The importance of originals in judgments of value. Journal of
Experimental Psychology. General, 141(3), 558–569. https://doi.org/10.1037/a0026035
Palmer, S. E.,  & Schloss, K. B. (2010). An ecological valence theory of human colour preference. Proceedings of the
National Academy of Sciences, 107(19), 8877–8882. https://doi.org/10.1073/pnas.0906172107
Pelowski, M., Markey, P. S., Forster, M., Gerger, G., & Leder, H. (2017). Move me, astonish me . . . delight my eyes
and brain: The Vienna integrated model of top-down and bottom-up processes in art perception (VIMAP) and cor-
responding affective, evaluative, and neurophysiological correlates. Physics of Life Reviews, 21, 80–125. https://doi.
org/10.1016/j.plrev.2017.02.003
Pihko, E., Virtanen, A., Saarinen, V. M., Pannasch, S., Hirvenkari, L., Tossavainen, T., . . . Hari. (2011). Experienc-
ing art: The influence of expertise and painting abstraction level. Frontiers in Human Neuroscience, 5, 94. https://doi.
org/10.3389/fnhum.2011.00094
Rauss, K., & Pourtois, G. (2013). What is bottom-up and what is top-down in predictive coding? Frontiers in Psychology,
4, 276. https://doi.org/10.3389/fpsyg.2013.00276
Schloss, K. B., Nelson, R., Parker, L., Heck, I. A., & Palmer, S. E. (2017). Seasonal variations in colour preference.
Cognitive Science, 41(6), 1589–1612. https://doi.org/10.1111/cogs.12429
Senzaki, S., Masuda, T., & Nand, K. (2014). Holistic versus analytic expressions in artworks: Cross-cultural differences
and similarities in drawings and collages by Canadian and Japanese school-age children. Journal of Cross-Cultural Psy-
chology, 45(8), 1297–1316. https://doi.org/10.1177/0022022114537704
Shea, N. (2014). Distinguishing top-down from bottom-up effects. Perception and Its Modalities, 73–91.
Silvia, P. J. (2005). Emotional responses to art: From collation and arousal to cognition and emotion. Review of General
Psychology, 9(4), 342–357. https://doi.org/10.1037/1089-2680.9.4.342

473
Aenne A. Brielmann

Steinbeis, N., & Koelsch, S. (2009). Understanding the intentions behind man-made products elicits neural activity in
areas dedicated to mental state attribution. Cerebral Cortex, 19(3), 619–623. https://doi.org/10.1093/cercor/bhn110
Strauss, E. D., Schloss, K. B., & Palmer, S. E. (2013). Colour preferences change after experience with liked/disliked
coloured objects. Psychonomic Bulletin and Review, 20(5), 935–943. https://doi.org/10.3758/s13423-013-0423-2
Szubielska, M., Imbir, K., & Szymańska, A. (2019). The influence of the physical context and knowledge of artworks on
the aesthetic experience of interactive installations. Current Psychology, 1–14.
Taylor, C., Clifford, A., & Franklin, A. (2013). Colour preferences are not universal. Journal of Experimental Psychology:
General, 142(4), 1015–1027. https://doi.org/10.1037/a0030273
van Paasschen, J., Bacci, F., & Melcher, D. P. (2015). The influence of art expertise and training on emotion and prefer-
ence ratings for representational and abstract artworks. PLoS One, 10(8), e0134241. https://doi.org/10.1371/journal.
pone.0134241
Vartanian, O., Navarrete, G., Chatterjee, A., Fich, L. B., Leder, H., Modroño, C., . . . Nadal. (2019). Preference for
curvilinear contour in interior architectural spaces: Evidence from experts and nonexperts. Psychology of Aesthetics,
Creativity, and the Arts, 13(1), 110–116. https://doi.org/10.1037/aca0000150
Vessel, E. A., Isik, A. I., Belfi, A. M., Stahl, J. L., & Starr, G. G. (2019). The default-mode network represents aesthetic
appeal that generalizes across visual domains. Proceedings of the National Academy of Sciences of the United States of America,
116, 19155–19164. https://doi.org/10.1073/pnas.1902650116
Vessel, E. A., Maurer, N., Denker, A. H., & Starr, G. G. (2018). Stronger shared taste for natural aesthetic domains than
for artifacts of human culture. Cognition, 179, 121–131. https://doi.org/10.1016/j.cognition.2018.06.009
Vessel, E. A., Starr, G. G., & Rubin, N. (2012). The brain on art: Intense aesthetic experience activates the default mode
network. Frontiers in Human Neuroscience, 6, 66. https://doi.org/10.3389/fnhum.2012.00066
Vogt, S., & Magnussen, S. (2007). Expertise in pictorial perception: Eye-movement patterns and visual memory in artists
and laymen. Perception, 36(1), 91–100. https://doi.org/10.1068/p5262
Wagner, V., Menninghaus, W., Hanich, J., & Jacobsen, T. (2014). Art schema effects on affective experience: The case of
disgusting images. Psychology of Aesthetics, Creativity, and the Arts, 8(2), 120–129. https://doi.org/10.1037/a0036126
Wallisch, P., & Whritner, J. A. (2017). Strikingly low agreement in the appraisal of motion pictures. Projections, 11(1),
102–120. https://doi.org/10.3167/proj.2017.110107
Yang, T., Silveira, S., Formuli, A., Paolini, M., Pöppel, E., Sander, T., & Bao, Y. (2019). Aesthetic experiences across
cultures: Neural correlates when viewing traditional Eastern or Western landscape paintings. Frontiers in Psychology,
10, 798. https://doi.org/10.3389/fpsyg.2019.00798
Zajonc, R. B. (1968). Attitudinal effects of mere exposure. Journal of Personality and Social Psychology Monographs, 9(2,
Pt.2), 1–27. http://dx.doi. https://doi.org/10.1037/h0025848

474
26
PREFERENCES NEED INFERENCES
Learning, valuation, and curiosity in aesthetic
experience

Sander Van de Cruys, Jo Bervoets and Agnes Moors

More than 40 years ago, pioneering social psychologist Robert Zajonc (1980) published his seminal work
titled “Preferences Need No Inferences,” in which he argued for the primacy of affect over cognition. Affec-
tive evaluation (the preference) comes first, he claimed, and only then do cognitive processes (the inferences)
kick in. The central piece of his evidence for this was the mere exposure effect: the finding that the mere
repeated presentation of a stimulus increases its liking, no matter whether the stimulus is (consciously) per-
ceived or categorized. Bracketing the discussion of the evidence for this effect for now (we will get to that),
watch how deceptive the mere exposure concept is in light of what we have come to understand about the
perceptual system in the past decades. “Mere exposure” perpetuates the myth that there is some raw sense
data to expose (“merely”) to the agent and to be “picked up” passively by the brain. The empirical evidence,
however, carried by the Bayesian view of perception, abundantly shows that perception is an active, (re)con-
structive process, biased from the very start by (implicit) hypotheses and expectations. Perceptual illusions are
obvious examples here, but they only illustrate the general principle, which is that the visual system makes
sense of ambiguous, imprecise sensory information by combining it with priors or probabilistic hypotheses.
If “mere perception” already consists of inferential processes, it is impossible for preferences to “need no
inferences,” as Zajonc had it. Moreover, his thesis is fundamentally at odds with the recent idea that the
brain performs all its functions through (approximate) Bayesian inference. This influential theory is known
as predictive processing (also called the Bayesian brain or active inference). It holds that perception and learning are
inference and, perhaps counterintuitively, valuation too.
In this chapter, we will tie these three—inference, learning and valuation—together to find answers to the
classical questions of what we look for (curiosity), what we appreciate (aesthetically) and why we prefer what
we prefer. We bracket “aesthetically” because we consider art an activity continuous with our more mundane
sensorimotor activities and experiences, so we should not expect separate answers with respect to the moti-
vational and affective principles that govern it. Our emphasis throughout will be on valuation—understood
as the process of how we come to value, prefer or appreciate—as a function of learning and inference.
Importantly, we will focus on so-called non-reinforced preference, meaning preferences that are not due to
pairing of stimuli with rewards or punishments (e.g., Boddez et al., 2019). Indeed, one of the counterintui-
tive aspects of the predictive processing theory is that it has no conventional concept of rewards, goals or
values. It eschews the classical sharp schism between the epistemic (beliefs, representations, cognitions) and
the conative (desires, preferences, motivations). However, what it puts in its stead is key to tackle traditional

DOI: 10.4324/9781003008675-28 475


Van de Cruys, Bervoets and Moors

conundrums in the science of aesthetic experience. It forces us to radically rethink how value emerges and
to reinterpret conventional strands of thinking on (aesthetic) valuation.
In what follows, we will set up our problem by way of a brief review of the major theories concerning
appreciation and curiosity in psychology. Curiosity and appreciation seem an inseparable pair when trying
to understand aesthetic experience, and yet, we still miss good integrated accounts. As we will argue in the
subsequent section, this is due to the lack of a language for articulating the beholder’s share, the precise kind
of active involvement of the subject with the artwork. Next, we give a brief overview of the state of the art
in the predictive processing theory and propose that it provides such a language. The subsequent section on
mechanisms of valuation unpacks this and provides the core of our argument. In the final part of the chapter,
we will illustrate this with a new take on the findings on mere exposure as well as (aesthetic) valuation in
general. The final sections try to broach the most gripping but ineffable capacities of art with the same tack.

A brief history of appreciation and curiosity in psychology

Appreciation
The Gestalt psychologists were among the first to put the question of what we prefer to look at and why
we have those perceptual preferences on the scientific agenda (Wagemans et al., 2012). Their answer can
be summarized as the perceptual parsimony thesis, where “parsimony” comprised such characteristics as order,
simplicity and symmetry (e.g., Koffka, 1935). The Gestalt psychologists argued that we tend to organ-
ize visual inputs in the simplest, most orderly way, and this is also what we prefer perceptually. While the
Gestalt tradition considered parsimony largely determined by the input properties, much later the emphasis
was shifted toward parsimony of processing, in the so-called processing fluency tradition (e.g., Reber et al.,
2004). Many of the same characteristics (order, symmetry, etc.) are thought to increase the ease or fluency
of processing, in addition to several factors that did not enter the equation for Gestalt psychologists, such
as familiarity (repetition) and prototypicality. This marked an important move towards a subject-dependent
definition of parsimony, away from purely stimulus-bound characteristics. Appreciation now is very much
dependent on individual processing and learning.
Of course, the processing fluency theory was proposed in large part to explain the swath of studies on the
mere exposure effect (Bornstein, 1989; Zajonc, 1968), under the assumption that, with repeated encounters,
the processing of a stimulus happens more fluently. To further explain the fluency phenomenon, and to rec-
oncile it with the work on reinforcement learning, the positive affective mark of fluency is sometimes seen as
a safety signal, a quelling of our assumed innate fear of the unknown with repeated “uneventful” experiences
(i.e., fluency as learned inhibition of fear; e.g., Winkielman et al., 2003). Another explanation of the posi-
tive effect of fluency considers it a signal that one has been able to (cognitively) deal with the stimulus in the
past. Here, liking would be the result of misattributing a characteristic of the subjective processing (i.e., the
ease) to the external stimulus. It is the stimulus that is liked, even though what is monitored and appraised is
a property of processing it. The liking of fluency is then a metacognitive signal of processing quality (Alter &
Oppenheimer, 2009): how successful I am (or have I been) in processing the current inputs. We will return
to this important idea later on.
Empirical findings have cast serious doubt on the parsimony thesis, whether in the “objective” form pro-
posed by the Gestalt tradition or in the “subjective” fluency/mere exposure tradition. People often seem to
prefer or appreciate medium complexity or medium orderliness (termed the Goldilocks principle) rather than
the most fluently processed, ordered or simple stimuli. This is of course apparent in art (Figure 26.1), where
certain violations of familiarity or order are often conducive to aesthetic appreciation, but it also holds in eve-
ryday life and controlled experiments. For example, a large meta-analysis of mere exposure studies (Montoya
et al., 2017) found that while appreciation increases with repeated presentations, most studies also found an

476
Preferences need inferences

Figure 26.1 Art, a prime example of appreciated stimuli, often breaks order or simplicity. For example, in these works by
Jan Vanriet (top; Séance, oil on canvas) and Gustav Klimt (bottom right; Reclining Woman) familiar shapes are
not depicted in their simplest, most recognizable forms. Similarly, the computer rendition by Neil Dodgson
(2009) inspired by Bridget Riley’s op-art (bottom left), includes clear violations of symmetry. The work by
Vanriet is reprinted with permission from the artist. The drawing by Klimt is reproduced from Wikipedia
Commons (https://commons.wikimedia.org/wiki/File:Klimt_-_Ausgestreckte_Frau_001.jpg). The work
by Dodgson is reprinted with permission from the artist.

inverted U-curve function, indicating that there is an optimum in liking for a medium number of repeti-
tions, after which liking goes down again (sometimes described as a boredom effect; Bornstein et al., 1990).
Of course, this directly links to Berlyne’s (1960) work on preference for stimuli of intermediate complexity
(see also Chmiel & Schubert, 2017).
Walker (1981, p. 40) summarizes Berlyne’s theory in two basic postulates: “1. There is an optimal level of
psychological complexity for a psychological event that will be preferred to either simpler or more complex
events. 2. Repeated experience of an event will lead to progressive simplification of that event,” thereby
making the connection between repeated exposure and complexity (simplicity; cf. Gestalt school) or fluency

477
Van de Cruys, Bervoets and Moors

explicit. Walker goes on to review the evidence for both postulates, concluding that there is reasonable sup-
port for both but that many methodological challenges hamper the empirical identification of an inverted
U-curve (provided there is one). For example, stimulus sets or contexts usually have a strong effect, where
appreciation (and complexity) ratings greatly depend on the particular number, order (serial dependency) and
types of stimuli in the test set. This problem is compounded by possible nonlinearities in both the complexity
and the liking dimension and by the fact that people do not even seem to process all information in complex
stimuli. Together these factors bias how much of the complexity continuum participants have experienced.
Even more problematic, participants may be evaluating different things (from the intended dimensions), pos-
sibly depending on the language used by the experimenter when describing the task of rating complexity or
appreciation. One counterintuitive consequence of all these confounds is that an (partial or shifted) inverted
U-curve pattern could be present in all of the participants but not (or even a regular, non-inverted U-curve)
in the shape of the average data, or vice versa (see Güçlütürk et al., 2016; Spehar et al., 2016; Walker, 1981).
Underlying these problems, is, as so often in psychology, the mere measurement effect clouding possible
U-curve relations. That is, the very fact of presenting, processing and measuring will substantially influence
the measurement. Every repeat of a stimulus with a certain level of complexity will in fact be a different
event, with a different psychological complexity, so averaging, even across the same stimulus (level), may bias
the shape of the curve. Indeed, the second postulate—progressive simplification with experience—urges us
to pay attention to subjective complexity/simplicity for a particular individual with her particular learning
history. Nonetheless, many early studies (e.g., Terwilliger, 1963) on the relation between complexity and
liking relied on objective complexity (measured for example as the number of angles in geometric shapes).
More recent elegant experiments in young infants managed to improve on this by quantifying (subjective)
complexity based on the presentation history of stimuli of a given participant (Kidd et al., 2012, 2014). With
this more individualized, computational approach, the authors could confirm a preference for intermediate
complexity (or predictability) in visual and auditory sequences. In parallel, studies on semantic-level com-
plexity also underline the need to account for subjective complexity that is not computable based on stimulus
features alone. For example, Nicki (1970) found that people prefer photographs of objects with intermedi-
ate blur rather than lower or higher blur but that this is likely due to the fact that images with intermediate
blur elicit a higher number of guesses about the content of those images, with even confidence about the
guesses (see also Van de Cruys et al., 2021). In other words, while intermediate uncertainty or blur seemed
to be preferred, this was actually the consequence of a preference for the most (semantically) unpredictable
or uncertain images. Indeed, research in artworks shows that low-level image statistics of complexity or
uncertainty explain little of the variance in appreciation (Van Geert & Wagemans, 2020) but that semantic
uncertainty (ambiguity) is a crucial factor in appreciation (Muth et al., 2016; Wang et al., 2020).
In recent years, the importance of taking into account the individual learning history has been bolstered
by studies showing that particularly relative fluency or exposure matters for liking rather than the absolute,
momentary fluency of a stimulus or trial. For example, Wänke and Hansen (2015) have shown that the typi-
cal mere exposure effect could be replicated only when the old stimuli were mixed with new stimuli. Also
emphasizing individual learning, Forster et al. (2015) showed that an effect of ease of processing could only
be found in within-participant (not between-subject) comparisons. Similarly, relative exposure, that is, being
presented more often than other stimuli in the set, increases liking more than absolute exposure (Mrkva &
Van Boven, 2020). Together, such findings suggest that participants implicitly track the expected fluency
across the recent history of experiences and attribute value to deviations from this expectation, a notion that
will become important later on.
An emphasis on learning and processing dynamics may hold the key to a reconciliation of the “simplicity-
fluency” camp and the “intermediate complexity” camp as well. In particular, one could describe what per-
ceivers like in terms of the subjective progress made in dealing with the stimulus (or activity), in other words,
a reduction in disfluency (Graf & Landwehr, 2015; Muth et al., 2015; Schmidhuber, 2009; Van de Cruys &

478
Preferences need inferences

Wagemans, 2011) or an increase in the system’s compression of a given stimulus, going from unpredictable
(or complex) to predictable (or simple). If it seems as if humans like simple stimuli, we are confusing a static
end product with the crucial process. If it seems as if we prefer medium complexity, we are confusing a static
starting point with the crucial process (Van de Cruys, 2018) because the chance of learning progress is the
highest for environments of medium subjective complexity. Under those conditions, we have some idea as
to what type of regularities rule the situation or environment, but a level of active (mental) engagement or
learning is required to resolve the remaining uncertainty.

Curiosity
While curiosity and intrinsic motivation clearly direct our engagement with and shape our appreciation
of art, few behavioural studies have explicitly examined the interplay of curiosity and appreciation in the
aesthetic experience. This is even more remarkable given that discussions on appreciation are partly mir-
rored in those about curiosity, especially with regard to the role of complexity and uncertainty. This may
in part be due to the fact that the same dependent measures, namely choice or preference behaviour, have
been used to infer participants’ appreciation in some studies and their curiosity in other studies, so they are
hard to disentangle in practice, even though they are cognitively and affectively clearly distinct. For good
measure, we understand curiosity as “a tendency or motive to acquire or transform information under
circumstances that offer no immediate adaptive value for such activity” (Livson, 1967, p. 75). It is a form
of motivation that is “inherent within information processing and action” (Hunt, 1981). Broadly speaking,
there are four (types of ) theories of curiosity. Optimal level theories (e.g., Berlyne, 1978) argue, in parallel with
the inverted U-curve discussion for appreciation, that medium complexity or uncertainty elicits the most
curiosity. Although this is a descriptive theory, Berlyne proposed that medium complexity corresponds to an
optimal, medium arousal that animals would gravitate towards in a homeostatic sense. However, it remains a
puzzle why a system would strive to be in a medium arousal state in the first place, and, indeed, subsequent
work has largely discredited the connection with arousal in Berlyne’s thesis (Silvia, 2005; Walker, 1981).
The second and currently most popular theory of curiosity is Loewenstein’s (1994) account of curiosity as
our sense of a gap in our knowledge, quantified using information theory (as the entropy, i.e., the number of
alternative options and their probability). The gap theory clearly captures something important, namely that
curiosity can be an increasing function of uncertainty relative to one’s current knowledge and expectations
rather than reaching a peak for intermediate uncertainty, something that later empirical studies have also
confirmed (e.g., Van de Cruys et al., 2021; van Lieshout et al., 2018). However, the computational tools of
classical information theory (entropy) are limited in truly shedding light on curiosity. For example, entropy
is only defined once particular predictions (or questions) are formulated, but Loewenstein provides few clues
on how our system generates those. Another problem in Loewenstein’s theory stems from his framing of the
curiosity gap as a classical drive: something that needs to be “removed” by action. Because people want to
resolve it, Loewenstein (and many others after him) infer curiosity must be aversive. In the same contorted
way, Freud counted sexual arousal among the aversive drives because the orgasm terminates the arousal
(Hunt, 1981). Obviously, curiosity, like sex, is usually pleasurable and appetitive, even though the uncertainty
at its core (the “gap”) is indeed often a cause or catalyst of anxiety (Hirsh et al., 2012). So what is missing
here? When is uncertainty not a bad thing?
The missing element may very well be provided by a third theory on curiosity that is based on the
appraisal theory of emotions. Silvia (2005) proposed that curiosity (or “interest,” as he prefers to call the emo-
tion associated with intrinsic motivation and learning) is the result of a combination of two different implicit
cognitive appraisals. The first is the novelty appraisal, which, analogously to Berlyne and Loewenstein, is
an evaluation of whether something is new, unfamiliar, uncertain, complex or unexpected (i.e., the “gap”).
The second crucial ingredient is an appraisal of coping potential. It involves an estimation of one’s ability to

479
Van de Cruys, Bervoets and Moors

understand or deal with the new, unexpected event, in the sense of rendering it predictable or meaningful
again. Like (relative) fluency previously, this is a metacognitive evaluation, but here it is about the extent to
which one expects to be able to reduce uncertainty. One might describe “promised insight” (Muth et al.,
2015) as an expected instead of an actual fluency. As long as there is this component to curiosity, uncertainty
can be a source of joy and “approach” behaviours, instead of just anxiety. Although we may be more likely to
be able to reduce uncertainty for medium levels of uncertainty, there is nothing that strictly binds curiosity
to just this level of uncertainty.
A fourth and final strand of theorizing on curiosity (Keller & Voss, 1983; Livson, 1967; McReynolds,
1971) dates from the 60s but has been largely forgotten, at least in comparison to Berlyne’s influential work
from the same era. Based on animal research, McReynolds (1971) developed a view of curiosity as “an
expected rate of cognitive structuring that an individual tries to maintain by exploration (or lack thereof  ).” Cog-
nitive structuring is a somewhat dated term, but McReynolds used it to refer to the process of assimilation by
which new percepts are made to fit with the existing mental schemata or models. Like Berlyne’s theory, this
view of curiosity has a strong adaptive, homeostatic flavour to it, not so much in the physiological domain
(cf. Berlyne’s arousal) but in the information processing domain. Originally, McReynolds used the term
perceptualization rate to denote the expected rate of structuring, inspired by Glanzer (1958) who described the
organism as “an information processing system that requires certain amounts of information per unit time.”
However, later on, McReynolds perceptively remarks that the essential variable to be tracked and regulated
is not the stimulation or some “raw” information rate but rather “the rate of cognitive structural change”
(McReynolds, 1971, p. 37). Hence, he preferred the concept of innovation rate centring on the assimilation
or learning process. The system’s core concern is to minimize “unassimilated material”—the uncertain,
unexpected inputs—and to optimize “innovation rate.” In more modern terms, curiosity is driven by the
expected rate of updating one’s mental models (the expectation of uncertainty reduction relative to one’s
mental models). In this way, the expected rate-based explanation of curiosity further specifies the previous
account centred on coping potential, or expected reducibility of uncertainty (i.e., a positive rate).
Using examples from rodent research, McReynolds reasoned that both long-term experience (e.g., “rear-
ing environment”) and short-term experience (cf. deprivation studies) could contribute to forming those
expected or “accustomed” rates of processing. If organisms form adaptive expectations on the rate of “struc-
turing” or uncertainty reduction that they try to maintain, this would require active exploration (as the
expression of curiosity) to fulfil: “an active seeking for new data to be digested” (McReynolds, 1971). It is
easy to see our creative, artistic explorations as (one of  ) the system’s efforts to upregulate the rate to match
the expected one. A concrete recent example was provided by the COVID-19 pandemic lockdown, during
which people apparently en masse took up learning to play an instrument (Hill, 2021). At the other end of
the spectrum, some environments may raise uncertainty levels too quickly too much, which also prevents
expected structuring rates from being re-established. Instead of exploration, here one would see (anxious)
avoidance or repetitive, rigidly structured (stereotypical) forms of self-stimulation to reinstate the expected
rates of uncertainty reduction (as also found in clinical disorders; e.g., Van de Cruys et al., 2014). The parsi-
mony of McReynolds’ view of anxiety and curiosity (and their behavioural expressions) as two sides of the
same coin is attractive and intuitively plausible and will become important for aesthetics as well.
Let us wrap up this historical overview by asking why we need a sense of curiosity in the first place.
A plausible answer is that we need it because learning is costly and fallible, so being able to sensitively direct
resources to where the best learning progress can be made has a considerable advantage for the system. It
entails learning not only the “subject matter” but also (meta-)learning what is learnable based on how your
mental models performed in similar environments (Gottlieb et al., 2013). Indeed, even young infants (in
artificial grammar learning) seem to already have a sense of where learnable inputs are situated rather than
just noise (Gerken et al., 2011), here investigated through a looking times paradigm. Hence, curiosity and
exploration can ensure that we remain in what Vygotsky (1962) has called the zone of proximal development

480
Preferences need inferences

(Metcalfe et al., 2020; Oudeyer et al., 2007), the optimal region for learning just above one’s current mental
models or abilities. Specifically, it allows the agent to avoid wasting resources on inputs that are either mere
noise (nothing to learn), too easy (all regularities are already learned) or too complex (for the current mental
models). One interesting way to implement this is through contextually adjusted expected rates of uncer-
tainty reduction, as McReynolds envisaged.

The quest for the beholder’s share


However, an underlying reason for the lack of work on the interplay of curiosity and appreciation is a lack
of a conceptual language to bring this to the fore. Most accounts of appreciation have been applications
of classical information theory and so adopted the vernacular of a passive receiver of a static artwork (the
sender’s “message”), leaving little room for the proactive involvement of the subject. While information
theory has clearly led to crucial insights (cf. Loewenstein), it also comes with serious limitations in articulat-
ing the beholder’s share (Gombrich, 1963; Seth, 2019). It misses a formal account of how the expectations
are built through learning and how they are brought to bear in experience. Indeed, if our goal is to explain
aesthetic experience, our analysis suggests those expectations are formed not only about the content level but
also about the meta-level of the content-generating processes (e.g., expected fluency or rates of uncertainty
reduction).
It must be said that the active involvement of the subject and the role of internal dynamics in the aes-
thetic experience have not been overlooked by philosophers and psychologists working on the subject
(e.g., Dewey, 2005; Hume, 2008). Notably, in 1942 Eysenck published his “law of aesthetic appreciation,”
with great ambition. The law states that “The pleasure derived from a percept as such is directly proportional
to the decrease of energy capable of doing work in the total nervous system, as compared with the original
state of the whole system” (Eysenck, 1942). There are three notable things in Eysenck’s law. First, aesthetic
pleasure is about the frugal use of energetic resources. Second, aesthetic appreciation is subjective in that the
energy changes are measured against a given system’s prior state. Third, the law is decidedly dynamic in
nature. Eysenck was clearly inspired by the early Gestalt psychologists like Köhler here, who had a pretty
dynamic view on perceptual experience already (using the language of fields and equilibrium processes;
Köhler, 1920; Wagemans, 2015). However, it failed to gain traction, as was also the case for Eysenck’s law.
Indeed, as we sketched earlier, the mistaken picture that survived from the Gestalt tradition is one centred
on static order and simplicity.
It is tempting to blame this disregard for internal dynamics on the cognitive revolution born in the slip-
stream of the rise of information theory and heralding the computer metaphor of the mind. Indeed, some
authors (e.g., Cupchik & Heinrichs, 1981) lament that the static, atomistic approach of tracking objective
probabilities of occurrence of singular elements or events in an artistic “stimulus” (as per information theory)
has limited import in understanding (aesthetic) appreciation. For example, Cupchik and Heinrichs (1981)
complain that “The notion that dabs of colour have a quantifiable uncertainty of appearing next to each
other does not reflect the structure and creative origin of the work” and that “information as a quantity is
different and independent from the meaning of a work of art.” On the face of it, this critique crushes the
hope that one could come up with a broadly applicable, formal account of aesthetic appreciation. However,
in the next sections, we will discuss the ways in which the predictive processing account helps to bridge the
distance from mere “atomistic” information to meaning. Information theory will remain central, although
on its own it indeed falls short in capturing the relevant processes.
That brings us to the actual reason that the dynamic view of Köhler, Eysenck and others (see also Arn-
heim, 1974; Pepperell, 2018) did not catch on at the time: They failed to articulate how their metaphoric
theorizing on forces, fields or energy dynamics mapped onto tractable psychological or neurophysiological
variables. For example, Eysenck (1942) writes: those external stimuli will be judged the most beautiful which

481
Van de Cruys, Bervoets and Moors

are most in agreement with the internal forces of perception” (our emphasis). But what are those internal forces
of perception? Have we made any progress in the last decades that could give us another shot at those richer
accounts of aesthetics of the days before the cognitive turn? We still do not have a good grasp on the coding
scheme, the “internal forces,” of the brain, which would be necessary to determine subjective (brain-based)
energetic or processing gains across time. However, one plausible, well-elaborated candidate for this coding
scheme is provided by hierarchical predictive processing.

Predictive processing in brief

Producing future
According to poet and philosopher Paul Valéry, the purpose of the brain is to produce future. In recent
decades, that very idea has been computationally fleshed out in the theory of predictive processing, known
under different guises such as the Bayesian brain or active inference (Clark, 2013; Friston, 2010; Hohwy, 2020).
To “produce future” efficiently means not just to respond to stimuli but to proactively model and predict
sensory inputs and the opportunities and challenges they represent. In its most basic sense, being (alive)
means creating those conditions that allow one to persist into the future and not dissipate under the second
law of thermodynamics. In information-theoretic terms, we can say that there is a limited set of expected or
“target” states that is consistent with continued existence and so has to be maintained (cf. homeostasis). More
concretely, an organism has become equipped through evolution with particular interoceptive expectations
(e.g., blood glucose or temperature at a particular level), which evolution has “discovered” to be consistent
with survival. For the most part, it cannot fulfil those interoceptive expectations on its own but only by going
via its environment (e.g., finding a food source) (Pezzulo et al., 2015).
To do so, the organism must possess or learn a (minimal) generative model of its environment. It is called
“generative” because it is a model about how the expected (interoceptive) observations can be generated or
recreated. We only have the sensory effects of the processes in the world around us to go on, but from the
regularities in those sensory inputs, we can infer their likely hidden causes. So we will experience a world
populated with objects such as rocks, animals and clouds. The pattern of inputs created by the cloud on our
senses is more alike than that of not-cloud, for example, that pattern follows a similar motion on the retina
(cf. grouping by “common fate”). The cloud “object” explains such input patterns, as well as the exact pat-
tern of darkness and temperature over time (and space) that I sense when the cloud covers the sun, and so
on. Ultimately, we see clouds and other hidden causes because it helps us predict exteroceptive (e.g., when
the sky turns dark grey, it’s likely to rain soon) and interoceptive states (e.g., I will get cold).
To turn events to our advantage and actually realize our target interoceptive states, we need to learn not
only about our surroundings but also about ourselves as a hidden cause. We are just another inferred hidden
cause in the world, albeit one with which we have very intimate and rich experience. By the kind of stable
correlated patterns in our different senses that our actions tend to cause, we discover who we are and how
our own actions can accomplish expected interoceptive observations (e.g., when putting on a raincoat, I will
not get cold). The same principle that allows us to see clouds and other objects—stable patterns of sensory
inputs can be used to successfully predict our surroundings—readily extends to the intentions and goals that
we perceive in ourselves and others: These inferred mental causes are as real as clouds or other objects we
perceive (Dennett, 1991), though they are situated at a hierarchically higher (or “deeper”) level summariz-
ing larger patterns in inputs. This does not mean that social perception is any less direct. We will perceive
coordinated motions of limbs immediately as intentions, just because of the predictive power of such inferred
causes that abstract away the concrete minor variations in the constituent motions.
We experience the world as populated with objects and intentions, because, irrespective of their veridical-
ity, this is the interface that works best in navigating our world and fulfilling our interoceptive expectations.

482
Preferences need inferences

By continually predicting incoming inputs based on current context, we actively construct our interface to
the world around us. Only in our failures, the mismatch between our constructed pattern and the observed
inputs, do we meet reality (von Glasersfeld, 1995). It is those prediction errors that keep our constructs in
check, so they continue to “work.” Input activity that is correctly predicted is suppressed (explained away),
while the errors are sent upwards to update predictions. This allows the system to reserve its resources for
unexpected observations (prediction error) while building increasingly efficient reconstructions of those
observations. By minimizing prediction errors, iterated across levels in the brain, we infer latent causes that
best explain the regularities in impinging inputs. Mathematically, the process is a form of gradient descent
optimization scheme, often described as descending an error “landscape” to find its minima (for computa-
tional details, see Aitchison & Lengyel, 2017; Gershman, 2019; G. B. Keller & Mrsic-Flogel, 2018; Smith
et al., 2021; Spratling, 2017).
Both perception (inference) and learning can be captured with this (approximate Bayesian) predictive
updating scheme. The speed and automaticity with which the visual system uses its generative model to
disambiguate the “hidden causes” of the sensory inputs and settles on the best explanation (e.g., this is figure,
that is ground) hides the underlying inferential process. We have to severely distort images, as we do to create
so-called Mooney images (Figure 26.2), to obstruct this process and block the use of our learned hierarchical
models. Even then, a very brief priming of the right semantic model is often sufficient to cause a dramatic
shift in how we see the image and as well as a near inability to unsee the right solution. The newly inferred
cause seems so strongly supported by image cues that we cannot but assume it was always there in the image
and not just in our experience.
The error-based updating scheme is repeated at each level of the cortical processing hierarchy, with each
higher level trying to predict activity in the region below. On a local level, “predicting” should be under-
stood in the technical sense of reconstructing input activity (from the lower region) quasi-simultaneously,
broader than the everyday notion of prediction as anticipation. However, across the hierarchy, predictions are
formed capturing statistical regularities that span more and more space and time. Take the example of read-
ing a book. While reading, we form expectations on event sequences across time in a story, which in turn

Figure 26.2 Two-tone or so-called Mooney images are created by blurring and thresholding greyscale photographs
(see the source photograph in Figure 26.3). They are examples of one-shot learning: Once you find or are
confronted with the solution, you cannot unsee it. “Discovery” of the familiar structure in the image usually
gives a positive feeling of insight or Aha-Erlebnis. Notice the similarity of this type of image degradation
with the technique used in the artworks in Figure 26.1.

483
Van de Cruys, Bervoets and Moors

creates expectations on high-level meanings that we will likely encounter, as well as the kind of words that
may appear. Reading a sentence, we have particular expectations about which words will likely appear in
which locations (based on learned syntax and semantics), which can further be unpacked in expected letters
and their compositing features (oriented lines). The hierarchy forms a cascade of interdependent predictions
going all the way down to the peripheral level where predictions can be “answered” at the level of the retina
(or receptors of other senses). Recent evidence supports the existence of such a predictive hierarchy in the
brain (Heilbron et al., 2021; Rohe & Noppeney, 2015; Wacongne et al., 2011). Using these predictions,
we exploit the redundancies in reading materials, which allows us to be more selective and efficient in our
sampling of the inputs. Indeed, eye movement patterns show we jump from word part to word part in a sen-
tence, without having to “read” each letter (Karl J. Friston et al., 2018). The whole process happens largely
flawlessly, though typos may be missed (especially when we edit our own text, because of strong and accurate
predictions) and unexpected “turns of sentences” might need to be read twice.

Part of the act


So far, we have only covered one way of reducing prediction errors, namely by changing your predictions
or our models, known as perceptual inference and learning. The other, complementary way to minimize
prediction errors is by changing the things predicted: by acting to bring the world (our observations) closer
to your models of it. Hence, adaptations to improve the fit between world and mind can have two different
directions: world-to-mind (assimilation) or mind-to-world (accommodation). This closes the perception-action
cycle at the core of this framework, using “fit” as quantified by prediction errors as the unified optimization
criterion.
Actions can be treated with the same predictive machinery described previously, when they are concep-
tualized not as motor commands but rather as their expected sensory consequences, in proprioception (the
state of our muscles and tendons) and exteroception (e.g., feeling the cup in my hand). This idea goes back
to classical work on ideomotor theory ( James, 1890) and the view of behaviour as the control of perception
(Powers, 1973). Like in perception, action models are hierarchically organized with abstract goals and inten-
tions at the top to be unpacked into policies (action sequences) and further into specific motor programs
represented as the exteroceptive and proprioceptive outcomes that they are expected to realize. At the lowest,

Figure 26.3 A photograph of a frog that has been used to create Figure 26.2.

484
Preferences need inferences

peripheral levels, proprioceptive predictions are confronted with the current state (before any movement) of
muscle and tendon receptors to form prediction errors. Instead of leading to model updates, here prediction
errors trigger classical reflex arcs executing the movement (K. J. Friston et al., 2010). The generative action
models can be exploited for action execution, that is, inferring the action sequences that accomplish intended
effects. However, in line with the mirror neuron system (Kilner et al., 2007), the same model can be applied
for action inference as well, that is, inferring your own or others’ intentions from sensory effects of actions—a
framework that is now referred to as active inference.
One important role for actions is foraging for more information (conversely, more information enriches
our action capabilities). Indeed, predicting well also means predicting when one cannot reliably predict
something—when more sensory sampling is needed (e.g., through head and/or eye movements). We con-
tinuously fine-tune our rhythm and direction of sampling based on the current context in the service of
better prediction error minimization. For example, when reading a difficult text (or one in a foreign lan-
guage), our predictions do not provide a lot of support, so reading becomes effortful (i.e., with frequent
model updates) and sampling intensive. But to actively direct our sampling of the world in an optimal way,
we need to have a way of gauging the quality of the evidence and of our predictions. Those “quality predic-
tions” are called expected precisions, which, in essence, are estimates of uncertainty or reliability attached to
observations and our first-order predictions. Since precisions are (meta-)predictions as well (predictions of
prediction errors), they can be updated in the same way as the “regular” predictions. The role of precision in
inference is to regulate the relative influence of new evidence (prediction errors) versus the predictions, for
example, to allow top-down predictions to dominate our percept when sensory information is noisy (unreli-
able). Estimating the quality (precision) of upcoming information also allows us to direct actions to seek out
informative (precise) prediction errors: Those future prediction errors that are relevant and reducible through
actions or model updates.
Here, we quickly need to dispel the criticism often raised at predictive processing that it would lead to
immobilization instead of action: An organism driven by prediction error minimization would retreat from
the world into a dark room where all its errors will eternally be minimized (if you just predict darkness)
(Seth et al., 2020; Sun & Firestone, 2020; Van de Cruys et al., 2020). It should be clear by now, however,
that minimizing prediction errors on those lowest sensory levels does nothing to resolve the multilevel and
multimodal expectations that an organism holds. We have seen that organisms often have to build a world
of higher-level constructs such as objects and intentions to be able to fulfil their interoceptive target states.
Staying in a dark room will typically not remain “expected” for very long, given the particular characteris-
tic states (the phenotype) of an organism (Friston et al., 2012). This implies that to ensure prediction error
minimization in the longer run, we need to seek out prediction errors in the here and now, especially the
most informative (high-precision) ones with regard to our preferred states.
The “dark room” confusion is born out of the misconception of a mental model as a purely epistemic
thing, a search for correspondence to some truth out in the world. Contrary to this conception, remember
that the generative model as used in predictive processing doubles as a specification of our preferences. To
be viable, our model has to be equipped with some prior expectations about the states we will tend to end
up in. That means our needs and goals will be described as an expectation (technically a probabilistic prior
distribution) on observations; for example, we strongly (i.e., with high precision) expect to sense a tem-
perature between certain parameter values. In other words, our model has to be optimistically biased (Sharot,
2011), in that it might not represent the current state of the world but rather a state that can, fallibly, be
attained by one’s actions. Indeed, pure optimistic hallucination of more favourable conditions without a
generative model of how to reliably accomplish those conditions would quickly be the death of a system.
Similarly, a perfectly accurate, exhaustive representation without concern for (the intrinsic regularities that
reproduce the limited resources of  ) the very system that does the representing would drive the system to
self-destruction. Any perceptual system can only exist if it translates sensory inputs into a format that says

485
Van de Cruys, Bervoets and Moors

something about the further persistence of that perceiving system. This format need not necessarily be
completely veridical (Hoffman et al., 2015; Tschantz et al., 2020), as long as it works: If it infers those hid-
den causes and actions that allow it to efficiently “generate” its preferred or expected states and so minimize
prediction errors. This makes room for a controlled optimism (Van de Cruys et al., 2020) that is continually
“negotiated” with the world.
With that in mind, we can see that the possibility to act in the world truly urges the organism to become
future-oriented. Being able to choose actions shifts the focus from just the current, momentary predic-
tion errors to expected prediction errors in the future and how our actions can be chosen optimally to resolve
them. Here, we take our models offline as it were, and we create “artificial” prediction errors by confront-
ing our preferred outcomes with the expected outcomes based on our action models. We use temporally
extended (“deep”) counterfactual models to ask “What if I do this, instead of that, how well does that reduce
my expected prediction errors (uncertainty)?” (Friston et al., 2021). We play out different models against
each other (largely implicitly), and we let our models or hypotheses die in our stead (Dennett, 2017). We
choose our actions such that they minimize the divergence (i.e., prediction errors, under some assumptions)
between two probability distributions (models): one describing our expected outcomes if we were to follow
a course of action and another describing our desired future (our prior or preferred outcomes).
To do this, we of course first need to resolve uncertainty or confidence about which actions lead to which
outcomes. So in choosing actions, the balance that an organism needs to strike entails consideration of both
epistemic value and pragmatic value (“rewards” or prior preferences) on equal footing (Friston et al., 2015).
Maximizing epistemic value (or information gain) means exploring the world to disclose its structure, to
reduce future prediction errors. Once this structure is known, one can exploit it to realize one’s prior prefer-
ences or expected observations (i.e., pragmatic value). Indeed, this intuition is vindicated by mathematical
treatments of active inference which show that minimizing uncertainty (expected prediction errors) relative
to the generative model can be decomposed into maximizing epistemic and pragmatic (goal-fulfilling or
reward-maximizing) value (Friston et al., 2015).
Psychologists in the 1970s have done experiments where people had to choose between tasks with a
range of different difficulty levels (Schneider  & Heckhausen, 1981; Trope, 1975). It turned out people
choose intermediate difficulty levels (somewhat biased in the optimistic direction, i.e., expecting success
levels slightly higher than their objective performance level), where subjective uncertainty with respect to
task outcome is greatest: I am sure I will win when I pick low difficulty and I will lose when I pick high
difficulty. Winning did not gain them anything in this task, so they chose the most uncertain option. Indeed,
when (pragmatic) value is equivocal, epistemic value determines action, according to predictive processing.
You first learn (to reduce uncertainty about) the structure of the world; then you will be able to exploit that
structure to generate the states of your strongest priors (preferences).
A sceptic might object that it is overly reductionist to posit that the mind (and all of our behaviour) is
governed by the minimization of uncertainty. It defies our intuitions, as well as the history of psychology,
where conative constructs like needs, desires and goals have always taken a central role. However, as we saw,
conative constructs are not made obsolete by predictive processing; they are merely absorbed in the model
that the organism embodies (Clark, 2020). In this view, what we colloquially call goals can be recast as stable
expectations in the sense that they usually have a longer temporal horizon, and a higher precision, mean-
ing that they are more robust to negating evidence, so they will be realized by actions (using precise action
models) instead of prediction updating. To put it more positively, minimizing uncertainty is equivalent to
maximizing the evidence for the model that defines us, often described as self-evidencing (Hohwy, 2014).
Generative models are not purely representational but bent towards desired states of the organism. It is a
record of representations (hidden states) that best explained past inputs, as well as a target to be realized in
future inputs with our actions (Hafner et al., 2020). The same model is used for both inference and action.

486
Preferences need inferences

Framed like this, uncertainty minimization is “merely” the underlying mechanism by which we accomplish
our needs and goals.

Mechanisms of valuation

Beyond mere exposure


With this brief overview of the predictive processing account, we can begin to consider why it is particularly
well equipped to address our questions about appreciation and curiosity. In the past 10 years, several authors
made important advances in articulating the beholder’s share using the framework, as well as in exploring
the mechanisms of affect generation in aesthetic experience (in music and visual art) (Forster, 2019; Kesner,
2014; Koelsch et al., 2019; Van de Cruys & Wagemans, 2011). In the remainder of this chapter, we sketch
the main lines of reasoning and zoom in on the mechanisms of valuation as they follow from predictive
processing.
As prefigured in the introduction, “mere exposure” theory is rendered meaningless in a predictive or
inferential view of perception. Instead, with more experience, we form better representations in the sense
that our generative models will explain (similar) inputs better. Capturing regularities in the inputs better
means that those inputs become subjectively simpler. Thanks to our adaptive, predictive models, commonly
occurring inputs also become least metabolically costly (Sengupta et al., 2013), simply because they gradu-
ally elicit fewer prediction errors. So increased predictive matching of a stimulus naturally goes along with
progressive simplification or increasingly efficient (compressed) coding of what appears often in the agent’s
environment. In this way, the account ties together predictability, (subjective) complexity and frugality.
So when we find that people like what is likely (“mere exposure”), typical or average, we should say
that they like the sensory inputs that their generative models can easily reconstruct (Brielmann & Dayan,
2022). This pleasantness of “generative success” can get quite literal: Chamberlain et al. (2021) showed that
drawings created by natural, human-like movement dynamics are preferred, at least by viewers with draw-
ing experience (and thus the necessary generative models). In a similar vein, we tend to prefer stimuli with
the same low-level statistical regularities (e.g., with respect to colour composition or 1/f spatial frequency
spectrum) as natural images (Graham & Redies, 2010; Nascimento et al., 2021), but this principle presum-
ably generalizes it to regularities on any level of abstraction (in the generative model). Additional support for
this comes from the finding that stimuli that are gazed upon more tend to be preferred, even though gaze is
manipulated by the experimenter, and there is no differential reward history whatsoever (Schonberg & Katz,
2020). Of course, all of this is consistent with processing fluency ideas, but we now have a mechanistic way
to conceptualize the beholder’s share (and how it evolves).
Since actions are naturally part of the predictive models, even regular action patterns (habits in layman’s
terms) can become intrinsically preferred patterns (as they indeed often do) following the same rationale.
Indeed, an unexpected inability to perform one’s habits (prediction error) can arouse intensely negative emo-
tions. But in isolation, habits do not normally come with intensely positive appreciations, similar to how the
most frequent or predictable stimuli are usually not liked much, except in special cases like faces (Ryali et al.,
2020) or wallpapers. The preference for regular wallpapers and average faces may have more to do with the
fact that we just want those to be the clean canvases for the actually interesting matter: social expressions or
interactions. This is similar to how habits just need to become the neutral background against which our
more burning deliberations can play out (e.g., wearing the same type of clothes every day to free up cognitive
space for more important decisions).
And yet we often take pleasure in returning to our familiar routines, for example, after a long, trying day,
even though there is nothing “objectively” rewarding about them. This “warm glow of familiarity” is not

487
Van de Cruys, Bervoets and Moors

necessarily just an expression of conservative taste. Remember that our whole generative model is built up
to fulfil our visceral predictions. It should not surprise us then that enacting the regimes inscribed with the
most precision in our generative model can regulate our homeostasis best. In some cases the force of habit
is so strong that even “objectively” painful stimuli can become attractive because they have become part of
one’s entrenched behavioural patterns.
The more general lesson here is that once we describe our preferences as prior expectations for particular
observations, they no longer need to be pre-specified genetically (indeed they cannot) but can be learned
as well: Our idiosyncratic expected patterns become attractors in themselves, irrespective of reward gains.
Instead of just learning what is in the world, we also learn what to want (Bem, 1972; Srivastava & Schrater,
2015). As social creatures, our parents, caregivers and peers are of course the main sources of familiar patterns
we are exposed to early in life. This type of learning usually happens in a behaviour-first fashion: Children
first merely ape the sensorimotor “rituals” and only later infer the goals and preferences (as hidden causes)
behind them. At that point, the sensory “evidence” of oneself behaving in a certain fashion is already so
irrefutably strong, that the best explanation to adopt will be “I have a preference to do X.” We see this prin-
ciple of putting the practice (ritual) before the ideology (“sacred values”) in acculturation of religions as well
(Heylighen et al., 2018).
Another telling example of how people often infer their preferences based on (past) behaviour is the
finding that our preferences will change when we have approached or chosen a particular stimulus before
(Kawakami et al., 2007; Schonberg & Katz, 2020; Van Dessel et al., 2019). The mere act of choosing some-
thing (even if the choice is in fact cunningly manipulated by an experimenter), similar to the mere act of per-
ceiving something, is shown to increase subsequent preference for it. Again, people seem to gravitate towards
perceptual or behavioural patterns that are already part of their generative model. Here, a preference appears
as just another hidden (high-level) cause explaining your sensorimotor interactions (minimizing prediction
errors). In our predictive nets, a preference is shorthand for the complex composite of multimodal sensory
cues and its associated reactive dispositions (Clark, 2019). All the things that make me feel X (and do Y)
have a hidden cause we call “value.” Similarly, we can learn from experience that when we have “desires,” we
tend to feel and do particular things, and that when we have similar desires/values, we end up doing similar
things. Again, preferences, desires and values are not eliminated in the predictive processing view; they are
just constructed. Preferences as high-level “empirical priors”—learned predictions of regular “packages” of
sensorimotor flow—can be readily recognized in new instances and, through verbal labelling, allow for effi-
cient communication and coordinated action (Clark, 2019).
This certainly does not imply that the usual, intuitive causal relation—going from preference to (choice)
behaviour—has no part to play. It merely says that the inferences go both ways. Often when we need to
determine value or preference, we rely on experience sampling; that is, we infer value based on past interac-
tions and behavioural dispositions (Gershman & Daw, 2012). We continually infer our own appreciation, just
like we infer other people’s preferences from their choice behaviour (except that we have additional intero-
ceptive data to infer our own preferences). This reminds us not to take preferences and values as “essences”
but rather as fluctuating and context-dependent constructs. That will come as no surprise to researchers in
aesthetic appreciation.
Once people can estimate (expect) the “value” (including the interoceptive observations it entails) that
they are going to experience, they can also form models on what factors generate or modulate that experi-
ence. The abstract construct of value is then just another tool to help us in our self-evidencing (so continued
existence), although animals without this level of abstracted causes can perfectly survive. Regardless, the
underlying mechanics remains that of uncertainty minimization. If I have a set of high precision expected
(interoceptive) states (a.k.a. the ones that “feel good”), and I have a model of how to generate those (e.g.,
through getting things that I value), active inference will get me to them. In the same way, I can have a
model of how to reach pleasure from a film or a book, namely by actually watching or reading it, instead of

488
Preferences need inferences

by just hearing the denouement. So we dislike spoilers even though that would make the whole experience
more predictable (in our colloquial sense of the word). Of course, our internal model of pleasure generation
might be wrong here. Research has shown that spoiling does not, in fact, substantially diminish the pleasure
of watching (Leavitt & Christenfeld, 2011). Indeed, people commonly reread the same book or rewatch the
same film.
To sum up, what we claim is not that people necessarily prefer the most predictable stimuli, person or
situation, by our colloquial sense of “predictable,” but rather that the way we become aware of our prefer-
ences (as well as our distastes) is as products of a thoroughly constructive, predictive process. Our preferences
are phenomenally transparent in Metzinger’s (2007) sense that we experience them as direct or given, with
no access to preference forming (valuation) processes. “The phenomenologically expressed preferences of a
conscious agent do not bear any evidence on the nature of the underlying brain processes that produce such
conscious preferences” (Weaver, personal communication).

The value of obstacles


Perhaps the reader still harbours some doubts about the picture of value so far. Perhaps an agent driven by
the proposed principles still too often seems to find value in echo chambers instead of museum halls. And
maybe the view of preference as just another inference falls short in capturing the specific phenomenology
of affective experience: the thing that determines the valence of value. What can our approach offer to ease
those worries?
For this we have to return to the way predictive processing puts value on epistemic concerns. Under
the account, an agent’s search to understand its surroundings is a crucial part of biology, not just limited to
human beings. Traditional scholars of motivation and psychological aesthetics have found themselves being
forced to postulate a separate “need for cognition” (e.g., Kagan, 1972) (known under many different guises
such as the need for “knowledge,” for “structure” or for “closure”) to adequately explain human behaviour.
In predictive processing, on the other hand, the tendency to explore and learn the structure of one’s envi-
ronment directly follows the principle of (expected) prediction error minimization. Bear in mind that this is
not a quest for some absolute truth (though it may have enabled one in humans): It operates on a pragmatic
principle of increasing coherence between expected and input patterns (and in the service of the body) rather
than towards increasing correspondence to some external ground truth.
The way that curious and explorative actions are driven by expected predictive progress (i.e., uncertainty
that is expected to be reducible) also aligns with recent theoretical, computational and empirical work casting
curiosity as expected learning progress, sometimes denoted as expected information or compression gains
(Gottlieb et al., 2013; Holm, 2017; Kidd & Hayden, 2015; Schmidhuber, 2009; Félix Schoeller et al., 2018;
Van de Cruys et al., 2021). What this means is that we are tuned to find regions in the input space that, given
our current mental models, hold the highest potential for—the best slopes of—prediction error (uncertainty)
reduction. As we saw in our overview of curiosity theories, we can say that we navigate the world guided by
learned (meta-)expectations on “cognitive structuring” rates, or expected rates of prediction error minimiza-
tion. The hypothesis is that people have a (imperfect) sense of where predictive progress can be made, which
we experience as curiosity. Actually making this predictive progress is experienced as pleasure or apprecia-
tion. Consequently, positive affect is determined by the rate of prediction error minimization, a change in
uncertainty over time, and not by the momentary absence of prediction errors (cf. familiarity previously).
This readily applies to our experience with art. Several seminal works in psychological aesthetics
(e.g., Huron, 2006; Meyer, 1961) have stressed the importance of expectation in aesthetic experience (espe-
cially music). Artists exploit learned (culturally dominant) predictions or establish new predictions in their
work (e.g., a rhythm or motif in music). These predictions can be based on lower-level regularities (e.g.,
perceptual symmetry) or on higher-level, semantic associations (e.g., story archetypes). Crucially, however,

489
Van de Cruys, Bervoets and Moors

they intuitively add prediction errors in their works to allow predictive progress in their audience: an active
experience of recovering structure or meaning, sometimes across levels of abstraction (Van de Cruys  &
Wagemans, 2011). For example, Picasso’s perceptually fragmented face in his famous work Weeping Woman
may be broken because it expresses sadness. Or Munch’s human figure in his painting Separation loses its
perceptual boundary with the background because it may express loss. But these simplistic examples betray
the complexity and uncertainties of the process. In practice, artworks will invite many cycles of curious
anticipation of prediction error resolution alternated by (local) instances of actual resolution, often without
reaching a final resolution, whatever that may mean (Muth & Carbon, 2016). A common way to put this is
that artists use errors to delay understanding, to allow their public to experience things anew, as if it is their
first encounter with the perceptual world (Shklovsky, 1917), when the biggest learning progress could still be
made, but always with the risk of leaving their audience in the dark. Indeed, artists mostly do not deliberately
search for the sweet spot of their audience (the result would be a gimmick instead of an artwork). An artist
may not be particularly popular in their own time. They often push the limits in terms of unexpectedness and
then have to rely on a cultural learning process to bring unpredictability to a level that is appreciated in terms
of perceived structure and reducibility of prediction errors. In fact, one could say that our account is infused
with the realization that, once one would elevate something (some predictable “rule”) to “art,” artists would
break this rule (introduce uncertainty). Indeed, they must do so, to be able to touch their audiences again.
Precise prediction errors can create curiosity in the sense that they allow expectations on error reduction
to grow, which spurs engagement through epistemic actions (e.g., eye movements on a painting). However,
as those prediction errors and the mental effort spent on matching them increase, the expectation of error
reduction decreases. In the limit, this may lead to anxiety, irritation or just looking away, which, in keep-
ing with our approach, should be understood as using overt action to reduce prediction errors, namely by
avoidance behaviour (instead of prediction updating). We shift attention to a more predictable setting, in
which expectations on rates of prediction error minimization can again be fulfilled. Just before this point of
abandonment, however, there is an opportunity for a higher than expected rate of error minimization. This
characterizes the intensely positive experience of sudden insight (“eureka”) or “aha.” Indeed, empirical
work has recently shown that aha moments arise when we are able to resolve a problem faster than expected
(Dubey et al., 2021).
The role of obstacles is also clear in the so-called generation effect in memory (Slamecka & Graf, 1978):
the finding that we remember things better if we can generate or infer them ourselves. Marketers exploit
this phenomenon to make their brands stick. They would give their potential customers only just enough
(or even slightly distorted) information to make the inference or generate the meaning themselves rather
than providing complete information. Again, this is a kind of anti-simplicity or disfluency principle that
causes people to remember and, we would say, appreciate a stimulus more (whether a painting or a brand;
see Figure 26.4).
Notice that predictive processing provides an account of the process here, as well as of the directionality:
One needs the overarching systemic principle of minimization of uncertainty (self-evidencing) to be able to
attach valence (so emotion) to it. More concretely, changes towards more uncertainty (e.g., dissolving fig-
ures) will be less pleasant than the reverse operation, unless the figures (unexpectedly) morph into some new
structure instead of mere disorder. The directionality is important but change is conditio sine qua non. Hence
the captivating power of ambiguous images that invite change (Figure 26.4).
As before, this analysis naturally extends to action. The action analogue of art is games. In playing games
we adopt rules, consisting of goals (expected outcomes) as well as “unnecessary obstacles [to our goals] to
make possible the activity of overcoming them” (Nguyen, 2017). The adopted ends are often arbitrary and
only adhered to for the duration of the game. But the “game” of increasing and decreasing uncertainties,
in this case about whether your actions can accomplish goals, is the same. Nguyen therefore calls games a
veritable “art of agency.” The aesthetic experience is associated with the increasing fit or harmony between

490
Preferences need inferences

Figure 26.4 Ambiguous figure adapted from Hebb (1949). Why do figures like this appeal? It is arguably not because
of the simplicity of the input per se (contra Gestaltists). Absences can become meaningful (as obstacles or
prediction errors) in a predictive processing account. The line is of course simple as such, which might
make you expect little meaning in it. But that changes once you look a fraction longer, probably because
your visual system registers that the line deviates (prediction error) from what you expect an average/
randomly drawn line looks like. This prediction error suggests an intention of the drawer, which in turn
creates an expected reducibility of the prediction errors, which is fulfilled (after a brief search) with the
discovery of a face. But this leaves some remaining errors: The one face has an odd contour, which leads to
the discovery of a second face. In what seem to be unassuming beginnings, there are already micro-cycles
of curiosity, discovery and appreciation. These would make up the “goodness” of a figure according to the
current account.

one’s abilities and the challenges of the situation (Nguyen, 2017). In art and games (as in life), the struggle
is inseparable from the joy. Of course, the dynamics of overcoming obstacles also provides the plot for much
of the (cultural) narratives that we enjoy so much, as many scholars and artists such as Kurt Vonnegut have
noticed (Felix Schoeller & Perlovsky, 2016; Singh, 2021). It is by incurring informational (and metabolic)
costs in the form of divergence between one’s models and the sensed world (Zénon et al., 2019) that a fallible
but pleasurable process of (re)harmonization is enabled.
Note that we carefully steer away from identifying aesthetic appreciation with just a resolved end product
(regularity) or mere effort/energy spent (obstacles or prediction error). Artworks (even static ones) that are
ambiguous or indeterminate are never free from regularities, however unstable they may be (and so regu-
larities are usually only temporarily, mentally reachable). Our mere failures in the encounter with art are
not the art, but artists are very creative in finding patterns in your failures, turning failure into art. Absolute
noise does not capture attention or please, except if there’s some story or rule (read: predictability) to it or
if it merges dynamically into something (identifiable). There is something pretty trivial to this idea, in the
sense that to be “identifiable” just means having a regularity. An aesthetic experience can definitely consist of
just lingering on one’s failing (prediction errors), but there has to be some curiosity still and so an expected

491
Van de Cruys, Bervoets and Moors

reducibility of (some of ) the “errors.” This in turn means that you have indeed been able to reduce them
in your past experience (in similar situations, in other art contexts, in other artworks from a particular artist
or the same artwork just moments ago). The artist cannot keep you hanging indefinitely or she loses your
engagement, precluding any further experience, aesthetic or not (of course her mileage may vary in the next
observer). In that sense, any experience is indeed aesthetic (else we would not have it), as Dewey (2005)
realized.

Learning dynamics and valuation


The general idea that takes shape here is that affective experience reflects the dynamics of inference and
learning (i.e., dynamics in prediction errors or uncertainty). This is a comprehensive hypothesis in the sense
that it subsumes curiosity and (aesthetic) appreciation, as well as our more conventional strivings: the pro-
gress (or regress) in reducing discrepancies relative to our goals as expected outcomes for the self. Positive
(vs. negative) rates of discrepancy reduction relative to one’s goals have long been proposed to be “what it’s
like” to experience positive (vs. negative) affect (Carver & Scheier, 1990; Kivetz et al., 2006). Since goals
are just one important type of (higher-level) expectations, we can readily generalize this view and situate
valence in the increasing versus decreasing rates of prediction error minimization. Perception and learning,
as well as discrepancy reduction by action, always take place against the backdrop of the model that includes
our goals and homeostatic expected states, so it is never divorced from valuation: Indeed, it is the very core of
valuation. With Skov and Nadal (2020), we emphasize that the basic mechanics of valuation are not specific
to our engagement with art. However, our particular hypothesis on the nature of valuation and affective
experience is relatively new, so its computational and psychological implications have only just begun to be
explored (Hesp et al., 2021; Joffily & Coricelli, 2013; Kiverstein et al., 2019; Solms, 2018; Van de Cruys,
2017; S. Wilkinson et al., 2019).
To see why the hypothesis is plausible, we first need to acknowledge that the notion of reward is not self-
evident or objectively determinable. So it cannot serve as the root cause of valence it is often presumed to
be. Reward (value) is not objectively given by the environment but instead always intrinsically determined
or inferred by the organism. Contrary to most reinforcement learning settings, there is no external “teacher”
that can configure the organism’s reward function ( Juechems  & Summerfield, 2019). Evolution is often
thought to take on this role by, as we saw, equipping an organism with desired or “expected” observations.
However, phenomena like satiation and the hedonic treadmill (the tendency to quickly return to a stable
level of pleasure; Brickman & Campbell, 1971) indicate that what is valuable is always crucially and very con-
textually dependent on internal states and learning. It is impossible for evolution to pre-specify a fixed reward
function for organisms that, like ourselves, are so dependent on learning for survival (we can become hunter-
gatherers or wage-earners-supermarket-goers). As we discussed, predictive processing naturally accommo-
dates learning of what is valuable (e.g., new goals) during our lifetime. What this theory suggests is that the
progress on goal attainment (including homeostatic set-points), can be evaluated by the change over time in
the distance between expected state and current state (change in prediction error). In other words, no (lack
of ) momentary discrepancy or prediction error should amount to affective value but rather the dynamics in
prediction errors: Are we in the process of decreasing or increasing them? This formulation reveals that the
epistemic or aesthetic emotions are not separate from the conventional “existential” or homeostatic ones but
rather rely on similar computational mechanisms.
Perhaps counterintuitively, the system is optimizing for learning instead of for rewards: It searches for
regions in the input space (the environment) of large gradients of prediction errors. Finding the pattern
is the actual reward, not getting a particular stimulus or substance. This isn’t to say that we engage with
art because we want (have a conscious goal) to experience (the positive feeling of ) uncertainty resolution.
The system just uses inference as uncertainty reduction mechanisms, so the appreciation could as well

492
Preferences need inferences

be epiphenomenal. However, the fact that we consciously experience these uncertainty dynamics not as
“content” but as affect means that the system can represent (predict) the changes in uncertainty per se, as
well as inferring the hidden causes not of inputs but of its own processing dynamics. It can build a generative
model of those changes as expected states. In other words, it represents “value” (and a generative model of
value) as a more abstract, universal currency, enabling comparisons among diverse situations or modalities.
In the previous sections, we described value as a hidden cause summarizing a regular pattern of sensory
consequences and behavioural reactions that is shared among instances of value. Here, we can add that
a core part of the “value complex” will be an expected dynamic in internal processes: the uncertainty
gradients.
In this view, people will seek out art because they have a model of how this type of environment pro-
vides them with good prediction error slopes. This is just a complicated way of saying they engage with
art because of the sense-making (regularity-revealing) value it generates (which as we saw crucially involves
sense-breaking). But for any specific artwork, additional experiences will reduce future uncertainty resolving
potential. This does not mean an artwork or piece of music will necessarily lose its emotional potential once
we are acquainted with it. For example, repeated experience with Picasso’s cubist faces will not overwrite
our overlearned beliefs on the proper structure of faces (lest it would break our everyday perception), so
prediction errors (and hence the potential for resolution) will continue to be present despite repeated expo-
sure. However, frequent engagement with a piece of art or music can detract from its emotionality under
this learning-based view, suggesting that we are often chasing an experience we cannot fully relive when
revisiting the same artwork.
Affective experience arises when we are able to build meta-models that monitor (model) how our mod-
elling efforts on the lower levels are doing. Hence, (first-order) learning is phylogenetically prior to affect
(Ginsberg & Jablonka, 2019). The need for meta-models may only have emerged from a powerful capac-
ity to learn and model our world in unprecedented temporal (hierarchical) depth. This capacity required
equally powerful feedback mechanisms on how well the system is doing in modelling, that is, coping with,
its environment. Indeed, in this conception, emotions track the fluctuations in uncertainty or confidence in
our models on longer timescales (spanning multiple glances or actions; Hesp et al., 2021). As an explanation
of emotions, this may have an overly intellectualistic or cognitivist ring to it, but in fact it merely reinforces
ideas that were always central in theories of emotion. Different prominent scholars describe emotions as a
form of nonconceptual monitoring of coping performance in our interaction with the world (Frijda, 2006;
Reisenzein, 2009). We just propose that prediction error dynamics are the form of feedback on the system’s
own functioning required to spell out those accounts in computational specificity (Hesp et al., 2021; Joffily &
Coricelli, 2013; Schillaci et al., 2020).
If we turn to the direct evidence for this view, we must conclude it is scarce so far. Of course, the view
is still very young, and the challenges to operationalize it are considerable. At its centre are unobservable,
hidden states and their dynamics which intrinsically depend on individual learning history. Advances in
disentangling and tracking neural activity associated with predictions versus prediction errors across time
and along the cortical hierarchy are modest so far (e.g., Corlett et al., 2021; Issa et al., 2018). Meanwhile,
in silico computational modelling implementations of these ideas provide an interesting proof of concept
(Hesp et al., 2021) but cannot make the hypothesized connections to the experiential dimensions of affec-
tive value. Still, the idea that positive affect is linked not to conventional reward (magnitudes) per se but to
learning dynamics—specifically one’s success in learning the structure of the environment—received sup-
port in a recent study that managed to cleanly disentangle those factors (Blain & Rutledge, 2020). More
fundamentally, the reward value of “mere” information gain (uncertainty resolution) is also well supported
(Bromberg-Martin & Monosov, 2020; Fiorillo et al., 2003; Fortes et al., 2016) and shown to be associated
with activity in the same dopaminergic brain regions as conventional rewards (Bromberg-Martin & Hiko-
saka, 2011). Indirect evidence for the link between positive affective experiences and learning comes from

493
Van de Cruys, Bervoets and Moors

studies reporting that a pleasurable aha experience is associated with better memory (Danek & Wiley, 2020;
Kizilirmak et al., 2016; see also Sarasso et al., 2021; Van de Cruys et al., 2021).
In the domain of psychological aesthetics, several recent findings nicely fit with a view that attaches pleas-
ure to learning (uncertainty) dynamics. For example, Cheung et al. (2019) showed that pleasure in music
is associated with the surprise (prediction error) of hearing a chord in a chord progression, when one actu-
ally had a strong (low uncertainty) prediction about the chord that would follow (but didn’t). The authors
managed to quantify uncertainty and surprise using a large corpus of chord progressions in popular music
(see also Gold et al., 2019; Pearce, 2018). We assume that the unexpected increase in prediction error (of
surprising tones, beats or chords in music) is what allows the pleasurable resolution of uncertainty when the
pattern is reinstated (in music). Similarly, the studies reporting a preference for intermediate predictability in
auditory or visual sequences (e.g., Delplanque et al., 2019; Witek et al., 2015) can be marshalled in support
of the proposed view, although the precise mechanisms usually remain beyond reach in those studies. In a
study using video art, Muth et al. (2015) found that dynamics in (self-reported) semantic uncertainty while
viewing animated sketches influence appreciation. Finally, the set effects discussed earlier (relative fluency
effect; Wänke & Hansen, 2015) also sit comfortably with our reasoning, because they suggest that affective
experience (appreciation) reflects a type of feedback or monitoring of one’s inferential processes across mul-
tiple stimuli or trials. Such studies underscore that even though we are asked to appreciate a single stimulus
and we ourselves also intuitively (mis)attribute our appreciation to a singular work or stimulus, in reality
our affective experience is a function of our own way of processing, in light of the more extensive context.
Similar to how it is hard to shake your confidence that the content of a perceptual experience is not just out
there in the world but the result of perceptual inference, the immediateness of the affective experience hides
the underlying (meta)inferences.

The touch of art

Attuning generative models


What is true for all perception holds for perceiving art: We build up a generative model of the artwork.
Because we know the artwork is human-made, the hidden causes that we infer will not only include the
objects portrayed or the medium used (e.g., brush strokes as latent causes of particular patterns of paint)
but also the emotions and intentions that the painter might have had (Freedberg & Gallese, 2007; Seth,
2019). For example, when we infer aggression as the hidden cause for an odd (unexpected) trace of a
brushstroke, we turn that sensory effect (brushstroke) into an expression. Of course, our inferences do
not necessarily reflect the meaning or emotion of the artist, but we can’t help but approach the work
in this (re)constructive sense (as we do for all sensory inputs), trying to see potential sense in the artist’s
apparent sense-breaking. Articulating these processes makes it seem a deliberate matter. In truth, we
automatically infer and readily perceive the causal history of visual stimuli, making irregularities regular,
as experiments have shown (Chen & Scholl, 2016; H. Leder et al., 2012; Pinna, 2010). Sometimes our
inferences rely on intuitive (possibly innate) mappings from “generative” emotions to auditory or visual
features, like when slow rhythmic and melodic contours express tranquillity in lullabies. Other times
a learned “language” of expressions is needed to “get” the artwork ( Jackendoff & Lerdahl, 2006). The
basic process is not unlike empathy (Wilkinson et al., 2021), but with a work of (personified) art instead
of with an actual person. We enter into a sort of dialogue in which we probe the work with predictions
that we “test” using targeted sampling—for example, with eye movements to regions with high expected
informativeness (precision). And the work will “respond” with inputs that become clues under our prob-
ing (Koenderink, 2010).

494
Preferences need inferences

When reading Aristotle on the origins of aesthetic experience (in his Poetics), one is struck by how much
of the current theorizing on inference and learning he already foreshadowed. Tracy (1946) summarizes Aris-
totle’s line of argument as follows:

in the case of a work of art, the observer is establishing a significant connection between the pres-
entation he sees (picture, play, etc.) and some original of which he has knowledge from his own
experience; the inferences drawn by an observer of a work of art have to do with, and are condi-
tioned by, the necessarily imperfect degree of adequacy it achieves; i.e., some effort on the part of
the observer is required to get him en rapport with the artist; satisfaction comes from the success-
ful integration between the artist’s way of presenting a given situation or object and the observer’s
power to interpret the artist’s procedures.

The key point here is the pleasure created by the increasing alignment (by surpassing obstacles; cf. effort)
between generative models of the work (or implicit artist) and the observer. Aristotle links this to the emo-
tional impact of tragedies or more generally the pleasure we can derive from negative emotions in art (see also
Menninghaus et al., 2017). For example, reading a good novel can provide an unexpected attunement with
core but largely unarticulated dynamics of the self (models), even (or especially) if those dynamics concern
negative emotions. A piece of art can provide validation or external evidence for the type of regularities in
our own mental life which we consider very personal and idiosyncratic (Van de Cruys et al., 2017) and in
this way supports self-evidencing at a very fundamental (and uniquely human) level. It is as if we can briefly
breach the epistemic and conative boundary between ourselves and the outside world (other agents). This
brings home the point that a predictive processing view of the aesthetic experience is about more than mere
“epistemic” problem-solving or striving for predictability. It captures the existential consolation of art as
well. Moreover, art puts in motion a self-reinforcing cycle, in the sense that validation of your models will
provide you with the confidence and safety—read: give you an expectation of a good rate of prediction
error minimization in a given context—to be curious again and to explore prediction errors that afford new
meaning. Those cycles may demarcate aesthetic experience from other experiences, more so than any of its
component processes separately.

Dislodge and roam free


With the great power of our modelling capacities—particularly the increased hierarchical depth of networks
of inferred causes—comes great vulnerability as well. Specifically, the (Bayesian) beliefs that make up our
models are at risk of becoming overly convoluted or even free-floating, impervious to evidence. Indeed,
in daily life, we often have only very sparse, biased or indirect data to speak to and constrain our predictive
models (Hoel, 2020). On top of that, because of the hierarchical model structure, rich auxiliary hypotheses
can be called upon to take the blame and explain away current prediction errors, thereby saving our “pre-
ferred” (high confidence) models from revision. In this way, we end up with overfitted models: models that
try to capture too much of a given set of sense data and as a result will not generalize to new situations (hence
creating prediction errors in the longer term). These models are too specific and inefficient in that they have
to use too many parameters (high complexity) in order to match every new situation encountered instead
of latching on to stable regularities across samples. Overfitted models capture incidental variability (noise)
in the “training” samples as putative regularities. Interestingly, a deliberate injection with noise inputs is a
proven technique in artificial intelligence to improve the generalizability of learned predictive models (see also
cybernetic phenomenon of “order from noise”; Von Foerster, 1960). Intuitively, it shakes the system out of
its habits that were too tuned to the particular features of its limited, biased sample of experiences.

495
Van de Cruys, Bervoets and Moors

Here, art may provide a safe, “offline” testing ground for our models: It can be a way of safely introducing
“chaos” to make our models more robust in the long term (Hoel, 2020). It may seem disrespectful to describe
art as mere noise, but it aligns with the freedom with which artists can upset our predictions, intentionally
creating ambiguities and fictions to shake up our categories (Hoel, 2019), often allowing our models to settle
down again, finding a more global and robust minimum in the prediction error landscape. Artists intuitively
direct their “noise” not only to hone existing models—reducing their complexity while retaining their pre-
dictive power—but also to open new pathways of future prediction error minimization.
At this point, we encounter the artist’s capability that seems diametrically opposed to the predictive pro-
cessing account with its sole focus on uncertainty minimization. Keats (Poetry Foundation, 2021b) called it
negative capability, that is, “when a man is capable of being in uncertainties, mysteries, doubts, without any
irritable reaching after fact and reason.” According to Keats and many after him, great artists have this apti-
tude that allows them to “bury self-consciousness, dwell in a state of openness to all experience, and identify
with the object contemplated” (Poetry Foundation, 2021a). It is the capacity to dwell in uncertainty—to
“stay with the trouble” (Haraway, 2016)—despite the existential threat it holds and despite the fact that our
whole being, per predictive processing, is fundamentally oriented towards uncertainty minimization. It is a
true capability because it is the openness for uncertainty or unpredictability that gives the opportunity for
radically new and different models to take hold. Instead of merely triggering a shift in probabilities of the
currently available hypotheses (prediction updates), it opens new hypothesis spaces. It is a capability that can-
not be approached as a capability, hence the qualifier “negative.”
The paradox is that expecting uncertainty and inviting chaos (what we could call radical curiosity) can
lead one to perceive new layers of regularities in reality. Such discoveries may rely on a recycling of existing
regularities (models) for new domains (cf. metaphors), but it crucially requires relinquishing our default,
prepotent models for that other domain, in other words, allowing uncertainty to rise first. Sometimes this is
just a matter of enlarging (the salience) the unexplained residue of our default models, as artists sometimes
do. However this “breaking of new ground” is done—what type of generative models allow for it (Williams,
2020)—it seems it is only possible in a very roundabout way. Like greedy reward optimization (exploitation
instead of exploration) makes you miss out on opportunities for better rewards, a focus on greedy informa-
tion gain or predictive progress would foreclose the discovery of new, more efficient regularities and so better
means for uncertainty minimization. We need to relinquish our focus on information gain relative to current
models to radically reconfigure the hypothesis space of our models and try on different (self-)models. No
artist (improviser) has to be convinced of this kind of generative power of errors and uncertainty, but it is hard
to form a model and plan for this generativity. However, it is something one recognizes when it happens. It
is marked by intensely positive emotion: Unexpected gains, or faster than expected rates of prediction error
reduction, feel better than mere gains. It is impossible to use information gain as a guide towards unexpected
gains, except by deliberately staying with or increasing one’s uncertainty. There is interesting work ahead in
exploring how this translates into the artist’s concrete (epistemic) practices.
A parallel can be drawn with evolution in single-celled organisms (Freddolino & Tavazoie, 2012). In harsh
conditions, these organisms upregulate mutation rate (randomness or uncertainty) to increase the chances for
the population to discover new useful regularities (hypotheses) and hence survive. In more complex organ-
isms, individuals are less expendable, so they have developed ways to ensure the organism does not need to
die with its hypotheses. However, as in single-celled organisms, the risk of introducing uncertainty seems to
be offset by the social group, albeit in a different way. A child takes more risks in the presence of the parent
that it trusts to resolve its uncertainty. Its freedom in making errors begets novel structure. Similarly, a con-
text of art, because it allows atypical amounts of certainty or randomness to be safely introduced, may be an
accelerator of novel predictive models. However, trust seems to be a precondition for this. Trust in oneself as
the metacognitive belief that one’s generative model is up to dealing with whatever is thrown at it (e.g., a jazz
musician who knows that, through her skills, she can deal with unexpected events or “errors”). And trust in

496
Preferences need inferences

the social context of art, understood here as our basic expectation that prediction errors are reducible when
one takes the effort to engage with it. I know there is a model; I just don’t know the model (yet). This trust
is rooted in the responsiveness of our sociocultural environment: The way it, from early on in development,
folds to our anticipatory activities (Hunt, 1981), thereby kickstarting our predictive senses and providing a
unique type of reliable validation of one’s models. If that trust or confidence has percolated into hierarchi-
cally deeper layers of our models, it seems to create room for the “letting be” of uncertainties on lower-level
branches of models, instigating novel, flexible model reconfigurations at those levels.
Allowing uncertainty to surge in an organism that owes its very existence to an uncertainty minimizing
process seems a perilous and even self-contradictory (literally and figuratively) undertaking. One way to
respond to this would be to say that uncertainty minimization is not a sufficient condition for existence; it can
only be used in retrospect to say that, if an organism is alive now, it must have been minimizing uncertainty
(Constant, 2021). This reasoning is reminiscent of evolutionary theory, where the emphasis on adaptation
as necessary for survival, risks obscuring the equally crucial role of non-adaptive variability (random muta-
tions). Weick (1979) remarks that “adaption precludes adaptability”: A built-in specialization (adaptation)
for a particular “goal,” while possibly more efficient in the short term, will get you trapped in a dead end
when the environment changes. Variability or uncertainty is the raw material for adaptive change and novel
regularities to be picked up. The example of single-celled organisms changing their mutation rates illustrates
the power of channelling uncertainty in volatile or harsh conditions. It is the active planning for uncertainty
in order to better minimize uncertainty (as the population), similar to what we saw with precision expecta-
tions in predictive processing (at the individual level). We can talk about “planning for uncertainty,” however
contradictory that sounds, because expected uncertainties become part of the model and the system benefits
from making uncertainty selective (e.g., limiting the increase of mutations to those parts of genetic material
most likely to solve unexpected environmental challenges but not to others). One could compare this to an
immune system that only works well when it has been challenged properly during its development (see also
the idea of hormesis and antifragility; Taleb, 2013). An even more plastic system like the brain similarly needs
those challenges, and so good uncertainty minimization slopes are what it should expect.
Humans, and artists all the more, seem specialized in selectively channelling and compartmentalizing
uncertainty in order to improve and better support their models. What enables this and allows us to find
novelty within an individual instead of a population is our “dissociative” capacity to accumulate evidence for
a new model without immediately having to relinquish existing models (Kelly, 1964). Art is a product of our
capacity to turn our models (including self-models) into objects themselves that can be taken “offline” and
explicitly questioned with fictive “data” (Miller et al., 2020) instead of just applied and updated in “online”
behaviour. This is what art does: asking and following through “what if ” stories. Such counterfactual think-
ing softens the threat of model invalidations. We can dry-run different models without (yet) committing. It
is the sand-boxing of a (self-)model to circumvent the logic of the law of the excluded middle, which says
that we can’t simultaneously affirm and deny a (hypo)thesis (Kelly, 1958). Indeed, given the constructivist and
pragmatist inspiration of our predictive modelling, nothing prevents us from violating that law of rationality
if it is to our benefit in coping with our world. The fascinating questions that have yet to be confronted by
predictive processing are precisely about these “dissociative” capacities and the competition between frac-
tionated (self )models that emerges. Some of these fractionated models are inferentially shielded so do not
have any way to cross-talk. The result is food for psychotherapists or for . . . artists.

Conclusions
Throughout this chapter, we developed the view that when we “artify”—to focus on the activity instead
of the products—we are, implicitly, optimizing for learning. We are curiously guided by expected pre-
dictive progress, we pleasurably make such progress, attune our models to the (social) world, and in the

497
Van de Cruys, Bervoets and Moors

process of introducing and sustaining (barely) controlled levels of uncertainty, we prevent overfitting
and facilitate the unmooring of old models and the discovery of new ones. Our view retells accounts of
aesthetic appreciation as old as Aristotle’s and extends recent predictive processing-based accounts of aes-
thetic experience (Pelowski et al., 2017; Sarasso et al., 2020; Felix Schoeller et al., 2015; Van de Cruys &
Wagemans, 2011).
Our artistic sense emerges as a neotenous trait with roots in playful behaviour in children and even non-
human animals (Bekoff, 2015). A relaxed, playful dog picks out the most unwieldy stick, the one that fights
back most. A child chooses the most unfamiliar, inappropriate object to play with (Andersen & Roepstorff,
2021; Bonawitz et al., 2012). At least in origin, this is not driven by self-conscious showmanship, let alone
a display of prowess or the “costly signaling” of one’s fitness, an evolutionary rationale that has been invoked
to explain art as well (Dissanayake, 2007; Helmut Leder & Nadal, 2014). Rather, the child is intrinsically
motivated to handle the more challenging objects, because the obstacles allow for predictive progress or
advances in the “grip” it has on its environment. So it goes in art as well, where artists throw up barriers and
constraints for their own performances to keep it genuinely interesting for themselves and their observers, in
other words: to make room for new learning.
The emphasis on learning and inference in this chapter may give the impression that art is a mere epis-
temic matter, serving the “world-disclosing” goal of uncertainty and complexity reduction (preventing over-
fitting, i.e., future prediction errors). This may align with the Kantian “disinterestedness” as precondition for
aesthetic judgment. However, as we saw, perceiving and learning is always dependent on “concepts” (predic-
tions) and is already valued in predictive processing; value is not something added to it. There is no “veridical
perception plus the values”; behavioural control and perception are one (against conventional modularity
ideas). Value functions are absorbed into our probabilistic models (priors), so utilities and probabilities are
not separately represented. This may explain why a positive aesthetic appreciation often goes together with
perceived truth (Reber et al., 2004) (incidentally, it may also explain the difficulty in separating fact from
value in everyday and ideological discussions). Of course, we humans can build explicit conceptual models
of value as something construed to be strictly divorced from probabilities or expectancies. But the underlying
processes do not honour that strict dissociation. On the one hand, the view reinforces the classical adagium
from Western philosophy that value (beauty) and truth coincide, if only in experience. On the other hand,
it overthrows the strict separation of cognition (probabilities, inferences) and emotion (values, preferences),
equally deeply rooted in Western thinking. Still, the approach does not explain away value but reveals the
mechanisms of valuation within an uncertainty minimizing agent.
It follows that we implicitly reconfigure value functions by exploring different models, taking on different
predictive “sets” or roles and moulding our models to better fit internal (self  ) and external sensorium. In art,
just as in play behaviour, our inferred models become targets in themselves. It is the context par excellence
for us to suspend the normal utility functions absorbed in our routine models and explore not just how to
attain our goals and values but rather what can be valuable in the first place. We do so by adopting arbitrary
goals (or expectations), “assigning arbitrary rewards and accepting unnecessary costs” (Chu & Schulz, 2020),
like we do in playing. We do so by relaxing the tight grip of prediction errors in the here and now, as
well as relaxing optimization of information gain relative to the current predictive set, to ultimately—and
fallibly—finding better long-term prediction error minimizing models. Far from being disinterested, it is
about exploring a surplus of interest.
Finally, the approach also allows us to find a middle ground between two classical, orthogonal views of
aesthetic appreciation. According to the “idiosyncratic” view, aesthetic appreciation is just a function of
personal taste and the vagaries of arbitrary individual emotions, while to the nomothetic view, our aesthetic
sense reflects some universal, objective (and human-independent) values. Philosophically, the latter view is of
course rooted in classical Platonic theories of aesthetic value (Hart, 1971; Zangwill, 2021). But the dispute
still plays out in current discussions in experimental aesthetics, with some studies reporting that aesthetic

498
Preferences need inferences

appreciation is largely dependent on individual rather than shared taste (A. A. Brielmann & Pelli, 2019),
while others try to establish objective stimulus features that determine appreciation (Graham  & Redies,
2010). The view espoused here follows the idiosyncratic view in that aesthetic value emerges in the relation
between a subject (with her particular models) and an object, much in line with the work of Dewey. Values
are dynamically inferred rather than objective and static, and dynamics in inferences bring about affective
values. However, as we have tried to show in this chapter, the same move gives us a window on the nomo-
thetic principles governing (aesthetic) valuation as well.

Acknowledgements
I (SVdC) thank my twin daughters Martha and Jacoba for reminding me of the value of radical curios-
ity. We also thank Eline van Geert, Pieter van Dessel, David Gijbels, Pietro Sarasso, Félix Schoeller, Pablo
Fernandez Velasco and Jacopo Frascaroli for helpful feedback on earlier versions of this chapter. This work
is partly supported by a Methusalem grant from the Flemish government (METH/14/02). JB is supported
by the NeuroEpigenEthics project that received funding from European Research Council (ERC) under
the European Union’s Horizon 2020 research and innovation program (grant agreement No 804881). AM
receives funding from the Research Foundation–Flanders (FWO, G073317N) and the Research Fund of KU
Leuven (C14/17/047).

References
Aitchison, L., & Lengyel, M. (2017). With or without you: Predictive coding and Bayesian inference in the brain. Current
Opinion in Neurobiology, 46, 219–227. https://doi.org/10.1016/j.conb.2017.08.010
Alter, A. L., & Oppenheimer, D. M. (2009). Uniting the tribes of fluency to form a metacognitive nation. Personality and
Social Psychology Review, 13(3), 219–235. https://doi.org/10.1177/1088868309341564
Andersen, M. M., & Roepstorff, A. (2021). Play in predictive minds: A cognitive theory of play. https://doi.org/10.31234/
osf.io/u86qy.
Arnheim, R. (1974). Entropy and art: An essay on disorder and order. University of California Press. http://books.google.
com/books?id=kL75af7hbXsC
Bekoff, M. (2015). Playful fun in dogs. Current Biology: CB, 25(1), R4–R7. https://doi.org/10.1016/j.cub.2014.09.007
Bem, D. J. (1972). Self-perception theory. Advances in Experimental Social Psychology, 6(1), 1–62. http://healthyin
fluence.com/wordpress/wp-content/uploads/2011/05/SP-Theory-Bem-Advances.pdf. https://doi.org/10.1016/
S0065-2601(08)60024-6
Berlyne, D. E. (1960). Conflict, arousal, and curiosity. http://psycnet.apa.org/psycinfo/2006-09643-000/
Berlyne, D. E. (1978). Curiosity and learning. Motivation and Emotion, 2(2), 97–175. https://doi.org/10.1007/BF00993037
Blain, B., & Rutledge, R. B. (2020). Momentary subjective well-being depends on learning and not reward. eLife, 9.
https://doi.org/10.7554/eLife.57977
Boddez, Y., Descheemaeker, M., Mertens, G., Truyts, A.,  & Van de Cruys, S. (2019). Like what you see: Gener-
alization of social learning determines art appreciation. Acta Psychologica, 196, 18–25. https://doi.org/10.1016/j.
actpsy.2019.04.001
Bonawitz, E. B., van Schijndel, T. J. P., Friel, D., & Schulz, L. (2012). Children balance theories and evidence in explora-
tion, explanation, and learning. Cognitive Psychology, 64(4), 215–234. https://doi.org/10.1016/j.cogpsych.2011.12.002
Bornstein, R. F. (1989). Exposure and affect: Overview and meta-analysis of research, 1968–1987. Psychological Bulletin,
106(2), 265–289. https://doi.org/10.1037/0033-2909.106.2.265
Bornstein, R. F., Kale, A. R., & Cornell, K. R. (1990). Boredom as a limiting condition on the mere exposure effect.
Journal of Personality and Social Psychology, 58(5), 791–800. https://doi.org/10.1037/0022-3514.58.5.791
Brickman, P., & Campbell, D. T. (1971). Hedonic relativism and planning the good science. In M. H. Appley (Ed.),
Adaptation level theory: A symposium (pp. 287–302). Academic Press.
Brielmann, A. A., & Pelli, D. G. (2019). Intense beauty requires intense pleasure. Frontiers in Psychology, 10, 2420. https://
doi.org/10.3389/fpsyg.2019.02420
Brielmann, A. A., & Dayan, P. (2022). Introducing a computational model of aesthetic value. https://doi.org/10.31234/osf.
io/eaqkc.

499
Van de Cruys, Bervoets and Moors

Bromberg-Martin, E. S., & Hikosaka, O. (2011). Lateral habenula neurons signal errors in the prediction of reward infor-
mation. Nature Neuroscience, 14(9), 1209–1216. https://doi.org/10.1038/nn.2902
Bromberg-Martin, E. S., & Monosov, I. E. (2020). Neural circuitry of information seeking. Current Opinion in Behavioral
Sciences, 35, 62–70. https://doi.org/10.1016/j.cobeha.2020.07.006
Carver, C. S., & Scheier, M. F. (1990). Origins and functions of positive and negative affect: A control-process view.
Psychological Review, 97(1), 19–35. https://doi.org/10.1037/0033-295X.97.1.19
Chamberlain, R., Berio, D., Mayer, V., Chana, K., Leymarie, F. F.,  & Orgs, G. (2021). A  dot that went for a walk:
People prefer lines drawn with human-like kinematics. British Journal of Psychology, 113(1), 105–130. https://doi.
org/10.1111/bjop.12527
Chen, Y. C., & Scholl, B. J. (2016). The perception of history: Seeing causal history in static shapes induces illusory
motion perception. Psychological Science, 27(6), 923–930. https://doi.org/10.1177/0956797616628525
Cheung, V. K. M., Harrison, P. M. C., Meyer, L., Pearce, M. T., Haynes, J. D., & Koelsch, S. (2019). Uncertainty and
surprise jointly predict musical pleasure and amygdala, hippocampus, and auditory cortex activity. Current Biology:
CB, 29(23), 4084–4092.e4. https://doi.org/10.1016/j.cub.2019.09.067
Chmiel, A., & Schubert, E. (2017). Back to the inverted-U for music preference: A review of the literature. Psychology of
Music, 45(6), 886–909. https://doi.org/10.1177/0305735617697507
Chu, J., & Schulz, L. E. (2020). Play, curiosity, and cognition. Annual Review of Developmental Psychology, 2(1), 317–343.
https://doi.org/10.1146/annurev-devpsych-070120-014806
Clark, A. (2013). Whatever next? Predictive brains, situated agents, and the future of cognitive science. Behavioral and
Brain Sciences, 36(3), 181–204. https://doi.org/10.1017/S0140525X12000477
Clark, A. (2019). Consciousness as generative entanglement. Journal of Philosophy, 116(12), 645–662. https://doi.
org/10.5840/jphil20191161241
Clark, A. (2020). Beyond desire? Agency, choice, and the predictive mind. Australasian Journal of Philosophy, 98(1), 1–15.
https://doi.org/10.1080/00048402.2019.1602661
Constant, A. (2021). The free energy principle: It’s not about what it takes, it’s about what took you there. In Biology and
Philosophy, 36(2). https://doi.org/10.1007/s10539-021-09787-1
Corlett, P. R., Mollick, J. A., & Kober, H. (2021). Substrates of human prediction error for incentives, perception, cognition, and
action. https://doi.org/10.31234/osf.io/pf89k.
Cupchik, G. C.,  & Heinrichs, R. W. (1981). Toward an integrated theory of aesthetic perception in the vis-
ual arts. In H. I. Day (Ed.), Advances in intrinsic motivation and aesthetics (pp.  463–485). Springer. https://doi.org/
10.1007/978-1-4613-3195-7_19
Danek, A. H.,  & Wiley, J. (2020). What causes the insight memory advantage? Cognition, 205, 104411. https://doi.
org/10.1016/j.cognition.2020.104411
Delplanque, J., De Loof, E., Janssens, C., & Verguts, T. (2019). The sound of beauty: How complexity determines aes-
thetic preference. Acta Psychologica, 192, 146–152. https://doi.org/10.1016/j.actpsy.2018.11.011
Dennett, D. C. (1991). Real patterns. Journal of Philosophy, 88(1), 27–51. https://doi.org/10.2307/2027085
Dennett, D. C. (2017). From bacteria to Bach and back: The evolution of minds. W. W. Norton, and Company. https://play.
google.com/store/books/details?id=PEp8DAAAQBAJ
Dewey, J. (2005). Art as experience. Penguin. http://books.google.com/books?id=aAbqAGo5MwwC
Dissanayake, E. (2007). What art is and what art does: An overview of contemporary evolutionary hypotheses. In Evolu-
tionary and neurocognitive approaches to aesthetics, creativity, and the arts (pp. 1–14). https://books.google.be/books?hl=nl
&lr=&id=np6RDwAAQBAJ&oi=fnd&pg=PT8&dq=artistic+aesthetic+evolutionary+adaptation+&ots=9epYxgU
zVm&sig=k7gulj_FA7I96RDmg8oaH8CRafU
Dubey, R., Ho, M. K., Mehta, H., & Griffiths, T. (2021). Aha! moments correspond to meta-cognitive prediction errors.
PsyArXiv. Retrieved June 22, from https://psyarxiv.com/c5v42/download/?format=pdf
Eysenck, H. J. (1942). The experimental study of the “good Gestalt”—A new approach. Psychological Review, 49(4),
344–364. https://doi.org/10.1037/h0057013
Fiorillo, C. D., Tobler, P. N., & Schultz, W. (2003). Discrete coding of reward probability and uncertainty by dopamine
neurons. Science, 299(5614), 1898–1902. https://doi.org/10.1126/science.1077349
Forster, M. (2019). Processing fluency. In M. N. A. Vartanian (Ed.), The Oxford handbook of empirical aesthetics. Oxford
University Press. https://doi.org/10.1093/oxfordhb/9780198824350.013.21
Forster, M., Gerger, G., & Leder, H. (2015). Everything’s relative? Relative differences in processing fluency and the
effects on liking. Everything’s Relative. PLoS One, 10(8), e0135944. https://doi.org/10.1371/journal.pone.0135944
Fortes, I., Vasconcelos, M.,  & Machado, A. (2016). Testing the boundaries of “paradoxical” predictions: Pigeons do
disregard bad news. Journal of Experimental Psychology. Animal Learning and Cognition, 42(4), 336–346. https://doi.
org/10.1037/xan0000114

500
Preferences need inferences

Freddolino, P. L.,  & Tavazoie, S. (2012). Beyond homeostasis: A  predictive-dynamic framework for understand-
ing cellular behaviour. Annual Review of Cell and Developmental Biology, 28, 363–384. https://doi.org/10.1146/
annurev-cellbio-092910-154129
Freedberg, D., & Gallese, V. (2007). Motion, emotion and empathy in esthetic experience. Trends in Cognitive Sciences,
11(5), 197–203. https://doi.org/10.1016/j.tics.2007.02.003
Frijda, N. H. (2006). The laws of emotion. Erlbaum.
Friston, K. J. (2010). The free-energy principle: A unified brain theory? Nature Reviews. Neuroscience, 11(2), 127–138.
https://doi.org/10.1038/nrn2787
Friston, K. J., Da Costa, L., Hafner, D., Hesp, C., & Parr, T. (2021). Sophisticated inference. Neural Computation, 33(3),
713–763. https://doi.org/10.1162/neco_a_01351
Friston, K. J., Daunizeau, J., Kilner, J., & Kiebel, S. J. (2010). Action and behaviour: A free-energy formulation. Biological
Cybernetics, 102(3), 227–260. https://doi.org/10.1007/s00422-010-0364-z
Friston, K. J., Rigoli, F., Ognibene, D., Mathys, C., Fitzgerald, T., & Pezzulo, G. (2015). Active inference and epistemic
value. Cognitive Neuroscience, 6(4), 187–214. https://doi.org/10.1080/17588928.2015.1020053
Friston, K. J., Rosch, R., Parr, T., Price, C., & Bowman, H. (2018). Deep temporal models and active inference. Neuro-
science and Biobehavioral Reviews, 90, 486–501. https://doi.org/10.1016/j.neubiorev.2018.04.004
Friston, K. J., Thornton, C., & Clark, A. (2012). Free-energy minimization and the dark-room problem. Frontiers in
Psychology, 3, 130. https://doi.org/10.3389/fpsyg.2012.00130
Gerken, L., Balcomb, F. K.,  & Minton, J. L. (2011). Infants avoid “labouring in vain” by attending more
to learnable than unlearnable linguistic patterns. Developmental Science, 14(5), 972–979. https://doi.
org/10.1111/j.1467-7687.2011.01046.x
Gershman, S. J. (2019). What does the free energy principle tell us about the brain? In arXiv [q-bio.NC]. arXiv. http://
arxiv.org/abs/1901.07945
Gershman, S. J., & Daw, N. D. (2012). Perception, action and utility: The tangled skein. Principles of brain dynamics:
Global state. Interactions, 293–312. http://books.google.be/books?hl=nl&lr=&id=KOWZ2sZNlWQC&oi=fnd&pg
=PA293&dq=Perception,+Action,+and+Utility:+The+Tangled+Skein+gershman&ots=e-7WaGMx4V&sig=CO2
ffgFSyMpRkWXR5l08YHTnGrc
Ginsberg, S., & Jablonka, E. (2019). The evolution of the sensitive soul. MIT Press.
Glanzer, M. (1958). Curiosity, exploratory drive, and stimulus satiation. Psychological Bulletin, 55(5), 302–315. https://
doi.org/10.1037/h0044731, PubMed: 13591450
Gold, B. P., Pearce, M. T., Mas-Herrero, E., Dagher, A., & Zatorre, R. J. (2019). Predictability and uncertainty in the
pleasure of music: A reward for learning? Journal of Neuroscience: the Official Journal of the Society for Neuroscience, 39(47),
9397–9409. https://doi.org/10.1523/JNEUROSCI.0428-19.2019
Gombrich, E. H. (1963). Meditations on a hobby horse: And other essays on the theory of art. Phaidon Press.
Gottlieb, J., Oudeyer, P. Y., Lopes, M., & Baranes, A. (2013). Information-seeking, curiosity, and attention: Computational
and neural mechanisms. Trends in Cognitive Sciences, 17(11), 585–593. https://doi.org/10.1016/j.tics.2013.09.001
Graf, L. K. M.,  & Landwehr, J. R. (2015). A  dual-process perspective on fluency-based aesthetics: The pleasure-
interest model of aesthetic liking. Personality and Social Psychology Review, 19(4), 395–410. https://doi.org/10.1177/
1088868315574978
Graham, D. J., & Redies, C. (2010). Statistical regularities in art: Relations with visual coding and perception. Vision
Research, 50(16), 1503–1509. https://doi.org/10.1016/j.visres.2010.05.002
Güçlütürk, Y., Jacobs, R. H. A. H.,  & van Lier, R. (2016). Liking versus complexity: Decomposing the inverted
U-curve. Frontiers in Human Neuroscience, 10, 112. https://doi.org/10.3389/fnhum.2016.00112
Hafner, D., Ortega, P. A., Ba, J., Parr, T., Friston, K., & Heess, N. (2020). Action and perception as divergence minimiza-
tion. In arXiv [cs.AI]. arXiv. http://arxiv.org/abs/2009.01791
Haraway, D. J. (2016). Staying with the trouble: Making kin in the Chthulucene. Duke University Press. https://play.google.
com/store/books/details?id=ZvDgDAAAQBAJ
Hart, S. L. (1971). Axiology—Theory of values. Philosophy and Phenomenological Research, 32(1), 29–41. https://doi.
org/10.2307/2105883
Hebb, D. O. (1949). The organization of behaviour. John Wiley.
Heilbron, M., Armeni, K., Schoffelen, J.-M., Hagoort, P.,  & de Lange, F. P. (2021). A hierarchy of linguistic predic-
tions during natural language comprehension. In bioRxiv (p. 2020.12.03.410399). https://doi.org/10.1101/2020.12.03.
410399
Hesp, C., Smith, R., Parr, T., Allen, M., Friston, K. J., & Ramstead, M. J. D. (2021). Deeply felt affect: The emer-
gence of valence in deep active inference. Neural Computation, 33(2), 398–446. https://doi.org/10.1162/neco_a_
01341

501
Van de Cruys, Bervoets and Moors

Heylighen, F., Kingsbury, K., Lenartowicz, M., Harmsen, T., & Beigi, S. (2018). Social systems programming: Behav-
ioural and emotional mechanisms co-opted for social control. Manuscript Submitted for Publication. http://
pespmc1. Vub. Ac. be/Papers/SSP2mechanisms. Pdf. [SVdC]. http://pespmc1.vub.ac.be/Papers/SSP2mecha
nisms.pdf
Hill, A. (2021, March 23). ‘A new obsession’: The people who learned to play instruments during lockdown. The Guard-
ian. www.theguardian.com/world/2021/mar/23/a-new-obsession-the-people-who-learnt-to-play-instruments-
during-lockdown
Hirsh, J. B., Mar, R. A., & Peterson, J. B. (2012). Psychological entropy: A framework for understanding uncertainty-
related anxiety. Psychological Review, 119(2), 304–320. https://doi.org/10.1037/a0026767
Hoel, E. (2019). Enter the supersensorium. The Baffler, 45, 118–133. www.jstor.org/stable/26639757?casa_token=s
qMDDLFrA5cAAAAA:ANKYuvUQ34o8hrXXXMNkU-nj74nTaZevqPv6h1_kB6q-TGNg8eto6pxmbGTaF_
Qz1D84e_05UhB2aae-kBFmEYYljoI17XQTfffvH7U5q-l4ePsVlEZi
Hoel, E. (2020). The Overfitted Brain: Dreams evolved to assist generalization. arXiv Preprint arXiv:2007.09560. http://
arxiv.org/abs/2007.09560
Hoffman, D. D., Singh, M., & Prakash, C. (2015). The interface theory of perception. Psychonomic Bulletin and Review,
22(6), 1480–1506. https://doi.org/10.3758/s13423-015-0890-8
Hohwy, J. (2014). The self-evidencing brain. Noûs, 50(2), 259–285. https://doi.org/10.1111/nous.12062
Hohwy, J. (2020). New directions in predictive processing. Mind and Language, 35(2), 209–223. https://onlinelibrary.
wiley.com/doi/abs/10.1111/mila.12281
Holm, L. (2017). Curiosity and expected information gain in word learning. Editors: Anders Arweström Jansson Uppsala
University Anders. Arwestrom. Jansson@ It. Uu. Se Anton Axelsson Uppsala University Anton. Axelsson@ It. Uu. Se Rebecca
Andreasson Uppsala University Rebecca. Andreasson@ It. Uu. Se, 43. www.diva-portal.org/smash/get/diva2:1156189/
FULLTEXT01.pdf#page=51
Hume, D. (2008). An enquiry concerning human understanding. https://doi.org/10.1093/owc/9780199549900.001.0001
Hunt, J. M. (1981). Experiential roots of intention, initiative, and trust. In H. I. Day (Ed.), Advances in intrinsic motivation
and aesthetics (pp. 169–202). Springer. https://doi.org/10.1007/978-1-4613-3195-7_8
Huron, D. (2006). Sweet anticipation: Music and the psychology of expectation (1st ed). MIT Press. http://www.amazon.com/
dp/0262083450
Issa, E. B., Cadieu, C. F., & DiCarlo, J. J. (2018). Neural dynamics at successive stages of the ventral visual stream are
consistent with hierarchical error signals. eLife, 7. https://doi.org/10.7554/eLife.42870
Jackendoff, R., & Lerdahl, F. (2006). The capacity for music: What is it, and what’s special about it? Cognition, 100(1),
33–72. https://doi.org/10.1016/j.cognition.2005.11.005
James, W. (1890). The principles of psychology. Harvard University Press. www.wwww4.com/read_book/_the_princi-
ples_of_psychology_1541318.pdf
Joffily, M., & Coricelli, G. (2013). Emotional valence and the free-energy principle. PLOS Computational Biology, 9(6),
e1003094. https://doi.org/10.1371/journal.pcbi.1003094
Juechems, K., & Summerfield, C. (2019). Where does value come from? Trends in Cognitive Sciences, 23(10), 836–850.
https://doi.org/10.1016/j.tics.2019.07.012
Kagan, J. (1972). Motives and development. Journal of Personality and Social Psychology, 22(1), 51–66. https://doi.
org/10.1037/h0032356
Kawakami, K., Phills, C. E., Steele, J. R., & Dovidio, J. F. (2007). (Close) distance makes the heart grow fonder: Improv-
ing implicit racial attitudes and interracial interactions through approach behaviours. Journal of Personality and Social
Psychology, 92(6), 957–971. https://doi.org/10.1037/0022-3514.92.6.957
Keller, G. B., & Mrsic-Flogel, T. D. (2018). Predictive processing: A canonical cortical computation. Neuron, 100(2),
424–435. https://doi.org/10.1016/j.neuron.2018.10.003
Keller, H., & Voss, H.-G. (1983). Curiosity and exploration: Theories and results. Academic Press.
Kelly, G. A. (1958). Man’s construction of his alternatives. Assessment of Human Motives, 33–64.
Kelly, G. A. (1964). The language of hypothesis: Man’s psychological instrument. Journal of Individual Psychology, 20(2),
137–152. www.aippc.it/wp-content/uploads/2019/04/2014.01.005.015.pdf
Kesner, L. (2014). The predictive mind and the experience of visual art work. Frontiers in Psychology, 5, 1417. https://
doi.org/10.3389/fpsyg.2014.01417
Kidd, C., & Hayden, B. Y. (2015). The psychology and neuroscience of curiosity. Neuron, 88(3), 449–460. https://doi.
org/10.1016/j.neuron.2015.09.010
Kidd, C., Piantadosi, S. T., & Aslin, R. N. (2012). The Goldilocks effect: Human infants allocate attention to visual
sequences that are neither too simple nor too complex. PLoS One, 7(5), e36399. https://doi.org/10.1371/journal.
pone.0036399

502
Preferences need inferences

Kidd, C., Piantadosi, S. T., & Aslin, R. N. (2014). The Goldilocks effect in infant auditory attention. Child Development,
85(5), 1795–1804. https://doi.org/10.1111/cdev.12263
Kilner, J. M., Friston, K. J., & Frith, C. D. (2007). Predictive coding: An account of the mirror neuron system. Cognitive
Processing, 8(3), 159–166. https://doi.org/10.1007/s10339-007-0170-2
Kiverstein, J., Miller, M., & Rietveld, E. (2019). The feeling of grip: Novelty, error dynamics, and the predictive brain.
Synthese, 196(7), 2847–2869. https://doi.org/10.1007/s11229-017-1583-9
Kivetz, R., Urminsky, O., & Zheng, Y. (2006). The goal-gradient hypothesis resurrected: Purchase acceleration, illu-
sionary goal progress, and customer retention. Journal of Marketing Research, 43(1), 39–58. https://doi.org/10.1509/
jmkr.43.1.39
Kizilirmak, J. M., Galvao Gomes da Silva, J., Imamoglu, F., & Richardson-Klavehn, A. (2016). Generation and the sub-
jective feeling of “aha!” are independently related to learning from insight. Psychological Research, 80(6), 1059–1074.
https://doi.org/10.1007/s00426-015-0697-2
Koelsch, S., Vuust, P., & Friston, K. J. (2019). Predictive processes and the peculiar case of music. Trends in Cognitive Sci-
ences, 23(1), 63–77. https://doi.org/10.1016/j.tics.2018.10.006
Koenderink, J. (2010). Vision and information. In L. Albertazzi, G. J. Van Tonder, & D. Vishwanath (Eds.), Perception
beyond inference: The information content of visual processes (p. 27). MIT Press. http://books.google.be/books?hl=nl&lr=
&id=Kaw100wfsBYC&oi=fnd&pg=PA27&dq=koenderink+vision+information&ots=_4TRc59NZY&sig=ItPvA
KWBOuI09r-nV-rUt8OcpVI
Koffka, K. (1935). Principles of Gestalt psychology. Brace.
Köhler, W. (1920). Die physischen Gestalten in Ruhe und im stationåren Zustand. Vieweg.
Leavitt, J. D., & Christenfeld, N. J. S. (2011). Story spoilers don’t spoil stories. Psychological Science, 22(9), 1152–1154.
https://doi.org/10.1177/0956797611417007
Leder, H., Bär, S., & Topolinski, S. (2012). Covert painting simulations influence aesthetic appreciation of artworks.
Psychological Science, 23(12), 1479–1481. https://doi.org/10.1177/0956797612452866
Leder, H., & Nadal, M. (2014). Ten years of a model of aesthetic appreciation and aesthetic judgments: The aesthetic
episode—Developments and challenges in empirical aesthetics. British Journal of Psychology, 105(4), 443–464. https://
doi.org/10.1111/bjop.12084
Livson, N. (1967). Towards a differentiated construct of curiosity. Journal of Genetic Psychology, 111(1st Half  ), 73–84.
https://doi.org/10.1080/00221325.1967.10533749
Loewenstein, G. (1994). The psychology of curiosity: A review and reinterpretation. Psychological Bulletin, 116(1), 75–98.
http://psycnet.apa.org/record/1994-41058-001. https://doi.org/10.1037/0033-2909.116.1.75
McReynolds, P. (1971). The three faces of cognitive motivation. In H. I. Day, D. E. Berlyne & D. E. Hunt (Eds.), Intrinsic
motivation: A new direction in education (pp. 33–45). Holt, Rinehart and Winston of Canada. https://market.android.
com/details?id=book-o27GQgAACAAJ.
Menninghaus, W., Wagner, V., Hanich, J., Wassiliwizky, E., Jacobsen, T., & Koelsch, S. (2017). The distancing-embracing
model of the enjoyment of negative emotions in art reception. Behavioral and Brain Sciences, 40, e347. https://doi.
org/10.1017/S0140525X17000309
Metcalfe, J., Schwartz, B. L., & Eich, T. S. (2020). Epistemic curiosity and the region of proximal learning. Current Opin-
ion in Behavioral Sciences, 35, 40–47. https://doi.org/10.1016/j.cobeha.2020.06.007
Metzinger, T. (2007). Self models. Scholarpedia, 2(10), 4174. www.scholarpedia.org/article/Self_models?&sa=U&ei=
Y8UNVPSgLMK_ygPWgYLADw&ved=0CFMQFjAO&usg=AFQjCNEiwQEhiPd-fSDqXm0_kTQEYkMnUg
Meyer. (1961). Emotion and meaning in music. University of Chicago Press. www.amazon.co.uk/dp/0226521397
Miller, M., Nave, K.,  & Clark, G. D. (2020, September  25). Use uncertainty to leverage the power of your predic-
tive brain—Mark Miller, Kathryn Nave. George Deane & Andy Clark. Aeon. https://aeon.co/essays/use-uncertainty-
to-leverage-the-power-of-your-predictive-brain
Montoya, R. M., Horton, R. S., Vevea, J. L., Citkowicz, M., & Lauber, E. A. (2017). A re-examination of the mere
exposure effect: The influence of repeated exposure on recognition, familiarity, and liking. Psychological Bulletin,
143(5), 459–498. https://doi.org/10.1037/bul0000085
Mrkva, K., & Van Boven, L. (2020). Salience theory of mere exposure: Relative exposure increases liking, extremity,
and emotional intensity. Journal of Personality and Social Psychology, 118(6), 1118–1145. https://doi.org/10.1037/
pspa0000184
Muth, C., & Carbon, C.-C. (2016). SeIns: Semantic instability in art. Art and Perception, 4(1–2), 145–184. https://doi.
org/10.1163/22134913-00002049
Muth, C., Raab, M. H., & Carbon, C. C. (2015). The stream of experience when watching artistic movies. Dynamic
aesthetic effects revealed by the continuous evaluation procedure (CEP). Frontiers in Psychology, 6, 365. https://doi.
org/10.3389/fpsyg.2015.00365

503
Van de Cruys, Bervoets and Moors

Muth, C., Raab, M. H., & Carbon, C. C. (2016). Semantic stability is more pleasurable in unstable episodic contexts.
On the relevance of perceptual challenge in art appreciation. Frontiers in Human Neuroscience, 10, 43. https://doi.
org/10.3389/fnhum.2016.00043
Nascimento, S. M. C., Marit Albers, A.,  & Gegenfurtner, K. R. (2021). Naturalness and aesthetics of colors—
Preference for color compositions perceived as natural. Vision Research, 185, 98–110. https://doi.org/10.1016/j.
visres.2021.03.010
Nguyen, C. T. (2017). Philosophy of games. Philosophy Compass, 12(8), e12426. https://doi.org/10.1111/phc3.12426
Nicki, R. M. (1970). The reinforcing effect of uncertainty reduction on a human operant. Canadian Journal of Psychology/
Revue Canadienne de Psychologie, 24(6), 389–400. https://doi.org/10.1037/h0082875
Oudeyer, P.-Y., Kaplan, F., & Hafner, V. V. (2007). Intrinsic motivation systems for autonomous mental development.
IEEE Transactions on Evolutionary Computation, 11(2), 265–286. https://doi.org/10.1109/TEVC.2006.890271
Pearce, M. T. (2018). Statistical learning and probabilistic prediction in music cognition: Mechanisms of stylistic encul-
turation. Annals of the New York Academy of Sciences. https://doi.org/10.1111/nyas.13654
Pelowski, M., Markey, P. S., Forster, M., Gerger, G., & Leder, H. (2017). Move me, astonish me . . . delight my eyes
and brain: The Vienna integrated model of top-down and bottom-up processes in art perception (VIMAP) and cor-
responding affective, evaluative, and neurophysiological correlates. Physics of Life Reviews, 21, 80–125. www.science-
direct.com/science/article/pii/S1571064517300325
Pepperell, R. (2018). Chapter 19. Art, energy, and the brain. In J. F. Christensen & A. Gomila (Eds.), Progress in brain
research (Vol. 237, pp. 417–435). Elsevier. https://doi.org/10.1016/bs.pbr.2018.03.022
Pezzulo, G., Rigoli, F., & Friston, K. (2015). Active Inference, homeostatic regulation and adaptive behavioural control.
Progress in Neurobiology, 134, 17–35. https://doi.org/10.1016/j.pneurobio.2015.09.001
Pinna, B. (2010). New Gestalt principles of perceptual organization: An extension from grouping to shape and meaning.
Gestalt Theory, 32, 1–67.
Poetry Foundation. (2021a). Negative capability. www.poetryfoundation.org/learn/glossary-terms/negative-capability
Poetry Foundation. (2021b). Selections from Keats’s letters by John Keats. www.poetryfoundation.org/articles/69384/
selections-from-keatss-letters
Powers, W. T. (1973). Behaviour: The control of perception, xi. Aldine.
Reber, R., Schwarz, N.,  & Winkielman, P. (2004). Processing fluency and aesthetic pleasure: Is beauty in the per-
ceiver’s processing experience? Personality and Social Psychology Review, 8(4), 364–382. https://doi.org/10.1207/
s15327957pspr0804_3
Reisenzein, R. (2009). Emotional experience in the computational belief–desire theory of emotion. Emotion Review,
1(3), 214–222. https://doi.org/10.1177/1754073909103589
Rohe, T., & Noppeney, U. (2015). Cortical hierarchies perform Bayesian causal inference in multisensory perception.
PLOS Biology, 13(2), e1002073. https://doi.org/10.1371/journal.pbio.1002073
Ryali, C. K., Goffin, S., Winkielman, P., & Yu, A. J. (2020). From likely to likable: The role of statistical typicality in
human social assessment of faces. Proceedings of the National Academy of Sciences of the United States of America, 117(47),
29371–29380. https://doi.org/10.1073/pnas.1912343117
Sarasso, P., Neppi-Modona, M., Sacco, K.,  & Ronga, I. (2020). ‘Stopping for knowledge’: The sense of beauty in
the perception-action cycle. Neuroscience and Biobehavioral Reviews, 118, 723–738. https://doi.org/10.1016/j.
neubiorev.2020.09.004
Sarasso, P., Perna, P., Barbieri, P., Neppi-Modona, M., Sacco, K.,  & Ronga, I. (2021). Memorisation and implicit
perceptual learning are enhanced for preferred musical intervals and chords. Psychonomic Bulletin and Review, 28(5),
1623–1637. https://doi.org/10.3758/s13423-021-01922-z
Schillaci, G., Ciria, A., & Lara, B. (2020). Tracking emotions: Intrinsic motivation grounded on multi-level prediction
error dynamics. In Joint IEEE 10th international conference on development and learning and epigenetic robotics (ICDL-
EpiRob) (pp. 1–8). https://doi.org/10.1109/ICDL-EpiRob48136.2020.9278106
Schmidhuber, J. (2009). Simple algorithmic theory of subjective beauty, novelty, surprise, interestingness, attention, curi-
osity, creativity, art, science, music, jokes. Journal of SICE, 48(1). www.idsia.ch/~juergen/sice2009.pdf.
Schneider, K., & Heckhausen, H. (1981). Subjective uncertainty and task preference. In H. I. Day (Ed.), Advances in
intrinsic motivation and aesthetics (pp. 149–167). Springer. https://doi.org/10.1007/978-1-4613-3195-7_7
Schoeller, F., & Perlovsky, L. (2016). Aesthetic chills: Knowledge-acquisition, meaning-making, and aesthetic emotions.
Frontiers in Psychology, 7, 1093. https://doi.org/10.3389/fpsyg.2016.01093
Schoeller, F., Perlovsky, L., & Arseniev, D. (2018). Physics of mind: Experimental confirmations of theoretical predic-
tions. Physics of Life Reviews, 25, 45–68. https://doi.org/10.1016/j.plrev.2017.11.021
Schoeller, F., Perlovsky, L., et al. (2015). Great expectations—Narratives and the elicitation of aesthetic chills. Psychology,
6(16), 2098. www.scirp.org/html/5-6901655_62308.htm

504
Preferences need inferences

Schonberg, T., & Katz, L. N. (2020). A neural pathway for nonreinforced preference change. Trends in Cognitive Sciences,
24(7), 504–514. https://doi.org/10.1016/j.tics.2020.04.002
Sengupta, B., Stemmler, M. B., & Friston, K. J. (2013). Information and efficiency in the nervous system—A synthesis.
PLOS Computational Biology, 9(7), e1003157. https://doi.org/10.1371/journal.pcbi.1003157
Seth, A. K. (2019). From unconscious inference to the beholder’s share: Predictive perception and human experience.
European Review, 27(3), 378–410. https://doi.org/10.1017/S1062798719000061
Seth, A. K., Millidge, B., Buckley, C. L., & Tschantz, A. (2020). Curious inferences: Reply to sun and firestone on
the dark room problem. Trends in Cognitive Sciences, 24(9), 681–683. cell.com. https://doi.org/10.1016/j.tics.2020.
05.011
Sharot, T. (2011). The optimism bias. Current Biology: CB, 21(23), R941–R945. https://doi.org/10.1016/j.
cub.2011.10.030
Shklovsky, V. (1917). Art as technique. Literary Theory: An Anthology, 3. https://books.google.be/books?hl=en&lr=&id
=XIL5DQAAQBAJ&oi=fnd&pg=RA1-PA8&dq=shklovsky+technique&ots=kldegdtA5O&sig=ktxgstenMqqOyp
FXcoMbIk577Yg.
Silvia, P. J. (2005). Emotional responses to art: From collation and arousal to cognition and emotion. Review of General
Psychology: Journal of Division 1, of the American Psychological Association, 9(4), 342. https://pdfs.semanticscholar.org/a87
a/95da82793dfe51b6f0a6d9d52c1c5d428f7c.pdf
Singh, M. (2021). The sympathetic plot, its psychological origins, and implications for the evolution of fiction. Emotion
Review, 13(3), 183–198. https://doi.org/10.1177/17540739211022824
Skov, M., & Nadal, M. (2020). A farewell to art: Aesthetics as a topic in psychology and neuroscience. Perspectives on
Psychological Science, 15(3), 630–642. https://doi.org/10.1177/1745691619897963
Slamecka, N. J., & Graf, P. (1978). The generation effect: Delineation of a phenomenon. Journal of Experimental Psychol-
ogy: Human Learning and Memory, 4(6), 592–604. https://doi.org/10.1037/0278-7393.4.6.592
Smith, R., Friston, K., & Whyte, C. (2021). A step-by-step tutorial on active inference and its application to empirical
data. https://doi.org/10.31234/osf.io/b4jm6
Solms, M. (2018). The hard problem of consciousness and the free energy principle. Frontiers in Psychology, 9, 2714.
https://doi.org/10.3389/fpsyg.2018.02714
Spehar, B., Walker, N., & Taylor, R. P. (2016). Taxonomy of individual variations in aesthetic responses to fractal pat-
terns. Frontiers in Human Neuroscience, 10, 350. https://doi.org/10.3389/fnhum.2016.00350
Spratling, M. W. (2017). A  review of predictive coding algorithms. Brain and Cognition, 112, 92–97. https://doi.
org/10.1016/j.bandc.2015.11.003
Srivastava, N., & Schrater, P. (2015). Learning what to want: Context-sensitive preference learning. PLoS One, 10(10),
e0141129. https://doi.org/10.1371/journal.pone.0141129
Sun, Z.,  & Firestone, C. (2020). The dark room problem. Trends in Cognitive Sciences, 24(5), 346–348. https://doi.
org/10.1016/j.tics.2020.02.006
Taleb, N. N. (2013). ‘Antifragility’as a mathematical idea. Nature, 494(7438), 430–430. www.nature.com/articles/494430e
Terwilliger, R. F. (1963). Pattern complexity and affective arousal. Perceptual and Motor Skills, 17, 387–395. https://doi.
org/10.2466/pms.1963.17.2.387
Tracy, H. L. (1946). Aristotle on aesthetic pleasure. Classical Philology, 41(1), 43–46. https://doi.org/10.1086/362919
Trope, Y. (1975). Seeking information about one’s ability as a determinant of choice among tasks. Journal of Personality
and Social Psychology, 32(6), 1004–1013. https://doi.org/10.1037/0022-3514.32.6.1004
Tschantz, A., Seth, A. K., & Buckley, C. L. (2020). Learning action-oriented models through active inference. PLOS
Computational Biology, 16(4), e1007805. https://doi.org/10.1371/journal.pcbi.1007805
Van de Cruys, S. (2017). Affective value in the predictive mind. In T. K. Metzinger & W. Wiese (Eds.), Philosophy and
predictive processing. MIND Group. https://doi.org/10.15502/9783958573253
Van de Cruys, S. (2018). Upgrading Gestalt psychology with variational neuroethology: The case of perceptual pleasures:
Comment on “Answering Schrödinger’s question: A free-energy formulation” by M.J. Desormeau Ramstead et al.
Physics of Life Reviews, 24(1), 21–23. https://lirias.kuleuven.be/retrieve/503242
Van de Cruys, S., Chamberlain, R., & Wagemans, J. (2017). Tuning in to art: A predictive processing account of negative
emotion in art. Behavioral and Brain Sciences, 40, e377. https://doi.org/10.1017/S0140525X17001868
Van de Cruys, S., Damiano, C., Boddez, Y., Król, M., Goetschalckx, L., & Wagemans, J. (2021). Visual affects: Linking
curiosity, Aha-Erlebnis, and memory through information gain. Cognition, 212, 104698. https://doi.org/10.1016/j.
cognition.2021.104698
Van de Cruys, S., Evers, K., Van der Hallen, R., Van Eylen, L., Boets, B., de-Wit, L., & Wagemans, J. (2014). Pre-
cise minds in uncertain worlds: Predictive coding in autism. Psychological Review, 121(4), 649–675. https://doi.
org/10.1037/a0037665

505
Van de Cruys, Bervoets and Moors

Van de Cruys, S., Friston, K. J., & Clark, A. (2020). Controlled optimism: Reply to Sun and Firestone on the dark room
problem. Trends in Cognitive Sciences, 24(9), 680–681. https://doi.org/10.1016/j.tics.2020.05.012
Van de Cruys, S.,  & Wagemans, J. (2011). Putting reward in art: A  tentative prediction error account of visual art.
i-Perception, 2(9), 1035–1062. https://doi.org/10.1068/i0466aap
Van Dessel, P., Hughes, S., & De Houwer, J. (2019). How do actions influence attitudes? An inferential account of the
impact of action performance on stimulus evaluation. Personality and Social Psychology Review, 23(3), 267–284. https://
doi.org/10.1177/1088868318795730
Van Geert, E., & Wagemans, J. (2020). Order, complexity, and aesthetic appreciation. Psychology of Aesthetics, Creativity,
and the Arts, 14(2), 135–154. https://psycnet.apa.org/journals/aca/14/2/135/
van Lieshout, L. L. F., Vandenbroucke, A. R. E., Müller, N. C. J., Cools, R., & de Lange, F. P. (2018). Induction and
relief of curiosity elicit parietal and frontal activity. Journal of Neuroscience: The Official Journal of the Society for Neurosci-
ence, 38(10), 2579–2588. https://doi.org/10.1523/JNEUROSCI.2816-17.2018
Von Foerster, H. (1960). On self-organizing systems and their environments (pp. 31–50). Self-Organizing Systems. www.
researchgate.net/profile/Mariana-Broens/publication/271732162_Complexity_and_information_technologies_an_
ethical_inquiry_into_human_autonomous_action/links/564a83ca08ae127ff986aa22/Complexity-and-information-
technologies-an-ethical-inquiry-into-human-autonomous-action.pdf
von Glasersfeld, E. (1995). Radical constructivism: A way of knowing and learning. Studies In Mathematics Education
Series: 6. Falmer press, Taylor and Francis Inc. http://eric.ed.gov/?id=ED381352
Vygotsky, L. S. (1962). Thought and word. https://psycnet.apa.org/record/2006-10268-007
Wacongne, C., Labyt, E., van Wassenhove, V., Bekinschtein, T., Naccache, L., & Dehaene, S. (2011). Evidence for a
hierarchy of predictions and prediction errors in human cortex. Proceedings of the National Academy of Sciences of the
United States of America, 108(51), 20754–20759. https://doi.org/10.1073/pnas.1117807108
Wagemans, J. (2015). Historical and conceptual background: Gestalt theory. The Oxford handbook of perceptual.
Organization, 3–20. www.gestaltrevision.be/pdfs/oxford/Wagemans-Historical_and_conceptual_background_Ges
talt_theory.pdf
Wagemans, J., Elder, J. H., Kubovy, M., Palmer, S. E., Peterson, M. A., Singh, M., & Von der, H. R. (2012). A century of
Gestalt psychology in visual perception: I. Perceptual grouping and figure-ground organization. Psychological Bulletin,
138(6), 1172–1217. https://doi.org/10.1037/a0029333
Walker, E. L. (1981). The quest for the inverted U. In Advances in intrinsic motivation and aesthetics (pp. 39–70). Springer.
https://link.springer.com/chapter/10.1007/978-1-4613-3195-7_3
Wang, X., Bylinskii, Z., Hertzmann, A., & Pepperell, R. (2020). Toward quantifying ambiguities in artistic images. ACM
Transactions on Applied Perception, 17(4), 1–10. https://doi.org/10.1145/3418054
Wänke, M., & Hansen, J. (2015). Relative processing fluency. Current Directions in Psychological Science, 24(3), 195–199.
https://doi.org/10.1177/0963721414561766
Weick, K. E. (1979). The social psychology of organizing. Random House. https://play.google.com/store/books/
details?id=gHJPAAAAMAAJ
Wilkinson, S., Deane, G., Nave, K., & Clark, A. (2019). Getting warmer: Predictive processing and the nature of emo-
tion. In L. Candiotto (Ed.), The value of emotions for knowledge (pp.  101–119). Springer International Publishing.
https://doi.org/10.1007/978-3-030-15667-1_5
Wilkinson, Z., Cunningham, R., & Elliott, M. A. (2021). The influence of empathy on the perceptual response to visual
art. Psychology of Aesthetics, Creativity, and the Arts. https://doi.org/10.1037/aca0000418
Williams, D. (2020). Predictive coding and thought. Synthese, 197(4), 1749–1775. https://doi.org/10.1007/
s11229-018-1768-x
Winkielman, P., Schwarz, N., Fazendeiro, T., Reber, R., et  al. (2003). The hedonic marking of processing fluency:
Implications for evaluative judgment. Psychology of Evaluation: Affective Processes in Cognition and Emotion, 189, 217.
https://books.google.com/books?hl=en&lr=&id=t1h6AgAAQBAJ&oi=fnd&pg=PA195&dq=fazendeiro+winkiel
man&ots=bDBy66NH8_&sig=lve9qpJ1K5569bs3U3cGksDne58
Witek, M. A. G., Clarke, E. F., Wallentin, M., Kringelbach, M. L., & Vuust, P. (2015). Correction: Syncopation, body-
movement and pleasure in groove music. PLoS One, 10(9), e0139409. https://doi.org/10.1371/journal.pone.0139409
Zajonc, R. B. (1968). Attitudinal effects of mere exposure. Journal of Personality and Social Psychology, 9(2, Pt.2), 1–27.
https://doi.org/10.1037/h0025848
Zajonc, R. B. (1980). Feeling and thinking: Preferences need no inferences. American Psychologist, 35(2), 151–175.
https://doi.org/10.1037/0003-066X.35.2.151
Zangwill, N. (2021). Aesthetic judgment. In E. N. Zalta (Ed.), Stanford encyclopedia of philosophy. Metaphysics Research
Laboratory, Stanford University. https://plato.stanford.edu/archives/spr2021/entries/aesthetic-judgment/
Zénon, A., Solopchuk, O., & Pezzulo, G. (2019). An information-theoretic perspective on the costs of cognition. Neu-
ropsychologia, 123, 5–18. https://doi.org/10.1016/j.neuropsychologia.2018.09.013

506
27
NEUROSCIENCE OF ARTISTIC
CREATIVITY
Oshin Vartanian

Historically, there has been longstanding interest in the biological bases of creativity (see Martindale, 1999).
Early scientific interest in this question can be traced to Sir Francis Galton’s (1869) landmark monograph
Hereditary Genius, which focused on the genetics of eminence. Specifically, Galton wondered whether genius
runs in families, an idea that was perhaps not too surprising given that his blood relations included his
grandfather Erasmus Darwin; his cousin Charles Darwin; and Charles’ scientifically eminent sons Francis the
botanist, Leonard the eugenicist, and Sir George the physicist, as well as the latter’s son, Sir Charles Galton
Darwin the physicist. To test his hypothesis that genius runs in families, Galton collected large volumes
of biographical data in various domains (literature, poetry, music, etc.), which he in turn analyzed statisti-
cally using methods that he invented specifically for this purpose (e.g., correlation coefficient). Despite his
painstaking efforts, the work was plagued with many difficulties. First and foremost, there was a criterion
problem because he defined genius in terms of achieved distinction or reputation, specifically “the opinion
of contemporaries, revised by posterity . . . the reputation of a leader of opinion, of an originator, of a man
to whom the world deliberately acknowledges itself largely indebted” (p. 37). This ambiguous definition
did not lend itself to an accurate characterization of people as born geniuses, because reputation could be
influenced as much by shared familial environment as by shared genetics (Simonton, 1994). In addition,
there were difficulties in ascertaining the actual familial relatedness of the people whom he studied, further
undermining the validity of his findings. As a result, overall, he was not able to provide convincing evidence
in support of his hypothesis. Despite these shortcomings, this book is considered a classic not only because
Galton collected and analyzed a large corpus of biographical data statistically but also because he treated intel-
lectual abilities as traits that obeyed the same laws as those that govern the rest of the natural world. As he
noted in his Introductory chapter, “I propose to show in this book that a man’s natural abilities are derived
by inheritance, under exactly the same limitations as are the form and physical features of the whole organic
world” (p. B). In this sense, he took great strides to naturalize the study of the mind as a topic worthy of
scientific study in its own right.
Interestingly, shortly after Hereditary Genius, Galton also conducted what is considered the very first inves-
tigation to study highly creative individuals directly (Simonton, 2001). Specifically, to figure out the causes of
their creativity, he used a questionnaire that he had constructed himself to survey the elected Fellows of the
Royal Society to measure the relative contributions of nature and nurture to eminence in science, addressing

DOI: 10.4324/9781003008675-29 507


Oshin Vartanian

a wide host of developmental issues, including the distribution of ability in the family, birth order, and edu-
cational experience, among others. Based on the analysis of his survey data, he noted that

It is, I believe, owing to the favourable conditions of their early training that an unusually large
proportion of the sons of the most gifted men of science become distinguished in the same career.
They have been nurtured in an atmosphere of free enquiry.
(Galton, 1874, p. 197)

suggesting an appreciation of the interplay between genetic and environmental factors in fostering emi-
nence. This monograph, published in 1874 and entitled English Men of Science: Their Nature and Nurture,
represents an important scientific milestone on the relative contributions of genetic vs. environmental factors
in the expression of creativity and demonstrated that one could collect data from creative persons to probe
the basis of their eminence.
Of course, we now know that genius does not run in families. Rather, studies have shown that creativ-
ity is an emergenic trait (Martindale, 1999; Waller et al., 1993), meaning that it is expressed if and only if
a number of independent subtraits or abilities are simultaneously present (Lykken et al., 1992). Because the
relationship among the subtraits is not additive but multiplicative in nature, the absence of any one compo-
nent is sufficient to block the occurrence of the emergenic trait altogether. Thus, even though many of the
subtraits may themselves may be normally distributed throughout the population (e.g., intelligence, openness
to experience), the emergenic trait itself exhibits a log-normal distribution (i.e., skewed distributions that
arise as the multiplicative product of many independent variables). This is because it is unlikely that any given
individual will simultaneously possess all the necessary components for the expression of the emergenic trait,
let alone different members of the same family. Consistent with this view, the distribution of real-life creative
achievement has been shown to be skewed based on various measures of real-life creativity (e.g., Diedrich
et al., 2018; Jauk et al., 2014).

From genetics to brain function


Throughout the 20th century, studies that investigated the biological bases of creativity maintained a focus
on its genetic bases (for reviews, see Martindale, 1999; Vartanian, 2011). However, this began to change
in the latter half of the century, likely because of the availability of technology that enabled studies of
brain function in vivo. Using electroencephalography (EEG), Colin Martindale and colleagues conducted
some of the earliest studies on the neurological bases of creativity. Importantly, they shifted the focus to
the investigation of brain function while participants were engaged in creative cognition. The theoretical
focus of this early work was on the relationship between cortical arousal and creativity. Specifically, because
higher arousal is associated with a narrowing of the focus of attention, it was hypothesized that creative
participants would exhibit lower levels of cortical arousal than non-creative participants, thereby allowing
the entry of more input into the focus of attention, which would in turn increase the probability of their
contribution to novel thought. The earliest studies focused on basal levels of arousal and did not collect
data in the course of creativity tasks. For example, Martindale and Armstrong (1974) administered the
Alternative Uses Task (AUT, Christensen et al., 1960) and the Remote Associates Test (RAT, Mednick &
Mednick, 1967) to their participants and recorded EEG separately. AUT is a test of divergent thinking in
which participants are presented with the names of ordinary objects and instructed to generate creative
uses for them. The output can be scored in relation to various criteria—in that case fluency (i.e., total
number of generated uses). In turn, for each item on the RAT, the participant is presented with three
words (e.g., cottage, Swiss, cake), and instructed to generate a fourth word that is common to all three
words (i.e., cheese). This test is considered to be a measure of one’s associative gradient, with more creative

508
Neuroscience of artistic creativity

people hypothesized to activate a larger number of more distant nodes in their semantic and episodic
memory networks than less creative people, and therefore more likely to stumble upon the common term
(Mednick, 1962; see also Kenett et al., 2014). Contrary to their predictions, the results of Martindale and
Armstrong (1974) demonstrated that creativity, as measured by a composite of AUT and RAT scores, was
associated with higher basal cortical activation as measured by the percent of time alpha waves were present
in the EEG recording. Indeed, Martindale’s (1990) review of seven EEG studies from that era revealed that
in only two of those studies was any difference observed between creative and non-creative participants in
basal cortical arousal, such that the overall pattern did not support the inference that creativity is associated
with this metric of brain activity.
Next, Martindale and Hines (1975) used a different design and recorded EEG activity while participants
completed the AUT, the RAT, and a culture-fair test of nonverbal intelligence (Cattell & Cattell, 1959).
The results demonstrated that creative participants exhibited a high degree of variability in cortical arousal
across the three tasks: They exhibited the lowest level of arousal on the AUT, slightly higher level of arousal
on the RAT, and an even higher level of arousal on the intelligence test. In turn, non-creative participants
exhibited a consistent and high level of arousal while taking all three tests. The authors argued that whereas
the AUT and the intelligence test represent “pure” measures of divergent and convergent thinking, respec-
tively, the RAT falls somewhere in the middle because it is associative in nature but also shares substantial
similarity with tests of intelligence. As such, performance on the RAT represents an admixture of divergent
and convergent thinking. Indeed, there is much evidence to show that the RAT exhibits stronger correlation
with convergent thinking as measured by intelligence tests than does the AUT (e.g., Lee et al., 2014). In this
sense, the results of Martindale and Hines (1975) suggested that in terms of cortical arousal, creative but not
non-creative participants exhibit sensitivity as cognition moves across a continuum from divergent thinking
to convergent thinking. Aside from its pioneering nature, that study is noteworthy for two reasons. First, it
demonstrated that rather than having stable levels of low arousal across all tasks, creative people exhibit more
variable levels of arousal as a function of task demands. Much experimental data has since supported the idea
that creative people exhibit greater variability and flexibility in cognition as a function of tasks demands—
regardless of the mechanism under examination (attention, cognitive control, etc.) (Dorfman et al., 2008;
Vartanian, 2009; Vartanian et al., 2007; Zabelina & Robinson, 2010). Second, it foreshadowed the main
finding of contemporary network neuroscience approaches to understanding creativity, which is that the
dynamics of thought rather than a single form of thought is the hallmark of the neural basis of creativity (see
Christoff et al., 2016). This key finding will be fleshed out next.

Early neuroscientific approaches to creativity across domains


When Martindale (1999) published his review of the biological bases of creativity right before the turn
of the century, no study had yet investigated the neural basis of creativity using a modern neuroimaging
modality—be it artistic creativity or otherwise. This changed soon thereafter, as independent investigators
in Russia (Bekhtereva et al., 2000) and Sweden (Carlsson et al., 2000) obtained measurements of regional
cerebral blood flow (rCBF) using positron emission tomography (PET) to demonstrate the involvement of
the frontal lobes in creative cognition. Thus began an intense period of research activity on the neural bases
of creativity that has now spanned nearly two decades.
The work over the first decade of this activity was not very systematic and typically involved a combining
a paradigm from the creativity literature with a neuroimaging modality to examine its neural correlates. The
typical types of tasks included creative story generation, open-ended problem solving (e.g., match problems),
AUT, RAT, drawing, analogy and metaphor, finding pragmatic links between incoherent sentences, and
anagrams and insight tasks, among others. The neuroimaging methods included functional and structural
magnetic resonance imaging (MRI), PET, EEG, event-related potentials (ERP), near infrared spectroscopy

509
Oshin Vartanian

(NIRS), diffusion tensor imaging (DTI), and single-photon emission computed tomography (SPECT). Not
surprisingly, when this body of work was reviewed a decade in, the added benefit of a neuroscience of creativity
did not seem apparent. For example, when Arden et al. (2010) conducted a review of this literature based
on a methodological categorization of the available studies (i.e., EEG, fMRI, PET, and SPECT), they noted
large variability and inconsistencies, prompting them to argue that

creativity research would benefit from psychometrically informed revision, and the addition of
neuroimaging methods designed to provide greater spatial localization of function. Without such
revision in the behavioral measures and study designs, it is hard to see the benefit of imaging.
(p. 143)

In a nutshell, the emergent neuroimaging and electrophysiological literature at that point did not seem to
have advanced our understanding of the neurology and/or psychology of creativity.
However, soon thereafter, three meta-analyses revealed important consistencies and associated insights
hidden within this work. First, Vartanian (2012) argued that when all manner of tasks and neuroimaging
modalities with different spatial and temporal resolutions are tossed together, it is not unreasonable to observe
a variegated pattern of findings. Instead, one should expect to observe consistent and theoretically derived
set of brain activations for a specific task (and its cognitive components) across studies using the same imaging
modality but dissociable sets of brain activations for closely related tasks that rely on different cognitive sub-
processes (to demonstrate discriminant validity). To test these two hypotheses, he conducted a meta-analysis
of twenty fMRI studies of analogy and metaphor—two processes that have been shown historically to play
a role in creativity. Analogical reasoning occurs when we aim to understand novel situations by drawing
parallels to earlier situations (Sternberg, 1977). Perhaps the most important example of analogical reasoning
in the history of scientific discovery is Bohr’s conceptualization of the motion of the electron around the
nucleus of the hydrogen atom (i.e., target domain) by drawing a parallel with the motion of planets around
the sun in the solar system (i.e., source domain). There is broad agreement that to draw such an analogy
requires two cognitive subcomponents (at minimum): The first is the ability to retrieve relevant content
from long-term memory for input into working memory, whereas the second is the ability to map (i.e.,
align) the representational content of cases in working memory to enable one to project inferences from
one case to another (i.e., source domain to target domain). There is now substantial evidence to suggest that
the core maintenance and manipulation functions of working memory are represented within the fronto-
parietal system in the brain (Baddeley, 2003), as well as converging neuropsychological (Waltz et al., 1999),
neuroimaging (Christoff et al., 2001), and developmental (Crone et al., 2009) evidence to suggest that the
rostrolateral prefrontal cortex (RLPFC) plays a critical role in structural alignment across relations—termed
relational integration (Bunge et al., 2009). As such, one would expect to observe activations in regions within
the fronto-parietal working memory system and RLPFC in analogy.
In contrast, metaphors serve as vehicles for contemplating concepts at higher levels of abstraction, thereby
making category membership more flexible. This can in turn contribute to a defining feature of creative
cognition that involves the flexible manipulation of concepts. For example, how does one understand an
utterance such as “lawyers are sharks”? According to classic standard pragmatic models of metaphor com-
prehension, people extract metaphoric meaning only after failure to extract a literal meaning (Grice, 1975).
In this case, given that the literal meaning (i.e., lawyers are marine creatures) is nonsensical, the metaphoric
meaning (i.e., lawyers are predatory animals) follows. According to this account, one should expect to
observe greater demand on working memory for the retrieval and manipulation of concepts in the focus of
attention, as well as on text comprehension resources for metaphorical rather than literal meaning. Neu-
rologically, these demands should be represented in the fronto-parietal working memory system and the
temporal lobe structures involved in linguistic comprehension (see Mashal et al., 2007).

510
Neuroscience of artistic creativity

Vartanian’s (2012) meta-analysis of the analogy and metaphor literatures supported both predictions.
Specifically, analogy was associated with reliable activation in left RLPFC and dorsolateral prefrontal cortex
(DLPFC) across studies. In contrast, metaphor was associated with reliable activation in left DLPFC and
temporal pole—a well-established hub for text comprehension (Ferstl et al., 2008). In addition, and contrary
to expectation, activation was also observed in the cingulate gyrus for metaphor. This structure is known to
be an important part of the brain’s frontal attentional control system (Carter et al., 1997), and its activation is
consistent with the suggestion that compared to the derivation of literal meaning, metaphoric comprehen-
sion requires more attention. The basic inference from this meta-analysis is that there is process-specificity in
the neural systems that support creativity: When specific cognitive processes are isolated, they have reliable
and dissociable neural correlates.
Another two meta-analyses that were published around the same time and covered much of the same
ground supported similar inferences. First, Gonen-Yaacovi et  al. (2013) conducted a meta-analysis of 34
neuroimaging studies (PET and fMRI) of creativity. When the data were analyzed based on whether the
tasks involved verbal or non-verbal stimuli, the results demonstrated dissociable patterns of activation. Spe-
cifically, verbal tasks engaged the left and right lateral prefrontal cortex (PFC), left anterior cingulate cortex
(ACC), left posterior superior temporal gyrus, right lingual gyrus, and left thalamus. Regions in left PFC
and temporal cortex that were activated more in verbal tasks are known to play important roles in linguistic
processing in terms of the generation and comprehension of language. In turn, regions more consistently
associated with non-verbal tasks included the right and left premotor regions, the left middle frontal gyrus,
and the left occipital cortex. Because most non-verbal tasks within this literature necessitate visuospatial pro-
cessing, the involvement of the occipital cortex and the middle frontal gyrus (which is known to lie in the
dorsal visual stream) would be expected. Furthermore, Gonen-Yaacovi et al. (2013) also found dissociations
when the data were parsed based on whether the tasks necessitated the generation of ideas (e.g., AUT) or the
combination of ideas (e.g., RAT). For example, combination tasks preferentially engaged left RLPFC, left
inferior and middle frontal gyri, right inferior frontal gyrus (IFG) and insula, left posterior middle temporal
gyrus, and left posterior parietal regions. Of particular interest here is the involvement of left RLPFC, given
its role in relational integration—a process that is considered relevant when conducting similarity judgments
of the type that would be necessary in RAT-like problems. In contrast, regions preferentially engaged by
generation tasks included bilateral cerebellum, bilateral thalamus, left inferior parietal lobule, right posterior
cingulate cortex (PCC), and left middle frontal gyrus. In contrast to combination tasks that more consist-
ently activated lateral aspects of the frontal lobes, generation tasks more consistently activated its medial
regions, including left middle frontal gyrus. This region is engaged by counterfactual thinking, prospective
memory, future thinking, mentalizing, and daydreaming. Its role in idea generation—which encompasses
the core feature shared by this list of thought processes—is likely the generation and simulation of new ideas
within the default-mode network. In addition, a large evidence base from lesion and neuroimaging studies
has supported the engagement of the inferior frontal gyrus (IFG) in cognitive, emotional, and behavioural
inhibition (Aron et al., 2003, 2004). Its involvement here signals a possible role for this region in relation to
the inhibition of inappropriate responses in generation tasks (i.e., responses that do not meet certain criteria).
Finally, Boccia et  al. (2015) conducted a meta-analysis of 45 fMRI studies of creativity. They parsed
the data based on domains (i.e., musical, verbal, and visuospatial) and found differences in the neural cor-
relates of creativity based on this categorization. Verbal creativity was associated with the engagement of
left-hemisphere regions in PFC, middle and superior temporal gyri, inferior parietal lobule, postcentral and
supramarginal gyri, middle occipital gyrus, and insula, as well as right IFG and lingual gyrus. The network
that supports verbal creativity is consistent with the linguistic processing that draws on semantic retrieval
and selection of stored knowledge largely localized within the left hemisphere (Badre et al., 2005; Badre &
Wagner, 2007; Moss et al., 2005; Seger et al., 2000; Thompson-Schill et al., 1997). In turn, musical creativity
was associated with activations in the bilateral medial frontal gyrus, left cingulate gyrus, middle frontal gyrus,

511
Oshin Vartanian

inferior parietal lobule, and right postcentral and fusiform gyri, whereas visuospatial creativity was associ-
ated with activations in right middle frontal gyrus, right IFG, bilateral thalamus, and left precentral gyrus.
Notably, both musical and visuospatial creativity engage regions involved in motor planning, consistent with
the idea that these regions contribute to motor and temporal sequencing in the service of musical creativ-
ity (Bengtsson et al., 2007; Berkowitz & Ansari, 2008; Brown et al., 2006; Limb & Braun, 2008), as well as
mental rotation of objects during visuospatial creativity (Milivojevic et al., 2009).
These meta-analyses confirmed that the neural bases of creativity are process specific. In other words, con-
sistent with componential models of creativity (Sternberg, 1980), it appears that various neural structures that
support specific aspects of cognition (e.g., attention, memory, etc.) are dynamically reconfigured in response
to task demands to support the generation of novel and useful output in given contexts. This is precisely
what has been found to be the case for other types of higher-order cognition such as reasoning (see Goel,
2007). Thus, rather than having a module dedicated to creativity, the neural architecture of creativity draws
on many of the same components that support other types of higher-order cognition. Furthermore, consist-
ent with Simon’s (1962) notion regarding the architecture of complexity in the mind, the neural architecture
of higher-order cognition is hierarchical in nature: There are domain-general components that contribute to
almost all instantiations thereof (e.g., memory), whereas there are also domain-specific components that con-
tribute to the process when necessary (e.g., motor planning to musical creativity but not verbal creativity).
The body of work that covered the first decade of research on the neuroscience of creativity set the stage for
network neuroscience approaches that have dominated the second decade of research on creative cognition.
That work will be discussed next.

Contemporary network neuroscience of creativity


One of the major technological advances in neuroimaging research has involved the use of resting-state
functional connectivity MRI to measure correlated patterns of fMRI activity across brain regions during
extended periods of awake rest (see Zabelina & Andrews-Hanna, 2016). Using this technique, one can iden-
tify brain regions that exhibit similar patterns of fMRI activity fluctuations, which can therefore be grouped
into large-scale brain systems called “networks.” In turn, one can study the interactions (i.e., dynamics) of
these large-scale brain networks in the service of various types of thinking, including creative cognition.
There is now substantial and growing evidence to demonstrate that three networks in particular play a key
role in creative cognition (for review, see Beaty et al., 2016). The first network is the default-mode network,
which is implicated in internally generated cognition and is reliably engaged during mind wandering, spon-
taneous thought, and daydreaming, among others. The second network is the executive control network,
which is implicated in the regulation of behaviour and cognitive control. The third is the salience net-
work, which is involved in the detection and allocation of neural resources to behaviourally relevant stimuli
(Bressler & Menon, 2010; Uddin, 2015). A key node within the salience network is the anterior insula,
which has been shown to be involved in higher-order cognitive control and attentional processes (Menon &
Uddin, 2010). Beaty et al. (2015) conducted an important early study that served to elucidate the dynamic
interactions among these networks in the course of engagement in the AUT. Specifically, a seed-based analy-
sis revealed direct functional connections between these three networks. PCC—a region that lies within the
default-mode network—exhibited increased functional coupling with regions of the executive control net-
work including the DLPFC, as well as regions within the salience network such as the bilateral insula. Fur-
thermore, a dynamic functional connectivity analysis conducted in the course of engagement with the AUT
revealed that the time course of the coupling between the PCC and regions within the salience and executive
control networks varied as a function of the phase of the task. Specifically, the PCC showed early coupling
with the insula and later coupling with the right DLPFC, among other regions. There is evidence to show
that one of the roles of the salience network is to facilitate switches between the default-mode and executive

512
Neuroscience of artistic creativity

control networks (Cocchi et al., 2013). As such, its early involvement in the AUT could be to facilitate later
coupling between the default-mode and the executive control networks. Importantly, from a theoretical
perspective, these set of findings are consistent with Campbell’s (1960) and Simonton’s (2010) “blind varia-
tion and selective retention” (BVSR) model of creative cognition according to which creative ideas emerge
in a two-step process consisting of the initial generation of novel ideas followed by the subsequent pruning of
those ideas to select the best output based on task demands. It appears that the default-mode network might
be particularly relevant for the generation of novel thought content, whereas the executive control networks
might play a critical role in executing top-down control over that generative process in the service of task
demands. In turn, the salience network triggers switches between the two networks by directing attention to
behaviourally relevant stimuli throughout the process.
Three bits of related evidence have strengthened the case for the involvement of these networks in crea-
tive cognition. First, Beaty et al. (2014) demonstrated that the close coupling between the default-mode
and the executive control networks is also present at rest. Based on performance on a battery of divergent
thinking tests, they were able to categorize participants into high- and low-creativity groups using behav-
ioural data. Seed-based functional connectivity analysis based on resting-state fMRI data revealed greater
connectivity between the left IFG and the entire default-mode network in the high-creative group, as well
as greater functional connectivity with bilateral inferior parietal cortex and left DLPFC. These data sug-
gested that the close coupling between these networks might be an intrinsic property of cortical function in
creative persons. Second, in a particularly important recent contribution to this literature, Beaty et al. (2018)
used connectome-based predictive modelling to demonstrate that the functional connectivity among the
default-mode, executive control, and salience networks predicts the creative quality of ideas generated by
novel participants within the sample. Furthermore, in a series of external validation analyses using data from
two independent task fMRI samples and a large task-free resting-state fMRI sample, they demonstrated that
the same pattern of functional connectivity could predict individual creative thinking ability in three other
samples. This study is noteworthy because it linked the functional connectivity among cortical hubs within
default-mode, salience, and executive control systems to the creative quality of thought rather than simply
engagement in a particular mode of thought. Third, using dynamic causal modelling (DCM), Vartanian et al.
(2018) demonstrated that the IFG, a node within the executive control system, unidirectionally controls
activity within the temporal and parietal lobes located within the default-mode network in the course of
engagement in the AUT. Thus, rather than revealing connectivity, this study demonstrated a causal pathway
consistent with BVSR according to which one would expect the executive control system to impose control
over neural activity in the default-mode network to ensure that the generation of novel thought content
meets task demands.

The neural bases of artistic creativity


So far, the discussion in this chapter has drawn from creativity research writ large. At this point, it is fair to
ask whether the way in which creativity emerges in the brain applies to all varieties of creativity (e.g., artis-
tic, scientific) or whether its neural architecture varies by the context in which it occurs. This touches on
one of the oldest questions in the scientific study of creativity, which revolves around whether the cognitive
machinery that supports creativity is domain-specific vs. domain-general—where a “domain” is defined by
a set of representations that govern thinking in a specific area of knowledge (Baer, 1998; Kaufman & Baer,
2005). As noted by Baer (2011), domain specificity is “a theory that argues that the skills, traits, or knowledge
that underlie creative performance in a given domain are largely unrelated to the skills, traits, or knowledge
that underlie creative performance in other domains” (p. 404). If domain specificity were true, then it would
follow that from a neurological perspective one would expect the corresponding neural architecture that
supports creativity to vary across domains. As is so often the case in psychological research, it appears that

513
Oshin Vartanian

the neural bases of creativity represent a blend of domain-specific and domain-general contributions. This
picture is consistent with a componential and hierarchical view of higher-order cognition: There is likely to
be generality in the less granular aspects of creativity across domains (e.g., attention, memory), whereas there
should be specificity in the more granular aspects of creativity across domains that are more context-specific
(e.g., imagery, motor control) (Simon, 1962). This is also consistent with contemporary hierarchical concep-
tions of creativity such as Baer and Kaufman’s (2005) amusement park theoretical (APT) model. According to
APT, the expression of creativity is influenced by four levels of factors that move progressively from gener-
ality to specificity. At the outset, there are initial requirements which consist of completely domain-general
factors that influence creativity across the board (e.g., intelligence). Next there are general thematic areas that
represent many related domains (e.g., math and science). Then there are domains that lie within larger gen-
eral thematic areas but refer to a more limited range of creative activities, such as fiction and poetry. Finally,
there are microdomains that refer to more specific activities within domains, such as haiku and sonnets within
the domain of poetry. According to this model, which is well supported empirically, one can expect to find
domain generality and/or specificity depending on which level one is focusing on. Armed with this theo-
retical background, the discussion will now turn to a handful of neuroscientific studies that have begun to
examine artistic creativity—defined as the generation of novel and useful products within artistic domains.
Until quite recently, technological limitations had made it difficult to examine the neural correlates of
artistic creativity in the fMRI scanner. This began to change as MRI-compatible devices enabled artists and
non-artists to engage in creative behaviour in ecologically valid ways inside scanners. An important early
study in this cannon was conducted by Limb and Braun (2008), who studied the neural correlates of jazz
improvisation based on data collected from full-time professional musicians playing jazz using an MRI-
compatible keyboard in the fMRI scanner. Importantly, all participants were highly proficient in playing jazz
piano. In the jazz improvisation condition, the participants were instructed to improvise using a composi-
tion’s underlying chord structure as the basis for spontaneous creative output. In contrast, for the jazz control
condition, they played a novel melody that was memorized prior to scanning. Their results demonstrated
that jazz improvisation compared to the control condition was associated with extensive deactivation of
DLPFC and lateral orbitofrontal regions, with focal activation in the medial prefrontal cortex. The locations
of deactivation fell within the executive control network, whereas the locations of activation fell within the
default-mode network. This led the authors to conclude that “Such a pattern may reflect a combination of
psychological processes required for spontaneous improvisation, in which internally motivated, stimulus-
independent behaviours unfold in the absence of central processes that typically mediate self-monitoring
and conscious volitional control of ongoing performance” (p. 1). In other words, given that jazz improvisa-
tion represents a quintessential example of spontaneous thought, it appears that the brain of improvisers was
characterized by a pattern that facilitated the production of novel content in the face of minimal top-down
control. In addition to the aforementioned pattern of activations and deactivations observed in the PFC,
there was also widespread activation of sensorimotor areas as well as the deactivation of limbic structures dur-
ing jazz improvisation. The activation in sensorimotor areas can be attributed to the motoric requirements
for the execution of music, whereas the deactivation in limbic structures could be related to the regulation of
the motivational and emotional aspects of jazz improvisation. Although this study did not involve an analysis
of the functional connectivity between the default-mode and executive control networks, the results are
consistent with the view that regions within those networks are engaged in musical creativity. Furthermore,
the study also revealed the engagement of domain-specific processes based on the activation observed in
sensorimotor areas in relation to the motoric requirements associated with jazz improvisation.
In turn, Ellamil et al. (2012) conducted another influential study on the neural bases of creative draw-
ing. Their study was designed to test predictions derived from BVSR (Campbell, 1960; Simonton, 2010).
Recall that the basic idea behind BVSR is that the creative process unfolds in two steps, where the first
step involves the generation of novel thought content, and the second step involves the evaluation of those

514
Neuroscience of artistic creativity

generated ideas for selecting the most optimal idea(s) for further development and exploitation. Ellamil et al.
(2012) hypothesized that the executive control network would contribute to the analytical evaluative pro-
cesses required during creative evaluation. They also hypothesized that the default-mode network would be
involved in creative generation because of its role in states of reduced cognitive control, but also perhaps in
creative evaluation because of its role in affective processing. Finally, they hypothesized that medial temporal
lobe (MTL) regions may contribute to associative processes that enable creative generation, because of the
MTL’s well-documented role in making semantic and episodic associations (Aminoff et al., 2007; Bar et al.,
2008; Henke et al., 1997, 1999; Rombouts et al., 1997), as well as during mental simulations of past, future,
and novel events that necessitate recombining stored associations (Addis et al., 2007; Botzung et al., 2008;
Hassabis et al., 2007; Okuda et al., 2003; Szpunar et al., 2007, 2009).
Ellamil et al. (2012) tested their predictions by instructing undergraduates recruited from the Emily Carr
University of Art and Design to design book covers using MRI-compatible tablets in the scanner. Specifi-
cally, the participants were presented with verbal descriptions of hypothetical books, following which they
made a drawing for the book covers. Then, having completed the drawings, they were given time to evaluate
the quality of their work. This design enabled the researchers to separate the neural correlates of generation
from the neural correlates of evaluation. The results demonstrated that compared to the evaluation of draw-
ings, their generation was correlated with relatively greater activation, most notably in the hippocampus and
the parahippocampus in the MTL, as well as other structures, including the inferior parietal lobule, premo-
tor area, left IFG, bilateral superior parietal lobule, bilateral fusiform gyrus, bilateral middle temporal gyrus,
and left cerebellum. In turn, compared to the generation of drawings, their evaluation was correlated with
relatively greater activation primarily in regions within the executive control network (DLPFC and dorsal
ACC), as well as regions within the default-mode network (medial prefrontal cortex, PCC/precuneus, and
temporoparietal junction), RLPFC, cerebellum, temporopolar cortex, and left anterior insula. Activations
were also observed in the supplementary motor area, bilateral IFG, bilateral superior parietal lobule, bilateral
middle temporal gyrus, bilateral lingual gyrus, bilateral middle occipital gyrus, and bilateral cuneus during
the evaluation phase.
This pattern of activations suggests that the evaluation of drawings activated regions within the default-
mode as well as the executive control networks. This finding is consistent with data derived from tests of
divergent thinking in which it has been shown that there is close functional connectivity between these
two networks, particularly in the later stages of creative cognition, presumably as solutions are worked
to completion (Beaty et al., 2015; Vartanian et al. 2018). At the same time, it was found that it is regions
within the MTL (i.e., hippocampus, parahippocampus) that play a prominent role in the generation of new
ideas. At first glance, this might seem counterintuitive, given that historically these regions have largely
been implicated in memory processing, such an encoding and retrieval of semantic and episodic memories.
However, more recent evidence has shown that the MTL performs associative and constructive functions
that may allow it to generate novel ideas, especially processes that support future event simulation (Addis
et al., 2007; Okuda et al., 2003); imagining novel, fictitious scenes (Hassabis et al., 2007); and mental simula-
tions (Schacter & Addis, 2009). This picture is consistent with the idea that the generation of new thought
content draws on enhanced associative processes that recombine and restructure pre-existing content within
episodic and semantic memory (Gabora, 2010; Hospers, 1985; see also Weisberg, 2004). In summary, the
study by Ellamil et al. (2012) demonstrated that when the neural bases of creative drawing are analyzed in
two phases, then the generation of novel ideas is supported by the MTL, whereas the evaluation of those
ideas is supported by structures within the default-mode and executive control networks. The involvement
of MTL structures in creative generation likely reflects their contribution to ideation via associative processes
in episodic and semantic memory, whereas the involvement of default-mode and executive control networks
in creative evaluation likely reflects the execution of top-down control by prefrontal structures on processes
supported by the default-mode network in the later phases of creative cognition. Aside from the involvement

515
Oshin Vartanian

of regions within the default-mode and executive control systems in the evaluation of creative drawings, the
evaluation and generation of drawings was also associated with activations of structures within the occipital
cortex (cuneus, middle occipital gyrus, lingual gyrus) and temporal lobes (fusiform gyrus) involved in visual
perception, including recognition of objects and their properties. This suggests that creative drawing, given
that it draws on vision, also recruits domain-specific components that extend beyond the default-mode and
executive control systems.
Another pioneering study on the neural correlates of artistic creativity was conducted Shah et al. (2013),
focusing on creative writing in non-expert writers. The study was designed to test Flower and Hayes’ (1981)
model of creative writing, according to which the writing process can be broken down into the three stages
of planning, translating, and reviewing. Planning involves generating ideas based on the contents of long-
term memory. Next, translating involves transcribing those ideas into text based on linguistic rules. Finally,
reviewing consists of the evaluation of the written text, possibly involving revision. The study design involved
four conditions (administered sequentially and repeatedly in the scanner). In the reading condition, the par-
ticipants read a text. In the copying condition, they copied the presented text. In the brainstorming condition,
they generated ideas for a “highly creative continuation” to the presented text (without writing anything
down). In the creative writing condition, they wrote down their generated creative ending, with the option to
go beyond what had been thought of in the brainstorming condition. The conditions were interspersed with
rest periods (used as baselines to calculate main effects). As with Ellamil et al. (2012), the study design enabled
the researchers to study the neural correlates of specific phases of the creative writing process.
For the present purposes, there were three particularly noteworthy contrasts. First, brainstorming (minus
rest) was associated with activation in a distributed but primarily left-lateralized network including the left
IFG (Broca’s area), left inferior parietal cortex (including Wernicke’s area), left superior temporal cortex and
the temporal pole, left anterior insula, left DLPFC extending into orbitofrontal cortex (OFC), left medial
superior frontal gyrus, and dorsal ACC. Activation was also observed in the right frontal, temporal, and pre-
motor cortex, as well as in bilateral occipital lobes and the cerebellum. These results revealed activation in a
core network of brain systems involved in the production and comprehension of language, including Broca’s
area, Wernicke’s area, and the left superior temporal cortex and the temporal pole. The temporal lobes also
play an important role in mnemonic processing, a relevant process in the production of novel literary con-
tent. Second, creative writing (minus rest) activated motor-associated areas including the primary motor
cortex, somatosensory areas, bilateral secondary motor areas, middle cingulum, bilateral superior parietal
lobules, the left thalamus, and cerebellum, with additional activations observed in the frontal, temporal, and
occipital lobes. This set of activations draws heavily on motor-associated areas, including the primary motor
cortex, somatosensory areas, and bilateral secondary motor areas, all of which would be expected to be
involved in the execution of writing. Third, creative writing (minus copying) revealed activation in bilateral
hippocampi, temporal pole, and PCC. Note that there is remarkable overlap between the results here and
those reported by Ellamil et al. (2012) when comparing generation to evaluation in the context of creative
drawing. As before, we observe activation in MTL, reinforcing the role of areas associated with episodic and
semantic memory in the generation of novel thought content. Furthermore, we see activation in temporal
lobe regions associated with linguistic processing, as well as the engagement of PCC—a key hub in the
default-mode network. As noted by Shah et al., it is known that the PCC has connections with both the
MTL as well as the temporal pole and that the coactivation of this network that draws on language, memory
and internally oriented cognition would appear to subserve many of the cognitive and affective functions
necessary for the generation of creative stories.
Finally, Shah and colleagues rated the output of their participants based on two indices. First, post-scan,
independent judges rated the creativity of each generated text. Second, participants completed a test to assess
their verbal fluency and production skills. The authors computed the correlation between brain activation

516
Neuroscience of artistic creativity

(derived from the creative writing–copying contrast) and each of those two indices separately. Whereas there
was no correlation between the judged creativity of each generated piece and brain activation, the results
revealed positive correlations between verbal fluency and production skills and activation in the left IFG and
superior temporal gyrus. Both regions are heavily implicated in linguistic and mnemonic processing, and
their correlations with verbal fluency and production skills is consistent with their contribution to those
abilities.
How do Shah et al.’s results shed light on Flower and Hayes’ (1981) model? Recall that according to that
model of creative writing, planning involves generating ideas based on the contents of long-term memory.
The main effect of brainstorming would appear to support this notion, given that it activated a core set of
structures involved in language and memory. In turn, translating was hypothesized to involve transcribing
those ideas into text based on linguistic rules, which appears to best reflect the observed activations in motor
areas during creative writing. Reviewing may be reflected in frontal activations observed during both brain-
storming and creative writing conditions. Overall, the results clearly support the engagement of domain-
general (i.e., default-mode, executive control, and salience network) and domain-specific (i.e., language,
motor) components in creative writing.
Although the results discussed previously are consistent with the engagement of domain-general and
domain-specific structures in jazz improvisation (Limb  & Braun, 2008), creative drawing (Ellamil et  al.,
2012), and creative writing (Shah et al., 2013), none of those early studies examined functional connectivity
during artistic creativity. Since then, a number of studies have tackled that important issue focusing on artis-
tic creativity. Liu et al. (2015) investigated the neutral correlates of spontaneous poetry composition in two
group: experts who had completed at least one year of a master of fine arts program and published in poetry
journals and novices who had no formal training or experience in poetry composition. In the fMRI scanner,
the participants were instructed to generate new poetry in one condition and to revise their self-generated
poems in another condition. In turn, in the course of functional network connectivity analysis, independent
component analysis (ICA) was used to derive 53 spatially independent components which were subsequently
organized into five clusters on the basis of statistical similarities in their temporal dynamics. Of those five
clusters, one included default-mode network regions (e.g., medial PFC), whereas another included regions
within the executive control network (e.g., DLPFC). The results demonstrated that medial PFC was active
during both the generation and revision phases, whereas the engagement of DLPFC and inferior parietal sul-
cus (another region within the executive control network) was phase dependent: Experts exhibited stronger
deactivation of DLPFC and inferior parietal sulcus during generation than did novices, although the same
pattern of activations, deactivations, and connectivity was largely observed in experts and novices alike. Note
that this pattern of deactivation in the executive control network is strikingly similar to what was observed
for jazz improvisation (Limb & Braun, 2008) and suggests a relaxation of cognitive control processes during
spontaneous generation of novel thought content. In turn, quality of poetry, assessed by an independent panel,
was associated with divergent patterns of connectivity in experts and novices, centred upon medial PFC (for
technical facility) and DLPFC and inferior parietal sulcus (for innovation), suggesting that the functional con-
nectivity between default-mode and executive control networks plays a key role in distinguishing the neural
architecture of creativity between expert (i.e., good) and novice (i.e., poor) performers (see also Beaty et al.,
2014, 2018). As noted by the authors, this set of results

can be understood in the context of a single neurocognitive model characterized by dynamic inter-
actions between medial prefrontal areas regulating motivation, dorsolateral prefrontal, and parietal
areas regulating cognitive control and the association of these regions with language, sensorimotor,
limbic, and subcortical areas distributed throughout the brain.
(p. 3351)

517
Oshin Vartanian

Similar results were obtained by Pinho et al. (2016), who instructed a sample of pianists with extensive
experience in classical or jazz piano playing to perform improvisations using a keyboard in the fMRI scan-
ner under two conditions: In one condition, they were instructed to express a certain emotional content
(e.g., happy, fearful) in their improvisations, whereas in the other condition, they were asked to use a specific
set of piano keys (i.e., tonal/atonal pitch sets). Their results demonstrated that compared to the pitch-set
condition, there was lower activity in right DLPFC, dorsal premotor, and inferior parietal cortex in the emo-
tional expression condition. This pattern is consistent with the idea that relative to the execution of pitch sets,
there is a reduction in cognitive control and motor control requirements when expressing emotions through
music. Furthermore, and notable from a functional connectivity perspective, DLPFC exhibited functional
connectivity with the default-mode network during emotional expression but with the premotor network in
the pitch-set condition. At face value, these results reveal that there are two dissociable strategies that support
musical creativity during improvisation: an interoceptive strategy involving DLPFC and the default-mode
network that underlies emotional expression and an exteroceptive strategy involving DLPFC and the pre-
motor network that underlies the execution of motor commands. More notably, the results suggest that the
exhibition of creative behaviour in the arts can be parsed into subtasks, each of which would likely be sup-
ported by a dynamic reconfiguration of supporting structures and networks in the service of task demands.
In summary, a review of the literature on artistic creativity suggests that its neural bases share many
similarities with the neural bases of creativity in other domains, especially insofar as they involve a dynamic
interaction between the default-mode and executive control networks. At the same time, it is also true that
artistic creativity involves the recruitment of structures that support domain-specific abilities and capacities,
such as language in creative writing, motor control in musical improvisation, and regions that support visual
processing in creative drawing, among others. Indeed, a recent meta-analysis that analyzed the available
neuroscientific data across domains reached a largely similar conclusion. Specifically, Chen et al. (2020) con-
ducted a meta-analysis of the neuroimaging literature for three forms of artistic creativity: musical improvisa-
tion, drawing, and literary creativity. Across all domains, creativity engaged left DLPFC, right IFG, and the
presupplementary motor area (pre-SMA). Although this overarching analysis did not reveal common nodes
within the default-mode and salience networks that were activated across all studies, many individual studies
did reveal the engagement of various but non-overlapping nodes within those networks in their analyses.
Then, broken down by domain, musical creativity was associated with the engagement of SMA, bilateral
IFG, left precentral gyrus, and left middle frontal gyrus. Creative drawing was associated with the engage-
ment of left fusiform gyrus, left precuneus, right parahippocampal gyrus, and right middle frontal gyrus. In
turn, literary creativity was associated with the engagement of left angular gyrus and right lingual gyrus. This
meta-analysis revealed quantitatively that there are both domain-general as well as domain-specific structures
that support artistic creativity.

Next steps
I will end this chapter by discussing two important questions in the neuroscientific study of artistic creativity
that the field is well poised to target in the near future. First, one of the oldest questions in creativity research
involves whether there are foundational differences in the cognitive and neural architectures supporting eve-
ryday (i.e., little-c) vs. extraordinary (Big-C) creativity (Kaufman & Beghetto, 2009). There are some data
to suggest that there might well be some differences in basic physiological processes associated with variable
expressions of creativity, for example, in relation to when creativity is measured via laboratory tasks vs. real-
world accomplishments. Zabelina et al. (2015) examined the relationship between individual differences in
divergent thinking performance as measured by the Abbreviated Torrance Test for Adults (ATTA: Goff &
Torrance, 2002), real-world creativity as measured by the Creative Achievement Questionnaire (CAQ; Car-
son et  al., 2005), and sensory gating as measured by the P50. The P50 is a form of neurophysiological

518
Neuroscience of artistic creativity

response (i.e., evoked potential) that occurs 50 ms  after stimulus onset. It is considered a very early and
automatic form of sensory gating, influencing which stimuli receive attention. Everything else being equal,
the P50 response tends to be attenuated upon the repeated presentation of the same stimulus due to latent
inhibition. Zabelina and colleagues demonstrated that whereas divergent thinking was associated with selec-
tive sensory gating, real-world creativity was associated with “leaky” sensory gating. In other words, the
P50 response was attenuated less upon repeated presentation of the same stimulus in people with high CAQ
scores. The authors concluded that

leaky sensory gating may help people integrate ideas that are outside of focus of attention, leading
to creativity in the real world; whereas divergent thinking, measured by divergent thinking tests
which emphasize numerous responses within a limited time, may require selective sensory process-
ing more than previously thought.
(p. 77)

This work suggests that varying forms of creativity might be associated with different patterns of basic
physiological responding, suggesting that little-c and Big-C creativity might be supported by different neural
bases.
Toward that end, there is now a small but important body of work examining the neural bases of little-c
vs. Big-C creativity. Japardi et al. (2018) examined the neural correlates of divergent and convergent think-
ing in a sample of high-level creative achievers who included internationally acclaimed creative achievers in
multiple disciplines spanning the visual arts (e.g., painting, drawing, sculpture, photography, graphic design,
and animation) and the sciences (e.g., biology, neuroscience, chemistry, and mathematics). Their results
demonstrated that compared to the control group, the Big-C group had less activation in frontal pole, right
frontal operculum, left middle frontal gyrus, and bilaterally in the occipital cortex during divergent think-
ing, suggesting that exceptionally creative people might depend less on brain regions that typically support
effortful cognition while they are thinking creatively. Overall, however, there was remarkable overlap in the
neural systems that support creativity in both groups (see also Chrysikou et al., 2020a). In turn, Chrysikou
et al. (2020b) examined associations between brain structure (cortical area and thickness) and creative think-
ing measures in high creative achievers selected from diverse fields of expertise vs. a “smart” comparison
group of age-, intelligence-, and education-matched average creative achievers. Their results demonstrated
that high and average creative achievers could be distinguished based on the relationship between posterior
parietal cortex morphometry and creativity, suggesting that creativity may be associated with measurable
structural brain differences within the parietal cortex. Simonton (1999, 2014) has argued that to truly under-
stand the workings of the mind in highly creative people, it is necessary to study such “significant samples”
that consist of highly creative people directly. Given scientific interest in understanding high-level artistic
creativity, coupled with the fact that functional connectivity can be used to assess the quality of creative ideas
(Beaty et al., 2018), it would seem that furthering our understanding of the neural basis of artistic eminence
is within grasp.
Second, unlike what is the case for intelligence, we know little about the relationship between creativity
and the trajectory of change in brain structure from a developmental perspective. For example, using a lon-
gitudinal design, Shaw et al. (2006) were able to show that in adults who exhibit superior intelligence there
was a predominantly negative correlation between intelligence and cortical thickness in early childhood,
which shifted to a positive correlation between intelligence and cortical thickness in late childhood and
beyond. Importantly, this developmental shift was particularly salient in the frontal lobes, which are impli-
cated heavily in intelligent behaviour in adults. Thus, the superior intelligence group had a thinner prefrontal
cortex earlier in development, but this was followed by a rapid increase in cortical thickness in the prefrontal
cortex, peaking around age 13 and waning in late adolescence. As noted by the authors, “ ‘Brainy’ children

519
Oshin Vartanian

are not cleverer solely by virtue of having more or less grey matter at any one age. Rather, intelligence is
related to dynamic properties of cortical maturation” (p. 678). Needless to say, such longitudinal data have
been invaluable in expanding our understanding of the developmental bases of the construct of intelligence.
Along similar lines, not only do we know that neuronal synaptic density changes in the course of develop-
ment (Corel, 1975) but also that brain regions with different functions appear to develop at different rates.
For example, sensory regions develop earlier than language regions, and those in turn develop earlier than
regions of the PFC that underlie higher-order cognition and executive functions (e.g., Huttenlocher, 1979;
Huttenlocher  & Dabholkar, 1997; see also Shonkoff  & Phillips, 2000). Given that synaptic development
occurs at different rates for brain regions involved in different functions (e.g., sensory, linguistic, executive),
one could probe whether creativity in different domains (e.g., visual arts vs. literature vs. science) exhibits
dissociable developmental trajectories, including variations in when and why it is most likely to peak. Such
data would be extremely useful not only for exploring what differentiates artistic creativity from creativity in
other domains but also to offer an explanatory window into its emergence.

Summary
Compared to the state of the literature at the turn of this century (see Martindale, 1999), it would be fair to
say that there has been noticeable progress in our understanding of the neural bases of creativity, including
artistic creativity, over the last two decades (for reviews, see Abraham, 2018; Jung & Vartanian, 2018). We
have a basic understanding of the key large-scale networks that are involved in creativity (Beaty et al., 2016),
ways to measure how functional connectivity predicts the quality of creative thought (Beaty et al., 2018),
and a sense of the causal pathways that determine the direction of control among brain regions that support
it (Vartanian et al., 2018). Perhaps the best indicator of progress in the domain is that we are at a point where
we can ask and test interesting questions, aided in part by computational networks science tools that can shed
quantitative light on the structure of cognition that support creative cognition (see Kenett & Faust, 2019). It
is hoped that this trajectory will lead to a growing understanding of the expression of exceptional creativity
in the arts, as well as the developmental pathways that support it.

References
Abraham, A. (2018). The neuroscience of creativity. Cambridge University Press.
Addis, D. R., Wong, A. T., & Schacter, D. L. (2007). Remembering the past and imagining the future: Common and
distinct neural substrates during event construction and elaboration. Neuropsychologia, 45(7), 1363–1377. https://doi.
org/10.1016/j.neuropsychologia.2006.10.016
Aminoff, E., Gronau, N., & Bar, M. (2007). The parahippocampal cortex mediates spatial and nonspatial associations.
Cerebral Cortex, 17(7), 1493–1503. https://doi.org/10.1093/cercor/bhl078
Arden, R., Chavez, R. S., Grazioplene, R., & Jung, R. E. (2010). Neuroimaging creativity: A psychometric view. Behav-
ioural Brain Research, 214(2), 143–156. https://doi.org/10.1016/j.bbr.2010.05.015
Aron, A. R., Fletcher, P. C., Bullmore, E. T., Sahakian, B. J., & Robbins, T. W. (2003). Stop-signal inhibition disrupted
by damage to right inferior frontal gyrus in humans. Nature Neuroscience, 6(2), 115–116. https://doi.org/10.1038/
nn1003
Aron, A. R., Robbins, T. W., & Poldrack, R. A. (2004). Inhibition and the right inferior frontal cortex. Trends in Cogni-
tive Sciences, 8(4), 170–177. https://doi.org/10.1016/j.tics.2004.02.010
Baddeley, A. (2003). Working memory: Looking back and looking forward. Nature Reviews. Neuroscience, 4(10), 829–839.
https://doi.org/10.1038/nrn1201
Badre, D., Poldrack, R. A., Paré-Blagoev, E. J., Insler, R. Z., & Wagner, A. D. (2005). Dissociable controlled retrieval
and generalized selection mechanisms in ventrolateral prefrontal cortex. Neuron, 47(6), 907–918. https://doi.
org/10.1016/j.neuron.2005.07.023
Badre, D., & Wagner, A. D. (2007). Left ventrolateral prefrontal cortex and the cognitive control of memory. Neuropsy-
chologia, 45(13), 2883–2901. https://doi.org/10.1016/j.neuropsychologia.2007.06.015

520
Neuroscience of artistic creativity

Baer, J. (1998). The case for domain specificity in creativity. Creativity Research Journal, 11(2), 173–177. https://doi.
org/10.1207/s15326934crj1102_7
Baer, J. (2011). Domains of creativity. In M. Runco & S. Pritzker (Eds.), Encyclopedia of creativity (2nd ed., pp. 404–408).
Academic Press.
Baer, J., & Kaufman, J. C. (2005). Bridging generality and specificity: The amusement park theoretical (APT) model of
creativity. Roeper Review, 27(3), 158–163. https://doi.org/10.1080/02783190509554310
Bar, M., Aminoff, E., & Schacter, D. L. (2008). Scenes unseen: The parahippocampal cortex intrinsically subserves con-
textual associations, not scenes or places per se. Journal of Neuroscience, 28(34), 8539–8544. https://doi.org/10.1523/
JNEUROSCI.0987-08.2008
Beaty, R. E., Benedek, M., Barry Kaufman, S. B., & Silvia, P. J. (2015). Default and executive network coupling supports
creative idea production. Scientific Reports, 5(1), article 10964. https://doi.org/10.1038/srep10964
Beaty, R. E., Benedek, M., Silvia, P. J., & Schacter, D. L. (2016). Creative cognition and brain network dynamics. Trends
in Cognitive Sciences, 20(2), 87–95. https://doi.org/10.1016/j.tics.2015.10.004
Beaty, R. E., Benedek, M., Wilkins, R. W., Jauk, E., Fink, A., Silvia, P. J., Hodges, D. A., Koschutnig, K., & Neubauer,
A. C. (2014). Creativity and the default network: A functional connectivity analysis of the creative brain at rest. Neu-
ropsychologia, 64, 92–98. https://doi.org/10.1016/j.neuropsychologia.2014.09.019
Beaty, R. E., Kenett, Y. N., Christensen, A. P., Rosenberg, M. D., Benedek, M., Chen, Q., Fink, A., Qiu, J., Kwapil,
T. R., Kane, M. J., & Silvia, P. J. (2018). Robust prediction of individual creative ability from brain functional con-
nectivity. Proceedings of the National Academy of Sciences of the United States of America, 115(5), 1087–1092. https://doi.
org/10.1073/pnas.1713532115
Bekhtereva, N. P., Starchenko, M. G., Klyucharev, V. A., Vorob’ev, V. A., Pakhomov, S. V., & Medvedev, S. V. (2000).
Study of the brain organization of creativity: II. Positron-emission tomography data. Human Physiology, 26(5), 516–
522. https://doi.org/10.1007/BF02760367
Bengtsson, S. L., Csíkszentmihályi, M., & Ullén, F. (2007). Cortical regions involved in the generation of musical struc-
tures during improvisation in pianists. Journal of Cognitive Neuroscience, 19(5), 830–842. https://doi.org/10.1162/
jocn.2007.19.5.830
Berkowitz, A. L., & Ansari, D. (2008). Generation of novel motor sequences: The neural correlates of musical improvisa-
tion. NeuroImage, 41(2), 535–543. https://doi.org/10.1016/j.neuroimage.2008.02.028
Boccia, M., Piccardi, L., Palermo, L., Nori, R., & Palmiero, M. (2015). Where do bright ideas occur in our brain? Meta-
analytic evidence from neuroimaging studies of domain-specific creativity. Frontiers in Psychology, 6, 1195. https://doi.
org/10.3389/fpsyg.2015.01195
Botzung, A., Denkova, E., & Manning, L. (2008). Experiencing past and future personal events: Functional neuroimaging
evidence on the neural bases of mental time travel. Brain and Cognition, 66(2), 202–212. https://doi.org/10.1016/j.
bandc.2007.07.011
Bressler, S. L., & Menon, V. (2010). Large-scale brain networks in cognition: Emerging methods and principles. Trends
in Cognitive Sciences, 14(6), 277–290. https://doi.org/10.1016/j.tics.2010.04.004
Brown, S., Martinez, M. J.,  & Parsons, L. M. (2006). Music and language side by side in the brain: A  PET study
of the generation of melodies and sentences. European Journal of Neuroscience, 23(10), 2791–2803. https://doi.
org/10.1111/j.1460-9568.2006.04785.x
Bunge, S. A., Helskog, E. H.,  & Wendelken, C. (2009). Left, but not right, rostrolateral prefrontal cortex meets a
stringent test of the relational integration hypothesis. NeuroImage, 46(1), 338–342. https://doi.org/10.1016/j.
neuroimage.2009.01.064
Campbell, D. T. (1960). Blind variation and selective retention in creative thought as in other knowledge processes.
Psychological Review, 67, 380–400. https://doi.org/10.1037/h0040373
Carlsson, I., Wendt, P. E., & Risberg, J. (2000). On the neurobiology of creativity. Differences in frontal activity between
high and low creative subjects. Neuropsychologia, 38(6), 873–885. https://doi.org/10.1016/s0028-3932(99)00128-1
Carson, S. H., Peterson, J. B., & Higgins, D. M. (2005). Reliability, validity, and factor structure of the creative achieve-
ment questionnaire. Creativity Research Journal, 17(1), 37–50. https://doi.org/10.1207/s15326934crj1701_4
Carter, C. S., Mintun, M., Nichols, T., & Cohen, J. D. (1997). Anterior cingulate gyrus dysfunction and selective atten-
tion deficits in schizophrenia: [15O]H2O PET study during single-trial Stroop task performance. American Journal of
Psychiatry, 154, 1670–1675. https://doi.org/10.1176/ajp.154.12.1670
Cattell, R. B., & Cattell, A. K. S. (1959). Handbook for the culture fair intelligence test. Institute for Personality and Ability
Testing.
Chen, Q., Beaty, R. E., & Qiu, J. (2020). Mapping the artistic brain: Common and distinct neural activations associated
with musical, drawing, and literary creativity. Human Brain Mapping, 41(12), 3403–3419. https://doi.org/10.1002/
hbm.25025

521
Oshin Vartanian

Christensen, P. R., Guilford, J. P., Merrifield, P. R., & Wilson, R. C. (1960). Alternate uses. Sheridan Press Psychological
Service.
Christoff, K., Irving, Z. C., Fox, K. C., Spreng, R. N., & Andrews-Hanna, J. R. (2016). Mind-wandering as spon-
taneous thought: A  dynamic framework.  Nature Reviews. Neuroscience,  17(11), 718–731. https://doi.org/10.1038/
nrn.2016.113
Christoff, K., Prabhakaran, V., Dorfman, J., Zhao, Z., Kroger, J. K., Holyoak, K. J., & Gabrieli, J. D. E. (2001). Rostrolat-
eral prefrontal cortex involvement in relational integration during reasoning. NeuroImage, 14(5), 1136–1149. https://
doi.org/10.1006/nimg.2001.0922
Chrysikou, E. G., Jacial, C., Yaden, D. B., van Dam, W., Kaufman, S. B., Conklin, C. J., Wintering, N. A., Abra-
ham, R. E., Jung, R. E.,  & Newberg, A. B. (2020a). Differences in brain activity patterns during creative idea
generation between eminent and non-eminent thinkers.  NeuroImage,  220, 117011. https://doi.org/10.1016/j.
neuroimage.2020.117011
Chrysikou, E. G., Wertz, C., Yaden, D. B., Kaufman, S. B., Bacon, D., Wintering, N. A., Jung, R. E., & Newberg,
A. B. (2020b). Differences in brain morphometry associated with creative performance in high- and average-creative
achievers. NeuroImage, 218, 116921. https://doi.org/10.1016/j.neuroimage.2020.116921
Cocchi, L., Zalesky, A., Fornito, A.,  & Mattingley, J. B. (2013). Dynamic cooperation and competition between
brain systems during cognitive control. Trends in Cognitive Sciences, 17(10), 493–501. https://doi.org/10.1016/j.
tics.2013.08.006
Corel, J. L. (1975). The postnatal development of the human cerebral cortex. Harvard University Press.
Crone, E. A., Wendelken, C., van Leijenhorst, L., Honomichl, R. D., Christoff, K.,  & Bunge, S. A. (2009).
Neurocognitive development of relational reasoning. Developmental Science, 12(1), 55–66. https://doi.
org/10.1111/j.1467-7687.2008.00743.x
Diedrich, J., Jauk, E., Silvia, P. J., Gredlein, J. M., Neubauer, A. C., & Benedek, M. (2018). Assessment of real-life crea-
tivity: The Inventory of Creative Activities and Achievements (ICAA). Psychology of Aesthetics, Creativity, and the Arts,
12(3), 304–316.  https://doi.org/10.1037/aca0000137
Dorfman, L., Martindale, C., Gassimova, V., & Vartanian, O. (2008). Creativity and speed of information processing:
A  double dissociation involving elementary versus inhibitory cognitive tasks. Personality and Individual Differences,
44(6), 1382–1390. https://doi.org/10.1016/j.paid.2007.12.006
Ellamil, M., Dobson, C., Beeman, M., & Christoff, K. (2012). Evaluative and generative modes of thought during the
creative process. Neuroimage, 59(2), 1783–1794. https://doi.org/10.1016/j.neuroimage.2011.08.008
Ferstl, E. C., Neumann, J., Bogler, C., & von Cramon, D. Y. (2008). The extended language network: A meta-analysis
of neuroimaging studies on text comprehension. Human Brain Mapping, 29(5), 581–593. https://doi.org/10.1002/
hbm.20422
Flower, L. S., & Hayes, J. R. (1981). A cognitive process theory of writing. College Composition and Communication, 32(4),
365–387. https://doi.org/10.2307/356600
Gabora, L. (2010). Revenge of the “neurds”: Characterizing creative thought in terms of the structure and dynamics of
memory. Creativity Research Journal, 22(1), 1–13. https://doi.org/10.1080/10400410903579494
Galton, F. (1869). Hereditary genius. Friedmann.
Galton, F. (1874). English men of science: Their nature and nurture. Macmillan, and Co.
Goel, V. (2007). Anatomy of deductive reasoning. Trends in Cognitive Sciences, 11(10), 435–441. https://doi.org/10.1016/j.
tics.2007.09.003
Goff, K., & Torrance, E. P. (2002). Abbreviated Torrance Test for Adults manual. Bensenvill. Scholastic Testing Service.
Gonen-Yaacovi, G., de Souza, L. C., Levy, R., Urbanski, M., Josse, G., & Volle, E. (2013). Rostral and caudal prefrontal
contribution to creativity: A meta-analysis of functional imaging data. Frontiers in Human Neuroscience, 7, article 465.
https://doi.org/10.3389/fnhum.2013.00465
Grice, H. P. (1975). Logic and conversation. In P. Cole & J. Morgan (Eds.), Syntax and semantics: Speech acts (Vol. 3,
pp. 41–58). Academic Press.
Hassabis, D., Kumaran, D.,  & Maguire, E. A. (2007). Using imagination to understand the neural basis of episodic
memory. Journal of Neuroscience, 27(52), 14365–14374. https://doi.org/10.1523/JNEUROSCI.4549-07.2007
Henke, K., Buck, A., Weber, B.,  & Wieser, H. G. (1997). Human hippocampus establishes associations in memory.
Hippocampus, 7(3), 249–256. https://doi.org/10.1002/(SICI)1098-1063(1997)7:3<249::AID-HIPO1>3.0.CO;2-G
Henke, K., Weber, B., Kneifel, S., Wieser, H. G., & Buck, A. (1999). Human hippocampus associates information in
memory. Proceedings of the National Academy of Sciences of the United States of America, 96(10), 5884–5889. https://doi.
org/10.1073/pnas.96.10.5884
Hospers, J. (1985). Artistic creativity. Journal of Aesthetics and Art Criticism, 43(3), 243–256. https://doi.org/
10.1111/1540_6245.jaac43.3.0243

522
Neuroscience of artistic creativity

Huttenlocher, P. R. (1979). Synaptic density in human frontal cortex—Developmental changes and effects of aging. Brain
Research, 163(2), 195–205. https://doi.org/10.1016/0006-8993(79)90349-4
Huttenlocher, P. R.,  & Dabholkar, A. S. (1997). Regional differences in synaptogenesis in human cerebral cortex.
Journal of Comparative Neurology, 387(2), 167–178. https://doi.org/10.1002/(sici)1096-9861(19971020)387:2<167::
aid-cne1>3.0.co;2-z
Japardi, K., Bookheimer, S., Knudsen, K., Ghahremani, D. G., & Bilder, R. M. (2018). Functional magnetic resonance
imaging of divergent and convergent thinking in Big-C creativity.  Neuropsychologia,  118(A), 59–67. https://doi.
org/10.1016/j.neuropsychologia.2018.02.017
Jauk, E., Benedek, M., & Neubauer, A. C. (2014). The road to creative achievement: A latent variable model of ability
and personality predictors. European Journal of Personality, 28(1), 95–105. https://doi.org/10.1002/per.1941
Jung, R. E., & Vartanian, O. (Eds.). (2018). The Cambridge handbook of the neuroscience of creativity. Cambridge University
Press.
Kaufman, J. C., & Baer, J. (Eds.). (2005). Creativity across domains: Faces of the muse. Lawrence Erlbaum.
Kaufman, J. C., & Beghetto, R. A. (2009). Beyond big and little: The four C model of creativity. Review of General Psy-
chology, 13(1), 1–12. https://doi.org/10.1037/a0013688
Kenett, Y. N., Anaki, D., & Faust, M. (2014). Investigating the structure of semantic networks in low and high creative
persons. Frontiers in Human Neuroscience, 8, article 407. https://doi.org/10.3389/fnhum.2014.00407
Kenett, Y. N.,  & Faust, M. (2019). A  semantic network cartography of the creative mind.  Trends in Cognitive Sci-
ences, 23(4), 271–274. https://doi.org/10.1016/j.tics.2019.01.007
Lee, C. S., Huggins, A. C., & Therriault, D. J. (2014). A measure of creativity or intelligence? Examining internal and
external structure validity evidence of the remote associates test. Psychology of Aesthetics, Creativity, and the Arts, 8(4),
446–460.  https://doi.org/10.1037/a0036773
Limb, C. J., & Braun, A. R. (2008). Neural substrates of spontaneous musical performance: An FMRI study of jazz
improvisation. PLoS One, 3(2), article e1679. https://doi.org/10.1371/journal.pone.0001679
Liu, S., Erkkinen, M. G., Healey, M. L., Xu, Y., Swett, K. E., Chow, H. M., & Braun, A. R. (2015). Brain activity and
connectivity during poetry composition: Toward a multidimensional model of the creative process. Human Brain
Mapping, 36(9), 3351–3372. https://doi.org/10.1002/hbm.22849
Lykken, D. T., McGue, M., Tellegen, A., & Bouchard, T. J., Jr. (1992). Emergenesis: Genetic traits that may not run in
families. American Psychologist, 47(12), 1565–1577. https://doi.org/10.1037//0003-066x.47.12.1565
Martindale, C. (1990). Creative imagination and neural activity. In R. Kunzendorf & A. Sheikh (Eds.), The psychophysiol-
ogy of mental imagery (pp. 89–108). Baywood Publishing.
Martindale, C. (1999). Biological bases of creativity. In R. J. Sternberg (Ed.), Handbook of creativity (pp. 137–152). Cam-
bridge University Press.
Martindale, C., & Armstrong, J. (1974). The relationship of creativity to cortical activation and its operant control. Journal
of Genetic Psychology, 124, 311–320. https://doi.org/ 10.1080/00221325.1974.10532293
Martindale, C., & Hines, D. (1975). Creativity and cortical activation during creative, intellectual and EEG feedback
tasks. Biological Psychology, 3(2), 91–100. https://doi.org/10.1016/0301-0511(75)90011-3
Mashal, N., Faust, M., Hendler, T., & Jung-Beeman, M. (2007). An fMRI investigation of the neural correlates underly-
ing the processing of novel metaphorical expressions. Brain and Language, 100(2), 115–126. https://doi.org/10.1016/j.
bandl.2005.10.005
Mednick, S. A. (1962). The associative basis of the creative process. Psychological Review, 69, 220–232. https://doi.
org/10.1037/h0048850
Mednick, S. A., & Mednick, M. T. (1967). Remote associates T&t. Examiners manual. Houghton Mifflin.
Menon, V., & Uddin, L. Q. (2010). Saliency, switching, attention and control: A network model of insula function. Brain
Structure and Function, 214(5–6), 655–667. https://doi.org/10.1007/s00429-010-0262-0
Milivojevic, B., Hamm, J. P., & Corballis, M. C. (2009). Functional neuroanatomy of mental rotation. Journal of Cognitive
Neuroscience, 21(5), 945–959. https://doi.org/10.1162/jocn.2009.21085
Moss, H. E., Abdallah, S., Fletcher, P., Bright, P., Pilgrim, L., Acres, K., & Tyler, L. K. (2005). Selecting among compet-
ing alternatives: Selection and retrieval in the left inferior frontal gyrus. Cerebral Cortex, 15(11), 1723–1735. https://
doi.org/10.1093/cercor/bhi049
Okuda, J., Fujii, T., Ohtake, H., Tsukiura, T., Tanji, K., Suzuki, K., Kawashima, R., Fukuda, H., Itoh, M., & Yamadori,
A. (2003). Thinking of the future and past: The roles of the frontal pole and the medial temporal lobes. NeuroImage,
19(4), 1369–1380. https://doi.org/10.1016/s1053-8119(03)00179-4
Pinho, A. L., Ullén, F., Castelo-Branco, M., Fransson, P., & de Manzano, Ö. (2016). Addressing a paradox: Dual strategies
for creative performance in introspective and extrospective networks. Cerebral Cortex, 26(7), 3052–3063. https://doi.
org/10.1093/cercor/bhv130

523
Oshin Vartanian

Rombouts, S. A., Machielsen, W. C., Witter, M. P., Barkhof, F., Lindeboom, J., & Scheltens, P. (1997). Visual associa-
tion encoding activates the medial temporal lobe: A functional magnetic resonance imaging study. Hippocampus, 7(6),
594–601. https://doi.org/10.1002/(SICI)1098-1063(1997)7:6<594::AID-HIPO2>3.0.CO;2-F
Schacter, D. L., & Addis, D. R. (2009). On the nature of medial temporal lobe contributions to the constructive simu-
lation of future events. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 364(1521),
1245–1253. https://doi.org/10.1098/rstb.2008.0308
Seger, C. A., Desmond, J. E., Glover, G. H., & Gabrieli, J. D. (2000). Functional magnetic resonance imaging evidence
for right hemisphere involvement in processing unusual semantic relationships. Neuropsychology, 14(3), 361–369.
https://doi.org/10.1037//0894-4105.14.3.361
Shah, C., Erhard, K., Ortheil, H. J., Kaza, E., Kessler, C., & Lotze, M. (2013). Neural correlates of creative writing: An
fMRI study. Human Brain Mapping, 34(5), 1088–1101. https://doi.org/10.1002/hbm.21493
Shaw, P., Greenstein, D., Lerch, J., Clasen, L., Lenroot, R., Gogtay, N., Evans, A., Rapoport, J., & Giedd, J. (2006).
Intellectual ability and cortical development in children and adolescents. Nature, 440(7084), 676–679. https://doi.
org/10.1038/nature04513
Shonkoff, J. P., & Phillips, D. A. (Eds.). (2000). From neurons to neighborhoods: The science of early childhood development.
National Academy Press.
Simon, H. A. (1962). The architecture of complexity. Proceedings of the American Philosophical Society, 106, 467–482.
Simonton, D. K. (1994). Greatness: Who makes history and why. Guilford Press.
Simonton, D. K. (1999). Significant samples: The psychological study of eminent individuals. Psychological Methods, 4(4),
425–451. https://doi.org/10.1037/1082-989X.4.4.425
Simonton, D. K. (2001). The psychology of creativity: A historical perspective. Presentation given at the Green College Lecture
Series on The nature of creativity: History, biology, and socio-cultural dimensions. University of British Columbia.
Simonton, D. K. (2010). Creative thought as blind-variation and selective-retention: Combinatorial models of excep-
tional creativity. Physics of Life Reviews, 7(2), 156–179. https://doi.org/10.1016/j.plrev.2010.02.002
Simonton, D. K. (2014). Significant samples—Not significance tests! The often overlooked solution to the replication
problem. Psychology of Aesthetics, Creativity, and the Arts, 8(1), 11–12. https://doi.org/10.1037/a0035849
Sternberg, R. J. (1977). Component processes in analogical reasoning. Psychological Review, 84(4), 353–378. https://doi.
org/10.1037/0033-295X.84.4.353
Sternberg, R. J. (1980). Sketch of a componential subtheory of human intelligence. Behavioral and Brain Sciences, 3(4),
573–584. https://doi.org/10.1017/S0140525X00006932
Szpunar, K. K., Chan, J. C. K., & McDermott, K. B. (2009). Contextual processing in episodic future thought. Cerebral
Cortex, 19(7), 1539–1548. https://doi.org/10.1093/cercor/bhn191
Szpunar, K. K., Watson, J. M., & McDermott, K. B. (2007). Neural substrates of envisioning the future. Proceedings of the
National Academy of Sciences of the United States of America, 104(2), 642–647. https://doi.org/10.1073/pnas.0610082104
Thompson-Schill, S. L., D’Esposito, M., Aguirre, G. K., & Farah, M. J. (1997). Role of left inferior prefrontal cortex
in retrieval of semantic knowledge: A reevaluation. Proceedings of the National Academy of Sciences of the United States of
America, 94(26), 14792–14797. https://doi.org/10.1073/pnas.94.26.14792
Uddin, L. Q. (2015). Salience processing and insular cortical function and dysfunction. Nature Reviews. Neuroscience,
16(1), 55–61. https://doi.org/10.1038/nrn3857
Vartanian, O. (2009). Variable attention facilitates creative problem solving. Psychology of Aesthetics, Creativity, and the Arts,
3(1), 57–59. https://doi.org/10.1037/a0014781
Vartanian, O. (2011). Nature and nurture. In M. Runco & S. Pritzker (Eds.), Encyclopedia of creativity (2nd ed., pp. 175–
178). Academic Press.
Vartanian, O. (2012). Dissociable neural systems for analogy and metaphor: Implications for the neuroscience of creativ-
ity. British Journal of Psychology, 103(3), 302–316. https://doi.org/10.1111/j.2044-8295.2011.02073.x
Vartanian, O., Beatty, E. L., Smith, I., Blackler, K., Lam, Q., & Forbes, S. (2018). One-way traffic: The inferior frontal
gyrus controls brain activation in the middle temporal gyrus and inferior parietal lobule during divergent thinking.
Neuropsychologia, 118(A), 68–78. https://doi.org/10.1016/j.neuropsychologia.2018.02.024
Vartanian, O., Martindale, C., & Kwiatkowski, J. (2007). Creative potential, attention, and speed of information process-
ing. Personality and Individual Differences, 43(6), 1470–1480. https://doi.org/10.1016/j.paid.2007.04.027
Waller, N. G., Bouchard, T. J. Jr., Lykken, D. T., Tellegen, A., & Blacker, D. M. (1993). Creativity, heritability, familiality:
Which word does not belong? Psychological Inquiry, 4(3), 235–237. https://doi.org/10.1207/s15327965pli0403_18
Waltz, J. A., Knowlton, B. J., Holyoak, K. J., Boone, K. B., Mishkin, F. S., de Menezes Santos, M., Thomas, C. R., &
Miller, B. L. (1999). A system for relational reasoning in human prefrontal cortex. Psychological Science, 10(2), 119–
125. https://doi.org/10.1111/1467-9280.00118

524
Neuroscience of artistic creativity

Weisberg, R. W. (2004). On structure in the creative process: A quantitative case-study of the creation of Picasso’s Guer-
nica. Empirical Studies of the Arts, 22(1), 23–54. https://doi.org/10.2190/EH48-K59C-DFRB-LXE7
Zabelina, D. L., & Andrews-Hanna, J. R. (2016). Dynamic network interactions supporting internally oriented cogni-
tion. Current Opinion in Neurobiology, 40, 86–93. https://doi.org/10.1016/j.conb.2016.06.014
Zabelina, D. L., O’Leary, D., Pornpattananangkul, N., Nusslock, R., & Beeman, M. (2015). Creativity and sensory gat-
ing indexed by the P50: Selective versus leaky sensory gating in divergent thinkers and creative achievers. Neuropsy-
chologia, 69, 77–84. https://doi.org/10.1016/j.neuropsychologia.2015.01.034
Zabelina, D. L., & Robinson, M. D. (2010). Creativity as flexible cognitive control. Psychology of Aesthetics, Creativity, and
the Arts, 4(3), 136–143. https://doi.org/10.1037/a0017379

525
28
EXPERTISE AND THE BRAIN OF
THE PERFORMING ARTIST
Fredrik Ullén

Music has become one of the most popular model domains in research on expertise and its brain mecha-
nisms. One reason for this is that a fairly large proportion of the population engages actively in music.
Accordingly, music is intrinsically interesting for many people, and there are many individuals with differ-
ent degrees of musical training from which one can recruit participants to scientific studies. Furthermore,
professional musicians typically start early and invest much time in goal-directed practice throughout life,
making them ideally suited for studies of neuroplasticity and phenomena such as sensitive periods for skill
acquisition. Finally, perceptual musical tasks, but also some forms of music performance (e.g., playing on a
piano keyboard), can be performed under relatively ecological conditions during brain scanning, making the
neural mechanisms of musical expert performance amenable to scientific study.
Consequently, the literature on musical expertise and its neural underpinnings has grown to a consid-
erable size during the last decades. The number of similar studies on other forms of performing art, such
as dance, is smaller, though interest is clearly growing rapidly, with many innovative studies published in
recent years (see Chapters 16 and 24). Research on different aspects of the brains of performing artists has
been reviewed previously on several occasions (see e.g., Münte et al., 2002; Olszewska et al., 2021; Ullén
et al., 2019; Zardi et al., 2021; Zatorre et al., 2013). Here, I will focus mainly on general conclusions from
work during the last decade, with a particular emphasis on the importance of gene-environment interplay
for expertise, an area where research on musical expertise has made considerable contributions to current
understanding.

Neuroanatomy of the performing artist


To date, more than 50 neuroimaging studies on the anatomical features of the brains of musicians have been
published. The literature largely consists of cross-sectional comparisons between groups or individuals with
different degrees of musical training and expertise, although there are also a few examples of longitudinal
studies of structural changes in the brain that accompany extensive periods (months–years) periods of musi-
cal training (see e.g., Habibi et al., 2018; Hyde et al., 2009). Overall, specific findings concerning individual
brain regions replicate relatively poorly across studies (Bermudez et al., 2009). The possible explanations for
this are well known: sample sizes are often relatively small, different studies use different inclusion criteria for
participants (musical instrument played, amount of musical training and experience, gender, etc.), and there
is methodological heterogeneity in terms of employed morphometrical techniques and statistical models

526 DOI: 10.4324/9781003008675-30


Expertise and the performing artist

(e.g., region-of-interest based or whole-brain analyses). Nevertheless, some interesting conclusions can be
drawn from the literature as a whole.
First of all, there is convincing evidence that brain regions involved in perception and performance of
music tend to be more well developed in musicians as a group than in people who are not actively involved
in music. One of the most consistent findings is that musicians have larger auditory cortical regions in
the superior temporal lobe. The first indication that this is the case came from work by one of the true
pioneers in the field, the German neurologist Siegmund Auerbach, who in the beginning of the 20th
century performed a series of post-mortem case studies of the brains of eminent musicians, including the
legendary pianist and conductor Hans von Bülow (for a review in English, see Meyer, 1977). Since then,
more well-developed auditory regions in musicians have been demonstrated in numerous magnetic reso-
nance imaging (MRI) studies using different morphometric techniques, including manual morphometry
(Schneider et  al., 2002, 2005), voxel-based morphometry (Bermudez et  al., 2009; Fauvel et  al., 2014;
Gaser & Schlaug, 2003; Vaquero et al., 2016), and estimations of cortical thickness (Bermudez et al., 2009;
de Manzano & Ullén, 2018; Karpati et al., 2017). The validity of these anatomical findings for musical
expertise is supported by studies finding correlations between auditory cortex morphology and perfor-
mance on musical auditory tasks (Hyde et al., 2009; Karpati et al., 2017; Schneider et al., 2002, 2005;
Seither-Preisler et al., 2014).
An interesting novel finding is that the auditory cortex of musicians differs from that of non-musicians
not only in measures of overall size but also in patterns of gyrification. Morphotypes with multiple gyri and
consequent multiplications of Heschl’s gyrus appear to be much more common in musicians (Benner et al.,
2017). The functional implications of these anatomical variants for musical expertise remain to be explored,
but initial experiments suggest that neurophysiological responses to tone stimuli differ between morphotypes
(Benner et al., 2017). Similarly, Bangert and Schlaug have reported differences between pianists and string
players in the gyral morphology of motor cortex (2006). These observations suggest that it may be important
for future studies of neuroanatomical correlates of expertise to consider macro-anatomical variants in addi-
tion to traditional measures of regional size.
Outside the auditory system, anatomical differences between musicians and non-musicians have, as one
may expect, frequently been found in motor and premotor areas, inferior frontal cortex, the basal ganglia,
and the cerebellum, that is to say in regions involved in auditory-motor integration and the learning, plan-
ning, execution, and online adjustment of sequential movements (Bermudez et al., 2009; de Manzano &
Ullén, 2018; Gaser & Schlaug, 2003; Hutchinson et al., 2003; James et al., 2014; Karpati et al., 2017; Slum-
ing et al., 2002; Vaquero et al., 2016). Several studies have reported correlations between anatomical features
of regions within the motor system, on the one hand, and measures of training or performance on motor
tasks, on the other hand, providing some further support for that the observed anatomical effects are relevant
for expertise (Amunts et al., 1997; Hutchinson et al., 2003; Hyde et al., 2009; Karpati et al., 2017; Sluming
et al., 2002). Some studies have also found associations between musical expertise and the regional anatomy
of the hippocampus (Groussard et  al., 2010, 2014; Vaquero et  al., 2016). These findings are potentially
interesting since the hippocampus, apart from its well-known involvement in episodic memory and spatial
cognition, has also been implicated in motor sequence learning ( Jacobacci et al., 2020; Schendan et al., 2003)
and expert working memory, that is, experts’ expanded working memory capacity specifically for stimuli
encountered in their domain of expertise (Guida et al., 2012; Ullén et al., 2016).
Several studies have also examined the structural correlates of musicianship in the brain’s white matter.
The corpus callosum—the major pathway connecting cortical regions of the two hemispheres, with an
important role in the bilateral coordination of movement ( Johansen-Berg et al., 2006)—has been shown in
numerous studies to have a larger cross-sectional area and more well-developed white matter microstructure
in musicians than in non-musicians (de Manzano & Ullén, 2018; Hyde et al., 2009; Lee et al., 2003; Öztürk
et al., 2002; Schlaffke et al., 2020; Schlaug et al., 1995; Steele et al., 2013). Some of these studies also report

527
Fredrik Ullén

significant correlations between corpus callosum structure and performance on rhythmic motor tasks (Hyde
et al., 2009; Steele et al., 2013), including drumming (Schlaffke et al., 2020). Interestingly, two longitudinal
studies on children have found structural changes in the corpus callosum to develop over time during 15–24
months of music training (Habibi et al., 2018; Hyde et al., 2009).
Two other white matter pathways that have been implicated repeatedly in musical expertise are the cor-
ticospinal tracts and the arcuate fasciculus. The corticospinal tracts, which constitute the major descending
pathway for the control of skilled hand and finger movements in humans (Armand et al., 1996), show a more
well-developed microstructure in musicians than in non-musicians (Bengtsson et al., 2005; de Manzano &
Ullén, 2018; Giacosa et al., 2016, 2019; Han et al., 2009; Rüber et al., 2015). The importance of these tracts
for musical expertise is further supported by demonstrations that corticospinal microstructure is correlated
with speed and accuracy of performance in hand motor tasks (Karahan et  al., 2019; Rüber et  al., 2015;
Steele et al., 2013). The arcuate fasciculus subserves communication between auditory regions in the supe-
rior temporal cortex, the temporo-parietal junction, and inferior parietal cortex and frontal cortical regions
(inferior frontal cortex, ventral premotor and prefrontal areas). It plays a key role for the processing of both
sounding music and spoken language (Hickok et al., 2003; Hickok & Poeppel, 2007; Vaquero et al., 2021).
The structure of this pathway has been related to musical expertise and training in several reports (Bengtsson
et al., 2005; de Manzano & Ullén, 2018; Halwani et al., 2011; Vaquero et al., 2020), and experimental stud-
ies have demonstrated associations between structural properties of the arcuate fasciculus and learning and
performance in musical tasks (Vaquero et al., 2018, 2021).
Apart from these differences in structure in specific pathways involved in musical processing, recent
findings indicate that musical expertise may also be associated with more global measures of connectivity.
Leipold and colleagues (2021) recently explored both structural white matter connectivity and resting state
functional connectivity in a sample of 103 musicians and 50 non-musicians. In line with the studies discussed
previously, significant relations were found between microstructural properties of the corpus callosum and
musicianship as well as age of onset of musical training. However, an additional finding was that age of onset
(but not musicianship as such) was correlated with betweenness-centrality, a graph-theoretical measure of
whole-brain structural network topology. The possible significance of this finding for musical expertise is
not clear, but the finding is interesting in that it suggests that future studies of neuroanatomical correlates of
expertise should pay attention not just to the structure of specific brain regions involved in the studied form
of expert performance but also to more global brain properties.
As mentioned at the outset, scientific work on the artistic brain has so far been mainly focused on musi-
cians, and the neuroanatomical literature on dancers remains comparatively small. A  reasonable a priori
expectation could be that the brains of dancers and musicians show both similarities and differences. On
the one hand, both music making and dance depend upon highly sophisticated sensorimotor integration,
where the performance of complex, precise movement sequences must be coordinated with ongoing musi-
cal stimuli. On the other hand, instrumental music performance and singing typically puts high demands on
the control of fine hand and finger movements and the vocal apparatus, whereas dance requires a high level
of skill in whole-body movements and postural control. Recent findings appear to be, at least in part, in
line with these predictions. Karpati and colleagues (2017) found that both dancers and musicians had higher
cortical thickness than controls in right temporal cortex, including auditory areas of the superior temporal
gyrus. Grey matter structure in this region was also correlated with performance on a dance imitation task as
well as a musical melody discrimination task. In contrast, musicians and dancers appear to show differences in
the white matter structure of the corticospinal tracts, which are revealed in measures of both regional volume
and microstructure (Giacosa et al., 2016, 2019; Hänggi et al., 2010). Interestingly, fine-grained analyses by
Giacosa and colleagues on corticospinal microstructure in these two expert groups revealed that dancers have
a greater radial diffusivity for the primary motor pathways for all body regions, whereas musicians showed
a higher anisotropy, suggesting increased fibre coherence, specifically in descending tracts to the right hand,

528
Expertise and the performing artist

arm, and trunk (Giacosa et al., 2019). Possibly, these group differences in white matter structure reflect the
different motor demands in the two forms of artistic expert performance.

Functional brain correlates of expertise in music and dance


It appears extremely plausible that many of the neuroanatomical differences between performing artists and
controls discussed in the previous section reflect functionally relevant differences in the organization of neural
circuitry, which allow information processing to be optimized for expert performance. Some support for this
is provided already by the already mentioned associations between structural properties of brain regions and
performance on tasks tapping into the studied form of expertise. Further insights into the functional brain
properties that allow high level performance in music and dance come from studies that analyse neural activ-
ity using functional neuroimaging or electrophysiological methods.
In general terms, long-term practice will result in a transition from controlled to automatic processing.
This means that performance will become less flexible but also less effortful, with a lower level of involve-
ment of neural systems subserving attention and working memory (Hill & Schneider, 2006). A surprising
finding in this regard is that even a highly complex and creative task such as musical improvisation appears
to be possible to perform with relatively little involvement of fronto-parietal executive circuits in musicians
with a high level of expertise in improvisation (Limb & Braun, 2008; Pinho et al., 2014, 2016). Possibly this
has the advantage that spontaneous generation of musical ideas can be driven by bottom-up processes with
little supervision from executive systems (Pinho et al., 2016). Recent studies have provided further support
for this notion. Rosen and colleagues thus found that transcranial direct current stimulation over prefrontal
cortex during improvisation had differential effects on performance quality in highly experienced and less
experienced improvisers (Rosen et al., 2016), while Vergara and colleagues analysed whole-brain functional
connectivity in skilled jazz improvisers and found weaker connectivity of prefrontal regions during improvi-
sation than during performance of pre-learned music (Vergara et al., 2021).
Both music and dance involve the action observation network, a set of brain areas that include the motor
and premotor areas, the superior and inferior parietal cortices, the inferior frontal cortex, and the cerebel-
lum, which are activated when an individual observes another person performing an action (Caspers et al.,
2010; Zardi et al., 2021). There is evidence for that the action observation network in musicians and danc-
ers, as well as in other expert groups such as athletes, has been tailored to process domain-specific actions
that the participant is familiar with (Haslinger et al., 2005; Liew et al., 2013; Zatorre et al., 2007). In an
influential early study, Calvo-Merino and colleagues compared brain responses to dance stimuli in dancers
specializing in either classical ballet or capoeira, as well as in a control group of non-experts (2005). They
found that activity in premotor and parietal cortical regions was higher specifically when dancers viewed
movements from dances that they had been trained to perform, as compared to dances they were unfa-
miliar with. Subsequent work on dance supports the notion that the action observation network plays an
important role for functions such as movement prediction (Gardner et al., 2015), and observational as well
as physical skill learning (Cross et al., 2009; for a review, see Zardi et al., 2021). In a similar vein, Haslinger
and colleagues found that pianists, as compared to non-musicians, displayed stronger activation of a cortical
fronto-parieto-temporal network, including auditory and premotor areas, during observation of silent piano
performance. Vice versa, piano sounds coupled with visually presented finger movements evoked increased
activation within sensorimotor areas in the pianists (Haslinger et al., 2005). Subsequent studies indicate that
such transmodal coupling can be established after relatively short periods of training. Lahav and co-workers
trained musically naïve participants to perform a melody on a keyboard for five days and found that after
training, listening to trained but not untrained music activated fronto-parietal regions, including the premo-
tor cortex and Broca’s area (Lahav et al., 2007). Recently, de Manzano and colleagues demonstrated similar
auditory-motor coupling after one day of training and, furthermore, that the voxel-wise patterns of activity

529
Fredrik Ullén

in premotor cortex during music listening represented the melodic structure of the presented melody (de
Manzano et al., 2020).
Expertise in the performing arts involves not just the acquisition of advanced sensorimotor skills but also,
as do other forms of expertise, more efficient domain-specific cognition. An important mechanism under-
lying superior domain-specific cognitive processing in experts is expert working memory, that is, an often
dramatically superior working memory for complex, ecological stimuli that are encountered in tasks per-
formed within the area of expertise. While initially characterized in seminal studies of chess experts (Chase &
Simon, 1973; Simon & Chase, 1973), expert working memory has since been documented in many domains
of expertise (Ullén et al., 2016), including music, with several behavioural studies demonstrating both more
efficient perceptual processing and higher memory capacity for notated music in musicians than in non-
musicians (Chang & Gauthier, 2020, 2021; Halpern & Bower, 1982; Meinz & Salthouse, 1998; Sloboda,
1976; Wong & Gauthier, 2012). A comprehensive discussion of different models of expert working memory
from a cognitive science perspective can be found in (Gobet, 2016). Neurophysiologically, expert working
memory presumably relies on efficient interactions between neural systems for working memory and long-
term memory (Ericsson & Kintsch, 1995; Gobet, 2016; Ullén et al., 2016, 2019). Although much progress
has been made through studies on adaptations in the visual system in response to complex visual stimuli (see
e.g., Bilalić et al., 2011; McGugin et al., 2014), the neural mechanisms of expert working memory in general
remain relatively poorly understood. Further research in this area will be an important task for future studies
on the neural basis of artistic expertise.

The performing arts as a model for gene-environment interplay


To recapitulate, there is abundant empirical evidence that the brains of musicians and dancers are specialized,
structurally as well as functionally, for optimal domain-specific performance. One of the most interesting
and challenging questions in the field at the moment is the etiological background to these findings. Why
do the brains and behavioural capacities of experts and non-experts differ? Traditionally, adherents of the
influential deliberate practice theory of expertise have essentially provided one answer to this question: They
differ because experts have engaged in long-term, deliberate practice (Ericsson, 2014; Ericsson et al., 1993).
However, empirical evidence accumulated during the last 10 years has made it increasingly clear that the
acquisition of expertise is a complex process that depends not just on practice but on many other variables
and their interplay.
One line of evidence consists of meta-analytical estimations of the effect sizes of associations between
measures of total practice and expert performance. Generally, these are more modest than would be expected
if practice were the sole determinant of expertise. In the case of music, comprehensive literature analyses
have estimated the percentage of variance in performance explained by cumulative measures of practice
to between 21% and 30% (Hambrick et al., 2014; Macnamara et al., 2014). An important point, which is
sometimes overlooked, is that these estimates do not address the question of causality, that is to say what
proportion of the effect that reflects causal influences of practice on performance, as stipulated by deliber-
ate practice theory, and how much reflects reverse causality and common influences from other factors,
such as genetic liability. Much controversy has been sparked in recent years by inconsistencies and confu-
sion around the definition of the term “deliberate practice” and related concepts (Hambrick et al., 2020).
Comprehensive discussions of these issues can be found elsewhere (Debatin et al., 2021; Hambrick et al.,
2020). Recent analyses do show, unsurprisingly, that practice characteristics matter and that effect sizes
of practice-performance associations may be higher when practice is individualized, structured, and goal
directed (Debatin et al., 2021; Miller et al., 2020). On the other hand, it is important to keep in mind that
such findings may not only reflect differences of efficacy of various forms of practice. It seems very likely that
there could be highly relevant constitutional differences in average motivation and aptitude between groups

530
Expertise and the performing artist

who choose to invest large amounts of time and energy in long-term goal-directed practice behaviours and
those who prefer less ambitious forms of engagement in the same domain. For the present purpose, it is suffi-
cient to note that while everyone agrees that well-organized practice is essential for expertise, the magnitude
of typical practice-performance relations clearly indicate that other factors also have an influence. To identify
these factors, more studies on expert performance that actually take into account both practice and other
relevant variables will be essential. For instance, in the musical domain, sight-reading ability is significantly
related to working memory capacity, even when controlling for practice (Meinz & Hambrick, 2010). In a
comprehensive study on real-life achievement in music, de Manzano and Ullén (2021) found that choosing a
musical occupation, as well as individual differences in accomplishment among professional musicians, show
a complex pattern of non-linear relations to several variables that include practice, musical aptitude, personal-
ity, motivational factors, and childhood musical environment. Interestingly, effects on musical achievement
also involve interactions between predictors. Mosing and colleagues (2019) found musical achievement in a
twin sample to be related to a positive interaction between practice and intelligence, while de Manzano and
Ullén (2021) documented that achievement differences among professional musicians depend on significant
interactions between psychosis proneness and openness, as well as between musical aptitude and intrinsic
motivation (musical flow proneness).
Last but not least, extensive behaviour genetic work during the last 10 years has provided ample evidence
that individual differences in musical expertise and related variables depend on an interplay between genetic
and environmental factors. Detailed reviews of this literature can be found elsewhere (Mosing et al., 2018;
Ullén et al., 2018; Wesseldijk et al., in press). Here, I would like to give a few examples to illustrate how
behaviour genetic analyses have transformed our view of musical expertise by going beyond mere demon-
strations that music-related variables are influenced by genetic factors, a fact which holds true of essentially
any complex human trait (Polderman et al., 2015). Studies combining multivariate classical twin modelling
with analyzes of causality based on co-twin control modelling have not only provided evidence for the
importance of active gene-environment co-variation in the acquisition of expertise, they have also led to the
surprising conclusion that associations between music practice and musical auditory discrimination (Mosing
et al., 2014), as well as intelligence (Mosing et al., 2016), are largely non-causal in nature and mainly driven
by common genetic influences. Sophisticated analyses of the relation between childhood environment and
musical achievement in adulthood, using moderation models in twins, have provided clear evidence for the
importance of gene-environment interactions (Wesseldijk et al., 2019). Studies of starting age of training and
its relation to musical abilities and achievements in adulthood clearly show that genetic factors are an impor-
tant confound to take into account when studying effects of early training and putative sensitive periods
(Wesseldijk, Mosing et al., 2021). Most recently, analyses using polygenic scores, based on weights derived
from a large-scale genome-wide association study (GWAS) on the self-reported ability to clap to a musical
beat (Niarchou et  al., 2021), have demonstrated co-variation between a musically stimulating childhood
environment and genetic factors involved in musical aptitude, as well as tentative evidence for assortative
mating for musicality (Wesseldijk et al., in press; Wesseldijk, Ullén et al., 2021).
These conclusions from studies at the phenotypic level have implications for the interpretation of the
neurobiological literature on performing artists. As already noted, research on musical training and the
brain mainly consists of cross-sectional and observational longitudinal studies. Schellenberg (2020) has
recently revealed a persistent and unfortunate bias in the research literature on musical training, particu-
larly in neuroscientific papers, where correlational findings tend to be discussed as the result of causal
effects of training, even though they could just as well reflect reverse causality, common influences of a
third factor (such as childhood environment or genetic liability), or some complex mixture of all three
causal scenarios.
In relation to the neuroanatomical findings on musicians and dancers summarized in the first section, one
should note that it is in fact far from obvious why training and skill acquisition should be accompanied by

531
Fredrik Ullén

monotonic increases in overall size of the involved brain regions. Despite much metaphorical rhetoric to the
contrary, the brain is not a muscle, and at the neuronal network level, fine-tuning of existing synapses and
elimination of unnecessary connections by pruning appear to be as important for learning as proliferation of
neural tissue (Wenger, Brozzoli et al., 2017; Xu et al., 2009). Wenger and colleagues have investigated the
temporal dynamics of the neuroanatomical changes accompanying skill learning, by scanning participants
repeatedly on up to 18 time-points during 7 weeks of training of a sensorimotor skill (2017). Interestingly, a
non-linear pattern was observed in motor cortex, with an expansion in size during the first weeks, followed
by a later partial renormalization. Nevertheless, cross-sectional studies have, as we have seen, fairly consist-
ently found that auditory-motor regions are larger in size in musicians. One explanation could be that even
though training a limited set of skills over the course of a few weeks may involve waves of plastic expansion
and reduction of neural tissue, the net effect of prolonged practice in experts—which occurs on a time scale
of many years and involves the acquisition of a vast repertoire of ever more advanced skills—is an overall
increase in regional size. This notion receives support from recent findings by de Manzano and Ullén (2018),
who scanned monozygotic twin pairs with a large intra-pair difference in piano training and found that the
twins who had played more piano had significantly more well-developed auditory-motor regions, both in
the grey and the white matter, than their co-twins.
However, such observations are, of course, perfectly compatible with the additional possibility that neu-
roanatomical differences between musicians and non-musicians also in part reflect differences in genetic
liability, perhaps related to aptitude and interest in music. In fact, several observations support that this is
likely to be the case. First, large-scale studies in genetically informative samples show that there is a substan-
tial genetic component to individual variation in brain structure (Zhao et al., 2019, 2021). Second, such
studies also consistently find that co-variation between brain structure and traits, such as cognitive abilities,
partly reflect shared genetic influences (Deary et al., 2022; Zhao et al., 2019, 2021). Given this, and the fact
that earlier research has found strong evidence for the heritability of music practice, as well as shared genetic
influences on music practice and musical aptitude (Mosing et al., 2014), it appears likely that associations
between musical aptitude and regional brain structure have a genetic component. Third, several studies have
found evidence for that structural and functional properties of auditory-motor areas measured before the
onset of musical training predict rate of learning (Herholz et al., 2016; Zatorre, 2013; Zatorre et al., 2012).
One interesting example is a study by Seither-Preisler and colleagues on the longitudinal relations between
brain measures, music practice, and musical aptitude in a sample of 111 children (2014). Neural measure-
ments were obtained at two time points separated by 13 months. The morphology of the auditory cortex
was remarkably stable over time, with no significant volumetric differences between time points 1 and 2.
Auditory cortex morphology showed a much stronger association with musical aptitude (β = 0.83) than with
intensity of practice (β = 0.17). Perhaps most strikingly, the volume of right Heschl’s gyrus at time point 1
significantly predicted music practice between time points 1 and 2, even when parcelling out practice before
time point 1. All in all, the results support that brain dispositions can have a considerable influence on an
individual’s musical aptitude and motivation to learn to play an instrument.

Conclusion
Research on expertise and its neural underpinnings is currently in a dynamic phase. The field is experiencing
a shift from a traditional view, where expertise was seen as essentially the result of practice, to a multifactorial
perspective that takes into account multiple influences on skill acquisition at the phenotypic level, as well
as the ubiquitous influences of gene-environment interplay. It is pleasing to see that studies using artistic
domains, such as music, as models of expertise have had a decisive impact on this development. Although
much progress has been made in recent years, many important questions remain to be investigated con-
cerning the brain basis of artistic expertise, such as the neural representation of acquired skills, interactions

532
Expertise and the performing artist

between working memory and long-term memory in expert problem solving, and multiperson coordination
in expert performances that require the interaction between several individuals, such as group dancing and
music making (see Chapter 24).
In differential studies of individual variation in expertise, adopting multifactorial designs that considers the
interaction between practice and other relevant variables appears essential for further progress. Twin research
on musical expertise highlights the importance of using genetically informative designs whenever possible
to enable an analysis of genetic and environmental factors and their interplay. Early gene-finding studies on
musical traits have been limited by small samples and replication problems, but the recent completion of the
first well-powered GWAS for a musical phenotype (Niarchou et al., 2021) suggests that interesting progress
will be made in future years on the molecular genetic basis of musical traits. For this purpose, additional large
GWASs using well-characterized musical phenotypes will certainly be needed to provide a better under-
standing of the genetic basis of individual variation in different musical behaviours.

Acknowledgements
I am grateful to Örjan de Manzano and Miriam Mosing for comments on an earlier version of this chapter.

References
Amunts, K., Schlaug, G., Jäncke, L., Steinmetz, H., Schleicher, A., Dabringhaus, A., & Zilles, K. (1997). Motor cor-
tex and hand motor skills: Structural compliance in the brain. Human Brain Mapping, 5(3), 206–215. https://doi.
org/10.1002/(SICI)1097-0193(1997)5:3<206::AID-HBM5>3.0.CO;2-7
Armand, J., Olivier, E., Edgley, S. A., & Lemon, R. N. (1996). The structure and function of the developing corti-
cospinal tract: Some key issues. In A. M. Wing, P. Haggard, & J. R. Flanagan (Eds.), Hand and brain (pp. 125–146).
Academic Press.
Bangert, M., & Schlaug, G. (2006). Specialization of the specialized in features of external human brain morphology.
European Journal of Neuroscience, 24(6), 1832–1834. https://doi.org/10.1111/j.1460-9568.2006.05031.x
Bengtsson, S. L., Nagy, Z., Skare, S., Forsman, L., Forssberg, H., & Ullén, F. (2005). Extensive piano practicing has region-
ally specific effects on white matter development. Nature Neuroscience, 8(9), 1148–1150. https://doi.org/10.1038/
nn1516
Benner, J., Wengenroth, M., Reinhardt, J., Stippich, C., Schneider, P., & Blatow, M. (2017). Prevalence and function of
Heschl’s gyrus morphotypes in musicians. Brain Structure and Function, 222(8), 3587–3603. https://doi.org/10.1007/
s00429-017-1419-x
Bermudez, P., Lerch, J. P., Evans, A. C., & Zatorre, R. J. (2009). Neuroanatomical correlates of musicianship as revealed
by cortical thickness and voxel-based morphometry. Cerebral Cortex, 19(7), 1583–1596. https://doi.org/10.1093/
cercor/bhn196
Bilalić, M., Langner, R., Ulrich, R., & Grodd, W. (2011). Many faces of expertise: Fusiform face area in chess experts
and novices. Journal of Neuroscience, 31(28), 10206–10214. https://doi.org/10.1523/JNEUROSCI.5727-10.2011
Calvo-Merino, B., Glaser, D. E., Grèzes, J., Passingham, R. E., & Haggard, P. (2005). Action observation and acquired
motor skills: An fMRI study with expert dancers. Cerebral Cortex, 15(8), 1243–1249. https://doi.org/10.1093/
cercor/bhi007
Caspers, S., Zilles, K., Laird, A. R., & Eickhoff, S. B. (2010). ALE meta-analysis of action observation and imitation in
the human brain. NeuroImage, 50(3), 1148–1167. https://doi.org/10.1016/j.neuroimage.2009.12.112
Chang, T. Y., & Gauthier, I. (2020). Distractor familiarity reveals the importance of configural information in musi-
cal notation. Attention, Perception and Psychophysics, 82(3), 1304–1317. https://doi.org/10.3758/s13414-019-01826-0
Chang, T. Y., & Gauthier, I. (2021). Domain-specific and domain-general contributions to reading musical notation.
Attention, Perception and Psychophysics, 83(7), 2983–2994. https://doi.org/10.3758/s13414-021-02349-3
Chase, W. G., & Simon, H. A. (1973). Perception in chess. Cognitive Psychology, 4(1), 55–81. https://doi.org/10.1016/
0010-0285(73)90004-2
Cross, E. S., Kraemer, D. J. M., Hamilton, A. F. D., Kelley, W. M., & Grafton, S. T. (2009). Sensitivity of the action obser-
vation network to physical and observational learning. Cerebral Cortex, 19(2), 315–326. https://doi.org/10.1093/
cercor/bhn083

533
Fredrik Ullén

de Manzano, Ö., Kuckelkorn, K. L., Ström, K., & Ullén, F. (2020). Action-perception coupling and near transfer: Listen-
ing to melodies after piano practice triggers sequence-specific representations in the auditory-motor network. Cerebral
Cortex, 30(10), 5193–5203. https://doi.org/10.1093/cercor/bhaa018
de Manzano, Ö., & Ullén, F. (2018). Same genes, different brains: Neuroanatomical differences between monozygotic
twins discordant for musical training. Cerebral Cortex, 28(1), 387–394. https://doi.org/10.1093/cercor/bhx299
de Manzano, Ö., & Ullén, F. (2021). Domain specific traits predict achievement in music and multipotentiality. Intel-
ligence, 89, 19, article 101584. https://doi.org/10.1016/j.intell.2021.101584
Deary, I. J., Cox, S. R., & Hill, W. D. (2022). Genetic variation, brain, and intelligence differences. Molecular Psychiatry,
27(1), 335–353. https://doi.org/10.1038/s41380-021-01027-y
Debatin, T., Hopp, M. D. S., Vialle, W., & Ziegler, A. (2021). The meta-analyses of deliberate practice underestimate the
effect size because they neglect the core characteristic of individualization-an analysis and empirical evidence. Current
Psychology, 11. https://doi.org/10.1007/s12144-021-02326-x
Ericsson, K. A. (2014). Why expert performance is special and cannot be extrapolated from studies of performance
in the general population: A response to criticisms. Intelligence, 45, 81–103. https://doi.org/10.1016/j.intell.2013.
12.001
Ericsson, K. A., & Kintsch, W. (1995). Long-term working memory. Psychological Review, 102(2), 211–245. https://doi.
org/10.1037/0033-295x.102.2.211
Ericsson, K. A., Krampe, R. T., & Tesch-Römer, C. (1993). The role of deliberate practice in the acquisition of expert
performance. Psychological Review, 100(3), 363–406. https://doi.org/10.1037/0033-295X.100.3.363
Fauvel, B., Groussard, M., Chételat, G., Fouquet, M., Landeau, B., Eustache, F., Desgranges, B., & Platel, H. (2014).
Morphological brain plasticity induced by musical expertise is accompanied by modulation of functional connectivity
at rest. NeuroImage, 90, 179–188. https://doi.org/10.1016/j.neuroimage.2013.12.065
Gardner, T., Goulden, N., & Cross, E. S. (2015). Dynamic modulation of the action observation network by movement
familiarity. Journal of Neuroscience, 35(4), 1561–1572. https://doi.org/10.1523/JNEUROSCI.2942-14.2015
Gaser, C., & Schlaug, G. (2003). Brain structures differ between musicians and non-musicians. Journal of Neuroscience,
23(27), 9240–9245. https://doi.org/10.1523/JNEUROSCI.23-27-09240.2003
Giacosa, C., Karpati, F. J., Foster, N. E. V., Hyde, K. L.,  & Penhune, V. B. (2019). The descending motor tracts
are different in dancers and musicians. Brain Structure and Function, 224(9), 3229–3246. https://doi.org/10.1007/
s00429-019-01963-0
Giacosa, C., Karpati, F. J., Foster, N. E. V., Penhune, V. B., & Hyde, K. L. (2016). Dance and music training have different
effects on white matter diffusivity in sensorimotor pathways. NeuroImage, 135, 273–286. https://doi.org/10.1016/j.
neuroimage.2016.04.048
Gobet, F. (2016). Understanding expertise: A multidisciplinary approach. Palgrave.
Groussard, M., La Joie, R., Rauchs, G., Landeau, B., Chételat, G., Viader, F., Desgranges, B., Eustache, F., & Platel, H.
(2010). When music and long-term memory interact: Effects of musical expertise on functional and structural plastic-
ity in the hippocampus. PLoS One, 5(10), e13225. https://doi.org/10.1371/journal.pone.0013225
Groussard, M., Viader, F., Landeau, B., Desgranges, B., Eustache, F., & Platel, H. (2014). The effects of musical prac-
tice on structural plasticity: The dynamics of grey matter changes. Brain and Cognition, 90, 174–180. https://doi.
org/10.1016/j.bandc.2014.06.013
Guida, A., Gobet, F., Tardieu, H., & Nicolas, S. (2012). How chunks, long-term working memory and templates offer
a cognitive explanation for neuroimaging data on expertise acquisition: A two-stage framework. Brain and Cognition,
79(3), 221–244. https://doi.org/10.1016/j.bandc.2012.01.010
Habibi, A., Damasio, A., Ilari, B., Veiga, R., Joshi, A. A., Leahy, R. M., Haldar, J. P., Varadarajan, D., Bhushan, C., &
Damasio, H. (2018). Childhood music training induces change in micro and macroscopic brain structure: Results
from a longitudinal study. Cerebral Cortex, 28(12), 4336–4347. https://doi.org/10.1093/cercor/bhx286
Halpern, A. R., & Bower, G. H. (1982). Musical expertise and melodic structure in memory for musical notation. Ameri-
can Journal of Psychology, 95(1), 31–50. https://doi.org/10.2307/1422658
Halwani, G. F., Loui, P., Rüber, T., & Schlaug, G. (2011). Effects of practice and experience on the arcuate fasciculus:
Comparing singers, instrumentalists, and non-musicians. Frontiers in Psychology, 2, 156. https://doi.org/10.3389/
fpsyg.2011.00156
Hambrick, D. Z., Macnamara, B. N.,  & Oswald, F. L. (2020). Is the deliberate practice view defensible? A  review
of evidence and discussion of issues. Frontiers in Psychology, 11, 1134, article 1134. https://doi.org/10.3389/
fpsyg.2020.01134
Hambrick, D. Z., Oswald, F. L., Altmann, E. M., Meinz, E. J., Gobet, F., & Campitelli, G. (2014). Deliberate practice: Is
that all it takes to become an expert? Intelligence, 45, 34–45. https://doi.org/10.1016/j.intell.2013.04.001

534
Expertise and the performing artist

Han, Y., Yang, H., Lv, Y. T., Zhu, C. Z., He, Y., Tang, H. H., Gong, Q. Y., Luo, Y. J., Zang, Y. F., & Dong, Q. (2009).
Gray matter density and white matter integrity in pianists’ brain: A combined structural and diffusion tensor MRI
study. Neuroscience Letters, 459(1), 3–6. https://doi.org/10.1016/j.neulet.2008.07.056
Hänggi, J., Koeneke, S., Bezzola, L.,  & Jäncke, L. (2010). Structural neuroplasticity in the sensorimotor network of
professional female ballet dancers. Human Brain Mapping, 31(8), 1196–1206. https://doi.org/10.1002/hbm.20928
Haslinger, B., Erhard, P., Altenmüller, E., Schroeder, U., Boecker, H., & Ceballos-Baumann, A. O. (2005). Transmodal
sensorimotor networks during action observation in professional pianists. Journal of Cognitive Neuroscience, 17(2),
282–293. https://doi.org/10.1162/0898929053124893
Herholz, S. C., Coffey, E. B. J., Pantev, C., & Zatorre, R. J. (2016). Dissociation of neural networks for predisposi-
tion and for training-related plasticity in auditory-motor learning. Cerebral Cortex, 26(7), 3125–3134. https://doi.
org/10.1093/cercor/bhv138
Hickok, G., Buchsbaum, B., Humphries, C.,  & Muftuler, T. (2003). Auditory-motor interaction revealed by fMRI:
Speech, music, and working memory in area Spt [Speech]. Journal of Cognitive Neuroscience, 15(5), 673–682. https://
doi.org/10.1162/089892903322307393
Hickok, G., & Poeppel, D. (2007). The cortical organization of speech processing. Nature Reviews. Neuroscience, 8(5),
393–402. https://doi.org/10.1038/nrn2113
Hill, N. M., & Schneider, W. (2006). Brain changes in the development of expertise: Neuroanatomical and neurophysi-
ological evidence about skill-based adaptations. In K. A. Ericsson, N. Charness, P. J. Feltovich, & R. R. Hoffman
(Eds.), The Cambridge handbook of expertise and expert performance (pp. 653–682). Cambridge University Press.
Hutchinson, S., Lee, L. H., Gaab, N., & Schlaug, G. (2003). Cerebellar volume of musicians. Cerebral Cortex, 13(9),
943–949. https://doi.org/10.1093/cercor/13.9.943
Hyde, K. L., Lerch, J., Norton, A., Forgeard, M., Winner, E., Evans, A. C.,  & Schlaug, G. (2009). Musical train-
ing shapes structural brain development. Journal of Neuroscience, 29(10), 3019–3025. https://doi.org/10.1523/
JNEUROSCI.5118-08.2009
Jacobacci, F., Armony, J. L., Yeffal, A., Lerner, G., Amaro, E., Jovicich, J., Doyon, J., & Della-Maggiore, V. (2020). Rapid
hippocampal plasticity supports motor sequence learning. Proceedings of the National Academy of Sciences of the United
States of America, 117(38), 23898–23903. https://doi.org/10.1073/pnas.2009576117
James, C. E., Oechslin, M. S., Van de Ville, D., Hauert, C. A., Descloux, C., & Lazeyras, F. (2014). Musical training
intensity yields opposite effects on grey matter density in cognitive versus sensorimotor networks. Brain Structure and
Function, 219(1), 353–366. https://doi.org/10.1007/s00429-013-0504-z
Johansen-Berg, H., Della-Maggiore, V., Behrens, T. E. J., Smith, S. M., & Paus, T. (2006). Integrity of white matter in
the corpus callosum correlates with bimanual co-ordination skills. NeuroImage, 36(Suppl. 2), T16–T21. https://doi.
org/10.1016/j.neuroimage.2007.03.041
Karahan, E., Costigan, A. G., Graham, K. S., Lawrence, A. D., & Zhang, J. X. (2019). Cognitive and white-matter com-
partment models reveal selective relations between corticospinal tract microstructure and simple reaction time. Journal
of Neuroscience, 39(30), 5910–5921. https://doi.org/10.1523/JNEUROSCI.2954-18.2019
Karpati, F. J., Giacosa, C., Foster, N. E. V., Penhune, V. B., & Hyde, K. L. (2017). Dance and music share gray matter
structural correlates. Brain Research, 1657, 62–73. https://doi.org/10.1016/j.brainres.2016.11.029
Lahav, A., Saltzman, E.,  & Schlaug, G. (2007). Action representation of sound: Audiomotor recognition net-
work while listening to newly acquired actions. Journal of Neuroscience, 27(2), 308–314. https://doi.org/10.1523/
JNEUROSCI.4822-06.2007
Lee, D. J., Chen, Y., & Schlaug, G. (2003). Corpus callosum: Musician and gender effects. NeuroReport, 14(2), 205–209.
https://doi.org/10.1097/00001756-200302100-00009
Leipold, S., Klein, C.,  & Jäncke, L. (2021). Musical expertise shapes functional and structural brain networks
independent of absolute pitch ability. Journal of Neuroscience, 41(11), 2496–2511. https://doi.org/10.1523/
JNEUROSCI.1985-20.2020
Liew, S. L., Sheng, T., Margetis, J. L., & Aziz-Zadeh, L. (2013). Both novelty and expertise increase action observation
network activity. Frontiers in Human Neuroscience, 7, 541. https://doi.org/10.3389/fnhum.2013.00541
Limb, C. J.,  & Braun, A. R. (2008). Neural substrates of spontaneous musical performance: An fMRI study of jazz
improvisation. PLoS One, 3(2), e1679. https://doi.org/10.1371/journal.pone.0001679
Macnamara, B. N., Hambrick, D. Z.,  & Oswald, F. L. (2014). Deliberate practice and performance in music,
games, sports, education, and professions: A  meta-analysis. Psychological Science, 25(8), 1608–1618. https://doi.
org/10.1177/0956797614535810
McGugin, R. W., Newton, A. T., Gore, J. C., & Gauthier, I. (2014). Robust expertise effects in right FFA. Neuropsycho-
logia, 63, 135–144. https://doi.org/10.1016/j.neuropsychologia.2014.08.029

535
Fredrik Ullén

Meinz, E. J., & Hambrick, D. Z. (2010). Deliberate practice is necessary but not sufficient to explain individual differ-
ences in piano sight-reading skill: The role of working memory capacity. Psychological Science, 21(7), 914–919. https://
doi.org/10.1177/0956797610373933
Meinz, E. J., & Salthouse, T. A. (1998). The effects of age and experience on memory for visually presented music.
Journals of Gerontology. Series B, Psychological Sciences and Social Sciences, 53(1), P60–P69. ://WOS:000071780300007.
https://doi.org/10.1093/geronb/53b.1.p60
Meyer, A. (1977). The search for a morphological substrate in the brains of eminent persons including musicians:
A historical review. In M. Critchley & R. A. Henson (Eds.), Music and the brain (pp. 255–281). William Heinemann
Medical Books Ltd.
Miller, S. D., Chow, D., Wampold, B. E., Hubble, M. A., Del Re, A. C., Maeschalck, C., & Bargmann, S. (2020). To
be or not to be (an expert)? Revisiting the role of deliberate practice in improving performance. High Ability Studies,
31(1), 5–15. https://doi.org/10.1080/13598139.2018.1519410
Mosing, M. A., Hambrick, D. Z., & Ullén, F. (2019). Predicting musical aptitude and achievement: Practice, teaching,
and intelligence. Journal of Expertise, 2(3), 184–197.
Mosing, M. A., Madison, G., Pedersen, N. L., Kuja-Halkola, R.,  & Ullén, F. (2014). Practice does not make per-
fect: No causal effect of music practice on music ability. Psychological Science, 25(9), 1795–1803. https://doi.
org/10.1177/0956797614541990
Mosing, M. A., Madison, G., Pedersen, N. L., & Ullén, F. (2016). Investigating cognitive transfer within the frame-
work of music practice: Genetic pleiotropy rather than causality. Developmental Science, 19(3), 504–512. https://doi.
org/10.1111/desc.12306
Mosing, M. A., Peretz, I., & Ullén, F. (2018). Genetic influences on musical expertise. In D. Z. Hambrick, G. Campi-
telli, & B. Macnamara (Eds.), The science of expertise: Behavioral, neural, and genetic approaches to complex skill (pp. 272–
282). Routledge.
Münte, T. F., Altenmüller, E., & Jäncke, L. (2002). The musician’s brain as a model of neuroplasticity. Nature Reviews.
Neuroscience, 3(6), 473–478. https://doi.org/10.1038/nrn843
Niarchou, M., Gustavson, D. E., Sathirapongsasuti, J. F., Anglada-Tort, M., Eising, E., Bell, E., McArthur, E., Straub, P.,
Team, T. A. R., McAuley, J. D., Capra, J. A., Ullén, F., Creanza, N., Mosing, M. A., Hinds, D., Davis, L. K., Jacoby,
N., & Gordon, R. L. (2021). Genome-wide association study of musical beat synchronization demonstrates high
polygenicity. bioRxiv, 836197 [Preprint]. https://doi.org/10.1101/836197
Olszewska, A. M., Gaca, M., Herman, A. M., Jednoróg, K., & Marchewka, A. (2021). How musical training shapes the
adult brain: Predispositions and neuroplasticity. Frontiers in Neuroscience, 15, article 630829. https://doi.org/10.3389/
fnins.2021.630829
Öztürk, A. H., Tasçioglu, B., Aktekin, M., Kurtoglu, Z., & Erden, I. (2002). Morphometric comparison of the human
corpus callosum in professional musicians and non-musicians by using in vivo magnetic resonance imaging. Journal of
Neuroradiology, 29(1), 29–34.
Pinho, A. L., de Manzano, Ö., Fransson, P., Eriksson, H., & Ullén, F. (2014). Connecting to create—Expertise in musical
improvisation is associated with increased functional connectivity between premotor and prefrontal areas. Journal of
Neuroscience, 34(18), 6156–6163. https://doi.org/10.1523/JNEUROSCI.4769-13.2014
Pinho, A. L., Ullén, F., Castelo-Branco, M., Fransson, P., & de Manzano, Ö. (2016). Addressing a paradox: Dual strategies
for creative performance in introspective and extrospective networks. Cerebral Cortex, 26(7), 3052–3063. https://doi.
org/10.1093/cercor/bhv130
Polderman, T. J. C., Benyamin, B., de Leeuw, C. A., Sullivan, P. F., van Bochoven, A., Visscher, P. M., & Posthuma, D.
(2015). Meta-analysis of the heritability of human traits based on fifty years of twin studies. Nature Genetics, 47(7),
702–709. https://doi.org/10.1038/ng.3285
Rosen, D. S., Erickson, B., Kim, Y. E., Mirman, D., Hamilton, R. H., & Kounios, J. (2016). Anodal tDCS to right dor-
solateral prefrontal cortex facilitates performance for novice jazz improvisers but hinders experts. Frontiers in Human
Neuroscience, 10, 579, article 579. https://doi.org/10.3389/fnhum.2016.00579
Rüber, T., Lindenberg, R., & Schlaug, G. (2015). Differential adaptation of descending motor tracts in musicians. Cer-
ebral Cortex, 25(6), 1490–1498. https://doi.org/10.1093/cercor/bht331
Schellenberg, E. G. (2020). Correlation = causation? Music training, psychology, and neuroscience. Psychology of Aesthet-
ics, Creativity, and the Arts, 14(4), 475–480. https://doi.org/10.1037/aca0000263
Schendan, H. E., Searl, M. M., Melrose, R. J.,  & Stern, C. E. (2003). An fMRI study of the role of the medial
temporal lobe in implicit and explicit sequence learning. Neuron, 37(6), 1013–1025. https://doi.org/10.1016/
s0896-6273(03)00123-5
Schlaffke, L., Friedrich, S., Tegenthoff, M., Güntürkün, O., Genç, E., & Ocklenburg, S. (2020). Boom bhack boom—A
multimethod investigation of motor inhibition in professional drummers. Brain and Behavior, 10(1), article e01490.
https://doi.org/10.1002/brb3.1490

536
Expertise and the performing artist

Schlaug, G., Jäncke, L., Huang, Y., Staiger, J. F., & Steinmetz, H. (1995). Increased corpus callosum size in musicians.
Neuropsychologia, 33(8), 1047–1055. https://doi.org/10.1016/0028-3932(95)00045-5
Schneider, P., Scherg, M., Dosch, H. G., Specht, H. J., Gutschalk, A., & Rupp, A. (2002). Morphology of Heschl’s
gyrus reflects enhanced activation in the auditory cortex of musicians. Nature Neuroscience, 5(7), 688–694. https://
doi.org/10.1038/nn871
Schneider, P., Sluming, V., Roberts, N., Scherg, M., Goebel, R., Specht, H. J., Dosch, H. G., Bleeck, S., Stippich, C., &
Rupp, A. (2005). Structural and functional asymmetry of lateral Heschl’s gyrus reflects pitch perception preference.
Nature Neuroscience, 8(9), 1241–1247. https://doi.org/10.1038/nn1530
Seither-Preisler, A., Parncutt, R., & Schneider, P. (2014). Size and synchronization of auditory cortex promotes musi-
cal, literacy, and attentional skills in children. Journal of Neuroscience, 34(33), 10937–10949. https://doi.org/10.1523/
JNEUROSCI.5315-13.2014
Simon, H. A., & Chase, W. G. (1973). Skill in chess. American Scientist, 61, 394–403.
Sloboda, J. A. (1976). Visual perception of musical notation: Registering pitch symbols in memory. Quarterly Journal of
Experimental Psychology, 28(1), 1–16. https://doi.org/10.1080/14640747608400532
Sluming, V., Barrick, T., Howard, M., Cezayirli, E., Mayes, A. R., & Roberts, N. (2002). Voxel-based morphometry
reveals increased gray matter density in Broca’s area in male symphony orchestra musicians. NeuroImage, 17(3), 1613–
1622. https://doi.org/10.1006/nimg.2002.1288
Steele, C. J., Bailey, J. A., Zatorre, R. J.,  & Penhune, V. B. (2013). Early musical training and white-matter plastic-
ity in the corpus callosum: Evidence for a sensitive period. Journal of Neuroscience, 33(3), 1282–1290. https://doi.
org/10.1523/JNEUROSCI.3578-12.2013
Ullén, F., de Manzano, Ö., & Mosing, M. A. (2019). Neural mechanisms of expertise. In P. Ward, J. M. Schraagen, J.
Gore, & E. Roth (Eds.), The Oxford handbook of expertise: Research & application. Oxford University Press. https://doi.
org/doi: 0.1093/oxfordhb/9780198795872.013.6
Ullén, F., Hambrick, D. Z., & Mosing, M. A. (2016). Rethinking expertise: A multi-factorial gene-environment inter-
action model of expert performance. Psychological Bulletin, 142(4), 427–446. https://doi.org/10.1037/bul0000033
Ullén, F., Mosing, M. A., & Hambrick, D. Z. (2018). The multifactorial gene-environment interaction model (MGIM)
of expert performance. In D. Z. Hambrick, G. Campitelli, & B. Macnamara (Eds.), The science of expertise: Behavioral,
neural, and genetic approaches to complex skill (pp. 365–375). Routledge.
Vaquero, L., Hartmann, K., Ripollés, P., Rojo, N., Sierpowska, J., François, C., Càmara, E., van Vugt, F. T., Moham-
madi, B., Samii, A., Münte, T. F., Rodríguez-Fornells, A., & Altenmüller, E. (2016). Structural neuroplasticity in
expert pianists depends on the age of musical training onset. NeuroImage, 126, 106–119. https://doi.org/10.1016/j.
neuroimage.2015.11.008
Vaquero, L., Ramos-Escobar, N., Cucurell, D., François, C., Putkinen, V., Segura, E., Huotilainen, M., Penhune,
V.,  & Rodríguez-Fornells, A. (2021). Arcuate fasciculus architecture is associated with individual differences in
pre-attentive detection of unpredicted music changes. NeuroImage, 229, article 117759. https://doi.org/10.1016/j.
neuroimage.2021.117759
Vaquero, L., Ramos-Escobar, N., François, C., Penhune, V., & Rodríguez-Fornells, A. (2018). White-matter structural
connectivity predicts short-term melody and rhythm learning in non-musicians. NeuroImage, 181, 252–262. https://
doi.org/10.1016/j.neuroimage.2018.06.054
Vaquero, L., Rousseau, P. N., Vozian, D., Klein, D., & Penhune, V. (2020). What you learn and when you learn it: Impact
of early bilingual and music experience on the structural characteristics of auditory-motor pathways. NeuroImage, 213,
article 116689. https://doi.org/10.1016/j.neuroimage.2020.116689
Vergara, V. M., Norgaard, M., Miller, R., Beaty, R. E., Dhakal, K., Dhamala, M.,  & V. D. Calhoun, V. D. (2021).
Functional network connectivity during jazz improvisation. Scientific Reports, 11(1), 12, article 19036. https://doi.
org/10.1038/s41598-021-98332-x
Wenger, E., Brozzoli, C., Lindenberger, U., & Lövdén, M. (2017). Expansion and renormalization of human brain struc-
ture during skill acquisition. Trends in Cognitive Sciences, 21(12), 930–939. https://doi.org/10.1016/j.tics.2017.09.008
Wenger, E., Kühn, S., Verrel, J., Mårtensson, J., Bodammer, N. C., Lindenberger, U., & Lövdén, M. (2017). Repeated
structural imaging reveals nonlinear progression of experience-dependent volume changes in human motor cortex.
Cerebral Cortex, 27(5), 2911–2925. ://WOS:000400461700014. https://doi.org/10.1093/cercor/bhw141
Wesseldijk, L. W., Mosing, M. A.,  & Ullén, F. (2019). Gene-environment interaction in expertise: The importance
of childhood environment for musical achievement. Developmental Psychology, 55(7), 1473–1479. https://doi.
org/10.1037/dev0000726
Wesseldijk, L. W., Mosing, M. A., & Ullén, F. (2021). Why is an early start of training related to musical skills in adult-
hood? A genetically informative study? Psychological Science, 32(1), 3–13. https://doi.org/10.1177/0956797620959014
Wesseldijk, L. W., Ullén, F., & Mosing, M. A. (2021). The predictive value and potential pathways of a polygenic score
for rhythm ability. Behavior Genetics, 51(6), 756–756.

537
Fredrik Ullén

Wesseldijk, L. W., Ullén, F., & Mosing, M. A. (in press). Music and genetics. In G. Bernatzky & G. Kreutz (Eds.), Music
and medicine.
Wong, Y. K., & Gauthier, I. (2012). Music-reading expertise alters visual spatial resolution for musical notation. Psycho-
nomic Bulletin and Review, 19(4), 594–600. https://doi.org/10.3758/s13423-012-0242-x
Xu, T. H., Yu, X. Z., Perlik, A. J., Tobin, W. F., Zweig, J. A., Tennant, K., Jones, T., & Zuo, Y. (2009). Rapid formation
and selective stabilization of synapses for enduring motor memories. Nature, 462(7275), 915–919-U108. https://doi.
org/10.1038/nature08389
Zardi, A., Carlotti, E. G., Pontremoli, A., & Morese, R. (2021). Dancing in your head: An interdisciplinary review.
Frontiers in Psychology, 12, 649121, article 649121. https://doi.org/10.3389/fpsyg.2021.649121
Zatorre, R. J. (2013). Predispositions and plasticity in music and speech learning: Neural correlates and implications.
Science, 342(6158), 585–589. https://doi.org/10.1126/science.1238414
Zatorre, R. J., Chen, J. L., & Penhune, V. B. (2007). When the brain plays music: Auditory-motor interactions in music
perception and production. Nature Reviews. Neuroscience, 8(7), 547–558. https://doi.org/10.1038/nrn2152
Zatorre, R. J., Fields, R. D., & Johansen-Berg, H. (2012). Plasticity in gray and white: Neuroimaging changes in brain
structure during learning. Nature Neuroscience, 15(4), 528–536. https://doi.org/10.1038/nn.3045
Zhao, B. X., Luo, T. Y., Li, T. F., Li, Y., Zhang, J. W., Shan, Y., Wang, X. F., Yang, L. Q., Zhou, F., Zhu, Z. L., & Zhu,
H. T. (2019). Alzheimers dis neuroimaging. genome-wide association analysis of 19,629 individuals identifies variants
influencing regional brain volumes and refines their genetic co-architecture with cognitive and mental health traits.
Nature Genetics, 51(11), 1637. https://doi.org/10.1038/s41588-019-0516-6
Zhao, B. X., Zhang, J., Ibrahim, J. G., Luo, T., Santelli, R. C., Li, Y., Li, T., Shan, Y., Zhu, Z., Zhou, F., Liao, H., Nich-
ols, T. E., & Zhu, H. (2021). Large-scale GWAS reveals genetic architecture of brain white matter microstructure and
genetic overlap with cognitive and mental health traits (n=17,706). Molecular Psychiatry, 26(8), 3943–3955. https://
doi.org/10.1038/s41380-019-0569-z

538
29
THE EMERGENCE OF SYMBOLIC
COGNITION
Francesco d’Errico and Ivan Colagè

All human cultures create symbolic systems to understand, organize, and modify the surrounding world.
Without the human ability to attribute meanings to natural features or artificial conventional signs, there
would be no art, faith, mathematics, writing, fashion, or computers. Recent paleoanthropological and
archaeological discoveries have changed our view of the emergence of this ability. Researchers were long
convinced that the complex cognition necessary to create and handle symbols appeared suddenly about
50,000  years ago and that the first visible products of that cognitive revolution—the famous European
painted caves, ornaments, and carvings—were produced by modern humans as they settled in Europe,
replacing resident Neanderthals (Chase & Dibble, 1987; Mellars, 1989; Pfeiffer, 1982).
A different view replaced this one during the early 21st century. From this new perspective, symbolic
cultures developed in Africa with the origin of our species in that continent, an origin that paleoanthro-
pological and genetic evidence dated to between 150,000 and 200,000  years ago (McBrearty  & Brooks,
2000; Henshilwood & Marean, 2003). This evidence has been generally interpreted as supporting a scenario
according to which the selective process that have given raise to our species in Africa would have endowed
our species with new cognitive functions that favoured the gradual and progressive emergence and accumula-
tion of innovations, including the production of a symbolic material culture. However, most of the cultural
innovations observed in the African Middle Stone Age (MSA) occurred dozens, if not hundreds of thousands,
years after selection of the adaptive mutations that led to our species. The absence of art and other unambigu-
ous symbolic mediated behaviours at African sites dated to that period was explained by arguing that modern
cognition proper of the new species was certainly there but that the process giving rise to symbolically fer-
tile human cultures must have been gradual and that, consequently, it may have taken between 50,000 and
100,000 years for the new symbolic mind to leave tangible expressions in the archaeological record.
These two scenarios assume that cognition is a species-specific trait that is shaped by natural selection.
In other words, a new form of cognition would stem from the classic Neo-Darwinian process of isolation,
random mutation, selection of advantageous characters, and speciation. Thus, natural selection would have
provided our species with a new form of cognition, one capable of producing symbols and embody them
in material culture. Such a mechanism would enable behavioural variability but would narrow its range of
expression to the species’ biological potential. The corollary of this assumption is that human populations
of the past with different morphological characters, recognized by paleoanthropologists as belonging to dif-
ferent fossil species, must have had different cognitions and were probably unable to produce a symbolic, or

DOI: 10.4324/9781003008675-31 539


Francesco d’Errico and Ivan Colagè

fully symbolic, material culture. If things were so, however, there should be a clear correspondence between
the emergence of our species and that of symbolic behaviour. But this is not the case.
Today, a growing body of new archaeological discoveries and a reappraisal of old ones are leading to yet
another view about the origin of symbolic culture. Many key cultural innovations were developed outside
the African continent before 70–60 ka (Colagè & d’Errico, 2020). The ability to ignite and control fire,
complex and changing lithic technologies, elaborated wooden weapons, intricate hafting techniques, var-
ied hunting strategies enabling archaic hominins to kill dangerous game, exploitation of aquatic and plant
resources, organization of living space, and seafaring are all innovations now attributed to Neanderthal cul-
tures in various regions of Europe, even before any contact with Modern Humans (Villa & Roebroek, 2014;
d’Errico & Stringer, 2011; Johansson, 2014). Other lines of evidence, such as burials and mortuary practices,
collection of rare items, pigment use, engravings and paintings on objects and cave walls, and the extraction
of bird feathers and claws, show that Neanderthals were engaged in symbolically mediated behaviour well
before any contact with anatomically modern humans (Villa & Roebroeks, 2014; Zilhão, 2016; Pettitt, 2011;
Majkić et al., 2017; Rodríguez-Vidal et al., 2014; Stiner, 2017; Hoffman et al., 2018). The production of
formal bone tools, sometimes decorated with notches and a varied array of personal ornaments, found at late
Neanderthal sites, demonstrate that the production of items characteristic of the Upper Palaeolithic was fully
in the realm of Neanderthal cognition (d’Errico et al., 2003; Soressi et al., 2013; Vanhaeren et al., 2019). In
addition, isolated, puzzling evidence for mortuary practices (Pettitt, 2011; Zilhao, 2016; Dirks et al., 2015,
2017) and abstract engravings (Mania & Mania, 1988; Joordens et al., 2015) occurs in and outside Europe
at sites that are hundreds of thousand years older than the Mousterian cultures of Europe and the Near East.
Paleogenetic data demonstrating interbreeding between Neandertals, Modern Humans, and Denisovans (Fu
et al., 2016), and suggesting gene flow from an archaic African population into the modern human genome
(Hammer et al., 2011), show that possible differences in cognition produced by genetic drift and natural
selection clearly did not prevent successful social integration and reproduction of hybrids.

Two views of the origins of human symbolic cognition


The paleoanthropological and archaeological records show that the emergence of symbolic thought was a
complex process that cannot be accounted for simply by considering symbolic culture as the product of a
species-specific form of symbolic cognition. A thorough understanding of the evolution of cognition will
require accounting for the relationships among four factors: genetic endowment, brain anatomy and physiol-
ogy, cognitive skills, and cultural practices.

The “bottom-up only” view


A view we will call the “bottom-up only” view assumes that a species’ genetic endowment sets the species-
specific brain gross functional anatomy and physiology, which enables the array of cognitive skills available
to the individuals of that species, which in turn determine the “typical” behavioural patterns expressed by
individuals and populations of that species (Herculano-Houzel, 2017; Miklosi, 2014; Moss, 2018; Santolin &
Saffran, 2018; Shettleworth, 2010; Tinbergen, 1951). In the case of humans, those behavioural patterns
include the “key cultural innovations”—that is, innovations that are so momentous that they spawn an array
of secondary products or behaviours (as defined by Reader & Laland, 2003)—that make us different from our
phylogenetically closest relatives (e.g., making composite tools, creating symbolic items, developing literacy
and numerical symbol systems, or even speaking an articulate language). Such a view implies a strong “chain
of dependence” from genes to brains, to cognitive skills, to some key adaptations shared by all human cultures
(Geschwind & Rakic, 2013; Somel et al., 2013).

540
The emergence of symbolic cognition

It follows, from this view, that a change in the genetic endowment was ultimately and necessarily required
for those typically human behaviours to emerge. Indeed, according to the “bottom-up only” view, the
related genetic changes are the cause of—or, at least, the necessary condition for—the emergence of the
(key) cultural novelties. From this view, this causal relation between genes and culture can take several forms.
Cultural novelties deriving from genetic changes need not always be “adaptations”; they can also be (biologi-
cal) “exaptations” (Gould & Lewontin, 1979; Gould & Vrba, 1982) and even the result of long-term “direc-
tional selection” (e.g., Price et al., 1988). In all these cases, indeed, underling and causally related genetic
evolution is involved (see also Colagè & D’Ambrosio, 2014).
Additionally, the “bottom-up only” view does not entail that the genetic changes responsible for the
emergence of key cultural innovations be of such a large an extent to determine speciation events. In other
words, genetic changes leading to key cultural innovations may well occur at the micro-evolutionary level
(i.e., roughly, “within” the evolution of a given species) and not only—as it is obviously the case—at the
macro-evolutionary level (i.e., at the level of evolution of new biological species). In this sense, the “bottom-
up only” view encompasses all those proposals according to which the earliest appearance of “cumulative
culture”—defined as the ability of a culture to accumulate and improve innovations over time—would have
been the outcome of either the speciation event leading to Homo sapiens in Africa ca. 260 ka (Bolhuis et al.,
2014; Bruner, 2014; Coolidge  & Wynn, 2004, 2007, 2018; McBrearty  & Brooks, 2000; Mithen, 2005;
Neubauer et al., 2018; Shea, 2011; Wynn et al., 2016) or of a key genetic change occurring ca. 50 ka and
inducing some form of alteration in the neural constitution of H. sapiens (e.g., Bar-Yosef, 1998; Klein, 1989,
2000; Mellars & Stringer, 1989; Tattersall, 1995).

The “top-down also” view


Empirical findings and theoretical advancements during the last few decades have cast serious doubts on
the “bottom-up-only” view and have begun to delineate an alternative, “top-down also” view on the origins
of symbolic thought (Colagè  & d’Errico, 2020). From this perspective, cultural innovations have direct
effects on the cognitive capabilities of hominin populations (via what can be called “cultural exaptation”)
and on their brains (via what may be labelled “cultural neural reuse”) (d’Errico & Colagè, 2018). Support
for these effects of cultural innovations on cognition and its neural underpinnings comes from four research
lines. The first of these is work on epigenesis and phenotypic plasticity ( Jablonka & Lamb, 2005; see also
D’Ambrosio & Colagè, 2017). Broadly speaking, this line of research shows that the phenotype should not
be considered merely the mechanistic outcome of a genotype. Phenotypic variation is not due solely to
genotypic variation: Novel phenotypic configurations can be induced by environmental factors without
requiring any change in the DNA nucleotide sequence. Many striking examples of this are known, including
the marked morphological and behavioural differences in genetically identical working versus queen honeybees
(Apis mellifera) due to diet differences during development (Lyko et al., 2010) or the effects on jaw size of a
modern soft diet, which requires less chewing force and time (Corruccini, 1984; Tang et al., 2004; Varrela,
1992, 2006). This implies that phenotypic change need not coincide with, or follow the same rate as, genetic
evolution. From the “top-down also” perspective, epigenesis and phenotypic plasticity play a crucial role in
the explanation of the origins of symbolic cognition because “phenotype” is understood as encompassing
cognition and behaviour, in addition to morphology, anatomy, and physiology and because “environment”
is understood as including cultural practices.
The second research line that lends support to the “top-down also” perspective on the emergence of sym-
bolic cognition is neural plasticity: the ability of the brain to modify physiological, functional, and/or struc-
tural features due to experience and practice or following brain lesions ( Johnson, 2001, 2011; Pascual-Leone
et al., 2005). The human brain is particularly plastic, and measurable brain changes (in both grey and white

541
Francesco d’Errico and Ivan Colagè

matter) can be detected even after relatively brief periods of practicing novel tasks (e.g., Draganski & May,
2008; Hecht et al., 2015). Neuroplasticity opens the possibility for novel “brain phenotypes” (potentially able
to support new cognitive skills and cultural practices) to be produced and—most important—stabilized in a
population without requiring concomitant genetic changes.
The third line is gene-culture co-evolution (e.g., Gintis, 2011; Laland et al., 2010; Richerson et al., 2010;
Varki et al., 2008). In the last 40 years, following Feldman and Cavalli-Sforza’s (1976) path-breaking work
on adult lactose tolerance in human populations with a history of dairy farming, several cases of a specific
cultural practice in a population directly influencing the genetic evolution (allele distribution) of individu-
als within that population have been identified. Although gene-culture co-evolution mainly involves, at
the genetic level, “classical” mechanisms of population genetics, the point is that cultural practices (and the
related cognitive skills) can emerge and (begin to) stabilize before genetic evolution takes place.
The fourth research line is the recent work on “culturally-driven evolution” (Dor  & Jablonka, 2014;
Heyes, 2012, 2018; Laland, 2017), that is to say, cultural evolution with “little if any” strictly biological
(namely, genetic) evolution. This idea is often applied to one of the most peculiarly human cognitive, cul-
tural, and symbolic abilities: spoken, articulate, and highly syntactical language. The consequences of articu-
late language on human cognition are enormous for sociality, communication, conceptual thinking, material
culture (mainly via teaching), theoretical culture, personal, social identity, and so on. It is also clear that
articulate language is the outcome of a long evolutionary process leading to a sufficiently complex brain to
support it and that this process entailed substantial genetic evolution (e.g., Fisher, 2017; Maricic et al., 2013).
Nevertheless, some authors suggest that at least the most recent stages of language evolution (mainly con-
cerning complex syntax and semantics) are culturally and not biologically driven (Anderson, 2014; Arbib,
2013, 2016; Bickerton, 2014; Dor & Jablonka, 2014; Evans & Levinson, 2009; see also Colagè, 2016).
These four research lines provide converging support for the “top-down also” view of the evolution of
symbolic cognition. The fundamental claim of this view is that cognitive evolution is not driven exclusively
by genetic evolution, not even as far as key cultural innovations are concerned. On the contrary, the “top-
down also” view maintains that truly cultural evolution can trigger the emergence of new cognitive skills
independently of any heritable change in the genetic make-up of organisms. In other words, the idea is that
cultural and cognitive novelty may emerge and stabilize even without any concomitant change at the genetic
level.
This view is compatible with the evidence on epigenesis and phenotypic plasticity noted previously, espe-
cially as phenotype is understood to include cognition and behaviour in addition to morphology, anatomy,
and physiology. It also fits with the suggestion that especially the recent stages of the evolution of genus Homo
and H. sapiens are mainly culturally driven. Moreover, the “top-down-also” view may add a new implication
to the issue of gene-culture co-evolution. It is part and parcel of gene-culture co-evolutionary approaches
that culture is the leading factor and that genetic evolution “just” follows. However, there might be cases in
which the ensuing genetic evolution is nonessential not only to the emergence of a key cultural innovation
but also to its spread and stabilization within and across populations. There is evidence suggesting that the
emergence of cultural novelties via genuinely cultural evolution not only may affect cognitive skills but also
the underlying neural substrates without requiring genetic evolution and that the novel cultural process,
related cognitive skills, and underpinning neural substrates may spread and stabilize much before any “genetic
accommodation” to those novelties appears.

Patterns of cultural innovations


The rich evidence yielded by the archaeological record offers robust support for the “top-down also” view
of the evolution of cognition. Figure 29.1 synthesizes what is currently known about the emergence, over
the last 800,000 years, of 29 cultural innovations related to subsistence strategies, technology, and symbolic

542
The emergence of symbolic cognition
543

Figure 29.1 Occurrence of cultural innovations in four regions of the world during the last 850 ky. Reproduced from Colagè and d’Errico (2020) with permission.
Francesco d’Errico and Ivan Colagè

behaviour in four geographical areas: Africa, the Near East, Europe, and Asia. The figure shows only inno-
vations with obvious links to cultural complexity and excludes those based on circumstantial evidence, that
is, evidence that relies on an inference to connect it to a conclusion. Recent innovations, such a pottery,
domestication, and writing, were not considered. Literature mining was based on previous compilations
(d’Errico, 2003; McBrearty & Brooks, 2000; Villa & Roebroek, 2014) and recent discoveries concerning
each innovation category (Colagè & d’Errico, 2020).
Eight main points emerge from the analysis of the graphic representation of the data and a critical analysis
of the literature (Colagè & d’Errico, 2020):

1 Isolated occurrences of key innovations (shell-fishing, microliths), including possible instances of sym-
bolic material culture (abstract engravings, carvings), appear in very ancient sites (800–300 ka) in the
four regions.
2 Several long-lasting innovations linked to subsistence (control of fire, hafting, blade production, grind-
stone use) and to possible symbolic activities (some mortuary practices, ochre use) are recorded in
Africa, Europe, the Near East, and, to a lesser extent, Asia between 500 and 250 ka.
3 A set of innovations (i.e., formal bone tools, pyrotechnology, use of mastics, marine and fresh-water
fishing, personal ornaments, engravings, objects colouring) appears in Africa between 100 and 70 ka.
However, these innovations are almost exclusively found at few sites in Northern and Southern Africa
and disappear in some cases for 5–10 ky to reappear again in different forms. Many are found in Europe
and the Near East before the arrival of modern humans.
4 Other innovations (rock art, possible systems of notations, use of poison) appear permanently every-
where only from 50 to 40 ka or later, with possible precursors in Europe 64 ka ago.
5 Most innovations recorded in Africa, Europe, and the Near East before 50 ka are found in Asia only after
that date.
6 Several isolated instances reported in the literature would require a critical reassessment before being
accepted.
7 In several cases, future discoveries may fill in the gap between early isolated and more recent consoli-
dated occurrences.
8 Since early occurrences of several key innovations were unknown a decade ago, we may expect that
future archaeological discoveries will almost certainly identify earlier occurrences of some innovations.

The emergence, adoption, loss, or reinvention of most these innovations cannot be connected with certainty
to any of the several different fossil populations that lived in those four areas during the considered time
span. These included Homo erectus, Homo antecessor, Homo neanderthalensis, Homo heidelbergensis, Homo sapiens,
and probably other populations of the genus Homo such as the Denisovans. The timing, location, and pace
of the appearance of cultural innovations are inconsistent with scenarios attributing the spread of modern
cultural traits to concomitant genetic evolutionary processes: (1) There is no increase in rate, amount, or
variety of innovations that could be linked to a sudden shift in human cognition produced by genetic change
or speciation at any given moment in the past, in a particular region or population. (2) No accumulation of
innovations is observed exclusively in only one single region. (3) Many subsistence-related cultural achieve-
ments previously considered the exclusive attributes of modern humans also occur among anatomically
“archaic” populations such as Neanderthals: Complex and changing technologies—including the ability to
produce blades, use of grindstones and mastic, heat treatment of rocks, hafting techniques, varied strategies
to hunt dangerous game, exploitation of marine and plant resources, the ability to ignite and control fire,
organization of living space, and seafaring—are among the innovations that are now recognized as inherent
to Neanderthal cultures in Europe before any contact with modern humans (d’Errico, 2003; d’Errico &
Stringer, 2011; Villa & Roebroek, 2014). (4) Several lines of evidence support the view that symbolically

544
The emergence of symbolic cognition

Figure 29.2 Frequency of cultural innovations in the last 850,000 years at the regional scale and (insert) at the global
scale. This graph summarizes the data in Figure 29.1. Reproduced from Colagè and d’Errico (2020) with
permission.

mediated behaviour also played a role in Neanderthal cultures: Treatment of the body after death, collection
of rare natural items, engraved and perforated objects, personal ornaments, pigment use, extraction of bird
feathers and claws probably for personal decoration, and abstract patterns engraved on cave walls are among
the significant symbolic achievements now firmly attributed to Neanderthals.
The virtual absence or discontinuous occurrence of such cultural innovations in Asia before anatomi-
cally modern populations colonized this region does not necessarily reflect built-in cognitive differences.
This could be due to different cultural trajectories less reliant on cultural transmission and accumulation of
knowledge or, alternatively, to a lack of research in, and data coming from, large regions of this continent.
In sum, cultural innovations seem to appear, disappear, and reappear (with different specific features) at
different times, in different places, and related to different human fossil species across the four geographical
areas. Such complex patterns of cultural innovation are incompatible with a scenario in which key cultural
innovations are concomitant and/or directly causally related to genetic changes. The overall trajectory (Fig-
ure 29.2) of cultural innovations in recent human evolution seems incompatible with the “bottom-up only”
view, and it arguably supports the idea that cultural dynamics may be the leading force of cognitive and
cultural evolution, consistently with the “top-down also” view.

From culture to cognition and brain

Cultural exaptation: how culture affects cognition


Cultural exaptation (Figure 29.3A) is a plausible explanation for the progression in cultural evolution noted
previously and illustrated in Figures 29.1 and 29.2 (Colagè & d’Errico, 2020). Gould and Vrba (1982) intro-
duced the notion of exaptation in the field of evolutionary biology to refer to traits that were not selected for
their current function but evolved either for another function or for no specific function (see also Gould &
Lewontin, 1979). Bird feathers are a classic and clear example of exaptation (Gould & Vrba, 1982): Feathers
first emerged due to the advantage in thermoregulation they conferred but were later “exapted” for flying
and successively refined and diversified for birds’ various flying needs. In the same vein, cultural exaptation
refers to the co-option of existing cultural features for new purposes. Cultural exaptation, as understood

545
Francesco d’Errico and Ivan Colagè
546

Figure 29.3 Schematic representation of the “top-down-also” view. (A) Culture may directly affect cognition via “cultural exaptation,” tr igger ing the emergence
of new cognitive skills without requir ing concomitant genetic changes. (B) Culture may also directly affect brain anatomy via “cultural neural reuse,”
thus favour ing the emergence of novel brain network without requir ing concomitant genetic changes. Reproduced from Colagè and d’Er r ico (2020)
with per mission.
The emergence of symbolic cognition

here, does not require concomitant biological exaptation or heritable genetic changes as a prerequisite but
occurs at the truly cultural level. In analogy with biological exaptation, however, the co-opted cultural
feature was originally devised for purposes quite different from those for which it later became “culturally
exapted.”
An illustrative example of cultural exaptation is the use of ochre (pigment containing hydrated iron
oxides). Compounds made with ochre offer effective photo-protection of the skin from harmful ultra-
violet radiation (Rifkin et al., 2015). There is evidence suggesting symbolic use of iron-rich pigments on
personal ornaments since at least 80 ka in Africa (d’Errico & Backwell, 2016; d’Errico et al., 2009). It is
possible that ochre was first exploited for its photo-protective function and, later, culturally exapted—for
example, in body painting—to reinforce cultural mechanisms related to intra- and inter-group self-iden-
tification, spurring the emergence of diversified symbolic material cultures. A subsequent cultural exapta-
tion may have occurred in an already fully symbolic context, with ochre purposely applied to ornaments
and clothes.
Another key example is provided by incised objects progressively acquiring more conventional symbolic
meanings, leading up to formal artificial memory systems (AMSs)—defined as devices specifically conceived
to store and recover information (e.g., d’Errico, 1998, 2002)—and the final development of number sym-
bols (d’Errico et al., 2017). We have proposed that the invention of number symbols required at least five
not necessarily successive stages or cultural exaptations. The first, “baseline” stage involved the motor and
cognitive skills necessary to produce durable and visible markings on bones with stone tools, an ability highly
likely related to butchery activities and as old as 2.6 Myr (Domínguez-Rodrigo & Alcalá, 2016; Domínguez-
Rodrigo et al., 2005). From this, a first cultural exaptation may have occurred when visible incised patterns,
composed of cut-mark-like lines, were purposely produced on bone or other materials to give rise to detect-
able abstract designs with iconic, indexical, or symbolic functions (Mellet et al., 2019a, 2019b). Engraved
fresh-water shells from Trinil ( Joordens et al., 2015) and the pattern engraved on the Bilzingsleben mam-
moth bone (Mania & Mania, 1988; but see Müller & Pasda, 2011) indicate that this practice may date back
even to 540 ka. An additional exaptation, exemplified by the pattern engraved on the Les Pradelles bone
(d’Errico et al., 2017), may have occurred when meaning was attributed to identical individual marks (rather
than to the whole pattern) produced during the same session. This type of AMS was possibly in use already
in 60 ka, and probably earlier (d’Errico et al., 2017). A third exaptation occurred when similar marks were
incised at different times, which gives the possibility of adding numerical information to an existing pattern.
The Border Cave notched bone (d’Errico et al., 2012, 2017), dated to 44–42 ka, is the earliest known exam-
ple of such a notation type. A fourth exaptation occurred when the morphology of the marks, their spatial
distribution, their number, and their accumulation over time were individually or conjointly given a role in
the code. The earliest known examples of such devices date back to 40–38 ka (d’Errico et al., 2012, 2017).
This fourth exaptation contains most of the features that eventually allowed some human populations to pro-
duce a further exaptation: the invention of the number symbols and numerical notations known historically
and still used today. The invention of symbolic number systems (and of arithmetic) clearly had enormous
consequences on human cognition and culture. This is because a symbolic and conventional system enables
processing exceedingly large quantities.
Cultural exaptation, as exemplified here, provides a mechanism for cultural evolution and for the ensuing
cognitive evolution. According to such a mechanism: (1) culture is a cause (sometimes, the primary or even
the only cause) of human cognitive evolution, and (2) human cognitive evolution does not need concomitant
genetic evolution as its necessary cause (Figure 29.3A). The concept of cultural exaptation differs from that
of cumulative culture in that, consistently with its biological counterpart, it refers to the use of an existing
cultural trait for a new purpose, whereas cumulative culture may proceed via refinement of solutions having
the same function. An additional difference is that cultural exaptation is independent from taxonomic affili-
ation, whereas many authors consider the ratchet effect associated with cumulative culture a characteristic

547
Francesco d’Errico and Ivan Colagè

feature of our own species (Tennie et al., 2009; Tomasello, 2009). Moreover, cultural exaptation allows for
losses of cultural innovations and is not inevitably cumulative.

From culture to the brain: cultural neural reuse


We have shown that the pattern of cultural innovation observed in the last 800,000 years is incompatible with
the idea that the evolution of human cognition was the mere product of genetic change (the “bottom-up
only” view). The cultural innovations related to subsistence, technology and symbolism in Africa, the Near
East, Europe, and Asia that we have reviewed previously, and illustrated in Figure 29.1, suggest an alternative
mechanism driving the evolution of human cognition: “cultural exaptation” (Figure 29.3A).
The emergence of cognitive skills, however, must be accompanied by changes in brain circuits, systems, or
networks able to implement the cognitive functions implicated in novel cultural practices. According to the
“bottom-up-only” view, new brain systems may emerge, fundamentally, only as a consequence of biological
(namely genetic) evolution. Indeed, the formation of new brain networks able to support novel cognitive
skills and cultural practices is usually assumed to require changes at the level of the species-specific genetic-
epigenetic program for the development of the brain gross functional anatomy (which indeed specifies the
distribution of different kinds of neurons in the brain, and the patterns of anatomical connectivity linking
distinct brain areas). As noted previously, recent advances in developmental neuroplasticity indicate that brain
development and maturation is significantly affected by environmental factors (including the social and cul-
tural environment) and the individual’s unique life history. This suggests that certain aspects of brain anatomy,
physiology, and functionality do not mechanistically follow only from the available genetic endowment,
though, of course, the peculiarly high brain plasticity of human beings is the outcome of a long biological
and genetic evolutionary path (Charrier et al., 2012; Somel et al., 2009).
There is increasing evidence that new brain networks (i.e., brain networks not determined by the species-
specific biological development program) may be favoured by cultural practices without the need of any con-
comitant or directly causally related changes in the heritable genetic-epigenetic endowment of individuals.
One such cultural practice is arithmetic (the ability to treat exact quantitative information in a symbolic and
conventional way, noted in the previous section). Another key case is that of literacy (the use of writing and
reading systems, i.e., systems able to translate an existing articulate language in visual symbols).
The development of these two recent cultural innovations seems to owe to the “cultural recycling of
cortical maps” (Dehaene & Cohen, 2007; Hannagan et al., 2015). According to this idea, as an individual
learns such cultural practices, existing brain areas, emerged in our evolutionary past for other purposes,
are reused (recycled or co-opted) as a basis for those new practices. There is consistent evidence that
both literacy and arithmetic recycle existing brain tissue that evolved (1) in the fusiform gyrus for pro-
cessing faces and high-resolution sharp-edged visual shapes in the case of literacy (Dehaene et al., 2015)
and (2) in the intra-parietal sulcus for rough estimation of amounts in the case of arithmetic (Nieder &
Dehaene, 2009). This recycling of existing neural networks to subserve novel cultural practices can involve
non-local, inter-regional re-arrangement of entire brain networks. Acquiring literacy requires not only
the formation of a new functional network that places the left fusiform face area (FFA) at the service of
literacy (Caspers et al., 2013, 2014) but also the strengthening of the anatomical connections of the mid-
fusiform gyrus with other perisylvian regions (like superior-middle temporal and inferior parietal areas)
involved in spoken language (Klingberg et al., 2000; Thiebaut de Schotten et al., 2014; Yeatman et al.,
2012). Acquisition of literacy would thus be a case in which a novel cultural practice not only prompts
new cognitive skills but also directly and significantly affects brain organization at the level of formation
of new anatomical networks.
We refer to this kind of effects of culture on the organization of brain networks as “cultural neural reuse”
(Figure 29.3B), in analogy with our notion of “cultural exaptation,” defined previously. Cultural neural reuse

548
The emergence of symbolic cognition

might provide a further mechanism, this time at the neural level, owing to which newly emerging cultural
practices (via cultural exaptation) may be stabilized within and across human populations by specifying
dedicated brain networks without the need for genetic evolution. Thus, cultural neural reuse contributes to
argue for culture as the driving force of recent human evolution at both the cognitive and the neural level
(Figures 29.3A and 29.3B).

Conclusion
The combination of cultural exaptation and cultural neural reuse provides a general mechanism through
which culture became the leading force of human recent cognitive and neural evolution. This is what we
have labelled the “top-down-also” view. In a nutshell, successive cycles of cultural exaptation and ensuing
cultural neural reuse represent the fundamental mechanism that has regulated the cultural evolution of our
lineage. This mechanism accounts for the asynchronous and patchy emergence, transformation, and disap-
pearance of innovations around the globe. It also explains trends toward complexification of cultural practices
recorded at different times and places in human history. We argue that symbolic cognition emerged following
the same path and coevolved with its material expressions and the different ways in which meanings attached
to symbols were transmitted from one generation to another. Symbolic systems of growing complexity and
their material expression may have stemmed from cultural exaptations. This led to more complex cultural
transmission strategies and novel cognitive tools, necessary to handle more complex, symbolically mediated
worldviews.
It is conceivable that at broader timescales, the innovations stabilized through the proposed mechanism
become somehow “genetically assimilated” (Waddington, 1942, 1953; see also Varki et al., 2008), and this
may well happen via gene-culture co-evolution. However, this would be a subsequent stage that is not
strictly required for significant cultural innovations to become stabilized, spread, and transmitted from gen-
eration to generation. Thus, we do not claim that genetic evolution is irrelevant to human cognitive and/or
cultural evolution. Both cultural exaptation and cultural neural reuse always build upon pre-existing cultural
practices, cognitive skills, and related neural substrates that generally are the outcome of previous biological
and genetic evolution stricto sensu. The review of the archaeological data suggests that biological setting on
which cultural exaptation and cultural neural reuse acted to promote cultural evolution are a characteristic
of our genus rather than of our species.
We have presented evidence showing that culture should be regarded as an effective causal factor in
human evolution, not only in the sense that cultural environments exert specific selective pressures on organ-
isms but also as a source of novelty independent of genetic changes and natural selection. This is compatible
with the idea that cultural evolution may affect genetic evolution (via gene-culture co-evolution) but also
suggests that cultural innovations may stabilize before (or even without) corresponding genetic changes, so
that the accelerating pace of human genetic evolution (Evans et al., 2005; Hawks et al., 2007) may be, at least
in some relevant cases, a nonessential consequence of cultural evolution.

Acknowledgements
We thank Marcos Nadal for his invaluable help during the preparation of this contribution. Francesco
d’Errico’s work is supported by the European Research Council through a Synergy Grant for the project
Evolution of Cognitive Tools for Quantification (QUANTA), No. 951388; the Research Council of Nor-
way through its Centres of Excellence funding scheme, SFF Centre for Early Sapiens Behaviour (SapienCE),
project number 262618, the Talents Programme of the Bordeaux University [grant number: 191022_001]
and the Grand Programme de Recherche ‘Human Past’ of the Initiative d’Excellence (IdEx) of the Bordeaux
University.

549
Francesco d’Errico and Ivan Colagè

References
Anderson, M. L. (2014). After phrenology: Neural reuse and the interactive brain. MIT Press.
Arbib, M. A. (2013). How the brain got language: The mirror system hypothesis. Oxford University Press.
Arbib, M. A. (2016). Towards a computational comparative neuroprimatology: Framing the language-ready brain. Physics
of Life Reviews, 16, 1–54. https://doi.org/10.1016/j.plrev.2015.09.003
Bar-Yosef, O. (1998). On the nature of transitions: The middle to upper palaeolithic and the neolithic revolution. Cam-
bridge Archaeological Journal, 8(2), 141–163. https://doi.org/10.1017/S0959774300001815
Bickerton, D. (2014). More than nature needs: Language, mind and evolution. Harvard University Press.
Bolhuis, J. J., Tattersall, I., Chomsky, N., & Berwick, R. C. (2014). How could language have evolved? PLOS Biology,
12(8), e1001934. https://doi.org/10.1371/journal.pbio.1001934
Bruner, E. (2014). Functional craniology, human evolution, and anatomical constraints in the Neanderthal braincase. In
T. Akazawa, N. Ogihara, H. C. Tanabe, & H. Terashima (Eds.), Dynamics of learning in Neanderthals and modern humans,
volume 2, cognitive and physical perspectives (pp. 121–129). Springer Japan.
Caspers, J., Zilles, K., Amunts, K., Laird, A. R., Fox, P. T., & Eickhoff, S. B. (2014). Functional characterization and
differential coactivation patterns of two cytoarchitectonic visual areas on the human posterior fusiform gyrus. Human
Brain Mapping, 35(6), 2754–2767. https://doi.org/10.1002/hbm.22364
Caspers, J., Zilles, K., Eickhoff, S. B., Schleicher, A., Mohlberg, H., & Amunts, K. (2013). Cytoarchitectonical analysis
and probabilistic mapping of two extrastriate areas of the human posterior fusiform gyrus. Brain Structure and Function,
218(2), 511–526. https://doi.org/10.1007/s00429-012-0411-8
Charrier, C., Joshi, K., Coutinho-Budd, J., Kim, J. E., Lambert, N., de Marchena, J., Jin, W. L., Vanderhaeghen,
P., Ghosh, A., Sassa, T., & Polleux, F. (2012). Inhibition of SRGAP2 function by its human-specific paralogs induces
neoteny during spine maturation. Cell, 149(4), 923–935. https://doi.org/10.1016/j.cell.2012.03.034
Chase, P. G., & Dibble, H. L. (1987). Middle Palaeolithic symbolism: A review of current evidence and interpretations.
Journal of Anthropological Archaeology, 6(3), 263–296. https://doi.org/10.1016/0278-4165(87)90003-1
Colagè, I. (2016). The cultural evolution of language and brain: Comment on “Towards a computational compara-
tive neuroprimatology: Framing the language-ready brain” by Michael A. Arbib. Physics of Life Reviews, 16, 61–62.
https://doi.org/10.1016/j.plrev.2016.01.013
Colagè, I., & D’Ambrosio, P. (2014). Exaptation and neural reuse: A research perspective into the human specificity.
Antonianum, 89, 333–358.
Colagè, I.,  & d’Errico, F. (2020). Culture: The driving force of human cognition. Topics in Cognitive Science, 12(2),
654–672. https://doi.org/10.1111/tops.12372
Coolidge, F. L., & Wynn, T. (2004). A cognitive and neuropsychological perspective on the Châtelperronian. Journal of
Anthropological Research, 60(1), 55–73. https://doi.org/10.1086/jar.60.1.3631008
Coolidge, F. L., & Wynn, T. (2007). The working memory account of Neandertal cognition—How phonological stor-
age capacity may be related to recursion and the pragmatics of modern speech. Journal of Human Evolution, 52(6),
707–710. https://doi.org/10.1016/j.jhevol.2007.01.003
Coolidge, F. L., & Wynn, T. (2018). The rise of Homo sapiens: The evolution of modern thinking. Oxford University Press.
Corruccini, R. S. (1984). An epidemiologic transition in dental occlusion in world populations. American Journal of Ortho-
dontics, 86(5), 419–426. https://doi.org/10.1016/S0002-9416(84)90035-6
D’Ambrosio, P., & Colagè, I. (2017). Extending epigenesis: From phenotypic plasticity to the bio-cultural feedback. Biol-
ogy and Philosophy, 32(5), 705–728. https://doi.org/10.1007/s10539-017-9581-3
d’Errico, F. (1998). Palaeolithic origins of artificial memory systems: An evolutionary perspective. In C. Renfrew & C.
Scarre (Eds.), Cognition and material culture: The archaeology of symbolic storage (pp. 19–50). The McDonald Institute
[Monographs].
d’Errico, F. (2002). Memories out of mind: The archaeology of the oldest artificial memory systems. In A. Nowell (Ed.),
In the mind’s eye. Archaeological Series 13 (pp. 33–49). International Monographs in Prehistory.
d’Errico, F. (2003). The invisible frontier. A multiple species model for the origin of behavioral modernity. Evolutionary
Anthropology, 12(4), 188–202. https://doi.org/10.1002/evan.10113
d’Errico, F., & Backwell, L. (2016). Earliest evidence of personal ornaments associated with burial: The Conus shells from
border cave. Journal of Human Evolution, 93, 91–108. https://doi.org/10.1016/j.jhevol.2016.01.002
d’Errico, F., Backwell, L., Villa, P., Degano, I., Lucejko, J. J., Bamford, M. K., Higham, T. F., Colombini, M. P., &
Beaumont, P. B. (2012). Early evidence of San material culture represented by organic artifacts from border cave,
South Africa. Proceedings of the National Academy of Sciences of the United States of America, 109(33), 13214–13219.
https://doi.org/10.1073/pnas.1204213109
d’Errico, F., & Colagè, I. (2018). Cultural exaptation and cultural neural reuse: A mechanism for the emergence of mod-
ern culture and behavior. Biological Theory, 13(4), 213–227. https://doi.org/10.1007/s13752-018-0306-x

550
The emergence of symbolic cognition

d’Errico, F., Doyon, L., Colagè, I., Queffelec, A., Le Vraux, E., Giacobini, G., Vandermeersch, B.,  & Maureille, B.
(2017). From number sense to number symbols. An archaeological perspective. Philosophical Transactions of the Royal
Society of London. Series B, Biological Sciences, 373(1740), 20160518. https://doi.org/10.1098/rstb.2016.0518
d’Errico, F., & Stringer, C. B. (2011). Evolution, revolution or saltation scenario for the emergence of modern cultures?
Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 366(1567), 1060–1069. https://doi.
org/10.1098/rstb.2010.0340
d’Errico, F., Vanhaeren, M., Barton, N., Bouzouggar, A., Mienis, H., Richter, D., Hublin, J. J., McPherron, S. P., &
Lozouet, P. (2009). Out of Africa: Modern human origins special feature: Additional evidence on the use of personal
ornaments in the middle paleolithic of North Africa. Proceedings of the National Academy of Sciences of the United States
of America, 106(38), 16051–16056. https://doi.org/10.1073/pnas.0903532106
Dehaene, S., & Cohen, L. (2007). Cultural recycling of cortical maps. Neuron, 56(2), 384–398. https://doi.org/10.1016/j.
neuron.2007.10.004
Dehaene, S., Cohen, L., Morais, J., & Kolinsky, R. (2015). Illiterate to literate: Behavioural and cerebral changes induced
by reading acquisition. Nature Reviews. Neuroscience, 16(4), 234–244. https://doi.org/10.1038/nrn3924
Dirks, P. H., Berger, L. R., Roberts, E.M., Kramers, J. D., Hawks, J., Randolph-Quinney, P. S., Elliott, M., Musiba, C.
M., Churchill, S. E., de Ruiter, D. J., & Schmid, P. (2015). Geological and taphonomic context for the new hominin
species Homo naledi from the Dinaledi Chamber, South Africa. eLife, 4, e09561.
Dirks, P. H., Roberts, E. M., Hilbert-Wolf, H., Kramers, J. D., Hawks, J., Dosseto, A., Duval, M., Elliott, M., Evans,
M., Grün, R., & Hellstrom, J. (2017). The age of Homo naledi and associated sediments in the Rising Star Cave,
South Africa. eLife, 6, e24231.
Domínguez-Rodrigo, M,  & Alcalá, L. (2016). 3.3-million-year-old stone tools and butchery traces? More evidence
needed. PaleoAnthropology, 2016, 46–53. https://doi.org/10.4207/PA.2016.ART99
Domínguez-Rodrigo, M., Pickering, T. R., Semaw, S., & Rogers, M. J. (2005). Cutmarked bones from Pliocene archae-
ological sites at Gona, Afar, Ethiopia: Implications for the function of the world’s oldest stone tools, Afar. Journal of
Human Evolution, 48(2), 109–121. https://doi.org/10.1016/j.jhevol.2004.09.004
Dor, D., & Jablonka, E. (2014). Why we need to move from gene–culture co-evolution to culturally driven co-evolution.
In D. Dor, K. Knight & J. Lewis (Eds.), Social origins of language (pp. 15–30). Oxford University Press.
Draganski, B.,  & May, A. (2008). Training-induced structural changes in the adult human brain. Behavioural Brain
Research, 192(1), 137–142. https://doi.org/10.1016/j.bbr.2008.02.015
Evans, N., & Levinson, S. C. (2009). The myth of language universals: Language diversity and its importance for cogni-
tive science. Behavioral and Brain Sciences, 32(5), 429–48; discussion 448. https://doi.org/10.1017/S0140525X099
9094X
Evans, P. D., Gilbert, S. L., Mekel-Bobrov, N., Vallender, E. J., Anderson, J. R., Vaez-Azizi, L. M., Tishkoff, S. A.,
Hudson, R. R., & Lahn, B. T. (2005). Microcephalin, a gene regulating brain size, continues to evolve adaptively in
humans. Science, 309(5741), 1717–1720. https://doi.org/10.1126/science.1113722
Feldman, M. W., & Cavalli-Sforza, L. L. (1976). Cultural and biological evolutionary processes, selection for a trait under
complex transmission. Theoretical Population Biology, 9(2), 238–259. https://doi.org/10.1016/0040-5809(76)90047-2
Fisher, S. E. (2017). Evolution of language: Lessons from the genome. Psychonomic Bulletin and Review, 24(1), 34–40.
https://doi.org/10.3758/s13423-016-1112-8
Fu, Q., Posth, C., Hajdinjak, M., Petr, M., Mallick, S., Fernandes, D., Furtwängler, A., Haak, W., Meyer, M., Mittnik,
A., & Nickel, B. (2016). The genetic history of Ice Age Europe. Nature, 534, 200–205.
Geschwind, D. H., & Rakic, P. (2013). Cortical evolution: Judge the brain by its cover. Neuron, 80(3), 633–647. https://
doi.org/10.1016/j.neuron.2013.10.045
Gintis, H. (2011). Gene-culture coevolution and the nature of human sociality. Philosophical Transactions of the Royal Society
of London. Series B, Biological Sciences, 366(1566), 878–888. https://doi.org/10.1098/rstb.2010.0310
Gould, S. J., & Lewontin, R. C. (1979). The spandrels of San Marco and the Panglossian paradigm: A critique of the
adaptationist programme. Proceedings of the Royal Society of London Series B, 205, 581–598.
Gould, S. J., & Vrba, E. S. (1982). Exaptation—A missing term in the science of form. Paleobiology, 8(1), 4–15. https://
doi.org/10.1017/S0094837300004310
Hammer, M., Woerner, A.E., Mendez, F. L., Watkins, J. C., & Wall, J. D. (2011). Genetic evidence for archaic admixture
in Africa. Proceedings of the National Academy of Sciences USA, 108, 15123–15128.
Hannagan, T., Amedi, A., Cohen, L., Dehaene-Lambertz, G., & Dehaene, S. (2015). Origins of the specialization for
letters and numbers in ventral occipitotemporal cortex. Trends in Cognitive Sciences, 19(7), 374–382. https://doi.
org/10.1016/j.tics.2015.05.006
Hawks, J., Wang, E. T., Cochran, G. M., Harpending, H. C., & Moyzis, R. K. (2007). Recent acceleration of human
adaptive evolution. Proceedings of the National Academy of Sciences of the United States of America, 104(52), 20753–20758.
https://doi.org/10.1073/pnas.0707650104

551
Francesco d’Errico and Ivan Colagè

Hecht, E. E., Gutman, D. A., Khreisheh, N., Taylor, S. V., Kilner, J., Faisal, A. A., Bradley, B. A., Chaminade, T., &
Stout, D. (2015). Acquisition of paleolithic toolmaking abilities involves structural remodeling to inferior frontopari-
etal regions. Brain Structure and Function, 220(4), 2315–2331. https://doi.org/10.1007/s00429-014-0789-6
Henshilwood, C. S.,  & Marean, C. W. (2003). The origin of modern human behavior. Current Anthropology, 44(5),
627–651. https://doi.org/10.1086/377665
Herculano-Houzel, S. (2017). Numbers of neurons as biological correlates of cognitive capability. Current Opinion in
Behavioral Sciences, 16, 1–7. https://doi.org/10.1016/j.cobeha.2017.02.004
Heyes, C. (2012). Grist and mills: On the cultural origins of cultural learning. Philosophical Transactions of the Royal Society
of London. Series B, Biological Sciences, 367(1599), 2181–2191. https://doi.org/10.1098/rstb.2012.0120
Heyes, C. (2018). Cognitive gadgets: The cultural evolution of thinking. Harvard University Press.
Hoffmann, D. L., Standish, C. D., García-Diez, M., Pettitt, P. B., Milton, J. A., Zilhão, J., Alcolea-González, J. J.,
Cantalejo-Duarte, P., Collado, H., de Balbín, R., Lorblanchet, M., Ramos-Muñoz, J., Weniger, G.-Ch., & Pike, A.
W. G. (2018). U-Th dating of carbonate crusts reveals Neandertal origin of Iberian cave art. Science, 359, 912–915.
Jablonka, E., & Lamb, M. J. (2005). Evolution in four dimensions: Genetic, epigenetic, behavioral, and symbolic variation in the
history of life. MIT Press.
Johansson, S. (2014). The thinking Neanderthals: What do we know about Neanderthal cognition? Wiley Interdisciplinary
Reviews. Cognitive Science, 5(6), 613–620. https://doi.org/10.1002/wcs.1317
Johnson, M. H. (2001). Functional brain development in humans. Nature Reviews. Neuroscience, 2(7), 475–483. https://
doi.org/10.1038/35081509
Johnson, M. H. (2011). Interactive specialization: A domain-general framework for human functional brain develop-
ment? Developmental Cognitive Neuroscience, 1(1), 7–21. https://doi.org/10.1016/j.dcn.2010.07.003
Joordens, J. C. A., d’Errico, F., Wesselingh, F. P., Munro, S., De Vos, J., Wallinga, J., Ankjærgaard, C., Reimann, T.,
Wijbrans, J. R., Kuiper, K. F., Mücher, H. J., Coqueugniot, H., Prié, V., Joosten, I., van Os, B., Schulp, A. S., Panuel,
M., van der Haas, V., Lustenhouwer, W., Reijmer, J. J. G., & Roebroeks, W. (2015). Homo erectus at Trinil on Java
used shells for tool production and engraving. Nature, 518, 228–231.
Klein, R. G. (1989). The human career. University of Chicago Press.
Klein, R. G. (2000). Archeology and the evolution of human behavior. Evolutionary Anthropology: Issues, News, and
Reviews, 9(1), 17–36. https://doi.org/10.1002/(SICI)1520-6505(2000) < 17:AID-EVAN3 > 3.0.CO.
Klingberg, T., Hedehus, M., Temple, E., Salz, T., Gabrieli, J. D. E., Moseley, M. E., & Poldrack, R. A. (2000). Micro-
structure of temporo-parietal white matter as a basis for reading ability: Evidence from diffusion tensor magnetic
resonance imaging. Neuron, 25(2), 493–500. https://doi.org/10.1016/S0896-6273(00)80911-3
Laland, K. N. (2017). Darwin’s unfinished symphony: How culture made the human mind. Princeton University Press.
Laland, K. N., Odling-Smee, J., & Myles, S. (2010). How culture shaped the human genome: Bringing genetics and the
human sciences together. Nature Reviews. Genetics, 11(2), 137–148. https://doi.org/10.1038/nrg2734
Lyko, F., Foret, S., Kucharski, R., Wolf, S., Falckenhayn, C., & Maleszka, R. (2010). The honey bee epigenomes: Differ-
ential methylation of brain DNA in queens and workers. PLOS Biology, 8(11), e1000506. https://doi.org/10.1371/
journal.pbio.1000506
Majkić, A., d’Errico, F., Milošević, S., Mihailović, D., & Dimitrijević, V. (2017). Sequential incisions on a cave bear bone
from the Middle Paleolithic of Pešturina Cave, Serbia. Journal of Archaeological Method and Theory, 25, 69–116.
Mania, D., & Mania, U. (1988). Deliberate engravings on bone artefacts of Homo erectus. Rock Art Research, 5(2), 9–107.
Maricic, T., Günther, V., Georgiev, O., Gehre, S., Curlin, M., Schreiweis, C., Naumann, R., Burbano, H. A., Meyer,
M., Lalueza-Fox, C., de la Rasilla, M., Rosas, A., Gajovic, S., Kelso, J., Enard, W., Schaffner, W., & Pääbo, S. (2013).
A recent evolutionary change affects a regulatory element in the human FOXP2 gene. Molecular Biology and Evolution,
25, 1257–1259. https://doi.org/10.1093/molbev/mss271
McBrearty, S., & Brooks, A. S. (2000). The revolution that wasn’t: A new interpretation of the origins of modern human
behavior. Journal of Human Evolution, 39(5), 453–563. https://doi.org/10.1006/jhev.2000.0435
Mellars, P. A. (1989). Major issues in the emergence of modern humans. Current Anthropology, 30(3), 349–385. https://
doi.org/10.1086/203755
Mellars, P. A., & Stringer, C. B. (Eds.). (1989). The human revolution: Behavioral and biological perspectives on the origins of
modern humans. Edinburgh University Press.
Mellet, E., Colagè, I., Bender, A., Henshilwood, C. S., Hugdahl, K., Lindstrøm, T. C., & d’Errico, F. (2019a). What pro-
cesses sparked off symbolic representations? A reply to Hodgson and an alternative perspective. Journal of Archaeological
Science: Reports, 28, 102043. https://doi.org/10.1016/j.jasrep.2019.102043
Mellet, E., Salagnon, M., Majkić, A., Cremona, S., Joliot, M., Jobard, G., Mazoyer, B., Tzourio Mazoyer, N., & d’Errico,
F. (2019b). Dog behaviour, evolution, and cognition. Neuroimaging supports the representational nature of the

552
The emergence of symbolic cognition

earliest human engravings. Royal Society Open Science. Oxford University Press, 6(7), 190086.Miklosi. https://doi.
org/10.1098/rsos.190086
Miklosi, A. (2014). Dog behaviour, evolution, and cognition. Oxford, UK: Oxford University Press.
Mithen, S. J. (2005). The singing Neanderthals: The origins of music, language, mind and body. Weidenfeld & Nicolson.
Moss, C. (2018). Neuroethological studies of cognitive and perceptual processes. Routledge.
Müller, W., & Pasda, C. (2011). Site formation and faunal remains of the Middle Pleistocene site Bilzingsleben. Quartär,
58, 25–49. https://doi.org/10.7485/QU58_02
Neubauer, S., Hublin, J. J., & Gunz, P. (2018). The evolution of modern human brain shape. Science Advances, 4(1),
eaao5961. https://doi.org/10.1126/sciadv.aao5961
Nieder, A., & Dehaene, S. (2009). Representation of number in the brain. Annual Review of Neuroscience, 32, 185–208.
https://doi.org/10.1146/annurev.neuro.051508.135550
Pascual-Leone, A., Amedi, A., Fregni, F., & Merabet, L. B. (2005). The plastic human brain cortex. Annual Review of
Neuroscience, 28, 377–401. https://doi.org/10.1146/annurev.neuro.27.070203.144216
Pettitt, P. (2011). The palaeolithic origins of human burial. Routledge: New York.
Pfeiffer, J. E. (1982). The creative explosion: An enquiry into the origins of art and religion. Harper & Row.
Price, T., Kirkpatrick, M., & Arnold, S. J. (1988). Directional selection and the evolution of breeding date in birds. Sci-
ence, 240(4853), 798–799. https://doi.org/10.1126/science.3363360
Reader, S., & Laland, K. N. (2003). Animal innovation: An introduction. In S. Reader & K. N. Laland (Eds.), Animal
innovation (pp. 1–9). Oxford University Press.
Richerson, P. J., Boyd, R., & Henrich, J. (2010). Colloquium paper: Gene-culture coevolution in the age of genomics.
Proceedings of the National Academy of Sciences of the United States of America, 107(Suppl. 2), 8985–8992. https://doi.
org/10.1073/pnas.0914631107
Rifkin, R. F., Dayet, L., Queffelec, A., Summers, B., Lategan, M., & d’Errico, F. (2015). Evaluating the photoprotective
effects of ochre on human skin by in vivo SPF assessment: Implications for human evolution, adaptation and dispersal.
PLoS One, 10(9), e0136090. https://doi.org/10.1371/journal.pone.0136090
Rodríguez-Vidal, J., d’Errico, F., Pacheco, F. G., Blasco, R., Rosell, J. J., Jennings, R. P., Queffelec, A., Finlayson, G.,
Fa, D. A., Gutiérrez-López, J. M., Carrión, J. S., Negro, J. J., Finlayson, S., Cáceres, L. M., Bernal, M. A., Jiménez,
S. F., & Finlayson, C. (2014). A rock engraving made by Neanderthals in Gibraltar. Proceedings of the National Academy
of Sciences USA, 111, 13301–13306
Santolin, C., & Saffran, J. R. (2018). Constraints on statistical learning across species. Trends in Cognitive Sciences, 22(1),
52–63. https://doi.org/10.1016/j.tics.2017.10.003
Shea, J. J. (2011). Homo sapiens is as homo sapiens was: Behavioral variability versus “behavioral modernity” in paleo-
lithic archaeology. Current Anthropology, 52(1), 1–35. https://doi.org/10.1086/658067
Shettleworth, S. J. (2010). Cognition, evolution, and behaviour. Oxford University Press.
Somel, M., Franz, H., Yan, Z., Lorenc, A., Guo, S., Giger, T., Kelso, J., Nickel, B., Dannemann, M., Bahn, S., Webster,
M. J., Weickert, C. S., Lachmann, M., P€a€abo, S., & Khaitovich, P. (2009). Transcriptional neoteny in the human
brain. Proceedings of the National Academy of Sciences of the United States of America, 106(14), 5743–5748. https://doi.
org/10.1073/pnas.0900544106
Somel, M., Liu, X.,  & Khaitovich, P. (2013). Human brain evolution: Transcripts, metabolites and their regulators.
Nature Reviews. Neuroscience, 14(2), 112–127. https://doi.org/10.1038/nrn3372
Soressi, M., McPherron, S. P., Lenoir, M., Dogandžić, T., Goldberg, P., Jacobs, Z., Maigrot, Y., Martisius, N. L., Miller,
C. E., Rendu, W., & Richards, M. (2013). Neandertals made the first specialized bone tools in Europe. Proceedings of
the National Academy of Sciences USA, 110, 14186–14190.
Stiner, M. C. (2017). Love and death in the stone age: What constitutes first evidence of mortuary treatment of the
human body? Biological Theory, 12, 248–261.
Tang, G. H., Rabie, A. B. M., & Hägg, U. (2004). Indian hedgehog: A mechanotransduction mediator in condylar car-
tilage. Journal of Dental Research, 83(5), 434–438. https://doi.org/10.1177/154405910408300516
Tattersall, I. (1995). The fossil trail: How we know what we think we know about human evolution. Oxford University Press.
Tennie, C., Call, J.,  & Tomasello, M. (2009). Ratcheting up the ratchet: On the evolution of cumulative culture.
Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 364(1528), 2405–2415. https://doi.
org/10.1098/rstb.2009.0052
Thiebaut de Schotten, M., Cohen, L., Amemiya, E., Braga, L. W., & Dehaene, S. (2014). Learning to read improves the
structure of the arcuate fasciculus. Cerebral Cortex, 24(4), 989–995. https://doi.org/10.1093/cercor/bhs383
Tinbergen, N. (1951). The study of instinct. Clarendon Press/Oxford University Press.
Tomasello, T. (2009). The cultural origins of human cognition. Harvard University Press.

553
Francesco d’Errico and Ivan Colagè

Vanhaeren, M., Julien, M., d’Errico, F., Mourer-Chauviré, C., & Lozouet, P. (2019). Les objets de parure. In M. Julien
(Ed.), Le Châtelperronien de la Grotte du Renne (Arcy-Sur-Cure) (pp. 259–285).
Varki, A., Geschwind, D. H., & Eichler, E. E. (2008). Explaining human uniqueness: Genome interactions with environ-
ment, behaviour and culture. Nature Reviews. Genetics, 9(10), 749–763. https://doi.org/10.1038/nrg2428
Varrela, J. (1992). Dimensional variation of craniofacial structures in relation to changing masticatoryfunctional demands.
European Journal of Orthodontics, 14(1), 31–36. https://doi.org/10.1093/ejo/14.1.31
Varrela, J. (2006). Masticatory function and malocclusion: A clinical perspective. Seminars in Orthodontics, 12(2), 102–109.
https://doi.org/10.1053/j.sodo.2006.01.003
Villa, P.,  & Roebroeks, W. (2014). Neandertal demise: An archaeological analysis of the modern human superiority
complex. PLoS One, 9(4), e96424. https://doi.org/10.1371/journal.pone.0096424
Waddington, C. H. (1942). Canalization of development and the inheritance of acquired characters. Nature, 150(3811),
563–565. https://doi.org/10.1038/150563a0
Waddington, C. H. (1953). Genetic assimilation of an acquired character. Evolution, 7(2), 118–126. https://doi.
org/10.1111/j.1558-5646.1953.tb00070.x
Wynn, T., Overmann, K. A., & Coolidge, F. L. (2016). The false dichotomy: A refutation of the Neandertal indistin-
guishability claim. Journal of Anthropological Sciences, 94, 201–221. https://doi.org/10.4436/JASS.94022
Yeatman, J. D., Dougherty, R. F., Ben-Shachar, M., & Wandell, B. A. (2012). Development of white matter and reading
skills. Proceedings of the National Academy of Sciences of the United States of America, 109(44), E3045–E3053. https://doi.
org/10.1073/pnas.1206792109
Zilhão, J. (2016). Lower and middle Palaeolithic mortuary behaviours and the origins of ritual burial. In C. Renfrew, M.
J. Boyd, & I. Morley (Eds.), Death rituals, social order and the archaeology of immortality in the ancient world (pp. 27–44).
Cambridge: Cambridge University Press.

554
30
NEUROPSYCHOLOGY OF ART
AND AESTHETICS
Alejandro Dorado and Marcos Nadal

Neuropsychology is a scientific discipline at the intersection of neurology, psychology, neuroanatomy, and


neurophysiology (Benton, 2000). Neuropsychology has two basic aims. On the one hand, it studies the
changes in behaviour and cognition brought on by brain damage and degenerative diseases with the purpose
of diagnosis and treatment. On the other hand, it studies these cases as windows into the architecture of cog-
nition and its relation to neural structure (Vallar, 2000). When applied to art and aesthetics, thus, neuropsy-
chology studies the effects of brain damage and diseases on artistic and aesthetic production and appreciation
with the goals of understanding their neuroanatomy and of using them for diagnosis and treatment.
Neuropsychology emerged in the second half of the 19th century from debates about the cortical locali-
zation of cognition and, specifically, from observations of the effects of brain lesions on language production
and comprehension (Benton, 2000; Gross, 1999; Vallar, 2000). This “golden age” of cerebral localization was
the fertile grounds for the first ideas about the neuroanatomical foundations of art and aesthetics (Benton,
2000; Finger, 1994). Most of neuropsychology’s founders, including Bouillaud, Broca, and John Hughlings
Jackson, recorded—and were struck by—cases in which people had lost the capacity to speak but could still
use words when singing. Their ideas about aphasia and the neuroanatomy of language were sharpened by
this apparent dissociation between song and speech.

The birth of the neuropsychology of art and aesthetics


Bring your hand to your temple, right next to your orbit, to the point at which, if you wore glasses, their
frame front would meet their temples. Feel around for a bulge or dimple. This is where Franz Joseph Gall,
in the early decades of the 19th century, would have searched for indications of your capacity to perceive
the relations among musical tones and your talent for music. A few centimetres to the back and to the top is
where he would feel about for your talent for poetry. He would assess your talent for painting and for colour
relations by feeling your orbital bones, under the middle of your eyebrow.
Gall speculated that the cerebral cortex could be divided into 27 faculties and that their development
in different people depended on the size of their corresponding cortical region, which could be measured
through prominences and depressions on the skull (Finger, 1994; Wickens, 2015). Among those 27, he
posited the existence of a faculty for perceiving the relation of tones, or the talent for music; a talent for
poetry; and a faculty of distinguishing the relation of colours, or the talent for painting (Eling et al., 2015).
Gall and other popularizers of phrenology in the early 19th century were met with hostility from the

DOI: 10.4324/9781003008675-32 555


Alejandro Dorado and Marcos Nadal

experimental physiologists, who criticized their reliance on anecdotal observations, lack of experimental
data, and sweeping claims about cortical localization. One of the most notable critics was Marie-Jean Pierre
Flourens (1824), who conducted a series of experiments on the effects of brain lesions on animal behaviour
that showed that lesions to the cerebral hemispheres produced detrimental effects on motivation, memory,
and perception. However, his observations led him to believe that, except for vision, the site of a lesion
was not determinant, indicating that these functions were not located in any particular region of the cortex
(Finger, 1994; Gross, 1999).
Although Gall’s method of observation of variations in behaviour and association with cranial measure-
ments rested on speculation rather than fact, his doctrine of organology had a substantial impact on the
study of the neuroanatomical foundations of behaviour and cognition: “while essentially wrong, [it] was just
enough right to further scientific thought” (Boring, 1950, p. 57). The accumulation of clinical observations
and electrical stimulation and ablation experiments on mammals throughout the 19th century cast the idea
of functional localization in a new light (Finger, 2010). Jean-Baptiste Bouillaud’s studies of the effects of
brain damage revealed that in most cases of speech disturbance, there was damage to the frontal lobe. These
patients could understand language and move their tongues and mouths normally when eating, so the dam-
age seemed to produce specific impediments to speech production. Bouillaud concluded, in 1825, that the
brain centre for speech production was in the frontal lobes (Finger, 2010; Wickens, 2015). The fact, however,
that there were many patients with frontal lobe damage who retained the capacity to produce language led
to questions about this conclusion.
Paul Broca continued Bouillaud’s line of research, studying the relation between brain damage and lan-
guage. In 1861, he saw a patient who had suffered from epilepsy for years, had a paralysis and loss of sensitiv-
ity on the right side of his body, and had lost his capacity to speak. Less than a week later, the patient died,
and the autopsy revealed brain damage in the third frontal convolution of the left hemisphere caused by a
stroke. Broca’s careful documentation of the symptoms and localization of the lesions in this and several other
cases, presented in 1863, together with Fritsch and Hitzig’s 1870 identification for a brain region for motor
control, persuaded the scientific community that certain cognitive functions indeed had a specific localiza-
tion in the brain (Finger, 2010; Wickens, 2015).
Gall’s speculations had a strong influence on 19th-century ideas about the cortical localization of capaci-
ties for art and aesthetics. Gall assumed that talent for music, poetry, and paining were distinct faculties, with
their own specific processes of perception and cognition, not the product of general cognitive processes. And
these distinct faculties were located in specific regions of the neocortex, nothing like a distributed network
of processing nodes. These views changed with work on aphasia and the new way in which cortical localiza-
tion of function was thought of.
Observations on patients with loss of speech revealed that musical abilities were sometimes spared, which
led to debates in the 1880s about the relation between language and speech and their neural correlates
(Eling et al., 2015). Interest in music initially resided in its usefulness in assessments of aphasia, as an aid to
determine the extent to which language was impaired (Eling et al., 2015; Johnson et al., 2010). It was strik-
ing to find patients who could sing words that they could not repeat or verbalize in conversation. Eminent
neurologists were intrigued by the relation between music and aphasia and published several cases of adults
and children who could verbalize only a few words but who could nonetheless sing parts of songs (Bouil-
laud, 1865; Jackson, 1871; Trousseau, 1865; Wernicke, 1874). These early cases showed that different aspects
of musical cognition could be impaired while preserving others and that music rested on a complex set of
abilities, some of which were related to language, while others were not ( Johnson et al., 2010). Jackson, for
instance, believed these cases showed that “intellectual language” could be impaired, while “emotional lan-
guage,” which included the ability to hum and sing, could be relatively preserved (1878, 1879).
Advances in the understanding of the neuroanatomy of language paved the way for new thinking about
the neural foundations of music. August Knoblauch (1888) followed Lichtheim’s (1885) ideas on language to

556
Neuropsychology of art and aesthetics

propose a model of the neural centres and connecting pathways underlying different aspects of music cogni-
tion, including music ideation, the motor and auditory centres for tones, and the motor and visual centres
for music notation (Graziano & Johnson, 2015; Johnson et al., 2010). He coined the term amusia to refer
to a specific type of impairment to the motor centre that produced the movements required for singing.
Subsequently, different forms of amusia (loss of the ability to sing, to read or write musical notation, to play
an instrument, and loss of recognition of melodies) were described, studied, and associated with damage to
different centres or fibres: “by the close of the nineteenth century, every form of amusia that is currently
recognized had been described” (Benton, 2000, p. 246).
By the end of the 19th century, ideas about music and its neuroanatomy had become much more sophis-
ticated than Gall’s. Music was now understood as composed of several perceptual, motor, and cognitive
systems that interacted to produce the creation and appreciation of music. It was also understood that these
were located in different regions that were connected forming a sort of network. As the 19th century gave
way to the 20th, there was a surge in studies examining the brains of accomplished musicians in search of
anatomical clues of musical talent, of studies on the relations between impairments of language and music,
and of studies on the impairment in musical experts and nonexperts to these components caused by brain
lesions. The German neurologist Siegmund Auerbach, for instance, described structural peculiarities in the
brains of accomplished musicians and composers, including surface expansion, markedness, and design of
gyri in temporal and parietal lobes (Auerbach, 1906, 1908, 1911, 1913).
Interest in the relations between the anatomy of language and music continued throughout the 20th
century. Dupré and Nathan (1911) reviewed the impact that aphasia and several forms of psychopathology
had on the production and appreciation of music. Their results seemed inconclusive: “The anatomo-clinical
method is unable to identify areas of musical language: the lesions observed in autopsies are often widespread
and do not allow precise localization” (Dupré & Nathan, 1911, p. 184). Two decades later, Souques and
Baruk (1930) described the case of a piano professor who suffered from a severe case of Wernicke’s aphasia.
His spontaneous musical execution was correct, though limited to few tunes, and his performance of basic
musical exercises that constitute the basis of professional automatism was generally accurate. His auditory
recognition of popular tunes was, however, very limited, and he could not reproduce them after hearing
them, playing a different tune, though with a similar rhythmical pattern. Overall, musical perception was
significantly disturbed. The patient could not read written words, yet he was still able to read music scores
and play them correctly. Souques and Baruk (1930) believed that this and similar cases allowed postulating
partially different neural correlates for language and music.
By comparing the production of a writer, a musician, and a painter before and after being struck by apha-
sia Alajouanine (1948) attempted to determine whether such conditions influence the work of great artists in
different domains. His analysis revealed that the brain damage had virtually wiped the writer and musician’s
creative capacity out, but not the painter’s. This suggests that aphasia is especially devastating when the artist’s
expressive means involves language or a symbolic system supported by language, such as musical notation.
Surprisingly, though, aesthetic sensibility was unimpaired in all three artists, as well as their capacity to detect
compositional problems in their own domains.
Luria and colleagues (1965) analysed the effects of a vascular lesion affecting left hemisphere speech
regions suffered by a great composer. Their results suggest that, in some cases at least, musical abilities and
creativity can be preserved despite severe linguistic impairment. Can similar expressive abilities in the visual
modality survive brain damage to language-related regions? Gourevitch’s (1967) and Zaimov et al.’s (1969)
reports suggested that they can. Gourevitch (1967) described the case of an art professor who suffered a left-
hemisphere stroke that left him aphasic, yet he continued to communicate actively with those around him
through graphic designs and symbols, suggesting that drawing is possible even when suffering a severe case
of aphasia. Zaimov and colleagues (1969) described the case of the Bulgarian painter Zlatyo Boyadjiev, who
suffered a left hemisphere stroke, which produced a right-sided hemiplegia and a mostly expressive aphasia.

557
Alejandro Dorado and Marcos Nadal

To resume his work, he was forced to learn to paint with his left hand, which led to a striking alteration in his
style. In addition to the simplification attributable to this change in hand, he began using bright colours and
switched from naturalistic themes to unrealistic and strange imagery, a new style that critics judged favour-
ably. Even 2 years after the stroke, however, his vocabulary was still under 100 words.

Neuropsychology of art production


Clinical case descriptions of artists suffering brain lesions or diseases have continued to grow over the last
decades, forming an extensive field. In fact, most of what is known about the effects of brain damage on
artistic production comes from single case reports. Together, these studies show that, despite their proficient
visuo-motor and musical skills, artists are vulnerable to the same visual, motor, auditory, and cognitive neu-
ropsychological deficits that affect other people. The difference, in Chatterjee’s (2004) words, is that artists
manifest these deficits in strikingly eloquent ways. Most of these artists continued to be artistically motivated,
productive, and expressive after the onset of their condition, probably due to the years devoted to practicing
their skills (Zaidel, 2005). In most of the cases described in the literature, neither left- nor right-hemisphere
damage completely impaired painters or composers’ ability to produce works of art. The most common
effect is a change in style (Heller, 1994). There is no direct relation between neuropsychological conditions
and quality of art production. In fact, Chatterjee (2004, 2006) notes several instances in which the effects the
condition had on the artworks were aesthetically surprising and pleasing.

Effects of focal brain damage on artistic production


Most studies of focal brain damage on artistic production report cases of stroke patients. Strokes are the result
of poor blood flow to certain brain regions that causes cells in those regions to die. The symptoms vary from
one brain region to another, but the most common include loss of vision in one field, impaired movement or
feeling on one side of the body, and problems producing or understanding language. These symptoms often
appear soon after the stroke, and they can be transient or permanent, depending on many factors. There
usually is a noticeable change in the work of most artists who have suffered a stroke. Many of them resume
their creations, though only after overcoming various forms of disability. Some of these artists had to change
working with their dominant to their non-dominant hand (Bäzner & Hennerici, 2006).
Although lesions in either hemisphere seem to leave traces on the production of artists (Zaidel, 2005),
spatial disorganization (perspective, third dimension, contours), neglect and distortion of representation is
more obvious in the artworks of artists who suffered a right hemisphere stroke (Bäzner & Hennerici, 2006),
such as the German artist Lovis Corinth (Blanke & Pasqualini, 2011). Corinth suffered a right hemisphere
stroke when he was 53, an age at which he was famous for his impressionist-like style of portraits, interiors,
still lives, and landscapes. The stroke caused left-side motor and visual problems, from which he recovered
over the years. Corinth’s style changed after the stroke, becoming more expressionistic, less balanced, and
flatter. Some of the paintings he completed soon after the stroke omit details on the left, and the spatial rela-
tions seem distorted (Blanke & Pasqualini, 2011; Chatterjee, 2015).
Loring Hughes, like Corinth, is an artist who continued painting after damage to the right side of her
brain. She was 28 when she experienced a series of strokes that lesioned close to a third of her right hemi-
sphere, and caused neglect of the left side of space and paralysis to the left side of her body (Heller, 1991,
1994). When she went back to drawing, she had difficulties with the left side of space, with rendering
perspective and spatial relations (Figure 30.1). At first, she struggled to draw in her previous realist style.
In time, however, she came to accept her new, more expressionist, style as a reflection of her own view of
reality.

558
Neuropsychology of art and aesthetics

Figure 30.1 Two examples of Loring Hughes’ artistic work. The one on the left was painted before right hemisphere
brain damage; the one on the right was painted afterward. Reproduced from Heller (1994) with permission.

Unilateral spatial neglect and visual agnosia


Patients with unilateral spatial neglect are unaware of one side of space (Heilman et al., 1993). Artists with
unilateral spatial neglect often neglect one side of the page or canvas in their drawings and paintings. This
was the case of film director, painter, and cartoonist Federico Fellini, who suffered a right parietal stroke
at the age of 73, which caused an almost complete left visual and motor neglect (Cantagallo & Della Sala,
1998). One of the tests the clinicians asked him to perform, known as a line bisection tests, required him
to bisect several horizontal lines at the midpoint. Fellini consistently placed the mark closer to the right of
the horizontal line, indicating his neglect of a good portion of the left visual field. Fellini added cartoons to
some of these lines (Figure 30.2), all of which are on the right side, which also evidences the left field neglect
(Cantagallo & Della Sala, 1998).
Visual agnosia is an impairment in the recognition of visual objects. The specific effect of visual agno-
sia on artists’ creations depends to a large extent on whether the object recognition problems are related
with their perceptual or conceptual features (Chatterjee, 2004). Patients with apperceptive agnosia have an
impaired ability to use visual information to represent objects coherently, and patients with associative agno-
sia have an impaired ability to represent semantic knowledge of objects. Artists with apperceptive agnosia
will often not be able to render the overall form and composition of the depicted objects but will include
some of their conspicuous features or use pictorial cues in fragmented ways. This was the case of an artist
who suffered an occipital stroke and lost his ability to visually recognize objects and scenes. He could still
use such pictorial techniques as perspective, texture, and shadowing to draw in a similar style to the work he
produced before the stroke. He was, however, unable to produce coherent scenes and figures differentiated

559
Alejandro Dorado and Marcos Nadal

Figure 30.2 Four examples of Federico Fellini’s performances in line-bisection tasks, with characteristic personal car-
toons. Reproduced from Cantagallo and Della Sala (1998) with permission.

from grounds (Wapner et al., 1978). Artists with associative agnosia are still adept at drawing when copying
but are incapable when asked to draw from memory and can only rely on their knowledge of the world. This
was the case of an artist with associative visual agnosia who was able to masterfully copy an original Botticelli
portrait or draw a live portrait but could only produce child-like schematic and simplified drawings of objects
when given only verbal labels (Franklin et al., 1992).

Aphasia
Bäzner and Hennerici (2006) argue that there is little evidence indicating a significant impact of aphasia on
the creation of visual art, which suggests that verbal and visual production may be related to distinct output
channels. However, Bogousslavsky (2005) and Chatterjee’s (2004) revisions indicated that whereas the pro-
duction of some aphasic artists seems to be largely unaffected, other artists become more expressive, and yet
others begin producing works with different content, suggesting that this condition may be too broad as a
window into the biological bases of art and aesthetic production.

Effects of degenerative diseases on artistic production


Danae Chambers was a well-known Canadian artist who, at the age of 49, began showing obvious symp-
toms of memory loss and disorientation. Her cognitive state continued to deteriorate, and she began having
trouble with self-care habits, following a movie plot, keeping her finances, and interacting with clients. She
was diagnosed with Alzheimer’s disease when she was 56. Alzheimer’s disease is a neurodegenerative disease
that usually begins with difficulty with remembering recent events. As the disease progresses, patients might

560
Neuropsychology of art and aesthetics

experience problems with language and orientation, emotional burst, lack of motivation, and neglect of
self-care.
Chambers’ evaluation revealed an impairment of verbal working memory, set shifting, verbal episodic
memory, and expressive language, but preservation of visuospatial working memory, visual attention, and
visual perception (Fornazzari, 2005). Despite her deteriorating cognitive state, which included disorienta-
tion, common object agnosia, and severe memory impairment, she remained creative and continued to
paint for almost 15 years after her diagnosis. Her work, however, gradually lost figure-ground definition
and proportion among elements, and the colours became darker (Fornazzari, 2005). Chambers’ case is not
uncommon. Artists suffering from Alzheimer’s disease seem to gradually lose the ability to represent the
world with precision. The production of these artists continues only as part of the general routines they
performed throughout their lifetimes and only if other people around them provide the necessary setting to
do so (Chatterjee, 2004). However, they can continue using colour and form in aesthetically appealing ways
(Miller & Hou, 2004).
William Utermohlen was an American artist who, at the age of 57, began having trouble tying his necktie,
handling household finances, remembering things, and writing. He became increasingly subdued emotion-
ally and dethatched from his surroundings. His cognitive and emotional state continued to deteriorate, and
he was diagnosed with Alzheimer’s disease when he was 61 (Crutch et al., 2001). He continued to paint for
about 7 more years, mostly self-portraits, his style changing and reflecting his cognitive decline. Figure 30.3A
shows one of these portraits painted when he was 60, before anyone noticed any change in his artistic skills
or style, and is representative of his earlier work. Subsequent portraits (Figure 30.3B–F) reveal changes in
the depiction of facial features and spatial relations among them that reflect the progressively deteriorating
visuoperceptual and visuospatial skills. His work, as a result, became increasingly disjointed, fragmented, sur-
realist, and abstract (Crutch et al., 2001; van Buren et al., 2013).
Lester Potts was another artist whose work shows the influence of Alzheimer’s disease. The first indica-
tions of the disease were loss of short-term memory and unusual emotional responses to events around him,
but he later developed expressive language and visuospatial dysfunctions. Shortly after being diagnosed with
Alzheimer’s disease, he began painting watercolours as part of a day care centre programme. When he died,
8 years later, he had painted over 100 watercolours (van Buren et al., 2013).
Is there a common pattern to the changes in artistic production following the onset of Alzheimer’s dis-
ease? This was one of the questions that van Buren and colleagues (2013) set out to answer. They asked
healthy participants to assess 5 of the self-portraits Utermohlen painted after the onset of his cognitive decline
and 25 of Potts’ watercolours on preference, interestingness, and the 12 attributes of Chatterjee et al.’s (2010)
Assessment of Art Attributes scale (AAA). This scale comprises the measurement of six perceptual features,
such as colour, balance, depth, or complexity, and six conceptual features, including abstraction, symbol-
ism, or emotional expressiveness. Because each of the two artists’ works had been dated, the researchers
could examine progressive changes to style (conceptual and formal aspects) and aesthetic value (preference
and interestingness). The study found that, as Utermohlen’s and Potts’ cognitive state declined, their work
gradually became more abstract, more symbolic, less accurate, and less realist. Potts’ watercolours also became
more saturated, warmer, simpler, less animate, and less emotional. Participants also found that their work
progressively became less interesting and less liked (van Buren et al., 2013). Together wither other single-case
studies, van Buren and colleagues’ (2013) chronological analysis of stylistic and aesthetic changes associated
with Alzheimer’s shows that visuospatial, semantic, and executive impairments lead to the increase in distor-
tions, abstraction, and loss of realism.
Frontotemporal dementias are diseases caused by the deterioration of regions of the frontal and tempo-
ral lobes. Patients with these dementias often undergo striking personality changes, becoming inappropri-
ate, disinhibited, disorganized, or compulsive, or begin to have problems with the use of language. Some
patients with a specific form of frontotemporal dementia, known as semantic dementia and caused by neural

561
Alejandro Dorado and Marcos Nadal

Figure 30.3 William Utermohlen. Series of Self-Portraits. Painted at age 60 years (A), at 62 years (B), at 63 years (C),
and at 64 years (D). Pencil drawing at age 66 years (E). Abstract self-portrait painted at age 65 years (F).
Reproduced from Crutch et al. (2001) with permission.

562
Neuropsychology of art and aesthetics

degeneration in the left anterior temporal lobe, develop a new interest in art. These patients approach their
pictorial activity compulsively, in spontaneous spouts of creative production, often painting repetitive motifs,
leading to progressive improvements to their work. These depictions are usually realistic, lacking in symbol-
ism or abstract elements (Miller et al., 1998; Miller & Hou, 2004). Similar cases of appearance of artistic
production have been described following sudden-onset brain damage. Lythgoe and colleagues (2005), for
instance, describe a patient who began drawing, painting, and sculpting faces and heads compulsively after a
subarachnoid haemorrhage caused frontal dysfunction. These new and prolific artistic productions appeared
to be particular manifestations of the general personality changes, an unusual pictorial expression of com-
pulsive behaviours. The preservation in these cases of posterior temporal and occipital cortices enables these
patients to continue representing faces, places, and objects and to use them as themes in their artistic produc-
tion (Chatterjee, 2015).
Changes in creativity and artistic production have also been observed in patients with Parkinson’s disease.
Parkinson’s disease is a neurodegenerative disorder that affects mainly the motor system, leading to slow move-
ments, stiffness, tremors, and walking problems. These symptoms are caused by dopamine deficits resulting
from cell death in the substantia nigra. As the disease progresses, cognition can also become impaired, leading
to executive dysfunction: problems with working memory, attention, planning, understanding rules, cogni-
tive flexibility, and action inhibition. In many patients, motor symptoms begin asymmetrically, reflecting a
decrease in dopaminergic neuron number in the contralateral substantia nigra. Drago et al. (2009) wished to
find out whether Parkinson’s disease was associated with changes in creativity and whether these were more
pronounced when the disease had a right or left hemibody onset. They asked 15 Parkinson’s patients and 7
healthy controls to take the Abbreviated Torrance Creativity Test and the Controlled Oral Word Associa-
tion Test. Of the patients, six had a left hemibody onset and nine a right hemibody onset. Both groups of
Parkinson’s patients scored lower on verbal fluency, as measured using the Controlled Oral Word Association
Test. Patients with right hemibody onset obtained lower verbal creative fluency in the Abbreviated Torrance
Creativity Test than patients with left hemibody onset and controls. No difference was found in visuospatial
creativity (Drago et al., 2009). These results might be related with the impaired executive functions of the
left frontal lobe in patients with right hemibody onset.
Parkinson’s patients are often treated with dopamine agonists to improve motor function. In some cases,
such dopamine agonists have produced changes in artistic production. Kulisevsky et  al. (2009) described
the case of a 47-year-old amateur painter who was diagnosed with Parkinson’s disease. His interest in paint-
ing decreased in the months before the diagnosis. Although still apathetic toward most everyday activities,
treatment with dopamine agonists reduced his motor symptoms and led him to resume painting. His style,
however, had changed. Before the onset of the illness, his style had been detailed and realist. His painting
was now impressionistic, with colour and light prevailing over shape and detail (Figure 30.4). His interest in
painting became so passionate that he would spend whole days engrossed in it and became irritable when
suggested to stop. As this intense focus on painting interfered with sleep and created problems with his family,
the doctors reduced the dose of dopamine agonists, which substantially decreased his artistic activity. One
possible interpretation of this effect of dopamine agonists is that they led to a tonic change in the activity of
the limbic basal ganglia, which would fit with the patient’s own statement that his change in style owed to a
more intense emotional experience while painting (Kulisevsky et al., 2009).

Neuropsychology of art appreciation


The clinical cases and experiments described in the studies surveyed in the preceding sections reveal that
focal brain damage and neurodegenerative diseases can have diverse—and maybe even distinctive—effects
on artistic production. But what about their influence on the other side of the artistic coin—appreciation?
What can similar cases tell us about the biological underpinnings of aesthetic appreciation? There are fewer

563
Alejandro Dorado and Marcos Nadal

Figure 30.4 Change in painting style observed after the start of dopaminergic treatment in a patient with Parkinson’s
disease. These paintings highlight the change from a realistic (A, B) to a more impressionistic representation
of nature (C, D). Reproduced from Kulisevsky and colleagues (2009) with permission.

studies that have been carried out on this issue, and, as with artistic production, most studies describe single
clinical cases (Chatterjee, 2015; Nadal, 2013). Nevertheless, the number of experimental studies comparing
clinical and control samples is growing, especially with regard to the effects of neurodegenerative diseases.

Effects of focal brain damage on the appreciation of art


Griffiths and colleagues (2004) reported the case a patient who suffered a stroke that left him unable to
experience emotion in response to music. The lesion mainly affected the left insula but extended into the
left frontal cortex and amygdala. He recovered his speech, which had also been initially impaired, after 12
months. However, even though his perception of diverse musical features was normal, 18 months after the
stroke, the patient was still emotionally unaffected by music, even though he did enjoy other activities. These
observations led the authors to suggest that perceptual and emotional component of music processing might
rest on functionally and anatomically distinct neural networks and that the insula is a crucial piece in the
neural underpinnings of the emotional response to music (Griffiths et al., 2004).
Sellal and colleagues (2003) described the case of an epilepsy patient who underwent left temporal lobe
resection, which only spared the hippocampus, the parahippocampal gyrus, and the amygdala. This case is
interesting because the brain region that was surgically removed corresponds roughly to that which typically

564
Neuropsychology of art and aesthetics

degenerates in the form of frontotemporal dementia mentioned previously. During the first year after sur-
gery, the patient became aware that he no longer enjoyed listening to rock music and that he now preferred
Celtic or Corsican polyphonic singing. His taste in literature also changed, in this case from science fiction to
Kafkian-inspired novels. The patient also began showing increased preference for realistic paintings, enjoying
the small details that previously went unnoticed to him. These changes in aesthetic preference are in contrast
with his unchanged preferences for food, fashion, or faces.
The crucial role of subcortical brain structures related to emotional processing in aesthetic appreciation
was revealed by two studies of the effects of damage to the amygdala. Adolphs and Tranel (1999) report that
there were differences between the preferences for visual stimuli of two patients with bilateral amygdala dam-
age and a group of healthy controls. Both patients expressed higher liking for three-dimensional geometrical
shapes, landscapes, and colour arrangements than the control participants. This difference was especially clear
for the stimuli least liked by controls. Similar results were obtained from the examination of musical prefer-
ences in a patient with almost exclusive bilateral damage to the amygdala (Gosselin et al., 2007). In this case,
whereas the patient was able to correctly process musical features, even tempo and mode, she showed selec-
tive impairment in the recognition of scary and sad music, but not of happy music. Thus, it seems that the
amygdala plays a role in the affective processes underlying aesthetic preference, especially in the experience
of disliking and in relation to negatively valanced stimuli.
Interesting as they are, it is unclear how representative the results of these single clinical cases are. Are these
the typical effects of the lesions described? Or are they exceptional cases? Answering these cases requires
studying larger samples of patients with lesions in similar brain regions. Such a study was conducted by
Bromberger et al. (2011). They used Chatterjee and colleagues’ (2010) AAA to examine the effects of right
hemisphere focal brain damage on the appreciation of six formal and six conceptual attributes in a sample of
patients. Their goal was to ascertain whether right hemisphere damage caused specific impairments in the
assessment of art attributes and to test whether certain locations in the right hemisphere were more likely to
cause these impairments. They asked 20 patients with right hemisphere damage caused by stroke and 30 age-
matched healthy controls to rate 24 images of paintings on the 12 AAA attributes using a 5-point Likert scale.
Voxel-lesion-symptom mapping allowed them to assess the way in which damage to a brain area correlates
with behavioural scores. They found that patients with damage to the right frontal lobe differed from the
controls in ratings of four of the six conceptual scales: abstractness, symbolism, realism, and animacy. Dam-
age to the right parietal lobe was associated with differences in ratings of animacy and symbolism. Patients
differed from controls in only one of the perceptual features: depth. Figure 30.5 shows the location of voxels
for which brain lesions correlated with changes in ratings of art attributes. Bromberger and colleagues’ (2011)
results, therefore, suggest that cortical regions in the right hemisphere play a crucial role in the appreciation
of conceptual attributes of art.

Effects of degenerative diseases on the appreciation of art


One of the symptoms of neurodegenerative diseases is cognitive decline and, more specifically, memory
loss. Does the inability to remember having seen an artwork change the way it is appreciated? Halpern et al.
(2008), Graham et al. (2013), and Halpern and O’Connor (2013) examined the effects that neurodegenera-
tive disorders have on the consistency of preferences for art. Their results showed that artistically untrained
patients diagnosed with Alzheimer’s disease (Graham et al., 2013; Halpern et al., 2008) and frontotemporal
dementia (Halpern & O’Connor, 2013) were as consistent as healthy controls in their preference for visual
artworks presented at different times, even though Alzheimer’s patients were unable to remember they had
seen the stimuli before. Thus, despite Alzheimer’s disease and frontotemporal dementia’s devastating effects
on general cognitive function, these disorders do not seem to prevent patients from experiencing art in a
personally meaningful and consistent way.

565
Alejandro Dorado and Marcos Nadal

Figure 30.5 Results of Bromberger and colleagues’ (2011) voxel lesion symptom mapping analyses showing areas where
damage was associated with significant deviations of art attribute judgments. Originally published in an
access article subject to a non-exclusive license between the authors and Frontiers Media SA, which permits
use, distribution, and reproduction in other forums, provided the original authors and source are credited.

Does this mean that neurodegenerative diseases do not alter the experience of art? The studies by Graham
et al. (2013), Halpern et al. (2008), and Halpern and O’Connor (2013) showed that consistency of art expe-
rience is preserved in the face of Alzheimer’s disease and frontotemporal dementia. They did not report,
however, differences in preference between patients and controls. Two reports of patients with frontotem-
poral dementia suggest that this neurological disorder could have profound effects on the experience of art.
Geroldi et al. (2000) and Boeve and Geda (2001) reported three patients who experienced marked changes in
preference for music after the onset of frontotemporal dementia. The patients began compulsively listening
to music they had not particularly enjoyed previously, playing it for hours on end. Thus, although fronto-
temporal dementia patients seem to preserve the capacity for enjoying art, and their preferences are consistent
over time, this disorder can alter the kind of art these patients enjoy.
Several recent studies have examined changes in the appreciation of art brought on by Parkinson’s disease.
Lauring and colleagues (2019) asked 21 Parkinson’s patients and 23 age- and gender-matched healthy con-
trols to rate 15 paintings using Chatterjee and colleagues’ (2010) AAA and to rate 60 other paintings on lik-
ing and beauty scales. The results showed no significant difference between Parkinson’s patients and healthy
controls on liking and beauty ratings and the AAA, with the exception of the AAA Emotionality scale, for
which the Parkinson’s patients gave higher scores than the control participants (Lauring et al., 2019).

566
Neuropsychology of art and aesthetics

One possible reason for this lack of effect of Parkinson’s disease on the appreciation of art could be the
motor specificity of this disease. It could be the case that Parkinson’s impairs the representation and apprecia-
tion of movement cues in painting but not other art attributes. Humphries et al. (2021) examined whether
art appreciation ratings, motion ratings, and the relation between them are altered in patients with Parkin-
son’s disease. They asked 43 Parkinson’s disease patients and 40 age- and education-matched healthy controls
to rate 10 action paintings by Jackson Pollock and 10 neoplastic paintings by Piet Mondrian on several
dimensions using 7-point Likert scales. There were four evaluative dimensions (liking, beauty, interest, and
familiarity), and five formal dimensions (motion, complexity, balance, colour hue, and colour saturation).
The analyses revealed that whereas control participants gave higher beauty, liking, and interest to Mondrian’s
paintings, Parkinson’s disease patients gave higher beauty, liking, and interest ratings to Pollock’s paintings.
They also showed that Parkinson’s patients gave lower motion ratings for both artists’ paintings. Both groups
of participants also differed in the effect of motion on liking and beauty ratings. For controls, the relation was
quadratic: with increasing motion, liking and beauty increased, but to a point. Beyond that point, movement
contributed negatively to liking and beauty. For Parkinson’s disease patients, motion linearly predicted liking
and beauty ratings. Humphries and colleagues (2021) concluded that patients with Parkinson’s disease are less
capable of translating pictorial motion cues into motor representations. This alters the way in which these
representations are used to rate liking and beauty.

Conclusions
Overall, the study of the impact of neurological disorders on the experience of art argues against the exist-
ence of specialized brain mechanisms underlying the creation and appreciation of art (Zaidel, 2005). There is
no single brain region—cortical or subcortical—that is particularly dedicated to the creation or appreciation
of art. Making and enjoying art rely on the same neural networks in the brain that orchestrate perception,
memory, imagination, language, pleasure, and movement in other domains of human experience. This is
probably the reason art creation and appreciation seem to be so resilient to brain lesions and diseases. The
distributed nature of the neuroanatomical foundations of these activities allows them to continue in the face
of these conditions, albeit with impairments of different kinds and degrees.
Focal brain lesions that affect specific nodes in these networks lead to impairments of certain aspects of art
making and appreciation. As we have seen, strokes can cause lesions to brain regions involved in processing
space and spatial relations. Artists’ difficulties in depicting these relations caused their styles to change, gener-
ally becoming more impressionistic or abstract, and, in some cases, becoming more emotionally evocative
and expressionist. Strokes can also alter liking for certain artforms or styles and impair the representation and
evaluation of certain art attributes, depending on whether the lesion affects brain regions involved primarily
in perception, conceptualization, or valuation.
Neurodegenerative diseases have profound effects on the creation of art. Painters with Alzheimer’s dis-
ease progressively lose the ability to paint accurately. Their work loses definition and proportion, becom-
ing disjointed, fragmented, and abstract. There is often also a change in the colour palette. Changes in the
production of art can occur as a result of the general changes in behaviour and personality that happen in
neurodegenerative diseases. This seems to be the case with some frontotemporal dementia patients, who start
compulsively drawing or painting repetitive motifs, often realistic and lacking in symbolism, and become
strikingly good at this. Changes in the production of art can be caused by the medication administered to
these patients, as we saw with the painter whose artistic creativity and style was boosted by dopamine ago-
nists. All these diseases can also change people’s appreciation for art, leading to changes in stylistic preferences
and loss of the ability to represent certain art attributes.
As in any scientific discipline, the strength of the conclusions depends on the strength of the evidence.
Much of what we know about the consequences of brain lesions and diseases on the creation and appreciation

567
Alejandro Dorado and Marcos Nadal

of art relies on few descriptive case reports of artists and other people. Single-case clinical studies are inter-
esting and illustrative. But the representativeness of any given case is uncertain. Artists and non-artists alike
differ enormously, so the effects of brain lesions or diseases on art production and appreciation in one case
may differ substantially from another.
Fortunately, a great deal of effort has been expended in the last decade to design systematic procedures
that allow assessing the impact of neurological conditions on the appreciation of distinct art attributes (Chat-
terjee et al., 2010) that can be used on samples of patients and healthy control participants (Graham et al.,
2013; Humphries et al., 2021; Lauring et al., 2019; van Buren et al., 2013) and procedures that allow sys-
tematic correlation of brain lesion location and symptoms (Bromberger et al., 2011). As the number of such
studies increases over the next decades, knowledge of the effects of brain lesions and diseases on art produc-
tion and appreciation will continue to grow and strengthen.

References
Adolphs, R., & Tranel, D. (1999). Preferences for visual stimuli following amygdala damage. Journal of Cognitive Neurosci-
ence, 11(6), 610–616. https://doi.org/10.1162/089892999563670
Alajouanine, T. (1948). Aphasia and artistic realization. Brain, 71(3), 229–241. https://doi.org/10.1093/brain/71.3.229
Auerbach, S. (1906). Beitrage zur Lokalisation des musikalischen Talentes im Gehirn un im Schädel. Archiv Für Anatomie
Un Physiologie, Anatomische Abteilung, 197–230.
Auerbach, S. (1908). Zur Lokalisation des musikalischen Talentes im Gehirn und am Schädel. Zweiter Beitrag. Archiv Für
Anatomie Un Physiologie, Anatomische Abteilung, 31–38.
Auerbach, S. (1911). Zur Lokalisation des musikalischen Talentes im Gehirn und am Schädel. Dritter Beitrag. Archiv Für
Anatomie Un Physiologie, Anatomische Abteilung, 1–10.
Auerbach, S. (1913). Zur Lokalisation des musikalischen Talentes im Gehirn und am Schädel. Vierter Beitrag. Archiv Für Anatomie
Un Physiologie (pp. 89–96). Anatomische Abteilung.
Bäzner, H., & Hennerici, M. (2006). Stroke in painters. In F. C. Rose (Ed.), The neurobiology of painting: International
review of neurobiology (Vol. 74, pp. 165–191). Academic Press. https://doi.org/10.1016/S0074-7742(06)74013-2
Benton, A. (2000). Exploring the history of neuropsychology. Oxford University Press.
Blanke, O., & Pasqualini, I. (2011). The riddle of style changes in the visual arts after interference with the right brain.
Frontiers in Human Neuroscience, 5, 154. https://doi.org/10.3389/fnhum.2011.00154
Boeve, B. F., & Geda, Y. E. (2001). Polka music and semantic dementia. Neurology, 57(8), 1485. https://doi.org/10.1212/
wnl.57.8.1485
Bogousslavsky, J. (2005). Artistic creativity, style and brain disorders. European Neurology, 54, 103–111.
Boring, E. G. (1950). A history of experimental psychology (2nd ed.). Appleton-Century-Crofts.
Bouillaud, J.-B. (1865). Discussion sur la faculté du langage articulé. Bulletin de l’Académie Impériale de Médecine, 30,
724–781.
Bromberger, B., Sternschein, R., Widick, P., Smith, W., & Chatterjee, A. (2011). The right hemisphere in esthetic per-
ception. Frontiers in Human Neuroscience, 5, 109. https://doi.org/10.3389/fnhum.2011.00109
Cantagallo, A., & Della Sala, S. (1998). Preserved insight in an artist with extraperonal spatial neglect. Cortex; A Journal
Devoted to the Study of the Nervous System and Behavior, 34(2), 163–189. https://doi.org/10.1016/s0010-9452(08)70746-9
Chatterjee, A. (2004). The neuropsychology of visual artistic production. Neuropsychologia, 42(11), 1568–1583. https://
doi.org/10.1016/j.neuropsychologia.2004.03.011
Chatterjee, A. (2006). The neuropsychology of visual art: Conferring capacity. International Review of Neurobiology, 74,
39–49. https://doi.org/10.1016/S0074-7742(06)74003-X
Chatterjee, A. (2015). The neuropsychology of visual art. In J. P M. Huston, F. M. Nadal, L. F. Agnati, & C. C. Cela-
Conde (Eds.), Art, aesthetics and the brain (pp. 341–356). Oxford University Press.
Chatterjee, A., Widick, P., Sternschein, R., Smith, W. B., & Bromberger, B. (2010). The assessment of art attributes.
Empirical Studies of the Arts, 28(2), 207–222. https://doi.org/10.2190/EM.28.2.f
Crutch, S. J., Isaacs, R., & Rossor, M. N. (2001). Some workmen can blame their tools: Artistic change in an individual
with Alzheimer’s disease. Lancet, 357(9274), 2129–2133. https://doi.org/10.1016/S0140-6736(00)05187-4
Drago, V., Foster, P. S., Skidmore, F. M., & Heilman, K. M. (2009). Creativity in Parkinson’s disease as a function of
right versus left hemibody onset. Journal of the Neurological Sciences, 276(1–2), 179–183. https://doi.org/10.1016/j.
jns.2008.09.026

568
Neuropsychology of art and aesthetics

Dupré, E.,  & Nathan, M. (1911). Le langage musical: Étude médico-psychologique. Maisons Félix Alcan et Guillaumin
Réunies.
Eling, P., Finger, S., & Whitaker, H. (2015). Franz Joseph Gall and music: The faculty and the bump. In Progress in brain
research (Vol. 216, pp. 3–32). https://doi.org/10.1016/bs.pbr.2014.11.001
Finger, S. (1994). Origins of neuroscience. A history of explorations into brain function. Oxford University Press.
Finger, S. (2010). The birth of localization theory. In S. Finger, F. Boller, & K. L. Tyler (Eds.), Handbook of clinical neurol-
ogy, history of neurology (Vol. 95, pp. 115–128). Elsevier.
Flourens, M.-J.-P. (1824). Recherches Expérimentales sur les Propriétés et les Fonctions du Système Nerveux dans les Animaux
Vertébrés (2nd ed.). Ballière.
Fornazzari, L. R. (2005). Preserved painting creativity in an artist with Alzheimer’s disease. European Journal of Neurology,
12(6), 419–424. https://doi.org/10.1111/J.1468-1331.2005.01128.X
Franklin, S., van Sommers, P., & Howard, D. (1992). Drawing without meaning? Dissociations in graphic performance
of an agnosic artist. In R. Campbell (Ed.), Mental lives. Case studies in cognition (pp. 179–198). Blackwell Publishing.
Geroldi, C., Metitieri, T., Binetti, G., Zanetti, O., Trabucchi, M., & Frisoni, G. B. (2000). Pop music and frontotemporal
dementia. Neurology, 55(12), 1935–1936. https://doi.org/10.1212/wnl.55.12.1935
Gosselin, N., Peretz, I., Johnsen, E., & Adolphs, R. (2007). Amygdala damage impairs emotion recognition from music.
Neuropsychologia, 45(2), 236–244. https://doi.org/10.1016/j.neuropsychologia.2006.07.012
Gourevitch, M. (1967). Un aphasique s’exprime par le dessin. Encéphale, 56(1), 52–68.
Graham, D. J., Stockinger, S., & Leder, H. (2013). An island of stability: Art images and natural scenes—But not natural
faces—Show consistent esthetic response in Alzheimer’s-related dementia. Frontiers in Psychology, 4, 107. https://doi.
org/10.3389/fpsyg.2013.00107
Graziano, A. B., & Johnson, J. K. (2015). Music, neurology, and psychology in the nineteenth century. Progress in Brain
Research, 216, 33–49. https://doi.org/10.1016/bs.pbr.2014.11.002
Griffiths, T. D., Warren, J. D., Dean, J. L., & Howard, D. (2004). ‘When the feeling’s gone’: A selective loss of musical
emotion. Journal of Neurology, Neurosurgery, and Psychiatry, 75(2), 344–345.
Gross, C. G. (1999). Brain, vision, memory. Tales in the history of neuroscience. MIT Press.
Halpern, A. R., Ly, J., Elkin-Frankston, S., & O’Connor, M. G. (2008). ‘I know what I like’: Stability of esthetic prefer-
ence in Alzheimer’s patients. Brain and Cognition, 66(1), 65–72. https://doi.org/10.1016/j.bandc.2007.05.008
Halpern, A. R., & O’Connor, M. G. (2013). Stability of art preference in frontotemporal dementia. Psychology of Aesthet-
ics, Creativity, and the Arts, 7(1), 95–99. https://doi.org/10.1037/a0031734
Heilman, K. M., Watson, R. T., & Valenstein, E. (1993). Neglect and related disorders. In K. M. Heilman & E. Valen-
stein (Eds.), Clinical neuropsychology (3rd ed., pp. 279–336). Oxford University Press.
Heller, W. (1991). New territory: Creativity and brain injury. Creative Woman, 11, 16–18.
Heller, W. (1994). Cognitive and emotional organization of the brain: Influences on the creation and perception of art.
In D. W. Zaidel (Ed.), Neuropsychology (pp. 271–292). Academic Press.
Humphries, S., Rick, J., Weintraub, D., & Chatterjee, A. (2021). Movement in aesthetic experiences: What we can learn
from Parkinson disease. Journal of Cognitive Neuroscience, 33(7), 1329–1342. https://doi.org/10.1162/JOCN_A_01718
Jackson, J. H. (1871). Singing by speechless (aphasic) children. Lancet, 98, 430–431
Jackson, J. H. (1878). On affections of speech from disease of the brain. Brain, 1(3), 304–330. https://doi.org/10.1093/
brain/1.3.304
Jackson, J. H. (1879). On affections of speech from disease of the brain. Brain, 2, 203–222, 323–356.
Johnson, J. K., Graziano, A. B., & Hayward, J. (2010). Historical perspectives on the study of music in neurology. In F.
C. Rose (Ed.), Neurology of music (pp. 17–30). Imperial College Press.
Knoblauch, A. (1888). Über Störungen der musikalischen Leistungsfähigkeit infolge von Gehirnläsionen. Deutsches
Archiv Für Klinische Medizin, 43, 331–352.
Kulisevsky, J., Pagonabarraga, J., & Martinez-Corral, M. (2009). Changes in artistic style and behaviour in Parkinson’s
disease: Dopamine and creativity. Journal of Neurology, 256(5), 816–819. https://doi.org/10.1007/s00415-009-5001-1
Lauring, J. O., Pelowski, M., Specker, E., Ishizu, T., Haugbøl, S., Hollunder, B., . . . Kupers. (2019). Parkinson’s disease
and changes in the appreciation of art: A comparison of aesthetic and formal evaluations of paintings between PD
patients and healthy controls. Brain and Cognition, 136, 103597. https://doi.org/10.1016/J.BANDC.2019.103597
Lichtheim, L. (1885). Ueber aphasie. Deutsches Archiv Für Klinische Medizin, 36, 204–268.
Luria, A. R., Tsvetkova, L. S., & Futer, D. S. (1965). Aphasia in a composer (V. G. Shebalin). Journal of the Neurological
Sciences, 2(3), 288–292. https://doi.org/10.1016/0022-510x(65)90113-9
Lythgoe, M. F. X., Pollak, T. A., Kalmus, M., De Haan, M., & Chong, W. K. (2005). Obsessive, prolific artistic output
following subarachnoid hemorrhage. Neurology, 64(2), 397–398. https://doi.org/10.1212/01.WNL.0000150526.09
499.3E

569
Alejandro Dorado and Marcos Nadal

Miller, B. L., Cummings, J., Mishkin, F., Boone, K., Prince, F., Ponton, M., & Cotman, C. (1998). Emergence of artistic
talent in frontotemporal dementia. Neurology, 51(4), 978–982. https://doi.org/10.1212/wnl.51.4.978
Miller, B. L., & Hou, C. E. (2004). Portraits of artists. Emergence of visual creativity in dementia. Archives of Neurology,
61(6), 842–844. https://doi.org/10.1001/archneur.61.6.842
Nadal, M. (2013). The experience of art. Insights from neuroimaging. Progress in Brain Research, 204, 135–158. https://
doi.org/10.1016/B978-0-444-63287-6.00007-5
Sellal, F., Andriantseheno, M., Vercueil, L., Hirsch, E., Kahane, P.,  & Pellat, J. (2003). Dramatic changes in artistic
preference after left temporal lobectomy. Epilepsy and Behavior: E&B, 4(4), 449–50; author reply 451. https://doi.
org/10.1016/s1525-5050(03)00146-x
Souques, A., & Baruk, H. (1930). Autopsie d’un cas d’amusie (avec aphasie) chez un professeur de piano. Revue Neu-
rologique, 1, 545–556.
Trousseau, A. (1865). Discussion sur la faculté du langage articulé. Bulletin de l’Académie Impériale de Médecine, 30,
647–657, 659–675.
Vallar, G. (2000). The methodological foundations of human neuropsychology: Studies in brain-damaged patients. In F.
Boller, J. Grafman, & G. Rizzolatti (Eds.), Handbook of neuropsychology (2nd ed., pp. 305–344). Elsevier.
van Buren, B., Bromberger, B., Potts, D., Miller, B., & Chatterjee, A. (2013). Changes in painting styles of two artists
with Alzheimer’s disease. Psychology of Aesthetics, Creativity, and the Arts, 7(1), pp, 89–94. https://doi.org/10.1037/
a0029332
Wapner, W., Judd, T., & Gardner, H. (1978). Visual agnosia in an artist. Cortex; A Journal Devoted to the Study of the Nerv-
ous System and Behavior, 14(3), 343–364. https://doi.org/10.1016/S0010-9452(78)80062-8
Wernicke, C. (1874). Der aphasische Symptomencomplex. Eine psychologische Studie auf anatomischer Basis. Max Cohn and
Weigert.
Wickens, A. P. (2015). A history of the brain. From Stone Age surgery to modern neuroscience. Psychology Press.
Zaidel, D. W. (2005). Neuropsychology of art: Neurological, cognitive, and evolutionary perspectives. Psychology Press.
Zaimov, K., Kitov, D., & Kolev, N. (1969). Aphasie chez un peintre. Encéphale, 58, 377–417.

570
INDEX

activation likelihood estimation (ALE) 114 androgen 46


activation observation network 180 anger 41
Adrian, E. 84 anhedonia 38, 67; specific musical anhedonia 38,
aesthetic appreciation 103, 118, 149, 177, 244, 307, 298, 139 – 140, 142 – 143, 255
449, 481, 563 – 567; neural correlates of 110 – 114, anterior cingulate cortex 37, 38, 66, 79, 116, 137,
136 – 143 511, 516
aesthetic emotions 3, 52, 400 – 401 aphasia 560 – 561
aesthetic evaluation 19 – 20, 52 – 53, 302, 351 – 353, 441, approach 31, 34, 40, 78, 149
468, 470 architecture 194 – 197, 346 – 359; and health 207; and
aesthetic experience 6, 14 – 15, 103, 201 – 203, 277, 398, liking 108f, 113, 114 – 117, 201 – 204; and movement
439, 441, 448, 481 253.355; and neuroscience 198 – 200; neural correlates
aesthetic faculty 255 of 202 – 206; 351 – 359
aesthetic fluency scale (AFS) 426 Aristotle 495, 498
aesthetic measurements 104 – 106, 242, 441 Arnheim, R. 174, 296
aesthetic pleasure 12, 139, 149, 481 art experience 295, 410 – 411, 423 – 432, 439, 448 – 456,
aesthetic preference 221, 277, 438 462, 466, 470 – 471, 494 – 495
aesthetic quality 52m 195 art making/production 411, 417 – 419, 526 – 532,
aesthetic sensitivity 240 – 248, 306f 555 – 563, 567; evolution of 539 – 540, 544 – 545,
aesthetic taste 7 – 8, 103, 242 547 – 548; neural correlates of 418 – 419, 514 – 515
aesthetic triad 120, 199f, 295, 301 – 302, 438, 464 – 465 artistic talent 242; general factor of 243
aesthetic value 103, 254, 277, 279 – 280 arousal 13, 40, 507 – 508
aesthetics 3 – 7, 15 – 16, 106 assessment of art attributes scale (AAA) 561, 565 – 566
aesthetics triumvirate 425, 427 asymmetry 41
affective sensitiveness 246 attention 47, 508
affective valence 63 attention restoration theory 200f
affordance 351, 354, 360 Auerbach, S. 527, 557
Alberti, L.B. 196 auditory agnosia 39
Alexander, C. 198 auditory nucleus 36
Allen, G. 12, 16 avoidance 31, 34, 40, 78, 149; pathogen avoidance,
alternative uses task (AUT) 508, 510 – 511 71 – 73, 77
amusement park theoretical (ATP) model 514
amusia; see music Barcelona Music Reward Questionnaire 141
amygdala 37, 38, 39, 41, 43 – 45, 48, 66, 73 – 74, 92, 115, Bard, P. 89
153, 158, 160, 255, 264 – 265, 270, 275, 454, 456, Bateman, A.J. 220
564 – 565 Baumgarten, A.G. 5, 8, 16
analgesia 415 – 416 beat 323 – 324

571
Index

beauty 50, 78 – 79, 382, 441; beauty-is-good 121; sexual diastole 92 – 93
beauty 218, 220, 226 diffusion tensor imaging (DTI) 38
beholder’s share 476, 481 direct benefit 221
Berger, H. 347 disgust 38, 40 – 41, 47, 71 – 80, 94; disgust sensitivity
Berlyne, D. 13, 109, 269, 463, 479 75 – 77; moral disgust 74 – 75, 77; pathogen disgust
Berridge, K.C. 39, 43 – 44 72 – 73, 77; sexual disgust 74 – 74; three-domain
biophilic design 194 – 195, 198, 207, 350 disgust scale 75
bitter taste 34, 43, 63, 77 disinterestedness 498
blood oxygen level-dependent (BOLD) signal 43, 46, disliking 31, 32, 34, 36, 37, 39, 40, 41, 43, 46 – 47, 51,
11, 114, 349 67; definition of 40 – 41
body aesthetics 179f displeasure 32, 38, 40 – 41
body temperature 40 dissonance 321 – 322
bottom-up processing 14, 461 – 466, 470 dopamine 46, 64, 115, 119, 136, 137, 232 – 233, 440, 563
brainstem 13, 42, 45, 92 dorsolateral prefrontal cortex 42, 118 – 119, 238, 305,
Bohr, N. 510 415, 470, 511 – 514, 516 – 517
Bouillaud, J.-P. 555 – 556
Brentano, F. 12 einfühlung 12
Broca, P. 555 – 556 electroencephalography (EEG) 347 – 348
built environment; see architecture embodiment 342, 353 – 354
Burke, E. 9 – 10 empirical aesthetics 11, 106, 108
Burt, C. 240, 242f entorhinal cortex 49
eudaimonia 67
Cannon, W. 89 evaluative event 32 – 33, 45, 49 – 50
Cavanagh, P. 14 evaluative system 34, 38, 42, 43
Chatterjee, A. 3, 8, 14 – 15 evaluative task 49 – 50
Child, I.L. 240, 244f executive control network 512 – 513
chills 40, 64, 137 – 138, 306, 439, 441, 451 expectation 109, 471; musical expectations 320 – 321,
cognitive penetration 441 469, 471, 489 – 490; and reward 47 – 48, 487 – 488
cognitive vitality model 412 expertise 339, 468 – 469, 526 – 530, 532; and liking
coldspots 43, 64 – 66 49, 341
collative variable 296 extrastriate body area (EBA) 179
colour 112, 222 – 223, 256, 303, 469 eye-tracking 17
common currency 66, 118, 137 Eysenck, H.J. 240, 243, 481
complexity 476 – 478
concert setting 451 – 454 facial attractiveness 36, 78 – 79, 107; neural correlates of
consonance 421 – 322, 412 112, 114 – 117; 119
contagion 428 facial disfigurement 41, 76
context 427 – 429, 438 – 439, 441 – 433, 448 familiarity 149, 468 – 469
contraposto 174 fear 38, 40 – 41, 52, 71
creativity 507 – 519, 529; artistic creativity 513 – 518; Fechner, G. 11, 13, 16, 295, 463
neural correlates of 510 – 518 film 454 – 456
curiosity 476 – 477, 479 – 482 fine art 4 – 6
curvature 106, 111, 257, 302, 352 flavour 148
Flourens, M.-J.P. 556
dance 113f, 119, 180, 337 – 342, 449, 529; and emotions fluency 109, 201, 229, 247, 296, 476, 508
341 – 343; and liking 180, 340 – 341; neural correlates of foregrounding 369, 371, 384
338 – 340, perception of 337 – 339 framing 49, 157, 160, 176, 300, 441 – 442, 466 – 467
Darwin, C. 30, 218, 220 – 221, 233, 507 Frege, G. 12
Darwin, E. 507 functional magnetic resonance imaging (fMRI) 43, 72,
Da Vinci, L. 174 103, 348 – 350
decision making 33, 35, 36, 73 fusiform face area (FFA) 202, 300 – 301
decoy effect 36 fusiform gyrus 49
default mode network (DMN) 119 – 120, 306, 381 – 382,
403, 512 – 513, 515 Gall, F.J. 555 – 556
denisovans 540 Galton, F. 242, 507 – 508
Descartes, R. 9 – 10 genius 508

572
Index

gestalt psychology 296, 476, 481 learning 44, 475 – 480, 492 – 494; reinforcement
golden ratio 196 learning 492
Gombrich, E. 174 Le Corbusier 196
Gray, A. 218 Leder, H. 298
grid cells 203f Leder model 14 – 15, 298 – 300, 464
G-PT3 368 levodopa 138
gut-brain axis 95f liking 31 – 32, 34, 36 – 37, 39 – 40, 43, 46, 47, 50,
475 – 478; aesthetic liking 51 – 52, 240, 254f; definition
von Haller, A. 9 of 33 – 34, 40; Berridge concept 44, 63 – 67, 232
health 22, 297 – 298, 402 – 404, 410 – 419 literature 366 – 374, 379 – 390; and liking 372 – 373; neural
heart beat 90, 92 correlates of 369 – 371, 373 – 374; reception of 368 – 373
heartbeat detection task 90 Livingstone, M. 14
hedonic evaluation 33, 39, 44 Locher, P. 17, 297, 464
hedonic evaluation 34, 38, 39, 40, 45, 47, 63 – 64, 103,
154 – 158, 183 – 184, 439 – 440 machine-learning assisted questionnaire 371 – 372
hedonic value mammalian brain evolution 267 – 268
Hegel, D. 6 Martindale, C. 508
homeostasis 45 – 46, 89 – 90 mate choice 35 – 36, 46, 179
hotspots 43, 64 – 66, 119, 440 Mayr, E. 254
Hughlings Jackson, J. 555 medial temporal cortex 178 – 179
human brain evolution 272 – 276, 540 – 542, Meier, N. 240, 242
548 – 549 Meier art test 242
Hume, D. 5,8 Melody 320 – 321
hunger 46 menstrual cycle 46
Husserl, E. 12 mere exposure 48, 206, 208, 475 – 476, 487
Hutcheson, F. 5, 8 mere measurement effect 478
hypothalamus 42, 45 – 46, 92 mesocorticolimbic system 32, 39, 44, 440; anatomy of
37 – 38, 42, 49
incentive salience 232 metabolism 46
indirect benefit 221 metaphor 369 – 370, 510 – 511
inferior frontal gyrus 118, 135, 140, 511, 513 metazoan brain evolution 260
insula 34, 37 – 39, 41 – 45, 66, 72, 74, 92, 118, 137, 153, meter 324f
158, 204f, 255, 414, 516, 564 – 565 Michelangelo 196
intelligence 242, 509 Milner, P. 136
international affective picture system (IAPS) 467 mirror model (Tinio) 297
interoception 89 – 90, 450 – 452; and cardiovascular system mirror neurons 180, 337 – 338
90, 92 – 95; and decision marking 95; definition of monetary loss 41
89; history of concept 89 – 90; and reward processing moral transgression 41
45 – 47; and visual attention motivation 44, 52, 63
interoceptive sensibility 90f motor cortex 516
interoceptive sensitivity 90f movement perception 178; and liking 112, 114 – 117,
interoceptor 89 177 – 183, 257, 303; neural correlates of 183 – 184, 303
museum/gallery 423 – 432, 448
James, W. 89 music 134, 315 – 327, 397 – 403, 413 – 416, 451; amusia
135f, 317 – 318 – 557; and 226 – 227; and emotions
Kant, I. 5 – 6, 8, 9 325 – 326, 399 – 400; and liking 144, 230 – 231,
Kaplan, S. 350 326 – 327, 398 – 399, 452; neural correlates of
Keats, J. 496 liking 136 – 143, 398 – 399; perception of 18 – 19,
Kellert, S. 198 39, 135 – 136, 230; neural correlates of perception
Knoblauch, A. 556f 316 – 325, 399 – 400
knowledge 48 – 49, 206, 425 – 426, 474 musical expertise 526 – 528, 530 – 532
Köhler, W. 481 music playing 514
Kristeller, P.O. 4 musicophilia 139 – 140

landscape 107 – 108, 113 – 117, 119, 181 – 182 Nadal, M. 52, 471


Lange, C. 89 naltrexone 39

573
Index

narrative 379 – 390; and liking 382 – 384; neural correlates recreational fear 41


of 380 – 384; reception of 380 – 382; and social reflex 90
cognition 385 – 387 remote associates task (RAT) 508, 510 – 511
natural selection 220 respiration 40
nausea 72 – 73, 77 reward prediction 38, 44, 255; reward prediction error
neanderthals 539 – 540, 544 44, 483 – 485, 489 – 490
nervous system 261 – 262 Reynolds, J. 6
neuroasthetics 1 – 21, 104, 120f; definition of 3; history of Richards, I.A. 366
1 – 2, 7 – 20 rhythm 322, 325
neurocognitive poetics model (NCPM) 372 – 373 risperidon 138
neuroimaging 38, 40 – 41 Rizzolatti, G. 337
Nodine, C. 17 Ruggles, D. 201
nostalgia 400 – 401 Ryan, M. 36
nucleus accumbens 37 – 38, 43 – 46, 48, 64 – 67, 74, 79,
114 – 115, 118 – 119, 136 – 138, 140, 143, 153, 315, sadomasochism 41, 74
440, 456 salience network 512 – 513
satiety 40, 45, 64
odour perception 149 – 153; and hedonics 154 – 158; and self 95
liking 158 – 161; neural correlates of 150 – 153 sensory bias/sensory drive 122, 221 – 223
odourant 149 – 150 sensory liking 31 – 32, 33, 35, 37, 51; definition of 31;
oestrogen 46 – 47, 232 mechanism 32
Olds, J. 136 sensory valuation; see hedonic valuation, hedonic value
olfactory system 150 – 153 and sensory liking
opioids 39, 64, 115, 440 serotonin 73
orbitofrontal cortex 37 – 39, 40, 42, 43 – 46, 49, 66, sexual selection 31, 218 – 219
116 – 119, 136, 143, 153, 158, 204, 255, 305, 470, sexy son hypothesis 221
514, 516 Sherrington, C. 89
oxytocin 77 Silva, P.J. 297f, 464
skin conductance 40
pain 38, 40 – 41, 403 – 403, 411, 413 – 416 Skov, M. 471
pallidum 37, 39, 42, 65 – 67, 255, 264, 440 somatic marker hypothesis 90
parahippocampal cortex Spencer, H. 12
parahippocampal place area (PPA) 201 300 – 301, 351 – 352 stress 402, 417 – 418, 451
parallelism 369 striatum 38 – 39, 41 – 44, 114 – 115, 136, 140, 158, 255,
Pavlov, I. 90 264, 456
peacock 218 – 220 subjective taste 107 – 108, 144
perceptual uncertainty 47 – 48, 149 superior temporal gyrus 38, 105, 134 – 135, 143
perirhinal cortex 49 sweet taste 34
physiological aesthetics 9 – 12 sympathetic system 92
pitch 315 – 317, 319 symmetry 106, 11, 257, 302
Plato 121 synchronous movement 449 – 450
pleasure 32, 38 – 39, 43 – 44, 63, 66 – 67, 122, 136, 397, 44; systole 92 – 93
definition of 40; pleasure cycle 64 – 65
poetry 366, 368, 516 – 517 taste 4, 31, 247; general factor, t 243; standard of taste 4,
positron emission tomography 136 – 137 244 – 245
prägnanz 296 testosterone 46, 74
predictive coding/active inference 47, 475, 482 – 497 thalamus 45, 92
preference 32, 44 theatre 449 – 450
progesterone 46 – 47 thirst 46
prospect theory 228 Thorndike, E. 24
primate brain evolution 270 – 271 timbre 322
punishment 45 time discounting 36
Pythagoras, school of 31 tonality 320
trans magnetic stimulation (TMS) 138
Ramachandran, V.S. 20 tribhanga 174
Reber, R. 201 túngara frogs 36, 122, 219, 224 – 225, 228 – 229, 233

574
Index

ugliness 9, 71, 76; ugliness judgments 78 – 80; ugly art 80 Vitruvius 196
Ulrich, R. 350 von Uexküll, J. 221

van der Rohe, M. 196 Wallace, A. 220 – 221


Vartanian, O. 8 wanting 44, 64 – 67, 232, 439
ventraltegmental area 46, 78, 95, 136, 158, 232, 440 Washburn, M. 246
ventromedial prefrontal cortex 32, 38, 42, 116 – 117 Weber, E. 11
vertebrate brain evolution 263 – 264 Weber’s law 228, 233
Vitruvial triad 196 well-being; see health
vienna art interest & art knowledge questionnaire 426 Wilson, E.O. 198
vienna integrated model of art perception (VIMAP) 300, Wölfflin, H. 174
463 – 465 working memory 3, 50, 510, 530,
virtual reality 350 562 – 563
visual aesthetic sensitivity test (VAST) 243 Wundt, W. 11
visual art 14 – 15, 18 – 19, 295 – 308; dynamic visual art
176; liking of 109, 113 – 117, 119; models of 297 – 301 Yarbus, A.L. 17
visual neglect 559
visual scenes; see landscapes Zajonc, R. 109, 475
visual system 13 – 14; anatomy of 48, 11 – 114; and liking Zeki, S. 14, 19 – 20
110 – 114 Zschokke, J.H. 8

575

You might also like