Sunteți pe pagina 1din 481

Emotion Regulation

and Psychopathology
in Children and Adolescents
ii
iii

Emotion Regulation
and Psychopathology
in Children and
Adolescents
Edited by

Cecilia A. Essau
Sara Leblanc
Thomas H. Ollendick

1
iv

1
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Oxford University Press 2017
The moral rights of the authors‌have been asserted
First Edition published in 2017
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2016963281
ISBN 978–​0–​19–​876584–​4
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Oxford University Press makes no representation, express or implied, that the
drug dosages in this book are correct. Readers must therefore always check
the product information and clinical procedures with the most up-​to-​date
published product information and data sheets provided by the manufacturers
and the most recent codes of conduct and safety regulations. The authors and
the publishers do not accept responsibility or legal liability for any errors in the
text or for the misuse or misapplication of material in this work. Except where
otherwise stated, drug dosages and recommendations are for the non-​pregnant
adult who is not breast-​feeding
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
v

Preface

This contemporary volume brings forefront research in emotion regulation and how processes
underlying emotion regulation have a bearing on the field of child and adolescent psychopathol-
ogy. The book shows continuity by initially introducing the topic of emotion and its regulation
and then narrowing its scope, analyzing the role emotion regulation plays in specific disorders
while critically examining current assessment and treatment strategies. In the concluding chap-
ters, emotion regulation in high risk, targeted groups is assessed and intervention and prevention
is explored.
This book has brought together an array of leading international scholars who specialize in
the emotional disorders. We have asked them to summarize the latest findings in their field
while assessing intervention through a comparative, critical lens in order to pass on this cru-
cial knowledge to the next generation of mental health professionals. Each chapter is unique, as
authors expose the reader to different approaches and outlooks from diverse specialties for diverse
problems.
This 20-╉chapter volume consists of four parts. In Part  1, broad issues are discussed such as
the biological, physiological and cultural factors underlying and impacting emotion regulation
and psychopathology in children and adolescents. In Part 2, specific disorders are delineated and
current treatment programs are discussed, including Attention Deficit Hyperactivity Disorder,
conduct disorder, anxiety disorders, depression, eating disorders, substance use disorders, autism
spectrum disorder, borderline personality disorder, and severe irritability and disruptive mood
dysregulation disorder. Part  3 assesses emotion dysregulation in specific targeted populations,
including children of abuse and neglect, children of divorce, children with incarcerated parents,
children exposed to traumatic stress, and adolescents who engage in nonsuicidal self-╉injury. It
investigates the interplay between environment, behavior and self-╉regulation and the etiology,
maintenance and propagation of psychopathology in these diverse environments. The final part of
this book conceptualizes emotional regulation as a transdiagnostic process and discusses innova-
tive approaches to treatment that arise when viewed through this lens.
This book combines the latest research from leading academics on a variety of clinical top-
ics with an emphasis on intervention from an applied perspective; this combination of appli-
cation and theory makes it a suitable reference for mental health professionals by providing
empirical review and current data on treatment efficacy. However, it was particularly designed
for graduate students taking advanced courses in clinical psychology and psychiatry who want to
remain abreast of current breakthroughs and leading treatment options for child and adolescent
psychopathology.
We wish to acknowledge the efforts of the contributors, whose expertise and dedication to the
project have been outstanding. Without them, a comprehensive coverage of the various topics
would not have been achieved. Additionally, we wish to acknowledge the support and cooperation
of the staff at Oxford University Press.
Cecilia A. Essau, Sara Leblanc, & Thomas H. Ollendick
vi
vi

Acknowledgments

I (Cecilia Essau) feel very honoured to have had this opportunity to co-╉edit this volume with
my highly respected colleague, Tom Ollendick, who’s been a great inspiration, mentor, scientist,
clinician and very good, patient and understanding friend to me, and with Sara Leblanc, who
introduced me to emotion regulation during her research. I wish to thank my family in Malaysia,
Canada and Germany, especially my husband, Juergen, and our daughter, Anna, for their continu-
ing support and inspiration. I dedicate this volume to my late parents, Essau Indit and Runyan
Megat, whose courage, love and belief in me have made me become who I am; had they still been
alive, they would have been most proud of this accomplishment and my choice of emotion regula-
tion strategies.
I (Sara LeBlanc) wish to express gratitude to my respected colleague Cecilia Essau for giving
me the opportunity to serve as a co-╉editor on this influential volume. Over the years Cecilia has
served as a role model, mentor and inspiration due to her humility, grace and impeccable ethic; it
was through her determination and vision that this work came to fruition. I also wish to express
my deepest thanks to Professor Ollendick, I am humbled and inspired by your contribution to
the field of Psychology, it was a privilege and honor to have the opportunity to work with you. I
also wish to thank my family and friends for their unwavering dedication and support, especially
my parents Blaine, Noreen and grandmother, Juanita. Finally, I also wish to thank my late sister
Amanda for our countless adventures; her valuable insights taught me to see the humor in all
things and have given me many memories I will eternally cherish, to her I dedicate this volume.
I (Tom Ollendick) wish to give thanks to my good friend and colleague, Cecilia Essau, who
invited me to serve as one of the co-╉editors of this important volume with her. This has been a
rewarding project and one that would not have been possible without her vision and dedication.
I also wish to thank Sara LeBlanc whom I have met through this project and with whom I would
very much like to work with in the future. Finally, I give thanks to my wife, Mary, our daughters,
Laurie and Katie, and our sons-╉in-╉law, David and Billy, as well as our six grandchildren, Braden,
Ethan, Calvin, Addison, Victoria and William. Without them, my life would be much less interest-
ing and enjoyable. I thank them for their love and support over the years. My own emotion regula-
tion has been much the better with them at my side. To them, I dedicate this work.
vi
ix

Table of Contents

List of Abbreviations╇ xi
List of Contributors╇ xv

Part I╇ Emotion Regulation: General Issues


1 Emotion Regulation: An Introduction╇ 3
Sara LeBlanc, Cecilia A. Essau, & Thomas H. Ollendick
2 The Relation of Self-╉Regulation to Children’s Externalizing
and Internalizing Problems╇ 18
Nancy Eisenberg, Maciel M. Hernández, & Tracy L. Spinrad
3 Biological and Physiological Aspects of Emotion Regulation╇ 43
Kateri McRae & Michelle Shiota
4 Cultural and Social Aspects of Emotion Regulation╇ 60
Selda Koydemir & Cecilia A. Essau
5 Research Domain Criteria (RDoC) and Emotion Regulation╇ 79
Michael Sun, Meghan Vinograd, Gregory A. Miller, & Michelle G. Craske

Part II╇Emotion Regulation and Child and Adolescent


Psychopathology
6 Emotion Regulation and Attention Deficit Hyperactivity Disorder╇ 113
Blossom Fernandes, Roseann Tan-╉Mansukhani, & Cecilia A. Essau
7 Emotion Regulation and Conduct Disorder: The Role
of Callous-╉Unemotional Traits╇ 129
Nicholas D. Thomson, Luna C. M. Centifanti, & Elizabeth A. Lemerise
8 Emotion Regulation and Anxiety: Developmental Psychopathology
and Treatment╇ 154
Dagmar Kr. Hannesdóttir & Thomas H. Ollendick
9 Emotion Regulation and Depression: Maintaining Equilibrium between
Positive and Negative Affect╇ 171
Frances Rice, Shiri Davidovich, & Sandra Dunsmuir
10 Emotion Regulation and Eating Disorders╇ 196
Julian Baudinet, Lisa Dawson, Sloane Madden, & Phillipa Hay
11 Emotion Regulation and Substance Use Disorders in Adolescents╇ 210
Thomas A. Wills, Jeffrey S. Simons, Olivia Manayan, & M. Koa Robinson
12 Emotion Regulation in Autism Spectrum Disorder╇ 235
Jonathan A. Weiss, Priscilla Burnham Riosa, Carla A. Mazefsky, & Renae Beaumont
13 Emotion Dysregulation in Adolescents with Borderline Personality
Disorder╇ 259
Carla Sharp & Timothy J. Trull
x

x Table of Contents

14 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation


Disorder  281
Katharina Kircanski, Ellen Leibenluft, & Melissa A. Brotman

Part III  Emotion Regulation in Specific Behavior/​Population


15 Children of Abuse and Neglect  305
Faye Riley, Anna Bokszczanin, & Cecilia A. Essau
16 Children of Divorce  331
Maria Caridad H. Tarroja, Ma. Araceli Balajadia-​Alcala,
& Maria Aurora Assumpta D. Catipon
17 Children’s and Adolescents’ Emotion Regulation in the Context of Parental
Incarceration  351
Janice Zeman & Danielle Dallaire
18 Children Exposed to Traumatic Stress  374
Brandon G. Scott & Carl F. Weems
19 Adolescents who Engage in Nonsuicidal Self-​injury (NSSI)  398
David Voon & Penelope Hasking

Part IV Epilogue
20 Transdiagnostic Approaches to Emotion Regulation: Basic Mechanisms
and Treatment Research  419
Brian C. Chu, Junwen Chen, Christina Mele, Andrea Temkin, & Justine Xue

Index  453
xi

List of Abbreviations

ACC anterior cingulate cortex DBT Dialectical Behavior Therapy


ACEs adverse child experiences DBT-​ST Dialectical Behavior Therapy
ADHD Attention Deficit Hyperactivity Skills Training
Disorder DERS Difficulties in Emotion
AG Agoraphobia Regulation Scale
ALS Affective Lability Scales DMDD disruptive mood dysregulation
disorder
AM autobiographical memory
DSM Diagnostic and Statistical
AN Anorexia Nervosa
Manual of Mental Disorders
ANS autonomic nervous system
DTS Distress Tolerance Scale
APA American Psychiatric
EA experiential avoidance
Association
EABT Emotion Acceptance Behavior
ARFID Avoidant/​Restrictive Food
Therapy
Intake Disorder
ECBT Emotion-​Focused Cognitive-​
ASD Autism spectrum disorder
Behavioral Therapy
BA behavioural activation
ED emotional dysregulation
BABCP British Association of
EF executive function
Behavioural and Cognitive
Psychotherapies EMA Ecological Momentary
Assessment
BP bipolar I disorder
EMDR eye-​movement sensitization
BPD borderline personality disorder
processing
BPS British Psychological Society
EMG electromyography
BSI Brief Symptom Inventory
EPM Extended Process Model
CAT Cognitive analytic therapy
ERPs event-​related potentials
CBITS Cognitive Behavioral
ERT Emotion Regulation Training
Intervention for Trauma in
Schools EUC Enhanced Usual Care
CBT cognitive behaviour therapy FBT family based treatment
CD conduct disorder FFCWB Fragile Families and Child
Well Being
CDC Centers for Disease Control
and Prevention FFT Functional family therapy
CERQ Cognitive Emotion Regulation fMRI functional magnetic resonance
Questionnaire, imaging
CGI Clinical Global Impression GAD generalized anxiety disorder
CIDI Composite International HCPC Health and Care Professions
Diagnostic Interview Council
CODIP Children of Divorce HED heavy episodic drinking
Intervention Program HF-​HRV high-​frequency heart rate
CP conduct problems variability
CSR clinical severity ratings HFASD high-​functioning autism
spectrum disorder
CU callous-​unemotional
HPA hypothalamic–p ​ ituitary–a​ drenal
CVR cardiovascular reactivity
xi

xii List of abbreviations

HRV heart rate variability PCMC-​A Parents and Children Making


ICD-​6 Sixth edition of the Connections—​Highlighting
International Classification of Attention
Diseases PD Panic Disorder
ID Intellectual disability PDs personality disorders
INS insomnia PFC prefrontal cortex
IPPA Inventory of Parent and Peer pgACC pregenual anterior
Attachment Inventory cingulate cortex
ISRE incarceration-​specific risk PMT Parent management training
experiences PNS peripheral nervous system
IY The Incredible Years PSST Problem-​Solving Skills
LHPA limbic-​hypothalamic-​pituitary-​ Training
adrenal PTS posttraumatic stress
LPE Limited Prosocial Emotions symptoms
LPP late positive potential PTSD posttraumatic stress disorder
MAAS Mindfulness Attention Scale RCTs randomized controlled trials
MACT Manual Assisted Cognitive-​ RDoC Research Domain Criteria
Behavioural Therapy RSA respiratory sinus arrhythmia
MANTRA Maudsley Model of Anorexia RVLPFC right ventral lateral
Nervosa Treatment for Adults prefrontal cortex
MBT mentalization based therapy RO-​DBT Radically-​Open DBT
MBT-​A MBT for adolescents SA sinoartial
MBT-​F MBT for families SAD Separation Anxiety Disorder
MDD major depressive disorder SAS-​OR Secret Agent Society-​Operation
MEAQ multi-​dimensional experiential Regulation
avoidance questionnaire SDQ Strengths and Difficulties
MPC medial prefrontal cortex Questionnaire
MST Multisystemic Therapy SHAPS Snaith-​Hamilton Pleasure
MTF The Monitoring the Future Scale
NBP New Beginnings Program SIB self-​injurious behavior
NC non-​clinical comparisons SM Selective Mutism
NICE The National Institute of SMD severe mood dysregulation
Clinical Excellence SNS sympathetic nervous system
NIMH National Institute for Mental SP Specific Phobias
Health’s SPACE Supportive Parenting for
NSDUH The National Survey on Drug Anxious Childhood
Use and Health Emotions
NSPCC National Society for the SSRIs selective serotonin reuptake
Prevention of Cruelty to inhibitors
Children SSRT Stop Signal Reaction
NSSI non-​suicidal self-​injury Time Task
ODD oppositional defiant disorder STAIR Skills Training in Affect and
OGM overgeneral Interpersonal Regulation
autobiographical memory STEPPS Systems Training for Emotional
OST One-​session treatment Predictability and Problem
Solving
PATHS Promoting Alternative
Thinking Strategies SUD substance use disorder
xi

List of abbreviations xiii

TAU treatment-​as-​usual UP-​Y Unified Protocol for the


TFP transference-​focused Treatment of Emotional
psychotherapy Disorders in Youth
TRY Thinking about Reward in WHO World Health Organization
Young People WPVA Word, Perception,
UP Unified Protocol Valuation, Action
xvi
xv

List of Contributors

Ma. Araceli Balajadia-​Alcala Blossom Fernandes


De La Salle University-​Manila, Philippines University of Roehampton, London, UK
Julian Baudinet Dagmar Kr. Hannesdóttir
Sydney Children’s Hospital Network Throska-​og hegdunarstod, Reykjavik, Iceland
(Westmead Campus), Sydney, Australia Penelope Hasking
Renae Beaumont Curtin University, Perth, Australia
University of Queensland, Australia Phillipa Hay
Anna Bokszczanin Western Sydney University, Sydney, Australia
University of Opole, Poland Maciel M. Hernández
Melissa A. Brotman Arizona State University,USA
National Institute of Mental Health, National Katharina Kircanski
Institutes of Health, Bethesda, MD, USA National Institute of Mental Health, National
Maria Aurora Assumpta D. Catipon Institutes of Health, Bethesda, MD, USA
InTouch Community Services, Makati City, Selda Koydemir
Philippines Middle East Technical University,
Luna C. M. Centifanti Northern Cyprus
University of Liverpool,UK Sara Leblanc
Junwen Chen College of New Caledonia, Quesnel, Canada
Flinders University, Adelaide, Australia Ellen Leibenluft
Brian C. Chu National Institute of Mental Health, National
Rutgers, The State University Institutes of Health, Bethesda, MD, USA
of New Jersey, USA Elizabeth A. Lemerise
Michelle G. Craske Western Kentucky University, USA
University of California, Los Angeles, USA Sloane Madden
Danielle Dallaire Sydney Children’s Hospital Network
College of William and Mary, USA (Westmead Campus), Sydney, Australia
Shiri Davidovich Olivia Manayan
University College London, London, UK University of Hawaii Cancer Center,
Lisa Dawson Honolulu, Hawaii, USA
University of Sydney, Sydney, Australia Carla A. Mazefsky
Sandra Dunsmuir University of Pittsburgh School of
University College London, London, UK Medicine,USA

Nancy Eisenberg Kateri McRae


Arizona State University, USA University of Denver, USA

Cecilia A. Essau
University of Roehampton, London, UK
xvi

xvi List of Contributors

Christina Mele Michael Sun
Rutgers, The State University University of California, Los Angeles, USA
of New Jersey, USA Roseann Tan-​Mansukhani
Gregory A. Miller De La Salle University-​Manila, Philippines
University of California, Los Angeles, USA Maria Caridad H. Tarroja
Thomas H. Ollendick De La Salle University-​Manila, Philippines
Virginia Polytechnic Institute and State Andrea Temkin
University, USA Rutgers, The State University
Frances Rice of New Jersey, USA
Cardiff University, UK & University College Nicholas D. Thomson
London, UK University of Durham, UK
Faye Riley Timothy J. Trull
University of Roehampton, London, UK University of Missouri, Columbia,
Priscilla Burnham Riosa Missouri, USA
York University, Canada Meghan Vinograd
M. Koa Robinson University of California, Los Angeles, USA
University of Hawaii Cancer Center, David Voon
Honolulu, Hawaii, USA Monash University, Melbourne, Australia
Brandon G. Scott Carl F. Weems
Arizona State University, REACH Iowa State University, USA
Institute, USA
Jonathan A. Weiss
Carla Sharp York University, Canada
University of Houston, Houston, Texas, USA
Thomas A. Wills
Michelle Shiota University of Hawaii Cancer Center,
Arizona State University, USA Honolulu, Hawaii, USA
Jeffrey S. Simons Justine Xue
University of South Dakota, USA Flinders University, Adelaide, Australia
Tracy L. Spinrad Janice Zeman
Arizona State University, USA College of William and Mary, USA
Part I

Emotion Regulation:
General Issues
2
3

Chapter 1

Emotion Regulation: An Introduction


Sara LeBlanc, Cecilia A. Essau, & Thomas H. Ollendick

Human emotions
Human emotions are an integral component of everyday life that influence cognitive functioning
(Bebko, Franconeri, Ochsner, & Chiao, 2011; Eysenck, 2004; Gross, 2013) memory (Christianson,
2014) and overall wellbeing (Kotsou, Gregoire, & Mikolajczak, 2011). Emotions impact both intra-
personal and interpersonal processes, and, when dysregulated, they may become destructive and
intrusive in daily life (Frijda, 1986; Slee, Arensman, Garnefski, & Spinhoven, 2007), contributing
to the development, maintenance, and propagation of psychopathology (Castella et al., 2013).
In general, emotion regulation competencies become differentiated as a function of develop-
ment. Children tend to seek support from adults or use behavioral techniques to regulate emo-
tions. As children reach adolescence, they become increasingly self-╉reliant, engaging in planful
problem solving and utilizing cognitive strategies (for example, reappraisal) more frequently
when faced with stressful life events (Zimmer-╉Gembeck & Skinner, 2011). Although the majority
of children and adolescents will successfully navigate these developmental stages by cultivating
adaptive coping skills, for some, this marks the beginning of lifelong challenges with emotion
regulation and resultant dysregulation (Kessler et al., 2005).
This introductory chapter will begin by discussing both the definition and functionality of emo-
tions; it will then turn to a discussion of emotion regulation and associated processes. Critically, it
will consider the importance of this topic as it pertains to emotional wellbeing, whilst also examin-
ing the crucial link between emotion dysregulation and psychopathology in children and adoles-
cents. Later in the chapter, various emotion regulation strategies will be described and categorized
according to their utility, emphasizing strategies that demonstrate adaptive social, cognitive and
physiological benefits. This information is critical in delineating the underlying mechanisms leading
to the development and propagation of psychopathology, which is crucial when tailoring effective
treatment and prevention programs specifically suited to both the child and adolescent populace.

Definition and function of emotions


Emotion is a dynamic and convoluted construct. A comprehensive definition of emotions must
consider three key aspects, which include:  The conscious experience elicited via the emotion;
the underlying neurological processes involved in emotion generation and finally, the observable
behavior and facial expression evoked by the emotion (Izard, 2013). Universally accepted within
the literature, emotion has been broadly defined as “[a]â•„person-╉situation transaction that compels
attention, has particular meaning to an individual, and gives rise to a coordinated, yet flexible,
multi-╉system response to the on-╉going person-╉situation transaction” (Thompson, 2007, p. 5).
Emotions are important due to their relative influence on cognition, appraisal processes, per-
ception, and ultimately, behavior. They impact decision-╉making (Cassotti, Habib, Poirel, Aïte, &
Moutier, 2012; Mikels, Maglio, Reed, & Kaplowitz, 2011; Oatley & Johnson-╉Laird, 1987), learning
(Ahmed, van der Werf, Kuyper, & Minnaert, 2013; Cahill, Prins, Weber, & McGaugh, 1994) and
4

4 Emotion Regulation: An Introduction

drive goal pursuits (Koole, 2009; Tice, Bratslavsky, & Baumeister, 2001). It has been hypothesized
that emotions evolved to promote the species by eliciting specific action patterned responses to
life threatening circumstance, thereby increasing the likelihood of survival. From this stand-╉point
a negative bias would be adaptive; for example, in prehistoric times hearing a rustling in the bush
if one was likely to interpret this as a threat, feel fear, and ultimately flee, one would be more likely
to survive than if a more positive appraisal was made, viewing the sound as innocuous, rather
than life threatening (Sapolsky, 2007). However, in post-╉industrial societies humans are often
faced with psycho-╉social stressors, which activate the fight-╉or-╉flight response (Hypothalamic–╉
pituitary–╉adrenal axis: HPA axis) in the same manner even though they are no longer placed in
life threatening circumstances. This chronic activation can have a deleterious impact on overall
wellbeing if stress levels are not regulated.
In terms of adaptive function, the positive emotions may be facilitative, as they broaden atten-
tional focus (Derryberry & Tucker, 1994) whilst concurrently enhancing the scope of cognition.
For example, a series of classical experiments demonstrated that when compared with a control
condition, those in a positive state were able to make more unique associations with neutral words
(Isen, Johnson, Mertz, & Robinson, 1985). This led researchers to conclude that positive affect
enhances cognitive processing via the promotion of cognitive flexibility, elaboration, and integra-
tion, whilst concomitantly fostering relatedness and interconnection between cognition, ideas,
and action (Isen, 1987; Isen & Daubman, 1984).

Emotions in childhood and adolescence


In terms of development, infants will vary in individual difference with regards to the intensity
and frequency at which they express emotion. Additionally, the way caregivers respond to their
expression of emotion is critical in the development of their emotion regulation competencies
and their attachment style (Izard, 2013). The famous strange situation experiment (Ainsworth
& Wittig, 1969) demonstrated varying attachment styles of children between the ages of 12 and
24 months based on their emotional response when placed in various stressful situations. Their
behavior towards their caregiver in these situations allowed their attachment to be categorized as
either secure, insecure avoidant, and insecure ambivalent/╉resistant. Insecure attachment occurred
when the infant’s emotional needs were not adequately met by the caregiver, this transactional
process impacted both the behavior of the child and caregiver. Accordingly, insecure attachment
has been associated with an increased risk of emotional, interpersonal, and behavioral problems
(Dang & Gorzalka, 2015; Kobak et al., 1993).
In terms of development, emotions impact personality development two-╉fold (John & Gross,
2004). Firstly, genetic predisposition plays an integral role in establishing core traits, propensities,
and thresholds for various emotive states (Hariri & Forbes 2007). The second key feature is the
child’s experiences (Campos, Walle, Dahl, & Main 2011) and key learnings relating to their emo-
tional health with particular importance placed on how the expression of emotion and regulation
is socialized (Izard, 2013). In addition, an individual’s development of their emotional traits will
play a critical role in their social development such that the child who is quick to anger, frightens
easily, frequently smiles, will attract and receive differentiated responses based on their behav-
ior (Van Reekum & Scherer, 1997). Thus, emotional development influences social development
and also plays a critical role in intellectual development. An infant who is frequently distressed
or afraid will be far less likely to explore their environment when compared with a child who is
content and curious. Tomkins (1962) asserts that the emotion “interest” is a critical component
required for intellectual development. Thus, adaptive emotional development serves a myriad of
functions that influence social, intellectual and interpersonal growth.
5

Emotion regulation 5

Emotion regulation
Emotions are complex and dynamic: They can be useful or deleterious. Thus, the key to optimum
emotional functioning is adaptive emotion regulation, which is characterized by implementing
effective strategies that are contextually appropriate and account for individual difference and
personal preference (Gross & John, 2003).
Varying definitions of emotion regulation exist within the developmental literature (Cole,
Martin & Dennis, 2004). For example, according to Gross (1998) emotion regulation refers to the
heterogeneous set of processes individuals implement to modulate their emotional experiences.
This definition subsumes both the “up” and “down” regulation of emotions, as an individual may
decrease, increase or maintain negative and positive emotions (Erber, Wegner, & Therriault, 1996;
Parrott, 1993). Alternatively, emotion regulation has been defined as a “[p]â•„rocess used to man-
age and change if, when, and how (e.g., how intensely) one experiences emotions and emotion-╉
related motivational and physiological states, as well as how emotions are expressed behaviorally”
(Eisenberg et al. 2007, p. 288). Eisenberg and Spinrad (2004) posit that although intrinsic and
extrinsic factors play a role in emotion regulation, it is advantageous to distinguish between exter-
nal and internal regulation. External regulation refers to external forces, such as parents, teachers,
and peers, which influence emotion regulation. This may be particularly pertinent in the early
childhood years, when support seeking from adults is a primary form of affect regulation in nor-
mative development (Zimmer-╉Gembeck & Skinner, 2011). In contrast, internal regulation refers
to effortful, self-╉regulation, which may include a variety of cognitive and behavioral strategies an
individual chooses to implement to modulate their emotional response.
The primary focus of this chapter will be internal self-╉regulation, as this type of regulation
is within the individual’s control and can be shaped through directed intervention, a topic that
will be discussed in greater detail in Chapter 2. The definition utilized within this chapter will
be consistent with the aforementioned definition put forth by Gross (1998), who views emotion
regulation as a varied set of processes individuals engage in to modify their emotional experience.
The ability to effectively regulate emotions is a critical and common place activity (Oschner
& Gross, 2005). Various strategies may be employed that are broadly categorized as antecedent-╉
focused or response-╉focused strategies (Gross, 1998). Antecedent strategies occur early in the
emotion generative process, altering the impact of emotion-╉eliciting cues; whereas, response-╉
focused emotion regulation occurs later in the process, impacting behavioral responses (Gross
& Thompson, 2007). Emotion regulation influences the intensity, duration, and expression of
emotions (Gross, 1999), occurring on a continuum from controlled to automatic, conscious to
unconscious (Koole, 2009).
Research has demonstrated the vast majority of emotional experience can be regulated (Canli,
Ferri & Dunman, 2009). There are a variety of different strategies which can be employed to regu-
late emotions which include: Reappraisal of the event (Hofmann, Heering, Sawyer, & Asnaani,
2009), situation modification (Gross, 1998), change of attentional focus (Rothermund, Voss, &
Wentura, 2008), and suppression (Dalgleish, Schweizer, & Dunn, 2009). The two strategies that
will be primarily focused on in this chapter are reappraisal and emotional suppression, as they
have received the most attention in the literature, with reappraisal primarily associated with posi-
tive health outcomes and suppression, primarily associated with negative health outcomes (Gross
& John, 2003).
Cognitive reappraisal is an antecedent technique that involves changing the interpretation of a
situation in order to reduce the emotional impact (Gross & Thompson, 2007). Perception is real-
ity and our thoughts are linked to our actions, which are linked to our behavior: Every situation
can be interpreted in a variety of different ways, and it is this interpretation, rather than the event
6

6 Emotion Regulation: An Introduction

itself, that impacts thoughts, behaviors, and emotions (Malooly, Genet, & Siemer, 2013; Wilding
& Milne, 2010). In children, reappraisal has been shown to be an adaptive method of managing
emotions, when compared to other strategies, such as suppression (Carthy et al., 2010; Garnefski
& Kraaij, 2009; McKrae, et  al., 2012). In addition, both longitudinally and cross-╉sectionally,
Garnefski and colleagues have demonstrated a strong negative relationship between the reported
use of reappraisal and depression in both adolescent and adult populations (Garnefski & Kraaij,
2006; Garnefski, Kraaij, & Spinhoven, 2001; Kraaij, Pruymboom, & Garnefski, 2002). Similarly,
research has shown children possessing a secure attachment style are more empathetic due to
their superior emotion regulation competencies (Panfile & Laible, 2012). These findings are mir-
rored in the adolescent populace, as adolescents demonstrating adaptive emotional regulation
competencies are more likely to achieve their goals and form strong interpersonal relationships;
whereas, adolescents with impaired emotion regulation skills, often manifest behavioral problems
and are less likely to achieve both long and short-╉term goals (Hum & Lewis, 2013).
In contrast to reappraisal, emotional suppression has been shown to have negative health out-
comes, as studies have linked the frequent use of suppression with depressive symptomology
in both children and adolescents (Betts, Gullone, & Allan, 2009; Hughes, Gullone, & Watson,
2011; Larsen et  al., 2013). Relatedly, in adolescents, deficits in emotion regulation have been
associated with substance abuse (Wilens et al., 2013), aggressive behavior (Herts, McLaughlin, &
Hatzenbuehler, 2012) and pathological gambling (Potenza, et al., 2011), topics which will be dis-
cussed in greater detail in subsequent chapters. In general, suppression has been associated with
increased negative affect (Srivastava et al., 2009), decreased positive affect (Gross & John 2003),
decreased social functioning (English & John, 2013), and enhanced levels of depressive sympto-
mology and obsessive thinking (Corcoran & Woody, 2009; Marcks & Woods, 2005). Furthermore,
suppression has been linked to decreased life satisfaction (Kashdan & Steger, 2006), decreased
interpersonal skills (Butler et al., 2003), enhanced sympathetic nervous-╉system activation (Egloff,
Schmuckle, Burns, & Schwerdtfeger, 2006), increased stress-╉related symptomology (Moore,
Zoellner, & Mollenholt, 2008)  and decreased memory recall (Richards, Butler, & Gross, 2003;
Richards & Gross, 2000). In conclusion, the frequent and inflexible use of emotional suppression
may be damaging as it prolongs the experience of negative affect (Campbell-╉Sills & Barlow, 2007),
makes excessive use of cognitive resources (Gross & John, 2003) and keeps physiological arousal
chronically activated (Eglof et al., 2006; Ohira et al., 2006). Thus, the cultivation of adaptive strat-
egies, such as cognitive reappraisal, is imperative during the formative years so that the use of
emotional suppression is minimized.

Function of ER
Historically, it was hypothesized that emotion regulation functioned to satisfy hedonic needs,
such that pleasure was maximized and pain minimized (Larsen, 2000; Westen, 1994). This may be
due in part to the realization that negative emotions drain an extensive amount of an individual’s
physical and mental resources (Sapolsky, 2004; 2007). However, even though hedonic needs may
fuel emotion-╉regulation in some circumstances, they are not the sole motivation for all regulatory
function (Erber & Erber, 2000; Erber, Wegner, & Therriault, 1996). For example, if one deems
their emotions to be beneficial they may choose to stay in that emotional state even though it is
associated with negative and unpleasant feelings (Gross, 2007).
Relatedly, goal pursuits may influence emotion regulation tendencies, leading to short-╉term
discomfort in the quest towards delayed gratification based on a strong commitment to long-╉term
goals (Mischel et al., 2010; Mischel & Ayduk, 2004). Delayed gratification is a common paradigm
employed in the investigation of emotion regulation competencies in children dating back to the
7

FACTORS INFLUENCING EMOTION REGULATION DEVELOPMENT 7

1970’s. Early research showed that some children were able to practice emotion-╉regulation strate-
gies, such as reframing and distraction, to delay gratification in the interest of garnering a greater
reward at a later time. A recent follow-╉up of the original studies conducted by Mischel and col-
leagues, demonstrated the predictive validity of the delayed gratification test across a wide range
of social, cognitive, and mental health indicators (Casey et al., 2011). Thus, one can infer from
this research that the absence of delayed gratification in children can be an early sign of emotion
dysregulation. In support, a study by Krueger and colleagues (1996) determined that, in pre-╉
adolescents, the inability to delay gratification was linked to the externalizing disorders. Relatedly,
the work of Shoda et al. (1990), determined that preschoolers’ performance on the delayed grati-
fication task accurately predicted behavioral problems from age five to eight. In part, this may be
attributed to deficiencies in attentional control and executive function.

Executive function and emotion regulation


Attentional processes play a key role in one’s ability to regulate motivation and emotional arousal;
therefore, executive function is considered a key component of effective emotion regulation.
Executive function (EF) is a multidimensional construct relating to the processes that exercise
control over cognition, attention, and behaviors (Blair, Zelazo, & Greenberg, 2005). EF is goal
oriented and involves higher order, self-╉regulatory processes (Nelson, Thomas, & Hann, 2006).
EF emerges during the end of infancy and shows striking changes during the preschool years,
continuing to develop throughout adolescence (Zelazo et al., 2008). The literature on EF asserts
that an important corollary of cognitive development in early childhood is the ability to diminish
the emotional impact of disruptive and distractive stimuli. A study by Ursache, Blair, Stifter, and
Voegtline (2013), for example, determined that high levels of executive function were associated
with children who exhibited high levels of emotional reactivity in conjunction with high levels of
emotion regulation competencies. In addition, children that rated high in both emotional reactiv-
ity and emotion regulation where more likely to have increased levels of adaptive parenting.
In adolescence, the cultivation of emotion regulation strategies shows an increase in the use of
reappraisal and a reduction in the use of suppression, a progression that mirrors developmental
changes occurring in executive functions during this period (Lantrip, Isquith, Koven, Welsh, &
Roth, 2015). More specifically, these researchers determined that the increased use of reappraisal
was associated with improved executive function; whereas, the increased use of suppression was
associated with worsened executive function in an adolescent sample. Deficiencies in EF are criti-
cal during this period, as they have been associated with a plethora of early onset psychiatric
disorders, such as conduct disorder and attention deficit/╉hyperactivity disorder (Willcutt, Doyle,
Nigg, Faraone, & Pennington, 2005), in addition to behavioral problems such as substance abuse
and physical aggression (Séguin & Zelazo, 2005). Thus, EF is intimately tied to emotion regulation.

Factors influencing emotion regulation development


Emotion regulation is influenced by a variety of genetic, biological and environmental factors.
Children demonstrate enhanced control of both affect and behavior, shifting control from the
brain’s orienting neuronal network during infancy to greater use of executive functions by the
age of three to four years. From infancy to toddlerhood connectivity changes in the following
way: During the early years, parietal and frontal areas play a crucial role in orienting; whereas,
executive function and emotion regulation is regulated via the midfrontal and anterior cingu-
late regions of the brain in the later years (Rothbart, Sheese, Rueda, & Posner, 2011). On a neu-
robiological level, individual variations in serotonin levels have been identified, which impact
8

8 Emotion Regulation: An Introduction

an individual’s emotional expressivity and regulation (Hariri & Forbes, 2007). Concordantly,
dysregulation of the dopaminergic system has been associated with major depressive disorder
(Kennedy, Koeppe, Young, & Zubieta, 2006). Moreover, there seems to be a distinct genetic com-
ponent to emotion regulation, as evidenced by twin studies, which have shown that identical
twins are more similar in emotional control, when compared with fraternal twins (Goldsmith,
Buss, & Lemery, 1997).
From a neurobiological stand-​point, the development of the prefrontal cortex, hippocampus,
and amygdala is associated with higher decision making processes, sustained attentional control
and the enhanced capacity to regulate one’s emotions (Ochsner & Gross, 2007) (see Chapter 3).
Environmental factors also play a key role in the development of emotion regulation competen-
cies, particularly in infancy and early childhood. By six months of age an infant’s primary form of
emotion regulation occurs through relative interactions with caregivers (Crockenberg & Leerkes,
2004). However, as the child ages, they are influenced by numerous factors such as the parents’
regulatory style, social referencing, peer influence, and parental reactions to their children’s dis-
plays of emotion (Zeman, Cassano, Perry-​Parrish, & Stegall, 2006).
Furthermore, culture impacts emotion regulation in a myriad of ways, by determining what
is valued (i.e., saving face, personal autonomy, etc.), which behaviors are socially acceptable and
what is deemed appropriate behavior in varying contexts (see Chapter 4). This was illustrated in
a study by Kagan (2003) who compared cultural norms in American and Chinese cultures. This
study determined that children in the American sample were socialized to be outgoing, assertive,
and bold; thus, children were taught to be highly expressive of both positive and negative affect.
In contrast, in the Chinese sample, shyness was seen as a positive attribute, as it demonstrated the
child was studious, hard-​working, and willing to prescribe to social norms. Relatedly, in some
Asian cultures, emotional suppression is considered an adaptive emotion-​regulation strategy,
unlike in autonomous cultures (e.g., Australia, America, and the UK); therefore, its consequences
do not manifest negatively in these cultures (Butler, Lee, & Gross, 2007) the way they do in cul-
tures subscribing to Western-​European value systems. Thus, adaptive emotion regulation is con-
textually specific and culturally motivated.

Emotion regulation development during childhood


and adolescence
Changes in emotion regulation strategies become evident during the first few years of life. At this
time, regulation becomes less reflexive (i.e., newborns) and more intentional, involving behav-
ioral control in the absence of external input or monitoring from parents (i.e., self-​regulation, see
Kopp & Neufeld, 2003). In young infants and toddlers, behavior such as self-​soothing (e.g., thumb
sucking; Ekas, Lickenbrock, Braungart-​Rieker, 2013), reorienting attention (Wiebe et al., 2011),
and distracting one’s gaze from negative stimuli (Crockenberg & Leerkes 2004) have been shown
to reduce negative affect. In addition, as noted, research has shown an increase in both executive
function and effortful self-​regulation as children age (Eisenberg et al., 2010).
As noted above, a common measure of emotional control in children is their ability to delay
gratification. Improvements in the ability to delay gratification are found from 24 months to four
years (Li-​Grining, 2007). Further improvements in self-​control and executive function occur dur-
ing the late pre-​school years (Mezzacappa, 2004); moreover, substantial development in emotion
regulation is manifested between six to ten years of age (Stegge & Terwogt, 2007). At this stage
of development, critical changes in regulatory competencies occur, as children learn to identify,
understand, and analyze emotion-​eliciting situations in a cause and effect way, whilst also discov-
ering alternative ways of expressing their feelings (Stegge & Terwogt, 2007). Thus, a large body
9

Emotion regulation and psychopathology in children and adolescents 9

of literature has demonstrated that a healthier emotion regulation profile is demonstrated as a


function of age and maturation (John & Gross, 2004; Silvers, McRae, Gabrieli, Gross, Remy, &
Ochsner, 2012; Tottenham, Hare, & Casey, 2011).
Sex differences in emotion regulation strategies also occur in line with the adult literature,
which have demonstrated that males employ emotional suppression more frequently than
females (Eisenberg, Spinrad, & Eggum, 2010). In part, this may be due to socialization processes.
Adolescent emotional development may be influenced by a variety of factors, such as interactions
with peers, parents, and teachers, as well as societal influences such as the Internet, media, and
contemporary culture (Klimes-​Dougan et al., 2007; Morris et al., 2007).
One line of scientific inquiry has investigated the importance of parental influence on ado-
lescents’ emotional development (Yap et al., 2008). This research determined that parental style
of emotion regulation (i.e., suppressive, hostile, controlling vs. caring and warm) (Jaffe et  al.
2010) and parental expression of emotions as well as their reactions to their children’s displays
of emotion (Morris et  al. 2007)  are all important factors influencing regulatory development.
Eisenberg and colleagues (1998) posit that socialization of emotion regulation occurs in three
primary ways, namely, 1) the socializers’ expression of their own emotions 2) the socializers’ reac-
tion to the children’s display of emotions and the 3) the socializers’ amenability towards discussing
emotion (see Chapter 2 by Eisenberg and her colleagues on the developmental aspects of emotion
regulation).

Emotion regulation and psychopathology in children


and adolescents
In children aged two to five years, general rates of psychopathology are 16.2 overall, 10.5% for
internalizing/​emotional disorders (see Chapters 8 and 9) and 9% for externalizing/​behavioral dis-
orders (see Chapters 6 and 7), rates which are comparable to those found among older children
(Egger & Angold, 2006). Alarmingly, in children, levels of anxiety and depression have increased
continually at a dramatic rate since the 1950s (Gray, 2011). Research has shown two broad types
of contributing factors: Environmental context/​events and child temperament. In relation to envi-
ronmental factors, this category includes both specific and global elements; in general, parental
supervision, peer problems and sexual abuse have been shown to predict externalizing disorders,
neglect has been linked to oppositional defiant disorder (ODD), lack of social support and expo-
sure to violence have been associated with the internalizing disorders, while childrearing factors
such as a controlling family environment have been linked to the anxiety and depressive disorders
(Dierker & Szatmari, 1998; Jaffee, et al., 2002; Merikangas, et al., 2010; Rapee, 1997). However, it
is important to note that risk factors are highly complex and it is likely that disorders do not have
a single cause but rather a causal chain or multiple causal chains that are influenced by the interac-
tion between various environmental, social, genetic, and biological risk factors (Kraemer, Stice,
Kazdin, Offord, & Kupfer, 2014).
Adolescence is a period characterized by marked changes occurring on a hormonal, neu-
rological, and developmental level. These changes directly impact emotionality, affecting both
the valence and intensity of negative and positive emotions, providing an opportune time for
regulatory skills to be cultivated and honed (Silk et  al., 2003). In general, research has shown
more extreme mood states (both positive and negative) are typical in adolescent daily life when
compared with the adult demographic (Larson, Moneta, Richards, & Wilson, 2002; Larson &
Richards, 1994).
Roughly 20% of adolescents have a psychiatric disorder (McLeod, Uemura, & Rohrman,
2012). An epidemiological study by Costello, Copeland, and Angold (2011) determined the
01

10 Emotion Regulation: An Introduction

prevalence rate for an anxiety disorder was 10.2%, with average onset occurring at eight
years of age and 50% of cases falling between six to twelve years of age. Similarly, an epi-
demiological study investigating general rates of psychopathology in high school students
found 10% of students currently had a clinical disorder and 33% had experienced one in their
life-​times. Additionally, within this sample, high relapse rates were found for both substance
abuse (15%) and depression (18%) (Lewinsohn, Hops, Roberts, Seeley, & Andrews, 1993).
Relatedly, a large scale, longitudinal study by Essau, Lewinsohn, Olaya, and Seeley (2014)
determined that adolescent anxiety predicted poor adjustment across a variety of domains
(work, family etc.); reduced life satisfaction, substance, alcohol abuse/​dependency, and anxi-
ety in adulthood in a large, community sample of 800 participants.
In the developmental literature, the primary area of interest has focused on children’s malad-
justed emotion regulation (Eisenberg, Spinrad, & Eggum, 2010). This research has investigated
when normal emotional development is compromised and identified the risk factors associated
with atypical development (Cicchetti & Cohen 2006). Factors influencing the development of
emotion regulation skills include inherent disposition in addition to social and environmental
resources available to the child. Additionally, genetic pre-​disposition and parental influences
have been shown to influence the development of psychopathology in adolescents (Rosenstein &
Horowitz, 1996).
From a genetic stand-​point, a twin study by Eaves (2006) measuring symptoms of psycho-
pathology, demonstrated that monozygotic twins were more strongly correlated than dizygotic
twins with most measures showing small to moderate genetic effects. Concordantly, in relation
to depression, the majority of twin studies suggest a moderate genetic influence, with heritability
rates ranging between 30–​80% (Eley & Plomin, 1997; Murray & Sines, 1996; Thapar & McGuffin,
1997). In general, these studies support genetic susceptibility to psychopathology across a broad
range of disorders in adolescent populations.
In relation to parental influences on emotion regulation propensities, research supports a para-
digm of adolescent psychopathology that is influenced by interpersonal interactions with parents
(Rosenstein & Horowitz, 1996). This is supported by the work of Grant (2006), who found con-
siderable evidence supporting the mediating role that family relationships play in the relation-
ship between stressors and psychological symptoms in both children and adolescents. A study
by Rosenstein and Horowitz (1996) determined, in a clinical sample of 60 adolescents psychiatri-
cally hospitalized, both child and maternal attachment style were highly concordant, manifesting
insecure attachments styles in both the adolescent and the parent. In general, when researching
parental influence on adolescent psychopathology, fathers have been highly underrepresented.
However, research shows there is substantial paternal influence; with particularly strong effects
found with relation to externalizing problems manifested in adolescents. In most cases these
effects were comparable to those associated with maternal psychopathology (Phares & Compas,
1992). A study by Achenbach (1991) determined that in four-​to eight-​year-​olds, externalizing
problems were associated with difficulties in emotion regulation including increased levels of
anger and impulsivity. Similarly, internalizing problems were associated with enhanced levels of
sadness, impulsivity, and reduced attentional control (Eisenberg et al., 2001). These relationships
were investigated via a longitudinal design and similar findings were obtained two years later
(Eisenberg et al., 2005). In children, certain components of emotion regulation have been associ-
ated with particular behavioral difficulties. For example, inhibiting anger or expressing anger in
a maladaptive way has been linked to internalizing problems (Zeman, Shipman, & Suveg, 2002).
Similarly, in a sample of eight-​to twelve-​year olds with various anxiety disorders, a significant
relationship was demonstrated between psychiatric disorder and maladaptive emotion regulation
as assessed via both self and parent report measures (Suveg & Zeman, 2004). More specifically,
1

Conclusion 11

children with anxiety disorders were more likely to be inflexible, demonstrating heightened worry,
anger and negative affect when compared with children in the control conditions. Likewise, a
recent study by Tortella-╉Feliu, Balle, and Sesé (2010), determined that adolescents scoring high in
negative affect were prone to implement dysfunctional emotion regulation coping styles.

Conclusion
Emotion dysregulation is strongly associated with psychiatric illness in youth. As mentioned pre-
viously, in both children (Hughes, Gullone, & Watson, 2011) and adolescents, the use of emotional
suppression has been linked to depressive and anxious symptomology (Betts, 2009; Hannesdottir
& Ollendick, 2007; Larsen et  al., 2013). Furthermore, in adolescents, deficits in emotion regu-
lation have been linked with aggressive behavior (Herts, McLaughlin, & Hatzenbuehler, 2012),
substance abuse (Wilens et al., 2013), and pathological gambling (Potenza, et al., 2011). Due to the
strong association between emotion dysregulation and psychopathology and related problems in
living, many studies have been conducted on this topic in the past 15–╉20 years. This book includes
a collection of these studies, touching on numerous contemporary topics, such as developmen-
tal psychology, developmental psychopathology, transdiagonostic issues, and cultural aspects of
emotion regulation with exciting incites from leading researchers in the field.

References
Achenbach, T. M. (1991). Child behavior checklist/╉4–╉18. Burlington: University of Vermont.
Ainsworth, M. D. S., & Wittig, B. A. (1969). Determinants of infant behavior. Attachment and exploratory
behavior of one-╉olds in a strange situation, 4.
Ahmed, W., Van der Werf, G., Kuyper, H., & Minnaert, A. (2013). Emotions, self-╉regulated learning,
and achievement in mathematics: A growth curve analysis. Journal of Educational Psychology, 105(1),
150–╉161.
Betts, J., Gullone, E., & Allen, J. S. (2009). An examination of emotion regulation, temperament, and
parenting style as potential predictors of adolescent depression risk status: A correlational study. British
Journal of Developmental Psychology, 27(2), 473–╉485.
Blair, C., Zelazo, P. D., & Greenberg, M. T. (2005). The measurement of executive function in early
childhood. Developmental Neuropsychology, 28(2), 561–╉571.
Butler, E. A., Lee, T. L., & Gross, J. J. (2007). Emotion regulation and culture: are the social consequences of
emotion suppression culture-╉specific? Emotion, 7(1), 30.
Cahill, L., Prins, B., Weber, M., & McGaugh, J. L. (1994). β-╉Adrenergic activation and memory for
emotional events. Nature, 371(6499), 702–╉704.
Carthy, T., Horesh, N., Apter, A., Edge, M. D., & Gross, J. J. (2010). Emotional reactivity and cognitive
regulation in anxious children. Behaviour Research and Therapy, 48(5), 384–╉393.
Campbell-╉Sills, L., & Barlow, D. H. (2007). Incorporating emotion regulation into conceptualizations and
treatments of anxiety and mood disorders. In Gross (Ed.), Handbook of Emotion Regulation (pp. 542–╉
559). New York: Guilford Press.
Campos, J. J., Walle, E., Dahl, A., & Main, A. (2011). Reconceptualizing emotion regulation. Emotion
Review, 3(1), 26–╉35.
Canli, T., Ferri, J., & Duman, E.A. (2009). Genetics of emotion regulation. Special Issue: Neurogenetics as
applied to systems and cognitive neuroscience. Neuroscience, 164, 43–╉54.
Casey, B. J., Somerville, L. H., Gotlib, I. H., Ayduk, O., Franklin, N. T., Askren, M. K., … & Shoda, Y.
(2011). Behavioral and neural correlates of delay of gratification 40 years later. Proceedings of the
National Academy of Sciences, 108(36), 14998–╉15003.
Cassotti, M., Habib, M., Poirel, N., Aïte, A., Houdé, O., & Moutier, S. (2012). Positive emotional context
eliminates the framing effect in decision-╉making. Emotion, 12(5), 926–╉931.
21

12 Emotion Regulation: An Introduction

Christianson, S. A. (2014). The handbook of emotion and memory: Research and theory. Psychology Press,
New York.
Cicchetti, D., & Cohen, D. J. (2006). Developmental Psychopathology. Wiley, Hoboken, New Jersey.
Cole, P. M., Martin, S. E., & Dennis, T. A. (2004). Emotion regulation as a scientific
construct: Methodological challenges and directions for child development research. Child
Development, 75(2), 317–​333.
Corcoran, K. M. & Woody, S. R. (2009). Effects of thought suppression and appraisals on thought
frequency and distress. Behaviour Research and Therapy, 47(12), 1024–​1031.
Costello, E. J., Egger, H. L., Copeland, W., Erkanli, A., & Angold, A. (2011). The developmental
epidemiology of anxiety disorders: phenomenology, prevalence, and comorbidity. Anxiety Disorders in
Children and Adolescents: Research, Assessment and Intervention, 14(4), 56–​75.
Crockenberg, S. C., & Leerkes, E. M. (2004). Infant and maternal behaviors regulate infant reactivity to
novelty at 6 months. Developmental Psychology, 40(6), 1123.
Dalgleish, T., Yiend, J., Schweizer, S., & Dunn, B. D. (2009). Ironic effects of emotion suppression when
recounting distressing memories. Emotion, 9(5), 744.
Dang, S. S., & Gorzalka, B. B. (2015). Insecure attachment style and dysfunctional sexual beliefs predict
sexual coercion proclivity in university Men. Sexual medicine, 3(2), 99–​108.
De Castella, K., Goldin, P., Jazaieri, H., Ziv, M., Dweck, C. S., & Gross, J. J. (2013). Beliefs about
emotion: Links to emotion regulation, well-​being, and psychological distress. Basic and Applied Social
Psychology, 35(6), 497–​505.
Derryberry, D., & Tucker, D. M. (1994). Motivating the focus of attention. In The heart’s eye: Emotional
influences in perception and attention, (pp. 167–​196). San Diego, CA, US: Academic Press, xiv, 289 pp.
http://​dx.doi.org/​10.1016/​B978-​0-​12-​410560-​7.50014-​4
Eaves, L. J. (2006). Genotype× environment interaction in psychopathology: fact or artifact? Twin Research
and Human Genetics, 9(01), 1–​8.
Egger H. L, Angold A. (2006). Common emotional and behavioural disorders in preschool
children: presentation, nosology, and epidemiology. Journal of Child Psychology and Psychiatry. 47(3–​4),
313–​37.
Egloff, B., Schmukle, S. C., Burns, L. R., & Schwerdtfeger, A. (2006). Spontaneous emotion regulation
during evaluated speaking tasks: associations with negative affect, anxiety expression, memory, and
physiological responding. Emotion, 6(3), 356–​366.
Eisenberg, N., Cumberland, A., & Spinrad, T. L. (1998). Parental socialization of emotion. Psychological
Inquiry, 9(4), 241–​273.
Eisenberg, N., Cumberland, A., Spinrad, T. L., Fabes, R. A., Shepard, S. A., Reiser, M., … & Guthrie, I.
K. (2001). The relations of regulation and emotionality to children’s externalizing and internalizing
problem behavior. Child Development, 72(4), 1112–​1134.
Eisenberg, N., Zhou, Q., Spinrad, T. L., Valiente, C., Fabes, R. A., & Liew, J. (2005). Relations Among
Positive Parenting, Children’s Effortful Control, and Externalizing Problems: A Three‐Wave
Longitudinal Study. Child Development, 76(5), 1055–​1071.
Eisenberg, D., Gollust, S. E., Golberstein, E., & Hefner, J. L. (2007). Prevalence and correlates of
depression, anxiety, and suicidality among university students. American Journal of Orthopsychiatry,
77(4), 534–​542.
Eisenberg, N., Spinrad, T. L., & Eggum, N. D. (2010). Emotion-​related self-​regulation and its relation to
children’s maladjustment. Annual Review of Clinical Psychology, 6, 495.
Ekas, N. V., Lickenbrock, D. M., & Braungart‐Rieker, J. M. (2013). Developmental Trajectories of Emotion
Regulation Across Infancy: Do Age and the Social Partner Influence Temporal Patterns. Infancy, 18(5),
729–​754.
Eley, T. C., & Plomin, R. (1997). Genetic analyses of emotionality. Current Opinion in Neurobiology, 7(2),
279–​284.
31

Conclusion 13

English, T., & John, O. P. (2013). Understanding the social effects of emotion regulation: The mediating role
of authenticity for individual differences in suppression. Emotion, 13(2), 314–​329.
Erber, R., Wegner, D. M., & Therriault, N. (1996). On being cool and collected: Mood regulation in
anticipation of social interaction. Journal of Personality and Social Psychology, 70(4), 757–​766.
Erber, R. & Erber, M. W. (2000). The self-​regulation of moods: Second thoughts on the importance of
happiness in everyday life. Psychological Inquiry, 11(3), 142–​148.
Essau, C. A., Lewinsohn, P. M., Olaya, B., & Seeley, J. R. (2014). Anxiety disorders in adolescents and
psychosocial outcomes at age 30. Journal of Affective Disorders, 163, 125–​132.
Eysenck, M. W. (2004). Applied cognitive psychology: Implications of cognitive psychology for clinical
psychology and psychotherapy. Journal of Clinical Psychology, 60, 393–​404.
Frijda, N. H. (1986). The emotions. Cambridge, UK: Cambridge University Press.
Garnefski, N., & Kraaij, V. (2006). Relationships between cognitive emotion regulation strategies and
depressive symptoms: A comparative study of five specific samples. Personality and Individual
Differences, 40(8), 1659–​1669.
Garnefski, N., & Kraaij, V. (2009). Cognitive Coping and Psychological Adjustment in Different Types of
Stressful Life Events. Individual Differences Research, 7(3).
Goldsmith, H. H., Buss, K. A., & Lemery, K. S. (1997). Toddler and childhood temperament: expanded
content, stronger genetic evidence, new evidence for the importance of environment. Developmental
Psychology, 33(6), 891.
Gray, P. (2011). The Decline of Play and the Rise of Psychopathology in Children and Adolescents.
American Journal of Play, 3(4), 443–​463.
Gross, J. J. (1998). The emerging field of emotion regulation: An integrative review. Review of General
Psychology, 2(3), 271–​299.
Gross, J. J. (1999). Emotion regulation: Past, present, future. Cognition and Emotion, 13(5), 551–​573.
Gross, J. J., & John, O. P. (2003). Individual differences in two emotion regulation processes: Implications
for affect, relationships, and wellbeing. Journal of Personality and Social Psychology, 85(2), 348–​362.
Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. In J. J. Gross,
Handbook of emotion regulation (pp. 3–​24): New York: Gilford Press.
Gross, J. J. (2013). Emotion regulation: taking stock and moving forward. Emotion, 13(3), 359.
Hannesdottir, D. K., & Ollendick, T. H. (2007). The role of emotion regulation in the treatment of child
anxiety disorders. Clinical Child and Family Psychology Review, 10, 275–​293.
Hariri, A. R., & Forbes, E. E. (2007). Genetics of emotion regulation. In J. J. Gross (Ed.), Handbook of
emotion regulation (pp. 110–​132), Guilford press, New York.
Herts, K. L., McLaughlin, K. A., & Hatzenbuehler, M. L. (2012). Emotion dysregulation as a mechanism
linking stress exposure to adolescent aggressive behavior. Journal of Abnormal Child Psychology, 40(7),
1111–​1122.
Hofmann, S. G., Heering, S., Sawyer, A. T., & Asnaani, A. (2009). How to handle anxiety: The effects of
reappraisal, acceptance, and suppression strategies on anxious arousal. Behaviour Research and Therapy,
47(5), 389–​394.
Hughes, E. K., Gullone, E., & Watson, S. D. (2011). Emotional functioning in children and adolescents
with elevated depressive symptoms. Journal of Psychopathology and Behavioral Assessment, 33(3),
335–​345.
Hum, K. M., Manassis, K., & Lewis, M. D. (2013). Neurophysiological Markers That Predict and Track
Treatment Outcomes in Childhood Anxiety. Journal of Abnormal Child Psychology, 41(8), 1243–​1255.
Isen, A. M., & Daubman, K. A. (1984). The influence of affect on categorization. Journal of personality and
social psychology, 47(6), 1206.
Isen, A. M., Johnson, M. M., Mertz, E., & Robinson, G. F. (1985). The influence of positive affect on the
unusualness of word associations. Journal of personality and social psychology, 48(6), 1413.
41

14 Emotion Regulation: An Introduction

Isen, A. M. (1987). Positive affect, cognitive processes, and social behavior. Advances in experimental social
psychology, 20, 203–​253.
Izard, C. E. (2013). Human emotions. Springer Science & Business Media, New York.
Jaffee, S. R., Moffitt, T. E., Caspi, A., Fombonne, E., Poulton, R., & Martin, J. (2002). Differences in early
childhood risk factors for juvenile-​onset and adult-​onset depression. Archives of General Psychiatry,
59(3), 215–​222.
Jaffe, M., Gullone, E., & Hughes, E. K. (2010). The roles of temperamental dispositions and perceived
parenting behaviours in the use of two emotion regulation strategies in late childhood. Journal of
Applied Developmental Psychology, 31(1), 47–​59.
John, O. P., & Gross, J. J. (2004). Healthy and unhealthy emotion regulation: Personality processes,
individual differences, and life span development. Journal of Personality, 72(6), 1301–​1333.
Kagan, J. (2003). Biology, context, and developmental inquiry. Annual Review of Psychology, 54(1),
1–​23.
Kashdan, T. B., Steger, M. F. (2006). Expanding the topography of social anxiety: An experience-​sampling
assessment of positive emotions, positive events, and emotion suppression. Psychological Science, 17(2),
120–​128.
Kennedy, S. E., Koeppe, R. A., Young, E. A., & Zubieta, J. K. (2006). Dysregulation of endogenous opioid
emotion regulation circuitry in major depression in women. Archives of General Psychiatry, 63(11),
1199–​1208.
Kessler, R. C., Chiu, W. T., Demler, O., & Walters, E. E. (2005). Prevalence, severity, and comorbidity of
12-​month DSM-​IV disorders in the National Comorbidity Survey Replication. Archives of General
Psychiatry, 62(6), 617.
Klimes-​Dougan, B., & Zeman, J. (2007). Introduction to the special issue of social development: Emotion
socialization in childhood and adolescence. Social Development, 16(2), 203–​209.
Kobak, R. R., Cole, H. E., Ferenz-​Gillies, R., Fleming, W. S., & Gamble, W. (1993). Attachment and
emotion regulation during mother-​teen problem solving: A control theory analysis. Child Development,
64(1), 231–​245.
Kopp, C. B., & Neufeld, S. J. (2003). Emotional development during infancy. In Handbook of Affective
Sciences, Oxford University Press, Oxford, 347–​374.
Koole, S. L. (2009). The psychology of emotion regulation: An integrative review. Cognition and Emotion,
23(1), 4–​41.
Kraaij, V., Pruymboom, E., & Garnefski, N. (2002). Cognitive coping and depressive symptoms in the
elderly: a longitudinal study. Aging & Mental Health, 6(3), 275–​281.
Kraemer, H. C., Stice, E., Kazdin, A., Offord, D., & Kupfer, D. (2014). How do risk factors work together?
Mediators, moderators, and independent, overlapping, and proxy risk factors. American Journal of
Psychiatry, 158(6), 848–​856.
Krueger, R. F., Caspi, A., Moffitt, T. E., White, J., & Stouthamer‐Loeber, M. (1996). Delay of Gratification,
Psychopathology, and Personality: Is Low Self‐Control Specific to Externalizing Problems? Journal of
Personality, 64(1), 107–​129.
Lantrip, C., Isquith, P/​. K., Koven, N. S., Welsh, K., & Roth, R. M. (2015). Executive Function and Emotion
Regulation Strategy Use in Adolescents. Applied Neuropsychology: Child, (ahead-​of-​print), 1–​6.
Larsen, R. J. (2000). Towards a science of mood regulation. Psychol/​ogical Enquiry, 11, 129–​141.
Larsen, J. K., Vermulst, A. A., Geenen, R., van Middendorp, H., English, T., Gross, J. J., … & Engels, R. C.
(2013). Emotion Regulation in Adolescence A Prospective Study of Expressive Suppression and
Depressive Symptoms. The Journal of Early Adolescence, 33(2), 184–​200.
Larson, R. W., Moneta, G., Richards, M. H., & Wilson, S. (2002). Continuity, stability, and change in daily
emotional experience across adolescence. Child Development, 73(4), 1151–​1165.
Lewinsohn, P. M., Hops, H., Roberts, R. E., Seeley, J. R., & Andrews, J. A. (1993). Adolescent
psychopathology: I. Prevalence and incidence of depression and other DSM-​III—​R disorders in high
school students. Journal of Abnormal Psychology, 102(1), 133.
51

Conclusion 15

Li-​Grining, C. P. (2007). Effortful control among low-​income preschoolers in three cities: Stability, change,
and individual differences. Developmental Psychology, 43(1), 208.
Malooly, A. M., Genet, J. J. & Siemer, M. (2013). Individual differences in reappraisal effectiveness: The role
of affective flexibility. Emotion, 13(2), 302.
Marcks, B. A., & Woods, D. W. (2005). A comparison of thought suppression to an acceptance-​based
technique in the management of personal intrusive thoughts: A controlled evaluation. Behaviour
Research and Therapy, 43(4), 433–​445.
McLeod, J. D., Uemura, R., & Rohrman, S. (2012). Adolescent mental health, behavior problems, and
academic achievement. Journal of Health and Social Behavior, 53(4), 482–​497.
Merikangas, K. R., Dierker, L. C., & Szatmari, P. (1998). Psychopathology among offspring of parents with
substance abuse and/​or anxiety disorders: a high-​risk study. Journal of Child Psychology and Psychiatry,
39(05), 711–​720.
Mezzacappa, E. (2004). Alerting, orienting, and executive attention: Developmental properties and
sociodemographic correlates in an epidemiological sample of young, urban children. Child
Development, 75(5), 1373–​1386.
Mikels, J. A., Maglio, S. J., Reed, A. E., & Kaplowitz, L. J. (2011). Should I go with my gut? Investigating the
benefits of emotion-​focused decision making. Emotion, 11(4), 743.
Mischel, W., & Ayduk, O. (2004). Willpower in a Cognitive-​Affective-​Processing System: The dynamics of
delay of gratification. In R. F. Baumeister & K. D. Vohs (Eds.). Handbook of self regulation: Research,
theory, and applications (pp. 99–​129). New York: Guilford Press.
Mischel, W., Ayduk, O., Berman, M. G., Casey, B. J., Gotlib, I. H., Jonides, J., … & Shoda, Y. (2010).
“Willpower” over the life span: decomposing self-​regulation. Social Cognitive and Affective
Neuroscience, nsq081.
Moore, S. A, Zoellner, L. A. & Mollenholt, N. (2008). Are expressive suppression and cognitive reappraisal
associated with stress-​related symptoms? Behaviour Research and Therapy, 46(9), 993–​1000.
Morris, A. S., Silk, J. S., Steinberg, L., Myers, S. S., & Robinson, L. R. (2007). The role of the family context
in the development of emotion regulation. Social Development, 16(2), 361–​388.
Murray, K. T., & Sines, J. O. (1996). Parsing the genetic and nongenetic variance in children’s depressive
behavior. Journal of Affective Disorders, 38(1), 23–​34.
Oatley, K., & Johnson-​Laird, P. N. (1987). Towards a cognitive theory of emotions. Cognition and Emotion,
1(1), 29–​50.
Ochsner, K. N., & Gross, J. J. (2007). The neural architecture of emotion regulation. Handbook of emotion
regulation, 1(1), 87–​109.
Ohira H., Nomura M., Ichikawa N., Isowa T., Iidaka T., Sato A., Fukuyama S., Nakajima T., Yamada
J. (2006). Association of neural and physiological responses during voluntary emotion suppression.
Neuroimage, 29(3), 721–​733.
Parrott, W. G. (1993). Beyond hedonism: Motives for inhibiting good moods and for maintaining bad
moods. In J. W. Pennebaker & D. M. Wegner (Eds.), Handbook of mental control (pp. 278–​305).
Englewood Cliffs, NJ: Prentice Hall.
Panfile, T. M., & Laible, D. J. (2012). Attachment security and child’s empathy: The mediating role of
emotion regulation. Merrill-​Palmer Quarterly, 58(1), 1–​21.
Phares, V. & Compas, B. E. (1992). The role of fathers in child and adolescent psychopathology: Make
room for daddy. Psychological Bulletin, 111(3), 387–​412.
Potenza, M. N., Wareham, J. D., Steinberg, M. A., Rugle, L., Cavallo, D. A., Krishnan-​Sarin, S., & Desai,
R. A. (2011). Correlates of at-​risk/​problem internet gambling in adolescents. Journal of the American
Academy of Child & Adolescent Psychiatry, 50(2), 150–​159.
Rapee, R. M. (1997). Potential role of childrearing practices in the development of anxiety and depression.
Clinical Psychology Review, 17(1), 47–​67.
Richards, J.M., Butler, E.A., & Gross, J.J. (2003). Emotion regulation in romantic relationships: The cognitive
consequences of concealing feelings. Journal of Social and Personal Relationships, 20(5), 599–​620.
61

16 Emotion Regulation: An Introduction

Rosenstein, D. S., & Horowitz, H. A. (1996). Adolescent attachment and psychopathology. Journal of
Consulting and Clinical Psychology, 64(2), 244.
Rothbart, M. K., Sheese, B. E., Rueda, M. R., & Posner, M. I. (2011). Developing mechanisms of self-​
regulation in early life. Emotion Review, 3(2), 207–​213.
Rothermund, K., Voss, A., & Wentura, D. (2008). Counter-​regulation in affective attentional biases: a basic
mechanism that warrants flexibility in emotion and motivation. Emotion, 8(1), 34.
Sapolsky, R. M. (2004). Why zebras don’t get ulcers. New York: Freeman.
Sapolsky, R. M. (2007). Stress, stress-​related disease, and emotional regulation. In J. J. Gross (Ed.),
Handbook of emotion regulation. New York: Guilford Press.
Séguin, J. R., & Zelazo, P. D. (2005). Executive function in early physical aggression New York, NY,
US: Guilford Press.
Shoda, Y., Mischel, W., & Peake, P. K. (1990). Predicting adolescent cognitive and self-​regulatory
competencies from preschool delay of gratification: Identifying diagnostic conditions. Developmental
Psychology, 26(6), 978.
Silk, J. S., Steinberg, L., & Morris, A. S. (2003). Adolescents’ emotion regulation in daily life: Links to
depressive symptoms and problem behavior. Child Development, 74(6), 1869–​1880.
Silvers, J. A., McRae, K., Gabrieli, J. D., Gross, J. J., Remy, K. A., & Ochsner, K. N. (2012). Age-​related
differences in emotional reactivity, regulation, and rejection sensitivity in adolescence. Emotion,
12(6), 1235.
Slee, N., Arensman, E., Garnefski, N., & Spinhoven, P. H. (2007). Cognitive behavioural therapy for
deliberate self-​harm. Crisis, 28(4), 175–​182.
Srivastava, S. Tamir, M., McGonigal, K. M., John, O. P. Gross, J. J. (2009). The social costs of emotional
suppression: A prospective study of the transition to college. Journal of Personality and Social
Psychology, 96(4), 883–​897.
Stegge, H., & Terwogt, M. M. (2007). Awareness and regulation of emotion in typical and atypical
development. Handbook of emotion regulation, 269–​286, Guilford Press, New York.
Suveg, C., & Zeman, J. (2004). Emotion regulation in children with anxiety disorders. Journal of Clinical
Child and Adolescent Psychology, 33(4), 750–​759.
Thapar, A., & McGuffin, P. (1997). Anxiety and depressive symptoms in childhood–​a genetic study of
comorbidity. Journal of Child Psychology and Psychiatry, 38(6), 651–​656.
Tice, D. M., Bratslavsky, E., & Baumeister, R. F. (2001). Emotional distress regulation takes precedence
over impulse control: If you feel bad, do it! Journal of Personality and Social Psychology, 80(1), 53–​67.
Tomkins, S. S. (1962). Affect imagery consciousness: Volume I: The positive affects (Vol. 1). Springer
publishing company, New York.
Tortella-​Feliu, M., Balle, M., & Sesé, A. (2010). Relationships between negative affectivity, emotion
regulation, anxiety, and depressive symptoms in adolescents as examined through structural equation
modeling. Journal of Anxiety Disorders, 24(7), 686–​693.
Tottenham, N., Hare, T. A., & Casey, B. J. (2011). Behavioral assessment of emotion discrimination,
emotion regulation, and cognitive control in childhood, adolescence, and adulthood. Frontiers in
Psychology, 2, 39.
Ursache, A., Blair, C., Stifter, C., & Voegtline, K. (2013). Emotional reactivity and regulation in infancy
interact to predict executive functioning in early childhood. Developmental Psychology, 49(1), 127.
Van Reekum, C. M., & Scherer, K. R. (1997). Levels of processing in emotion-​antecedent appraisal.
In G. Matthews (Ed.), Cognitive science perspectives on personality and emotion (pp. 259–​300).
Amsterdam: Elsevier Science.
Westen, D. (1994). Toward an integrative model of affect regulation: Applications to social-​psychological
research. Journal of Personality, 62(4), 641–​667.
71

Conclusion 17

Wiebe, S. A., Sheffield, T. D., Nelson, J. M., Clark, C. A. C., Chevalier, N., & Espy, K. A. (2011). The
structure of executive function in 3-​year-​old children. Journal of Experimental Child Psychology, 108(3),
436–​452. “Executive Function” Special Issue.
Wilding, C., & Milne, A. (2010). Cognitive behavior therapy. McGraw-​Hill, London.
Wilens, T. E., Martelon, M., Anderson, J. P., Shelley-​Abrahamson, R., & Biederman, J. (2013). Difficulties
in emotional regulation and substance use disorders: A controlled family study of bipolar adolescents.
Drug and alcohol dependence, 132(1), 114–​121.
Willcutt, E. G., Doyle, A. E., Nigg, J. T., Faraone, S. V., & Pennington, B. F. (2005). Validity of the executive
function theory of attention-​deficit/​hyperactivity disorder: a meta-​analytic review. Biological Psychiatry,
57(11), 1336–​1346.
Yap, M. B., Allen, N. B., & Ladouceur, C. D. (2008). Maternal socialization of positive affect: The impact
of invalidation on adolescent emotion regulation and depressive symptomatology. Child Development,
79(5), 1415–​1431.
Zelazo, P. D., Carlson, S. M., & Kesek, A. (2008). The development of executive function in childhood, MIT
Press, Cambridge.
Zeman, J., Shipman, K., & Suveg, C. (2002). Anger and sadness regulation: Predictions to internalizing and
externalizing symptoms in children. Journal of Clinical Child and Adolescent Psychology, 31(3), 393–​398.
Zeman, J., Cassano, M., Perry-​Parrish, C., & Stegall, S. (2006). Emotion regulation in children and
adolescents. Journal of Developmental & Behavioral Pediatrics, 27(2), 155–​168.
Zimmer-​Gembeck, M. J., & Skinner, E. A. (2011). Review: The development of coping across childhood
and adolescence: An integrative review and critique of research. International Journal of Behavioral
Development, 35(1), 1–​17.
81

Chapter 2

The Relation of Self-╉Regulation


to Children’s Externalizing and
Internalizing Problems
Nancy Eisenberg, Maciel M. Hernández,
& Tracy L. Spinrad

Self-regulation in children
The lack of emotional or behavioral regulation is often viewed as a component of psychopathology;
indeed, some types of problems are defined in part by the lack of self-╉regulation (e.g., some exter-
nalizing problems and depression/╉anxiety; American Psychiatric Association, 2013). However,
empirically, agreement has not been reached regarding what capacities are included in the con-
struct of “regulation” or “self-╉regulation” (e.g., see Eisenberg, Hofer, Sulik, & Spinrad, 2014). In
this chapter, we consider some useful conceptual distinctions in the domain of control, briefly
present heuristic hypotheses regarding the relations between regulation-╉relevant constructs and
externalizing and internalizing behaviors, and review representative empirical findings.

Conceptual issues
Eisenberg, Hofer, Sulik, and Spinrad (2014) defined emotion-╉related self-╉regulation as a process
used to “manage and change whether, when and how (e.g., how intensely) one experiences emo-
tions and emotional-╉related motivational and physiological states, as well as how emotions are
expressed behaviorally. Thus, it includes processes used to change one’s own emotional state, to
prevent or initiate emotion responding (e.g., by selecting or changing situations), to modify the
significance of an event for the self, and to modulate the behavioral expression of emotion (e.g.,
through verbal or nonverbal cues)” (p. 157). The term “emotion-╉related” self-╉regulation is used
because many of the processes/╉abilities that are part of emotion-╉related regulation can be involved
in regulating multiple aspects of functioning, which include not only the expression and experi-
ence of emotion, but also aspects of cognition, attention, and behavior that do not involve (or
secondarily or minimally involve) modulating the expression and experience of emotion. Thus,
emotion regulation, defined by Gross (2014) as “shaping which emotions one has, when one has
them, and how one experiences or expresses these emotions” (p. 6), can be viewed as occurring
when emotion-╉related self-╉regulatory skills are applied directly to the experience or expression of
emotion. Of course, external influences such as parents or providers of social support can contrib-
ute to the modulation of emotion and its expression, but for clarity, we have argued it is clearer to
differentiate such external controlling factors from self-╉regulation (Eisenberg & Spinrad, 2004).
Regardless of the specific terminology, we have suggested it is useful to distinguish between
self-╉regulatory processes that can readily become volitional when required to adapt or achieve
91

Conceptual issues 19

a goal and those “regulating” or controlling processes that affect emotion and behavior but are
harder to control volitionally. As has been discussed by researchers from multiple subdisciplines
of psychology (see Carver, 2005), many non-╉volitional processes have important modulating
(in a sense, regulating) effects on attention, behavior, cognition, and physiological responding.
Eisenberg et  al. (2014) used the term “self-╉regulation” to refer to “potentially volitional, self-╉
regulatory processes.”
The interconnectedness of these constructs makes it difficult to differentiate emotion from its
self-╉regulation; someone who expresses little emotion in a potentially evocative context may be
regulating his or her emotion or simply may not be responding emotionally. Thus, it is beneficial
to focus on the processes used to manage emotion, cognition, and associated behavior, rather than
to measure the amount of emotion experienced or expressed. Consequently, when studying the
regulation of emotion, rather than look for possible self-╉regulation of emotional displays, there
are advantages to focusing on and measuring aspects of executive functioning that contribute
to self-╉regulation (e.g., executive attention) and dispositional differences in self-╉regulation that
employ the skills used for the regulation of emotion and related cognitive, physiological, and
behavioral responses.

Effortful self-╉regulatory processes


The temperamental underpinnings of self-╉regulation are known as effortful control, defined by
Rothbart and Bates (2006, p. 129) as “the efficiency of executive attention, including the ability
to inhibit a dominant response and/╉or to activate a subdominant response, to plan, and to detect
errors.” Effortful control includes the capacities to effortfully (i.e., willfully) deploy attention (e.g.,
to focus and shift attention as needed) and to willfully inhibit or activate behavior, especially when
doing so is a non-╉preferred (subdominant) response (i.e., inhibitory control and activational con-
trol, respectively). Effortful control is often measured with parents’ and teachers’ ratings and with
a variety of behavioral measures, including those assessing the ability to delay gratification (e.g.,
wait until a bell rings to pick up a snack), inhibiting and activating similar behaviors based on
different commands (games like Simon says), and executive-╉functioning skills involving the man-
agement of attention and inhibiting behavior.
Although effortful control is believed to be the temperamental core of self-╉regulation, emerg-
ing self-╉regulation can be conceptualized broadly, including more than the basic executive func-
tioning-╉related skills involved in effortful control (e.g., the abilities to effortfully shift and focus
attention and to effortfully activate and inhibit behavior as needed for adaptation). For instance,
complex cognitive strategies (e.g., cognitive restructuring), seeking social support, persistence,
and motivational components, such as the desire to act in ways consistent with norms or expecta-
tions, may also be viewed as aspects of self-╉regulation.
Eisenberg and colleagues (Eisenberg et al., 2014) have argued that although effortful control
is defined as effortful or willful, individuals may not always be aware that they are modulating
emotion, attention, or behavior. Some aspects of effortful control undoubtedly become automatic
and executed without substantial conscious awareness in contexts with relevant triggering cues
(Mischel & Ayduk, 2011); however, an effortful control-╉related process can shift into a volitional
and more conscious mode of functioning when it is adaptive to move from an automatic to effort-
ful status (analogous to automobile driving becoming much more effortful and less automatic
when on ice).
Self-╉regulatory abilities are not necessarily inherently good or bad in terms of their outcomes.
People can use self-╉regulation to achieve goals that are maladaptive or adaptive, and whether a
consequence is positive or negative (socially, morally, or in normative terms) can differ in the
short-╉term versus long-╉term. Nonetheless, effortful self-╉regulatory processes are probably more
02

20 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

likely than some less volitional aspects of control (see below) to result in adaptive outcomes, or
at least in desired goals (regardless of whether or not they are actually socially or functionally
adaptive) because they can be flexibly applied when needed to accommodate contextual demands
rather than being applied in a rigid manner.

Reactive control processes


As previously noted, there are many processes that are controlling or regulating in the sense that
they modulate another system but which are also relatively non-╉volitional, nearly always auto-
matic, and less flexible than effortful control. Because both the volitional nature and the flexibility
of “regulatory” processes likely affect their effectiveness and outcomes, it can be argued that it is
useful to differentiate volitional self-╉regulation from less volitional processes involved in control
of emotion, physiological responding, and behavior.
Rothbart and Bates (2006) differentiated between temperamental regulation and reactivity.
They defined reactivity as “responsiveness to change in the external and internal environment”
(p. 100), including emotional reactivity and action tendencies. We use the term “reactive control”
to refer to the action tendencies, rather than the emotion, that are part of reactivity. Rothbart
and Bates (2006) defined self-╉regulation as “processes such as effortful control and orienting that
function to modulate reactivity” (p. 100). Although Rothbart and colleagues (e.g., Derryberry &
Rothbart, 1997) view emotional reactivity and behavioral reactivity as strongly linked processes,
it seems likely that reactive behaviors sometimes occur without being evoked by emotion because
they are part of a child’s characteristic way of responding in particular contexts.
In our view, overcontrolled and undercontrolled behaviors reflect two types of reactive control at
the extremes of a heuristic continuum. In regard to overcontrol, children’s inhibited behavior in cer-
tain contexts often is relatively involuntary and difficult to modulate willfully (Eisenberg & Morris,
2002). For example, children labeled as “behaviorally inhibited” by Kagan (1998) tend to be wary
and overly constrained in novel and/╉or stressful situations and appear to have difficulty willfully
modulating their inhibition. On the other extreme, undercontrol—╉the impulse to approach people
or inanimate objects in the environment (often quickly) without much thought—╉often appears to be
relatively involuntary. Such behavior is clearly reflected in at least some types of impulsive behavior.
Undercontrol and overcontrol map onto Gray’s (Pickering & Gray, 1999) behavioral activation (BAS;
which involves sensitivity to cues of reward or cessation of punishment) and behavioral inhibition
(BIS; activated in situations involving novelty and stimuli signaling punishment or frustrative non-
reward) systems. Both these systems are believed to be centered in subcortical regions of the brain.
In contrast, effortful control appears to be centered primarily in cortical regions of the brain such as
the anterior cingulate gyrus and prefrontal areas for the “cool” executive functioning components
(Cohen & Lieberman, 2010; Rothbart & Bates, 2006) and perhaps the ventromedial prefrontal cor-
tex and orbitofrontal cortex for “hot” tasks involving rewards/╉delays (Happaney, Zelazo, & Stuss,
2004). However, there undoubtedly are many connections between these cortical areas and the sub-
cortical systems involved in reactive control and emotion (Goldsmith, Pollak, & Davidson, 2008).
The distinction between volitional and non-╉volitional regulatory or controlling processes
has been discussed in diverse literatures, including work on coping (Compas, Connor-╉Smith,
Saltzman, Thomsen, & Wadsworth, 2001), and in the personality, clinical, social psychological,
and cognitive literatures (see Carver, 2005, for a review of similar perspectives, including dual
processing models). Supporting this distinction, Eisenberg and her colleagues (Eisenberg et al.,
2004, 2013; Valiente et al., 2003), with modeling procedures, found that they could differentiate
empirically between effortful and reactive control when assessed with a variety of adult-╉report
and/╉or behavioral measures of effortful and reactive control from 30 months to pre/╉early adoles-
cence. Thus, prediction of maladjustment is likely to be enhanced by considering both effortful
and reactive processes used in regulation/╉control.
12

RELATIONS OF REGULATORY PROCESSES TO MALADJUSTMENT: A FRAMEWORK 21

Relations of regulatory processes to maladjustment:


A framework
In 1992, Eisenberg and Fabes developed a heuristic model to guide predictions regarding the
relations of regulation (and emotional intensity) to an array of developmental outcomes, includ-
ing some problem behaviors. The regulation aspect of this model was updated by Eisenberg and
Morris (2002). Briefly, they hypothesized that externalizing problems tend to be related to low
levels of effortful control/​self-​regulation, including attentional, activational, and inhibitory con-
trol, and high levels of reactive undercontrol (impulsivity). In contrast, internalizing problems
were predicted to be associated with high reactive overcontrol, low attentional control (used to
modulate emotions associated with internalizing problems), and low activational control in some
contexts (e.g., when used to counter tendencies to withdraw socially); internalizing problems were
not expected to be associated with sizable deficits in effortful inhibitory control. Optimally regu-
lated children—​those without internalizing or externalizing problems—​were hypothesized to be
high in all types of effortful control (attentional, activational, inhibitory) and to be neither overly
controlled nor highly undercontrolled.
Nigg (2006) argued that there are at least two temperamental pathways to conduct disorders.
One is based on a low fear response and low affiliation (resulting in low empathy and sympa-
thy), often accompanied by high impulsivity (low reactive control), and sometimes involving psy-
chopathy. Low physiological arousal to potential punishment in these individuals makes efforts
to socialize the child difficult and often unsuccessful. Nigg’s second pathway leading to impulsive
conduct problems involves extreme levels of approach (e.g., to incentives), especially if combined
with average to high negative emotionality (also see Eisenberg & Fabes, 1992), and average or
low levels of reactive overcontrol and effortful control. Similarly, Nigg (2006) further suggested
that attention deficit hyperactivity disorder has at least two temperamental pathways, one involv-
ing very low effortful control (often co-​occurring with high emotionality) and another involving
strong approach tendencies.
It is important to note that investigators vary somewhat in what they conceptualize as impul-
sivity and this might affect theoretical predictions. For example, Johnson, Carver, and Joormann
(2013) measured impulsive responses to emotions versus non-​emotion-​relevant impulsivity.
They argued that people with sensitive incentive-​approach temperament (what many others
call impulsivity) and high reactivity to emotions may be overwhelmed by strong desires and
hence prone to sensation seeking or antisocial impulses. They further suggested that those with
low approach temperament along with high reactivity to emotions might be overly affected by
sadness and fatigue, resulting in lethargy and inaction that generally characterizes depression.
In contrast, those with sensitive threat-​avoidance temperament and high reactivity to emotions
may be especially prone to anxiety. This focus on control over emotion reactions differs from
that of Eisenberg and colleagues (e.g., 2002), who focused more on the approach/​incentive
behavioral component of impulsivity, which led to different predictions. We would argue that
Johnson et al.’s (2013) measure of emotion-​relevant impulsivity reflects both a lack of effortful
control of emotions and impulsivity when emotionally aroused.
Frick and Morris (2004) proposed that although deficits in self-​regulation contribute to
reactive, emotionally driven conduct problems (e.g., reactive aggression), they are unlikely
to be involved in covert externalizing problems (e.g., stealing) and proactive externalizing
problems (e.g., unprovoked, unemotional aggression that is used for personal gain or to influ-
ence and coerce others). They also suggested that children with reactive, emotionally driven
externalizing problems are prone to negative emotion and have difficulties regulating emotion
and inhibiting behavior when emotionally aroused. They further argued that these children’s
emotion dysregulation can impair the development and use of sociocognitive skills involved
2

22 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

in information processing and undermine the quality of socializing interactions. Similar to


Nigg (2006), Frick and Morris (2004) hypothesized that children prone to proactive aggres-
sion, including those with psychopathic (callous-╉unemotional) traits, are low in inhibition due
to fear, which undermines the development of the conscience, but are not consistently low in
self-╉regulation.
The role of self-╉regulation in internalizing problem behaviors also likely varies with the type of
problem. For example, social withdrawal is often a component of internalizing symptoms; how-
ever, socially withdrawn behavior can stem from social anxiety and/╉or fearfulness, social rejec-
tion (perhaps due to lack of self-╉regulation), and the mere preference of being alone (Coplan &
Armer, 2007). Attentional control may be particularly important for modulating the experience
of social anxiety and fearfulness, and effortful activational control may help fearful/╉anxious chil-
dren overcome their withdrawn behavior. Lack of effortful inhibitory control may be especially
related to social withdrawal due to peer rejection for inappropriate behavior (because of its role
in externalizing behaviors that elicit peer rejection). In contrast, various components of effortful
control may be irrelevant for social withdrawal due to the preference to be alone (which may
not reflect an internalizing problem). Moreover, the attentional component of effortful control
may be more highly related to depressive and anxious symptoms than is inhibitory or activa-
tional control. Unfortunately, there are few studies in which various aspects of effortful con-
trol/╉self-╉regulation have been examined as separate predictors of psychological symptoms, both
broadband (externalizing or internalizing more generally) or specific types of internalizing or
externalizing problems.

Externalizing problems
Externalizing behaviors—╉“behaviors that violate the rights of others (e.g., aggression, destruc-
tion of property) and/╉or that bring the individual into significant conflict with societal
norms or authority figures” (American Psychiatric Association, 2013)—╉are associated with
adjustment problems in the academic, social, and emotional domains across the lifespan.
Externalizing problem behaviors, unified by a common theme of outward behaviors, con-
stitute a variety of behaviors:  Aggression, delinquency, hyperactivity, defiance (American
Psychiatric Association, 2013), and subtypes of reactive aggression (e.g., emotionally-╉driven
aggression), covert externalizing (e.g., stealing, lying), proactive externalizing (e.g., aggression
for self-╉gain), and callous-╉unemotional trait conduct problems (e.g., unprovoked and unemo-
tional aggression; Frick, Ray, Thornton, & Kahn, 2014). Growing empirical evidence supports
the premise that emotion regulation is negatively associated—╉and impulsivity is positively
associated—╉with externalizing problem behaviors across development (Eisenberg, Spinrad, &
Eggum, 2010).

Self-regulation and externalizing problems


Diverse measures of effortful emotion-╉related self-╉regulation have been negatively associated with
externalizing behaviors, especially emotionally-╉driven reactive externalizing (Eisenberg, Spinrad,
& Eggum, 2010; e.g., Valiente et al., 2006). For example, attention refocusing (i.e., shifting atten-
tion from an emotion-╉eliciting stimulus) during a disappointment task was negatively associated
with externalizing two years later, particularly for children who expressed higher anger than other
children (assessed in preschool/╉kindergarten; Morris, Silk, Steinberg, Terranova, & Kithakye,
2010). Also, performance on executive control (i.e., inhibition and attention) and delay tasks at
36 to 40 months has predicted lower hyperactivity and externalizing problems at 63 to 67 months
(Lengua et  al., 2015). Similarly, from ages four to seven, emotion regulation (assessed with a
32

EXTERNALIZING PROBLEMS 23

combined measure of emotionality and regulation) was consistently negatively associated with
externalizing across time, but not vice versa (Blandon, Calkins, Grimm, Keane, & O’Brien, 2010).
Associations between effortful control and externalizing behaviors have also held across longer
spans of time and/​or for older children (e.g., Eisenberg, Zhou, et al., 2005; Lengua, 2008). Effortful
control (i.e., attention control, inhibitory control, low activity levels) at four-​and-​a-​half years old
was associated with lower externalizing and risk-​taking behaviors at 15 years old (Honomichl &
Donnellan, 2011). Similarly, Belsky, Pasco Fearon, and Bell (2007) found that at 54 months, first
grade, and fifth grade, attention problems (measured with a continuous performance test) were
positively associated with externalizing behaviors even when controlling for their prior levels.
Wang, Brinkworth, and Eccles (2013) observed that misconduct behaviors decreased substan-
tially from 13 to 18 years of age for adolescents with higher effortful control (i.e., attention shift-
ing, activation control) at age 13. Relatedly, among nine-​and-​a-​half-​year-​olds, individual growth
in effortful control (but not growth in impulsivity) predicted lower externalizing problems three
years later (King, Lengua, & Monahan, 2013).
In a recent study, among 36-​month-​olds assessed four times until 90 months of age in a cross-​
lagged panel design, executive functioning (assessed with a set of behavioral tasks closely related
to effortful control) consistently predicted lower externalizing behaviors (Sulik et al., 2015). In
one instance, however, externalizing behaviors also predicted lower executive functioning from
48 to 60 months of age. In contrast, Eisenberg, Spinrad, Eggum, et al. (2010) found that external-
izing (and internalizing) at 30 months old significantly and negatively predicted effortful control
at 42 months old (measured with a delay task and parent/​caregiver reports). However, effortful
control (although correlated with) did not significantly predict externalizing across time while
controlling for the stability of all measures (also see Eisenberg, Taylor, Widaman, & Spinrad, 2015,
with the same sample at 30 to 54 months). Thus, although executive functioning/​effortful control
frequently predicts low levels of later externalizing problems, there may be age-​or context-​depen-
dent periods (e.g., in the transition to formal schooling) when externalizing behaviors and execu-
tive functioning abilities form a reciprocal process or, perhaps, when externalizing problems have
a stronger effect on self-​regulation.
Indeed, although effortful control frequently has been associated with subsequent externalizing
problems, the relation of self-​regulatory abilities to concurrent and later externalizing problems
varies across studies and samples, especially when controlling for initial levels (e.g., Eisenberg,
Spinrad, Eggum et al., 2010; Spinrad et al., 2012). Among children of Chinese immigrant parents
in the United States, effortful control (i.e., parent-​and teacher-​reported inhibitory control, atten-
tion focusing) measured among first and second graders was positively associated with social
competence but not with externalizing behaviors in fifth and sixth grade (Zhou, Main, & Wang,
2010). Also, among Dutch children, effortful control (i.e., parent-​reported inhibitory control and
attention focusing) and delayed gratification among preschoolers (36-​month-​olds) were nega-
tively associated with concurrent hyperactivity and conduct problems but did not significantly
predict later measures of maladjustment in kindergarten when controlling for initial levels at 36
months old (Gusdorf, Karreman, van Aken, Dekovic, & van Tuijl, 2011). Similarly, Lengua (2003)
found that difficulty in delay of gratification was positively associated with externalizing symp-
toms concurrently, but not one year later, among third through fifth graders; however, inhibitory
control did predict lower levels of later externalizing symptoms (Lengua, 2003). These results echo
findings from a meta-​analysis showing that inhibitory control, compared to executive function-
ing, was more strongly associated with externalizing behaviors among preschoolers (Schoemaker,
Mulder, Deković, & Matthys, 2013).
In examining effortful control, some researchers have also distinguished “hot” and “cool”
aspects based on the emotional and cognitive demands of the different tasks used to assess
42

24 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

effortful control (“cool” tasks often are executive functioning tasks). Di Norcia, Pecora, Bombi,
Baumgartner, and Laghi (2014) found that hot (i.e., delayed gratification), but not cool (e.g., slow
down, reverse categorization) effortful control, was negatively associated with concurrent aggres-
sion and anger among Italian preschoolers. Similarly, Kim, Nordling, Yoon, Boldt, and Kochanska
(2013) found that only “hot” effortful control (i.e., delayed gratification) was negatively associ-
ated with behavioral problems at 67–╉100 months of age. Effortful control measured with “cool”
tasks (e.g., day/╉night, motor inhibition) did not significantly predict behavioral problems unless
estimated together with delayed gratification (Kim et  al., 2013). Thus, although most research
has examined effortful control as one construct given conceptual concordance and measurement
properties (Eisenberg et  al., 2013), continued examination of the aspects of different effortful
control measures is warranted given that in some studies not all components of effortful self-╉
regulation have significantly predicted maladjustment (Di Norcia et al., 2014; Kim et al., 2013).

Callous-unemotional, covert, and proactive externalizing


Most of the studies cited thus far have evaluated models predicting reactive externalizing (aggres-
sive emotional responses to blocked goal/╉provocation; Frick et  al., 2014)  or undifferentiated
externalizing problems. Less is known about the extent to which regulatory processes are associ-
ated with covert (i.e., secretive externalizing such as lying, cheating, stealing) and proactive (i.e.,
aggression for self-╉gain) externalizing across development. Verbal ability may differentially pre-
dict reactive and proactive aggressive tendencies, suggesting a social-╉cognitive pathway; Arsenio,
Adams, and Gold (2009) found that among adolescents, verbal ability was negatively associated
with reactive aggression and positively associated with proactive aggression. Also, in that study,
attention problems (an indicator of low effortful control) were more strongly and positively asso-
ciated with reactive than with proactive aggressive tendencies. In another study, impulsivity and
inattention measures were also positively associated with overt antisocial behaviors and positively
associated with covert behaviors particularly for children with higher verbal ability (McEachern
& Snyder, 2012). Relatedly, White, Jarrett, and Ollendick (2012) found that behavioral regulation
was associated with reactive but not proactive aggression among children and adolescents. These
studies suggest that proactive and reactive aggression have different correlates and that emotion
regulation may be most associated with reactive externalizing difficulties. Furthermore, these
studies imply that impulsivity and verbal ability are different risk factors for covert versus overt
externalizing behaviors in childhood.
However, some findings are inconsistent with those just reviewed. Evidence regarding the rela-
tion of reactive and proactive aggression to self-╉regulation is not very clear. Marsee and Frick
(2007) reported that reactive aggression was uniquely associated with poor emotion regula-
tion when controlling for proactive aggression but not in the zero-╉order correlation. Moreover,
other researchers (de Castro, Merk, Koops, Veerman, & Bosch, 2005; Xu, Farver, & Zhang, 2009)
found that reactive and proactive aggression were both inversely associated with effortful con-
trol. Perhaps different domains of self-╉regulation (e.g., regulation of emotional experience versus
behavior) are associated with reactive and proactive aggression. Alternatively, deficits in “cool”
executive functioning skills, which could undermine integration and decision making, may con-
tribute to proactive aggression, whereas self-╉regulation of more emotionally tinged behavior may
be particularly related to problem behaviors that are impulsive and appear to be emotionally
driven. Consistent with that view, the tendency to experience negative emotions (e.g., anger) has
been more consistently associated with reactive than proactive aggression (e.g., Hubbard et al.,
2002). In any case, the research suggests that attention to the type of externalizing problem may
be important when examining associations with self-╉regulation (and emotionality).
52

Externalizing problems 25

A type of externalizing problem that is generally proactive is conduct problems with callous
unemotional traits (i.e., unprovoked/╉unemotional aggression, lack of guilt and concern for oth-
ers; Frick et al., 2014). Temperamental fearlessness (assessed with parent-╉report measures) has
been positively associated with conduct problems or callous-╉unemotional traits among older
children and adolescents (Barker, Oliver, Viding, Salekin, & Maughan, 2011; Lengua, 2003); how-
ever, compared to children without conduct problems, first graders with conduct problems and
callous-╉unemotional behaviors exhibited more intense fear reactions (during a mask task) and
higher baseline cortisol levels at the age of two (Mills-╉Koonce et al., 2015). Perhaps intense fear
observed in toddlers, a sign of emotion dysregulation and heightened sensitivity, is a precursor to
or marker of conduct problems with callous-╉unemotional traits, and fearlessness develops later.
Alternatively, relations of fearfulness (and its regulation) may vary as a function of method of
assessing fear (e.g., observations versus parents’ reports; Mills-╉Koonce et al., 2015).

Impulsivity and externalizing problems


Impulsivity—╉characterized by unplanned and sudden reactions without concern for negative
consequences—╉also confers a risk for deviant behaviors, is related to temperamental surgency
(Rothbart & Bates, 2006) and, as previously argued, has been viewed as a distinct element of emo-
tion dysregulation separate from effortful control (Eisenberg et al., 2013; Spinrad et al., 2012).
Marmorstein (2013) identified five types of impulsive behavior: Sensation seeking, lack of plan-
ning, lack of perseverance, negative urgency, and positive urgency; however, it could be argued
that some of the forementioned actually tap effortful control as much or more than impulsivity.
Impulsivity is concurrently and prospectively associated with externalizing in childhood and
adolescence. Youths who report more impulsive behaviors (e.g., act without thinking, need a lot
of self-╉control to stay out of trouble) also report more risky behavior, including externalizing
problems (e.g., Johnson et al., 2013; Romer et al., 2011). Among middle school students, negative
urgency (i.e., impulsive behavior when distressed or in a negative mood) was positively associ-
ated with attention deficit/╉hyperactivity disorder and conduct disorder symptoms (Marmorstein,
2013). Also, impulsivity predicted rank-╉order change in adolescents’ alcohol use, conduct prob-
lems, and hyperactive/╉inattention behaviors, and sensation seeking predicted higher alcohol
problems one year later (Wang, Chassin, Geiser, & Lemery-╉Chalfant, 2016). However, Lengua
(2003) found that mother-╉reported impulsivity was positively correlated with, but did not signifi-
cantly predict, children’s externalizing symptoms among third through fifth graders one year later.
Similarly, impulsivity may not be uniquely associated with externalizing problems in older chil-
dren once the predictive effects of effortful control are taken into account (Eisenberg, Spinrad et
al., 2004; Valiente et al., 2006; Wang, Chassin, Eisenberg, & Spinrad, 2015). Such findings suggest
that impulsivity may not always be uniquely associated with externalizing and that prospective
associations vary with age and/╉or context.
Other work suggests that different aspects of impulsivity predict externalizing problems,
although this work is difficult to interpret because some measures of impulsivity may actually
tap effortful control. Based on a cross-╉sectional study, Settles et al. (2012) found that among fifth
grade students, negative urgency (but not lack of planning) positively predicted alcohol use and
smoking status (behaviors that often are considered to be externalizing problems). Furthermore,
lack of perseverance (which might assess low effortful control) positively predicted smoking sta-
tus among boys, but not among girls (Settles et al., 2012). Associations between components of
impulsivity and types of externalizing may vary by age. For instance, lack of planning (which
might reflect low effortful control or high impulsivity) and negative urgency both positively pre-
dicted alcohol and drug use among college students, whereas lack of planning positively predicted
62

26 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

delinquent behavior, and negative urgency positively predicted aggression (Settles et al., 2012).
Continued examination of the predictive validity of impulsivity subscales and combined mea-
sures of emotion self-╉regulation will help inform various pathways to externalizing across devel-
opment (e.g., Eisenberg, Spinrad et al., 2004). For example, Morales, Beekman, Blandon, Stifter,
and Buss (2015) found that exuberance (i.e., combined measures of impulsivity and activity) was
positively associated with later externalizing in kindergarten, particularly for children with high
physiological dysregulation (e.g., low baseline respiratory sinus arrhythmia [RSA], higher RSA
suppression during an emotion stimulus).

The role of culture


Researchers have found that effortful control is negatively associated with externalizing difficulties
in children from Indonesia (Eisenberg, Liew, & Pidada, 2004), China (Zhou, Eisenberg, Wang, &
Reiser, 2004; Zhou, Lengua, & Wang, 2009), Portugal (Conceição & Carvalho, 2013), and Latino
and African American backgrounds in the United States (Loukas & Roalson, 2006). However,
a limited number of studies have found concurrent but not longitudinal empirical support for
the association between effortful control and externalizing difficulties across children of Chinese
immigrant parents (Zhou et al., 2010) and Dutch children (Gusdorf et al., 2011). What cultural
processes inform the association between self-╉regulation and externalizing problems across time
are less clear.

Research and praxis: Targeting regulation


Given the empirical support that emotion self-╉regulation predicts lower externalizing problems,
interventions targeting self-╉regulation to prevent the development of externalizing behaviors have
been designed and increasingly implemented (Eisenberg, Spinrad, & Eggum, 2010). Research of
this sort can be a stringent test of causal relations between self-╉regulatory skills and externalizing
problems, although often more than self-╉regulation is targeted by these interventions. Moreover,
the results from school-╉based interventions intended to improve children’s behavioral regulation
and reduce subsequent problem behaviors are promising, although some show mixed results,
sometimes as a function of participant background characteristics (Morris et al., 2014).
Among preschoolers from low-╉income backgrounds, children participating in The Incredible
Years intervention (which includes curriculum on regulation and emotion knowledge) had higher
emotion knowledge post-╉intervention, but they showed similar executive functioning and prob-
lem behaviors compared to children in the control group (Morris et al., 2014). However, among
children who had higher levels of behavioral problems at baseline, fewer problems were reported
after the intervention compared to children in the control condition. Children participating in
Tools of the Mind, a similar intervention, showed increased levels of executive functioning, par-
ticularly for those from high-╉poverty schools (Blair & Raver, 2014).
Programs that integrate parent-╉and child-╉focused components in the implementation of
interventions are also likely to increase the prospect of favorable outcomes. Parents likely help
children develop self-╉regulation skills through their parenting strategies and modeling of self-╉
regulation (Belsky et  al., 2007). Moreover, some researchers have found that self-╉regulation
mediates the association between environmental factors (e.g., parenting, socioeconomic sta-
tus) and externalizing behaviors (e.g., Lengua et al., 2015). Thus, targeting parenting may have
downstream effects on problem behaviors. For example, the Parents and Children Making
Connections—╉Highlighting Attention (PCMC-╉A) program includes parenting-╉focused (i.e.,
stress management, parent discipline, language use, and child attention exercises) and child-╉
specific (i.e., modifying attention regulation and emotion states curriculum for children)
72

Internalizing problems 27

training components for low-╉income parents and their preschool children attending Head Start
(Neville et al., 2013). Compared to no-╉preschool/╉Head Start-╉only control participants, PCMC-╉
A children showed increased attention regulation (measured with event-╉related potentials
[ERP] during attention tasks) and decreased problem behaviors after the eight week interven-
tion. Parents in PCMC-╉A also showed decreased parenting stress. Physiological measures, such
as those assessed with ERP, provide additional support for the regulatory mechanisms modi-
fied by intervention. Similarly, among homeless youth participating in an intervention with
emotional self-╉regulation and parenting guidance components, intervention effects on reduced
conduct problems were mediated by improvements in children’s executive functioning (Piehler
et al., 2014).
Some interventions have addressed specific needs of children. For example, Fast Track
Promoting Alternative Thinking Strategies (PATHS) integrated the classroom-╉ based socio-╉
emotional curriculum from Fast Track and additional intervention components (i.e., parent-
ing support classes, home visits) for children identified by teachers at higher risk for aggression
(Conduct Problems Prevention Research Group, 2010). Compared to controls, participants in
Fast Track PATHS showed lower aggression and higher prosocial behavior from first to third
grade, with some effects strongest for boys and children originally high in aggression (Conduct
Problems Prevention Research Group, 2010). Thus, programs addressing socio-╉emotional learn-
ing through curriculum in the classroom and home (especially for children identified at higher
risk for behavior problems) show promising results.
Across development, children increase in their regulatory capabilities but also encounter dif-
ferent sets of challenges that may require varied skills and support; for example, risk taking and
problem behaviors often show increases in the transition to adolescence (Duckworth, Gendler,
& Gross, 2014). Family-╉based interventions intended to address system-╉level factors (e.g., fam-
ily, peers, school) have found reduced problem behaviors among children and adolescents (e.g.,
Chang, Shaw, Dishion, Gardner, & Wilson, 2014; Prado et al., 2013). Together, empirical results
suggest that interventions that target different levels of the child’s system (e.g., child-╉, parent-╉,
peer-╉, and/╉or school-╉level factors) modify emotion self-╉regulation and adjustment.

Internalizing problems
Effortful control (particularly attentional control) is expected to reduce the internalizing symp-
toms of depression, anxiety, and social withdrawal. In contrast, children who are rigid, con-
strained, and behaviorally inhibited (i.e., discomfort with novel stimuli, including people; Kagan
& Fox, 2006) may be prone to internalizing problems. Thus, as discussed previously, it may be
hypothesized that children with internalizing symptoms are somewhat low in effortful control
(particularly attentional control) and reactive undercontrol (i.e., impulsivity), but high in reactive
overcontrol (i.e., behavioral inhibition).

Self-╉regulation and internalizing problems


In fact, the empirical findings regarding the relations of effortful control and reactive control
to children’s internalizing problems are somewhat mixed. For the most part, investigators have
found the predicted negative relation between effortful control and internalizing problems,
anxiety, and/╉or depression (Buckner, Mezzacappa, & Beardslee, 2009; Eisenberg et  al., 2001;
Eisenberg et al., 2007; Emerson, Mollet, & Harrison, 2005; Hopkins, Lavigne, Gouze, LeBailly,
& Bryant, 2013; McCoy & Raver, 2011; Morris et  al., 2013; Muris, 2006; Muris, de Jong, &
Engelen, 2004; Oldehinkel, Hartman, De Winter, Veenstra, & Ormel, 2004; Verstraeten, Vasey,
Raes, & Bijttebier, 2009; Zalewski, Lengua, Wilson, Trancik, & Bazinet, 2011). Longitudinal
82

28 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

data also support the negative relation (Eisenberg et al., 2005; 2009; Kiff, Lengua & Bush, 2011;
King, Lengua, & Monohan, 2013; Lengua, 2006; Oldehinkel, Hartman, Ferdinand, Verhulst,
& Ormel, 2007; Valiente, Eisenberg, Spinrad et  al., 2006). Nonetheless, in a sample of tod-
dlers, effortful control was negatively related to toddlers’ separation distress at both 18 and
30  months of age. However, the relations became nonsignificant once stability in separation
distress was controlled (Spinrad et al., 2007; see Eisenberg, Spinrad et al., 2010, for similar find-
ings at 42 months of age).
Moreover, some investigators have found a positive relation between self-​regulation and inter-
nalizing problems. For example, Murray and Kochanska (2002) reported that children with high
effortful control exhibited higher internalizing symptoms than did those with moderate effort-
ful control, although very few children in the study had severe internalizing problems. Other
researchers have reported no relations (see Lengua, 2008).
In most research, investigators have used continuous measures of internalizing and external-
izing and have not dealt with the reality of frequent co-​occurrence of the externalizing and
internalizing symptoms. Thus, relations of low regulation to internalizing symptoms could be
due to the co-​occurrence of externalizing symptoms. In research that examined pure internal-
izing problems, Eisenberg and colleagues (2001, 2005; Eisenberg, Valiente, et al., 2009; Wang,
Eisenberg, Valiente, & Spinrad, 2016) found little evidence of relations of effortful control to
pure (non-​co-​occurring) internalizing problems except in the early school years. Oldehinkel et
al. (2004) found that preadolescents with pure internalizing symptoms were lower in effortful
control than nondisordered children, but that difference was not nearly as great as for youths
with co-​occurring problems. In a more at-​risk adolescent sample, low effortful control pre-
dicted higher pure (non-​co-​occurring) depression (Wang et al., 2015). Thus, co-​occurrence
and age may affect the strength of relations between indices of self-​regulation and internalizing
symptoms.
It is important to differentiate between aspects of effortful control in understanding the rela-
tions. It is likely that some aspects of effortful control, such as attentional control, may be par-
ticularly important for children with internalizing problems. Specifically, attentional control is
thought to reduce bias toward negative emotions such as sadness and anxiety and may serve to
move attention from negative to neutral or positive thoughts (Derryberry & Rothbart, 1997). On
the other hand, other components of effortful control, such as inhibitory control, may be related
to internalizing problems due only to its negative associations with externalizing problems; that is,
relations of inhibitory control to internalizing problems may be inflated because of co-​occurring
externalizing problems (see Eisenberg, Spinrad, & Eggum, 2010).
Indeed, children with pure internalizing symptoms, compared to non-​disordered children,
exhibited deficits in attentional control at 55 to 97 months (Eisenberg et al., 2001), although
not two or four years later (Eisenberg, Sadovsky, et al., 2005; Eisenberg, Valiente, et al., 2009).
Attentional control also has been negatively associated with anxiety disorder symptoms (Muris
et al., 2004) and with boys’ anxiety and depression (Emerson et al., 2005). In longitudinal
research, low attentional control has been associated with more internalizing problems from
four-​and-​a-​half to eleven years (or consistently high internalizing problems; Kim & Deater-​
Deckard, 2011) and with withdrawal that was high and declined over six years (Eggum et al.,
2009).
Another aspect of effortful control, inhibitory control, has been examined in relation to inter-
nalizing problems, although it is conceptually less related to internalizing than attentional con-
trol. Findings have been somewhat mixed. Some researchers have reported a negative relation
between the two constructs, but potential co-​occurring externalizing problems were not con-
trolled in these studies (Lengua, 2003; Rhoades, Greenberg, & Domitrovich, 2009; Riggs, Blair, &
92

Internalizing problems 29

Greenberg, 2003). On the other hand, Eisenberg and colleagues (2001) found that internalizing
children (without externalizing symptoms), compared to nondisordered children, had similar
levels of inhibitory control.
As previously noted, children’s age may moderate the relations between effortful control/╉reac-
tive control and internalizing problems. In a series of studies, Eisenberg and colleagues (Eisenberg
et al., 2001; Eisenberg, Sadovsky, et al., 2005; Eisenberg, Valiente, et al., 2009) found that whereas
the relation between effortful control and pure internalizing problems was evident in young
children, it was not found in mid-╉to late-╉elementary school. Effortful control may be linked
to children’s internalizing in younger age groups because effortful control at younger ages may
be particularly important when effortful control is rapidly developing. Further, effortful control
may be particularly important in early development when internalizing problems may require the
regulation of negative emotions of sadness and anxiety as opposed to more cognitive processes.
As further evidence for moderation by age, Dennis, Brotman, Huang, and Gouley (2007) reported
that children’s observed effortful control was negatively related to internalizing problems at age
four but not at ages five and six. However, it is clear that internalizing problems are associated with
deficits in executive functioning—╉including overlapping self-╉regulatory capacities—╉in adulthood
(see Snyder, 2013, for a meta-╉analysis). Continued research on age-╉related changes in the relations
of effortful control to internalizing problems is needed.
Other emotion regulation strategies, such as cognitive reappraisal or suppression of emotion,
have been studied in association with internalizing symptoms. Lougheed and Hollenstein (2012)
found that adolescents who had a range of emotion regulation strategies were lower in internal-
izing problems than those with limited emotion regulation strategies. In addition, adolescents’
cognitive reappraisal has been found to predict lower levels of depression, whereas suppressing
emotional expressions has been related to higher social anxiety (Eastabrook, Flynn, & Hollenstein,
2014). Similarly, rumination, defined as “the process of thinking perseveratively about one’s feel-
ings and problems rather than in terms of the specific content of thoughts” (p. 400), often has been
viewed as reflective of the lack of effective coping or self-╉regulation (Nolen-╉Hoeksema, Wisco, &
Lyubomirsky, 2008), is thought to be related to maladaptive suppression, and has been related
to a variety of internalizing problems (Nolen-╉Hoeksema, Stice, Wade, & Bohon, 2007; Nolen-╉
Hoeksema et al., 2008). Thus, it appears that internalizing problems are associated with specific
maladaptive methods of regulating emotion.

Reactive control and internalizing problems


Researchers have less frequently examined the associations of impulsivity (i.e., reactive under-
control) to internalizing problems. Eisenberg et  al. (2001) found that exclusively internalizing
children (i.e., those without externalizing problems) were characterized by low levels of impulsiv-
ity (also see Eisenberg et al., 2007; Eisenberg, Valiente, et al., 2009). However, Lengua et al. (1998)
found a positive relation between impulsivity and depression after removing overlapping items
among the constructs, but not before removing overlapping items (note, however, that they did
not differentiate children with pure from co-╉occurring internalizing symptoms). Similarly, Stifter
and colleagues (2008) reported that exuberant children (i.e., those high on approach and posi-
tive affect) were rated by their parents as higher in internalizing behaviors, as well as externaliz-
ing problems, than were low-╉reactive children. Given that exuberant children were differentiated
from inhibited children, it is likely that the exuberant children’s impulsivity was related to their
high level of externalizing problems.
On the other hand, reactive overcontrol (i.e., behavioral inhibition) tends to show consistent
positive relations to internalizing symptoms (Caspi, Henry, McGee, Moffitt, & Silva, 1995; Degnan,
Almas, & Fox, 2010; Kagan & Snidman, 1999; Schwartz, Snidman, & Kagan, 1999). Eggum and
03

30 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

colleagues (2012) reported that children’s shyness (a construct highly related to behavioral inhibi-
tion) was positively related to internalizing problems, even after controlling for earlier levels of
internalizing symptoms. Further, Spinrad et al. (2007) found that toddlers’ inhibition to novelty
was positively correlated with separation distress at both 18 and 30 months, controlling for the
effect of effortful control.

Culture and the self-regulation of internalizing problems


It is likely that cultural norms play a role in how temperamentally based self-╉regulation capaci-
ties impact children’s outcomes because certain traits may be more acceptable in certain cultures.
For example, in collectivist cultures, such as China, where group harmony is emphasized and
the expression of negative emotion is discouraged, children’s characteristics such as regulation
and attention may be highly valued and reinforced. Further, it is possible that shyness and social
withdrawal have been viewed as less problematic in Asian, compared to Western cultures (Chen,
Cen, Li, & He, 2005).
Although more cross-╉cultural work in the area needs to be conducted, it appears that the rela-
tions between effortful control and internalizing problems in current day China are similar to
findings in the US (Eisenberg et al., 2007; Muhtadie, Zhou, Eisenberg, & Wang, 2013; Zhou et al.,
2004). In a within-╉culture study of Chinese children, first and second graders with low effortful
control were found to have relatively high levels of pure internalizing problems (Eisenberg et al.,
2007); these children were also low in impulsivity. In another study, childhood delay ability at age
two negatively predicted Chinese children’s loneliness and depression (Chen, Zhang, Chen, & Li,
2012). In a cross-╉cultural study, Zhou et al. (2009) found that country of origin (China versus US)
did not moderate the findings when predicting children’s membership in an internalizing (versus
nondisordered) group.
Similar findings have been obtained in Western European countries. In a series of studies,
Muris and colleagues found that in the Netherlands, children’s effortful control was negatively
related to anxiety or more global internalizing symptoms (Muris et  al., 2004; Muris, Meesters,
& Blijlevens, 2007; Muris, Meesters, & Rompelberg, 2007; Muris, van der Pennen, Sigmond,
& Mayer, 2008; see Oldehinkel et  al., 2004, for similar findings). Further, as found in the US,
behavioral inhibition was positively correlated with internalizing problems (Muris, Meesters, de
Kanter, & Timmerman, 2005). Rydell and colleagues (2003) reported that poorly regulated fear
was associated with internalizing problems in a sample of Swedish children. Finally, in a multi-╉
ethnic sample of adolescents from the Netherlands, de Boo and Kolk (2007) found that the inverse
relations between effortful control and depressive mood were consistent across Dutch, Turkish,
Moroccan, and mixed ethnic participants. Thus, findings for Western European samples appear
quite consistent with US samples.
Cross-╉cultural differences between the US and Russia were found in one study examining
the role of infant temperament to toddlers’ behavior problems. Gartstein and colleagues (2013)
found that infants’ falling reactivity (i.e., the ability to calm following arousal) predicted lower
internalizing problems in the US but not in Russia. It should be noted that there were also mean-╉
level differences in internalizing problems in the Russian sample, such that Russian toddlers
were significantly lower than US toddlers on internalizing problems, regardless of regulatory
skills in infancy. Thus, these findings may be attributed to the fact that internalizing symptoms
were quite rare in the Russian sample.
Of course, there may be cultural differences in relations within the United States. Because
most studies have focused on White, non-╉Hispanic children in the US, it is important to
13

Internalizing problems 31

understand whether relations of effortful and reactive control to internalizing problems dif-
fer across ethnic groups. In one of the few studies to examine such relations, Loukas and
Roalson (2006) reported that effortful control was negatively related to depression in both
European American and Latino adolescents. Consistent with these findings, in a sample of
Head Start preschoolers, the negative relation between regulation and internalizing problems
was not moderated by ethnicity/╉race (Hispanic versus African American; McCoy & Raver,
2011). Thus, the evidence thus far suggests that the relations of effortful and reactive control
to children’s internalizing problem behaviors are quite similar across countries and cultural
groups.

Moderating processes
Perhaps some of the inconsistencies in the relations of effortful control and reactive control to
internalizing problems are due to interactions between effortful control (or impulsivity) and other
aspects of temperament (i.e., negative emotionality, shyness) when predicting internalizing prob-
lems. Indeed, there is evidence that effortful control moderates the positive relations between
negative affect and depression/╉internalizing problems (Muris, 2006; Oldehinkel et  al., 2007;
Verstraeten et al., 2009; Yap et al., 2011). Specifically, the positive relations between negative emo-
tionality and internalizing problems appear to be stronger for children low in effortful control. In
contrast, Eisenberg and colleagues (2004) found no evidence of an interaction between negative
emotionality and effortful control when predicting internalizing problems.
Relations between reactive control and children’s internalizing problems also may be mod-
erated by effortful control. White, McDermott, Degnan, Henderson and Fox (2011) found that
behavioral inhibition (reactive overcontrol) at 24 months predicted parent-╉reported anxiety dur-
ing preschool, but only for children with poor attention shifting. Similarly, in a study conducted
with adolescents in the Netherlands, behavioral inhibition predicted higher internalizing prob-
lems, particularly for children with low attentional control (Sportel, Nauta, de Hullu, de Jong, &
Hartman, 2011). Contrary to the findings for attentional control, the positive relations between
behavioral inhibition and anxiety have been found for children high in inhibitory control but
not for those with low inhibitory control, suggesting that children who are high in both inhibi-
tory control and behavioral inhibition may be somewhat overcontrolled and anxious (White
et al., 2011).

Mediating processes
The relation between effortful control and internalizing symptoms may also be mediated by
dispositional factors such as ego-╉resiliency. That is, effortful control may allow for flexible and
adaptive behavior in the face of challenge, and this flexibility may counter the development
of internalizing problems. In a number of studies, ego-╉resiliency mediated the relations of
effortful control to low levels of internalizing (but usually not externalizing) problems (e.g.,
Eisenberg, Spinrad, et  al., 2004; Valiente et  al., 2006). This mediated relation also has been
found in Chinese (Eisenberg, Chang, Ma, & Huang, 2009)  and French (Hofer, Eisenberg, &
Reiser, 2010) samples.
In sum, the relations of effortful control and reactive control to children’s internalizing problems
are somewhat complex. Although most literature supports a negative relation between effortful
control and internalizing problems, findings are somewhat mixed. A  more nuanced approach
indicates that perhaps attentional components of effortful control are more strongly related to
internalizing problems than are behavioral components (i.e., inhibitory control). Further attention
23

32 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

to moderating and mediating processes might help to clarify the relation between self-╉regulation
and internalizing problems.

Self-regulation and co-occurring problems


Internalizing and externalizing behaviors are positively associated (Bornstein, Hahn, & Haynes,
2010) and comorbidity can develop concurrently and/╉or sequentially (Boylan, Vaillancourt, Boyle,
& Szatmari, 2007). However, we know less about the underlying regulatory mechanisms involved
in the development and maintenance of their co-occurrence. Unfortunately, most researchers
have used measures of internalizing and externalizing problems without differentiating children
with pure problems of either type from those with co-╉occurring problems.
In early work in which children with co-╉occurring symptoms (at a borderline clinical level
or higher) were differentiated from those with pure externalizing or internalizing problems,
findings for children with pure externalizing problems and co-╉occurring problems were
combined when looking at relations with effortful control and impulsivity (Eisenberg et al.,
2001, 2005). In one longitudinal study (Eisenberg et  al., 2001, 2005), children in the early
and mid-╉elementary school years with both pure externalizing and co-╉occurring problems
(teacher-╉and parent-╉reported) were low in adult-╉reported effortful control (inhibitory con-
trol, attention focusing, and attention shifting) and high in impulsivity. When Eisenberg and
colleagues followed up the children at, on average, mid-╉to late-╉elementary school (or middle
school) age, they separated pure externalizing from co-╉occurring externalizing and internal-
izing (Eisenberg, Valiente, et al., 2009). Children with co-╉occurring problems were still low in
attentional and inhibitory control and high in impulsivity compared to nondisordered chil-
dren (as were children with pure externalizing problems). These researchers did not directly
compare youth with pure symptoms versus co-╉occurring ones. However, Oldehinkel et  al.
(2004) constructed similar groups and found that preadolescents with co-╉occurring symptoms
were lower in effortful control than children with pure internalizing or pure externalizing
symptoms or non-╉disordered children.
In the aforementioned studies, children were assigned to one maladjustment group or another
and the relations of effortful control/╉impulsivity to maladjustment were examined in relation to
a control group and, sometimes, another maladjustment group; the relations of effortful control/╉
impulsivity to the various maladjustment groups were not necessarily unique. In a recent study,
Wang et  al. (2015) examined the unique prediction of levels of pure internalizing (depressive
symptoms), pure externalizing (antisocial/╉aggressive behaviors), and co-╉occurring internalizing
and externalizing (depressive and antisocial/╉aggressive behaviors) in adolescence from effortful
control and impulsivity (simultaneously) five to six years earlier. Using the bi-╉factor modeling
technique, they first computed factors for pure externalizing, pure internalizing, and co-╉occurring
symptoms that were orthogonal from each other (e.g., variance unique to pure externalizing after
covarying out variance attributable to pure internalizing or co-╉occurring symptoms). Then they
predicted three symptom factors from both effortful control and impulsivity. Low effortful control
uniquely predicted pure externalizing problems and both low effortful control and low impul-
sivity uniquely predicted pure internalizing and co-╉occurring problems. According to an inter-
action among effortful control, impulsivity, and age, for older adolescents only, lower effortful
control predicted more pure externalizing and co-╉occurring symptoms at average and high levels
of impulsivity. Controlling for the relation of effortful control to maladjustment likely eliminated
the expected positive relation of impulsivity to co-╉occurring problems; indeed, in zero-╉order
correlation, impulsivity was positively related to externalizing problems (but was not related to
depression).
3

Summary 33

Other research suggests that children and adolescents with co-╉occurring internalizing
and externalizing symptoms, like those with pure externalizing problems, have difficulties
with self-╉regulation. For example, in research with a behavioral measure of inhibitory con-
trol administered to aggressive children, eight-╉to 12-year-olds with pure externalizing prob-
lems, had somewhat greater problems with slowing down and monitoring responding when
needed than those with co-╉occurring internalizing problems, although the latter group still
exhibited marginally less response slowing than did nondisordered children (Stieben et al.,
2007). In a study of adolescents, Garnefski, Kraaij, and van Etten (2005) found that those
with co-╉occurring symptoms or pure internalizing were higher on the maladaptive emotion
regulation strategies of self-╉blame and rumination than those with pure externalizing or no
symptoms.
Moreover, Pang and Beauchaine (2013) found that eight-╉to 12-╉year-╉old children with co-╉
occurring internalizing (diagnosed depression) and externalizing (conduct disorder) problems
exhibited lower resting RSA and greater RSA withdrawal to emotion evocation beyond com-
promised RSA responding for pure diagnoses. Baseline RSA is often viewed as reflecting physi-
ological self-╉regulation and excessive RSA withdrawal is viewed as an index of emotional lability.
However, results on RSA and co-╉occurring internalizing and externalizing problems vary across
studies (e.g., see Hinnant & El-╉Sheikh, 2013). In contrast, in research with eight-╉to 12-╉year-╉old
aggressive children (Stieben et al., 2007), ERP responding to an emotional induction supported
the conclusion that children with pure externalizing problems had greater regulatory difficul-
ties. Results in regard to regulatory deficits may differ depending on whether the measure of
regulation taps inhibition of behavior (which may be worse to pure externalizers) or modula-
tion of emotion and related attention (which may be exacerbated by co-╉occurring internalizing
problems).

Summary
In summary, investigators have frequently found that children’s self-╉regulation is related to their
externalizing and internalizing symptoms. However, research on co-╉morbidity is limited, and
relatively little is known about differential relations with self-╉regulatory processes for various
types of internalizing problems (e.g., anxiety, depression, and social withdrawal) and externaliz-
ing problems (e.g., overt versus covert, behaviors that vary in destructiveness to others). Similarly,
more research is needed to determine if some aspects of self-╉regulation (e.g., delay skills versus
“cooler” executive attention abilities) relate to externalizing and/╉or internalizing symptoms more
than others (Kim et al., 2013).
In addition, it would be useful to examine additional moderators (besides negative emotional-
ity) of the association between self-╉regulation and children’s externalizing and internalizing symp-
toms. For example, impulsivity may interact with self-╉regulatory skills to predict maladjustment,
perhaps more at periods of development when individual differences in impulsivity are relatively
marked (e.g., in the early years and in adolescence; see Wang et al., 2015). More attention to the
mediators of the relation between self-╉regulation and children’s symptoms would also be useful;
coping efficacy or self-╉efficacy come to mind as potential mediators.
Moreover, there is limited research examining potential causal relations between self-╉regulation
and maladjustment, either in experimental designs or in longitudinal research that controls for
prior levels of maladjustment. The research suggests that relations are bi-╉directional and may vary
in causal predominance at different ages (Sulik et al., 2015).
The results from interventions suggest that children, especially those high in baseline levels
of maladjustment or other risk factors, improve more in regulation/╉lower externalizing than
43

34 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

do children with lesser levels of risk. However, there is less work on which aspects of interven-
tions (e.g., training in self-╉regulation, teaching an understanding of emotions) have the obtained
effects on reducing maladjustment. In addition, factors in children’s ecologies that might mod-
erate the association between maladjustment and effortful control, such as stress from poverty,
poor schools, or neighborhood violence, merit attention. Moreover, such contextual factors, if
they affect self-╉regulation and/╉or maladjustment, could play a role in mediated relations (e.g.,
from poverty to self-╉regulation to maladjustment; see Lengua et al., 2015).
Finally, although there seem to be more similarities than differences across cultural groups
in the relations of self-╉regulation to problem symptoms, researchers have emphasized the need
to understand the cultural processes (e.g., acculturation, cultural values) associated with child
development and demographic heterogeneity within cultural groups (García Coll et  al., 1996;
Li-╉Grining, 2012). For example, Telzer and colleagues (2011) found that adolescents of European-╉
and Latino-╉American origin who endorsed strong familism values (e.g., cultural attitudes of fam-
ily respect and obligation) showed more neural activation of regions involved with self-╉control
(i.e., ventral striatum regions associated with reward processing) when making costly contribu-
tions to family in a “family assistance task” in early adulthood. That is, cultural processes (rather
than cultural group differences) were associated with self-╉regulatory processes in the context of
a culturally relevant reward task. Moreover, cultural values may affect the degree to which self-╉
regulation is valued and the degree to which externalizing or internalizing symptoms are deemed
problematic, with the consequence that acceptance of these values affects the relation between
self-╉regulation and problem behaviors. Future research examining the cultural processes associ-
ated with the development of self-╉regulation, as well as the relation between self-╉regulation to
externalizing and internalizing symptoms, could contribute to an understanding of the role of
cultural factors in self-╉regulation and maladjustment.

References
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Arlington, VA: American Psychiatric Association.
Arsenio, W. F., Adams, E., & Gold, J. (2009). Social information processing, moral reasoning, and emotion
attributions: Relations with adolescents’ reactive and proactive aggression. Child Development, 80,
1739–╉1755. doi:10.1111/╉j.1467-╉8624.2009.01365.x
Barker, E. D., Oliver, B. R., Viding, E., Salekin, R. T., & Maughan, B. (2011). The impact of prenatal
maternal risk, fearless temperament and early parenting on adolescent callous-╉unemotional
traits: A 14-╉year longitudinal investigation. Journal of Child Psychology and Psychiatry, 52, 878–╉888.
doi: 10.1111/╉j.1469-╉7610.2011.02397.x
Belsky, J., Pasco Fearon, R. M., & Bell, B. (2007). Parenting, attention and externalizing problems: Testing
mediation longitudinally, repeatedly and reciprocally. Journal of Child Psychology and Psychiatry, 48,
1233–╉1242. doi: 10.1111/╉j.1469-╉7610.2007.01807.x
Blair, C., & Raver, C. C. (2014). Closing the achievement gap through modification of neurocognitive
and neuroendocrine function: results from a cluster randomized controlled trial of an innovative
approach to the education of children in kindergarten. PLoS One, 9(11). e112393. doi:10.1371/╉journal.
pone.0112393
Blandon, A. Y., Calkins, S. D., Grimm, K. J., Keane, S. P., & O’Brien, M. (2010). Testing a developmental
cascade model of emotional and social competence and early peer acceptance. Development and
Psychopathology, 22, 737–╉748. doi:10.1017/╉S0954579410000428
Bornstein, M. H., Hahn, C. S., & Haynes, O. M. (2010). Social competence, externalizing, and internalizing
behavioral adjustment from early childhood through early adolescence: Developmental cascades.
Development and Psychopathology, 22, 717–╉735. doi:10.1017/╉S0954579410000416
53

Summary 35

Boylan, K., Vaillancourt, T., Boyle, M., & Szatmari, P. (2007). Comorbidity of internalizing disorders in
children with oppositional defiant disorder. European Child and Adolescent Psychiatry, 16, 484–​494. doi
10.1007/​s00787-​007-​0624-​1
Buckner, J. C., Mezzacappa, E., & Beardslee, W. R. (2009). Self-​regulation and its relations to adaptive
functioning in low income youths. American Journal of Orthopsychiatry, 29, 19–​30. doi:10.1037/​
a0014796
Carver, C. S. (2005). Impulse and constraint: Perspectives from personality psychology, convergence with
theory in other areas, and potential for integration. Personality and Social Psychology Review, 9,
312–​333. doi:10.1207/​s15327957pspr0904_​2
Caspi, A., Henry, B., McGee, R. O., Moffitt, T. E., & Silva, P. A. (1995). Temperamental origins of child and
adolescent behavior problems: From age three to fifteen. Child Development, 66, 55–​68. doi:10.2307/​
1131190
Chang, H., Shaw, D. S., Dishion, T. J., Gardner, F., & Wilson, M. N. (2014). Direct and indirect effects of
the family check-​up on self-​regulation from toddlerhood to early school-​age. Journal of Abnormal Child
Psychology, 42, 1117–​1128. doi:10.1007/​s10802-​014-​9859-​8
Chen, X., Cen, G., Li, D., & He, Y. (2005). Social functioning and adjustment in chinese children: The
imprint of historical time. Child Development, 76, 182–​195. doi:10.1111/​j.1467-​8624.2005.00838.x
Chen, X., Zhang, G., Chen, H., & Li, D. (2012). Performance on delay tasks in early childhood predicted
socioemotional and school adjustment nine years later: A longitudinal study in Chinese children.
International Perspectives in Psychology: Research, Practice, Consultation, 1, 3–​14. doi:10.1037/​a0026363
Cohen, J. R., & Lieberman, M. D. (2010). The common neural basis of exerting neural self-​control in
multiple domains. In R. R. Hassin, K. N. Oscher & Y. Trope (Eds.), Self control in society, mind, and
brain. New York: Oxford University Press.
Compas, B. E., Connor-​Smith, J. K., Saltzman, H., Thomsen, A. H., & Wadsworth, M. E. (2001). Coping
with stress during childhood and adolescence: Problems, progress, and potential in theory and
research. Psychological Bulletin, 127, 87–​127. doi:10.1037/​0033-​2909.127.1.87
Conceição, A., & Carvalho, M. (2013). Emotional and behavioral problems in preadolescents: Relationships
with temperament, coping and emotional regulation strategies, and facial expressions identification.
Psychologica, 56, 83–​100. doi:10.14195/​1647-​8606_​56_​5
Conduct Problems Prevention Research Group. (2010). The effects of a multiyear universal social-​
emotional learning program: The role of student and school characteristics. Journal of Consulting and
Clinical Psychology, 78, 156–​168. doi:10.1037/​a0018607
Coplan, R. J., & Armer, M. (2007). A “multitude” of solitude: A closer look at social withdrawal
and nonsocial play in early childhood. Child Development Perspectives, 1, 26–​32. doi:10.1111/​
j.1750-​8606.2007.00006.x
de Boo, G. M., & Kolk, A. M. (2007). Ethnic and gender differences in temperament, and the relationship
between temperament and depressive and aggressive mood. Personality and Individual Differences, 43,
1756–​1766. doi:10.1016/​j.paid.2007.05.012
de Castro, B. O., Merk, W., Koops, W., Veerman, J. W., & Bosch, J. D. (2005). Emotions in social
information processing and their relations with reactive and proactive aggression in referred
aggressive boys. Journal of Clinical Child and Adolescent Psychology, 34, 105–​116. doi:10.1207/​
s15374424jccp3401_​10
Degnan, K. A., Almas, A. N., & Fox, N. A. (2010). Temperament and the environment in the etiology
of childhood anxiety. Journal of Child Psychology and Psychiatry, 51, 497–​517. doi:10.1111/​
j.1469-​7610.2010.02228.x
Dennis, T. A., Brotman, L. M., Huang, K., & Gouley, K. K. (2007). Effortful control, social competence,
and adjustment problems in children at risk for psychopathology. Journal of Clinical Child and
Adolescent Psychology, 36, 442–​454. doi:10.1080/​15374410701448513
Derryberry, D., & Rothbart, M. K. (1997). Reactive and effortful processes in the organization of
temperament. Development and Psychopathology, 9, 633–​652. doi:10.1017/​S0954579497001375
63

36 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

Di Norcia, A., Pecora, G., Bombi, A. S., Baumgartner, E., & Laghi, F. (2014). Hot and cool inhibitory
control in Italian toddlers: associations with social competence and behavioral problems. Journal of
Child and Family Studies, 24, 909–​914. doi:10.1007/​s10826-​014-​9901-​z
Duckworth, A. L., Gendler, T. S., & Gross, J. J. (2014). Self-​control in school-​age children. Educational
Psychologist, 49, 199–​217. doi: 10.1080/​00461520.2014.926225
Eastabrook, J. M., Flynn, J. J., & Hollenstein, T. (2014). Internalizing symptoms in female
adolescents: Associations with emotional awareness and emotion regulation. Journal of Child and
Family Studies, 23, 487–​496. doi:10.1007/​s10826-​012-​9705-​y
Eggum, N. D., Eisenberg, N., Reiser, M., Spinrad, T. L., Michalik, N. M., Valiente, C., … Sallquist, J.
(2012). Relations over time among children’s shyness, emotionality, and internalizing problems. Social
Development, 21, 109–​129. doi:10.1111/​j.1467-​9507.2011.00618.x
Eggum, N. D., Eisenberg, N., Spinrad, T. L., Reiser, M., Gaertner, B. M., Sallquist, J., & Smith, C. L.
(2009). Development of shyness: Relations with children’s fearfulness, sex, and maternal behavior.
Infancy, 14, 325–​345. doi: 10.1080/​15250000902839971
Eisenberg, N., Chang, L., Ma, Y., & Huang, X. (2009). Relations of parenting style to Chinese children’s
effortful control, ego resilience, and maladjustment. Development and Psychopathology, 21, 455–​477.
doi:10.1017/​S095457940900025X
Eisenberg, N., Cumberland, A., Spinrad, T. L., Fabes, R. A., Shepard, S. A., Reiser, M., … Guthrie, I.
K. (2001). The relations of regulation and emotionality to children’s externalizing and internalizing
problem behavior. Child Development, 72, 1112–​1134. doi:10.1111/​1467-​8624.00337
Eisenberg, N., Edwards, A., Spinrad, T. L., Sallquist, J., Eggum, N. D., & Reiser, M. (2013). Are effortful
and reactive control unique constructs in young children? Developmental Psychology, 49, 2082–​2094.
doi:10.1037/​a0031745
Eisenberg, N., & Fabes, R. A. (1992). Emotion, regulation, and the development of social competence. In
M. S. Clark (Ed.), Review of personality and social psychology (Vol. 14, pp. 119–​150). Newbury Park,
CA: Sage.
Eisenberg, N., Hofer, C., Sulik, M. J., & Spinrad, T. L. (2014). Effortful control and its socioemotional
consequences. In J. J. Gross (Ed.), Handbook of emotion regulation (2nd ed.). New York: Guilford Press.
Eisenberg, N., Liew, J., & Pidada, S. U. (2004). The longitudinal relations of regulation and emotionality to
quality of Indonesian children’s socioemotional functioning. Developmental Psychology, 40, 790–​804.
doi:10.1037/​0012-​1649.40.5.790
Eisenberg, N., Ma, Y., Chang, L., Zhou, Q., West, S. G., & Aiken, L. (2007). Relations of effortful control,
reactive undercontrol, and anger to Chinese children’s adjustment. Development and Psychopathology,
19, 385–​409. doi: 1017/​S0954579407070198
Eisenberg, N., & Morris, A. S. (2002). Children’s emotion-​related regulation. In R. V. Kail (Ed.), Advances
in child development and behavior (Vol. 30, pp. 189–​229). San Diego, CA: Academic Press.
Eisenberg, N., Sadovsky, A., Spinrad, T. L., Fabes, R. A., Losoya, S. H., Valiente, C., … Shepard, S. A.
(2005). The relations of problem behavior status to children’s negative emotionality, effortful control,
and impulsivity: Concurrent relations and prediction of change. Developmental Psychology, 41,
193–​211. doi: 10.1037/​0012-​1649.41.1.193
Eisenberg, N., & Spinrad, T. L. (2004). Emotion-​related regulation: Sharpening the definition. Child
Development, 75, 334–​339. doi:10.1111/​j.1467-​8624.2004.00674.x
Eisenberg, N., Spinrad, T. L., & Eggum, N. D. (2010). Emotion-​related self-​regulation and its relation to
children’s maladjustment. Annual Review of Clinical Psychology, 6, 495–​525. doi: 10.1146/​annurev.
clinpsy.121208.131208
Eisenberg, N., Spinrad, T. L., Eggum, N. M., Silva, K. M., Reiser, M., Hofer, C., … Michalik, N. (2010).
Relations among maternal socialization, effortful control, and maladjustment in early childhood.
Development and Psychopathology, 22, 507–​525. doi:10.1017/​S0954579410000246
Eisenberg, N., Spinrad, T. L., Fabes, R. A., Reiser, M., Cumberland, A., Shepard, S. A., … Murphy, B.
(2004). The relations of effortful control and impulsivity to children’s resiliency and adjustment. Child
Development, 75, 25–​46. doi:10.1111/​j.1467-​8624.2004.00652.x
73

Summary 37

Eisenberg, N., Taylor, Z. E., Widaman, K. F., & Spinrad, T. L. (2015). Externalizing symptoms, effortful
control, and intrusive parenting: A test of bi-​directional longitudinal relations during early childhood.
Development and Psychopathology, 27, 953–​968. doi:10.1017/​S0954579415000620.
Eisenberg, N., Valiente, C., Spinrad, T. L., Cumberland, A., Liew, J., Reiser, M., … Losoya, S. H. (2009).
Longitudinal relations of children’s effortful control, impulsivity, and negative emotionality to their
externalizing, internalizing, and co-​occurring behavior problems. Developmental Psychology, 45, 988–​
1008. doi:10.1037/​a0016213
Eisenberg, N., Zhou, Q., Spinrad, T. L., Valiente, C., Fabes, R. A., & Liew, J. (2005). Relations among
positive parenting, children’s effortful control, and externalizing problems: A three-​wave longitudinal
study. Child Development, 76, 1055–​1071. doi:10.1111/​j.1467-​8624.2005.00897.x
Emerson, C. S., Mollet, G. A., & Harrison, D. W. (2005). Anxious-​depression in boys: An evaluation of
executive functioning. Archives of Clinical Neuropsychology, 20, 539–​546. doi:10.1016/​j.acn.2004.10.003
Frick, P. J., & Morris, A. S. (2004). Temperament and developmental pathways to conduct problems.
Journal of Clinical Child and Adolescent Psychology, 33, 54–​68. doi:10.1207/​S15374424JCCP3301_​6
Frick, P. J., Ray, J. V., Thornton, L. C., & Kahn, R. E. (2014). Annual research review: A developmental
psychopathology approach to understanding callous-​unemotional traits in children and adolescents with
serious conduct problems. Journal of Child Psychology and Psychiatry, 55, 532–​548. doi: 10.1111/​jcpp.12152
García Coll, C., Lamberty, G., Jenkins, R., Pipes McAdoo, H., Wasik, B. H., & Vásquez García, H.
(1996). An integrative model for the study of developmental competencies in minority children. Child
Development, 67, 1891–​1914. doi:10.2307/​1131600
Garnefski, N., Kraaij, V., & van Etten, M. (2005). Specificity of relations between adolescents’ cognitive
emotion regulation strategies and internalizing and externalizing psychopathology. Journal of
Adolescence, 28, 619–​631. doi:10.1016/​j.adolescence.2004.12.009
Gartstein, M. A., Slobodskaya, H. R., Kirchhoff, C., & Putnam, S. P. (2013). Cross-​cultural differences
in the development of behavior problems: Contributions of infant temperament in Russia and U.S.
International Journal of Developmental Science, 72, 95–​104. doi: 10.3233/DEV-1312104
Goldsmith, H. H., Pollak, S. D., & Davidson, R. J. (2008). Developmental neuroscience perspectives on
emotion regulation. Child Development Perspectives, 2, 132–​140. doi:10.1111/​j.1750-​8606.2008.00055.x
Gross, J. J. (2014). Emotion regulation: Conceptual and empirical foundations. In J. J. Gross (Ed.),
Handbook of emotion regulation (2nd ed., pp. 3–​20). New York: Guilford Press.
Gusdorf, L. M., Karreman, A., van Aken, M. A., Dekovic, M., & van Tuijl, C. (2011). The structure
of effortful control in preschoolers and its relation to externalizing problems. Brittish Journal of
Developmental Psychology, 29, 612–​634. doi:10.1348/​026151010X526542
Happaney, K., Zelazo, P. D., & Stuss, D. T.. (2004). Development of orbitofrontal function: Current themes
and future directions. Brain and Cognition, 55, 1–​10. doi 10.1016/​j.bandc.2004.01.001
Hinnant, J. B., & El-​Sheikh, M. (2013). Codevelopment of externalizing and internalizing symptoms in
middle to late childhood: Sex, baseline respiratory sinus arrhythmia, and respiratory sinus arrhythmia
reactivity as predictors. Development and Psychopathology, 25, 419–​436. doi:10.1017/​S0954579412001150
Hofer, C., Eisenberg, N., & Reiser, M. (2010). The role of socialization, effortful control, and ego resiliency
in French adolescents’ social functioning. Journal of Research on Adolescence, 20, 555–​582. doi:10.1111/​
j.1532-​7795.2010.00650.x
Honomichl, R. D., & Donnellan, M. B. (2011). Dimensions of temperament in preschoolers predict risk
taking and externalizing behaviors in adolescents. Social Psychological and Personality Science, 3, 14–​22.
doi:10.1177/​1948550611407344
Hopkins, J., Lavigne, J. V., Gouze, K. R., LeBailly, S. A., & Bryant, F. B. (2013). Multi-​domain models of
risk factors for depression and anxiety symptoms in preschoolers: Evidence for common and specific
factors. Journal of Abnormal Child Psychology, 41, 705–​722. doi:10.1007/​s10802-​013-​9723-​2
Hubbard, J. A., Smithmyer, C. M., Ramsden, S. R., Parker, E. H., Flanagan, K. D., Dearing, K. E., …
Simons, R. F. (2002). Observational, physiological, and self-​report measures of children’s
anger: Relations to reactive versus proactive aggression. Child Development, 73, 1101–​1118.
doi:10.1111/​1467-​8624.00460
83

38 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

Johnson, S. L., Carver, C. S., & Joormann, J. (2013). Impulsive responses to emotion as a transdiagnostic
vulnerability to internalizing and externalizing symptoms. Journal of Affective Disorders, 150, 872–​878.
doi:10.1016/​j.jad.2013.05.004
Kagan, J. (1998). Biology and the child. In W. Damon & N. Eisenberg (Eds.), Social, emotional and
personality development. Vol. 3. Handbook of child psychology (5th ed., Vol. 3, pp. 177–​235). Hoboken,
NJ: Wiley.
Kagan, J., & Fox, N. A. (2006). Biology, culture, and temperamental biases Handbook of child psychology: Vol.
3, social, emotional, and personality development (6th ed., pp. 167–​225). Hoboken, NJ: Wiley.
Kagan, J., & Snidman, N. (1999). Early childhood predictors of adult anxiety disorders. Biological
Psychiatry, 46, 1536–​1541. doi:10.1016/​S0006-​3223(99)00137-​7
Kiff, C. J., Lengua, L. J., & Bush, N. R. (2011). Temperament variation in sensitivity to parenting: Predicting
changes in depression and anxiety. Journal of Abnormal Child Psychology, 39, 1199–​1212. doi:10.1007/​
s10802-​011-​9539-​x
Kim, J., & Deater-​Deckard, K. (2011). Dynamic changes in anger, externalizing and internalizing
problems: Attention and regulation. Journal of Child Psychology and Psychiatry, 52, 156–​166.
doi:10.1111/​j.1469-​7610.2010.02301.x
Kim, S., Nordling, J. K., Yoon, J. E., Boldt, L. J., & Kochanska, G. (2013). Effortful control in “hot” and
“cool” tasks differentially predicts children’s behavior problems and academic performance. Journal of
Abnormal Child Psychology, 41, 43–​56. doi:10.1007/​s10802-​012-​9661-​4
King, K. M., Lengua, L. J., & Monahan, K. C. (2013). Individual differences in the development of self-​
regulation during pre-​adolescence: Connections to context and adjustment. Journal of Abnormal Child
Psychology, 41, 57–​69. doi:10.1007/​s10802-​012-​9665-​0
Lengua, L. J. (2003). Associations among emotionality, self-​regulation, adjustment problems, and
positive adjustment in middle childhood. Journal of Applied Developmental Psychology, 24, 595–​618.
doi:10.1016/​j.appdev.2003.08.002
Lengua, L. J. (2006). Growth in temperament and parenting as predictors of adjustment during children’s
transition to adolescence. Developmental Psychology, 42, 819–​832. doi:10.1037/​0012-​1649.42.5.819
Lengua, L. J. (2008). Anxiousness, frustration, and effortful control as moderators of the relation between
parenting and adjustment in middle-​childhood. Social Development, 17, 554–​577. doi:10.1111/​
j.1467-​9507.2007.00438.x
Lengua, L. J., Moran, L., Zalewski, M., Ruberry, E., Kiff, C., & Thompson, S. (2015). Relations of growth in
effortful control to family income, cumulative risk, and adjustment in preschool-​age children. Journal of
Abnormal Child Psychology, 43, 705–​720. doi 10.1007/​s10802-​014-​9941-​2.
Lengua, L. J., West, S. G., & Sandler, I. N. (1998). Temperament as a predictor of symptomatology in
children: Addressing contamination of measures. Child Development, 69, 164–​181. doi:10.2307/​
1132078
Li-​Grining, C. P. (2012). The role of cultural factors in the development of Latino preschoolers’ self-​
regulation. Child Development Perspectives, 6, 210–​217. doi:10.1111/​j.1750-​8606.2012.00255.x
Lougheed, J. P., & Hollenstein, T. (2012). A limited repertoire of emotion regulation strategies is
associated with internalizing problems in adolescence. Social Development, 21, 704–​721. doi:10.1111/​
j.1467-​9507.2012.00663.x
Loukas, A., & Roalson, L. A. (2006). Family environment, effortful control, and adjustment among
European American and Latino early adolescents. The Journal of Early Adolescence, 26, 432–​455.
doi:10.1177/​0272431606291939
Marmorstein, N. R. (2013). Associations between dispositions to rash action and internalizing and
externalizing symptoms in children. Journal of Clinical Child and Adolescent Psychology, 42, 131–​138.
doi:10.1080/​15374416.2012.734021
Marsee, M. A., & Frick, P. J. (2007). Exploring the cognitive and emotional correlates to proactive and
reactive aggression in a sample of detained girls. Journal of Abnormal Psychology, 35, 969–​981.
doi:10.1007/​s10802-​007-​9147-​y
93

Summary 39

McCoy, D. C., & Raver, C. C. (2011). Caregiver emotional expressiveness, child emotion regulation, and
child behavior problems among Head Start families. Social Development, 20, 741–​761. doi:10.1111/​
j.1467-​9507.2011.00608.x
McEachern, A. D., & Snyder, J. (2012). The relationship of impulsivity-​inattention and verbal ability to
overt and covert antisocial behaviors in children. Journal of Youth and Adolescence, 41, 984–​994.
doi:10.1007/​s10964-​011-​9710-​2
Mills-​Koonce, W. R., Wagner, N. J., Willoughby, M. T., Stifter, C., Blair, C., Granger, D. A., & Family Life
Project Key Investigators. (2015). Greater fear reactivity and psychophysiological hyperactivity among
infants with later conduct problems and callous-​unemotional traits. Journal of Child Psychology and
Psychiatry, 56, 147–​154. doi:10.1111/​jcpp.12289
Mischel, W., & Ayduk, Ö. (2011). Willpower in a cognitive-​affective processing system: The dynamics of delay
of gratification (2nd ed.). New York: Guilford Press.
Morales, S., Beekman, C., Blandon, A. Y., Stifter, C. A., & Buss, K. A. (2015). Longitudinal associations
between temperament and socioemotional outcomes in young children: The moderating role of RSA
and gender. Developmental Psychobiology, 57, 105–​119. doi:10.1002/​dev.21267
Morris, A. S., John, A., Halliburton, A. L., Morris, M. D. S., Robinson, L. R., Myers, S. S., … Terranova, A.
(2013). Effortful control, behavior problems, and peer relations: What predicts academic adjustment in
kindergartners from low-​income families? Early Education & Development, 24, 813–​828. doi:10.1080/​
10409289.2013.744682
Morris, A. S., Silk, J. S., Steinberg, L., Terranova, A. M., & Kithakye, M. (2010). Concurrent and
longitudinal links between children’s externalizing behavior in school and observed anger regulation in
the mother–​child dyad. Journal of Psychopathology and Behavioral Assessment, 32, 48–​56. doi:10.1007/​
s10862-​009-​9166-​9
Morris, P., Mattera, S. K., Castells, N., Bangser, M., Bierman, K., & Raver, C. (2014). Impact findings
from the Head Start CARES demonstration: National evaluation of three approaches to improving
preschoolers’ social and emotional competence. OPRE Report 2014-​44. Washington, DC: Office of
Planning, Research and Evaluation, Administration for Children and Families, U.S. Department of
Health and Human Services.
Muhtadie, L., Zhou, Q., Eisenberg, N., & Wang, Y. (2013). Predicting internalizing problems in chinese
children: The unique and interactive effects of parenting and child temperament. Development and
Psychopathology, 25, 653–​667. doi:10.1017/​S0954579413000084
Muris, P. (2006). Unique and interactive effects of neuroticism and effortful control on psychopathological
symptoms in non-​clinical adolescents. Personality and Individual Differences, 40, 1409–​1419.
doi:10.1016/​j.paid.2005.12.001
Muris, P., de Jong, P. J., & Engelen, S. (2004). Relationships between neuroticism, attentional control, and
anxiety disorders symptoms in non-​clinical children. Personality and Individual Differences, 37, 789–​
797. doi:10.1016/​j.paid.2003.10.007
Muris, P., Meesters, C., & Blijlevens, P. (2007). Self-​reported reactive and regulative temperament in early
adolescence: Relations to internalizing and externalizing problem behavior and “Big Three” personality
factors. Journal of Adolescence, 30, 1035–​1049. doi:10.1016/​j.adolescence.2007.03.003
Muris, P., Meesters, C., de Kanter, E., & Timmerman, P. E. (2005). Behavioural inhibition and behavioural
activation system scales for children: Relationships with Eysenck’s personality traits and psychopathological
symptoms. Personality and Individual Differences, 38, 831–​841. doi:10.1016/​j.paid.2004.06.007
Muris, P., Meesters, C., & Rompelberg, L. (2007). Attention control in middle childhood: Relations to
psychopathological symptoms and threat perception distortions. Behaviour Research and Therapy, 45,
997–​1010. doi: 10.1016/​j.brat.2006.07.010
Muris, P., van der Pennen, E., Sigmond, R., & Mayer, B. (2008). Symptoms of anxiety, depression, and
aggression in non-​clinical children: Relationships with self-​report and performance-​based measures
of attention and effortful control. Child Psychiatry and Human Development, 29, 455–​467. doi:10.1007/​
s10578-​008-​0101-​1
04

40 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

Murray, K. T., & Kochanska, G. (2002). Effortful control: Factor structure and relation to externalizing
and internalizing behaviors. Journal of Abnormal Child Psychology, 30, 503–​514. doi:10.1023/​
A:1019821031523
Neville, H. J., Stevens, C., Pakulak, E., Bell, T. A., Fanning, J., Klein, S., & Isbell, E. (2013). Family-​based
training program improves brain function, cognition, and behavior in lower socioeconomic status
preschoolers. PNAS, 110, 12138–​12143. doi:10.1073/​pnas.1304437110
Nigg, J. T. (2006). Temperament and developmental psychopathology. Journal of Child Psychology and
Psychiatry, 47, 395–​422. doi:10.1111/​j.1469-​7610.2006.01612.x
Nolen-​Hoeksema, S., Stice, E., Wade, E., & Bohon, C. (2007). Reciprocal relations between rumination
and bulimic, substance abuse and depressive symptoms in female adolescents. Journal of Abnormal
Psychology, 116, 198–​207. doi:10.1037/​0021-​843X.116.1.198
Nolen-​Hoeksema, S., Wisco, B. E., & Lyubomirsky, S. I. (2008). Rethinking rumination. Perspectives on
Psychological Science, 3, 400–​424. doi 10.1111/​j.1745-​6924.2008.00088.x
Oldehinkel, A. J., Hartman, C. A., De Winter, A. F., Veenstra, R., & Ormel, J. (2004). Temperament
profiles associated with internalizing and externalizing problems in preadolescence. Development and
Psychopathology, 16, 421–​440. doi:10.1017/​S0954579404044591
Oldehinkel, A. J., Hartman, C. A., Ferdinand, R. F., Verhulst, F. C., & Ormel, J. (2007). Effortful control
as modifier of the association between negative emotionality and adolescents’ mental health problems.
Development and Psychopathology, 19, 523–​539. doi:10.1017/​S0954579407070253
Pang, K. C., & Beauchaine, T. P. (2013). Longitudinal patterns of autonomic nervous system responding
to emotion evocation among children with conduct problems and/​or depression. Developmental
Psychobiology, 55, 698–​706. doi:10.1002/​dev.21065
Pickering, A. D., & Gray, J. A. (1999). The neuroscience of personality. In L. A. Pervin & O. P. John (Eds.),
Handbook of personality: Theory and Research (pp. 277–​299). New York: Guilford Press.
Piehler, T. F., Bloomquist, M. L., August, G. J., Gewirtz, A. H., Lee, S. S., & Lee, W. S. (2014). Executive
functioning as a mediator of conduct problems prevention in children of homeless families residing
in temporary supportive housing: A parallel process latent growth modeling approach. Journal of
Abnormal Child Psychology, 42, 681–​692. doi:10.1007/​s10802-​013-​9816-​y
Prado, G., Huang, S., Cordova, D., Malcolm, S., Estrada, Y., Cano, N., … Brown, C. H. (2013).
Ecodevelopmental and intrapersonal moderators of a family based preventive intervention for Hispanic
youth: A latent profile analysis. Prevention Science, 14, 290–​299. doi:10.1007/​s11121-​012-​0326-​x
Rhoades, B. L., Greenberg, M. T., & Domitrovich, C. E. (2009). The contribution of inhibitory control to
preschoolers’ social-​emotional competence. Journal of Applied Developmental Psychology, 30, 310–​320.
doi:10.1016/​j.appdev.2008.12.012.
Riggs, N. R., Blair, C. B., & Greenberg, M. T. (2003). Concurrent and 2-​year longitudinal relations between
executive function and the behavior of 1st and 2nd grade children. Child Neuropsychology, 9, 267–​276.
doi:10.1076/​chin.9.4.267.23513
Romer, D., Betancourt, L. M., Brodsky, N. L., Giannetta, J. M., Yang, W., & Hurt, H. (2011). Does
adolescent risk taking imply weak executive function? A prospective study of relations between
working memory performance, impulsivity, and risk taking in early adolescence. Developmental
Science, 14, 1119–​1133. doi:10.1111/​j.1467-​7687.2011.01061.x
Rothbart, M. K., & Bates, J. E. (2006). Temperament. In N. Eisenberg (Ed.) and W. Damon & R. M. Lerner
(Series Eds.), Handbook of child psychology: Vol. 3. Social, emotional, and personality development (6th
ed., pp. 105–​176). New York, NY: Wiley.
Rydell, A., Berlin, L., & Bohlin, G. (2003). Emotionality, emotion regulation, and adaptation among 5-​to
8-​year-​old children. Emotion, 3, 30–​47. doi:10.1037/​1528-​3542.3.1.30
Schoemaker, K., Mulder, H., Deković, M., & Matthys, W. (2013). Executive functions in preschool children
with externalizing behavior problems: A meta-​analysis. Journal of Abnormal Child Psychology, 41, 457–​
471. doi: 10.1007/​s10802-​012-​9684-​x
14

Summary 41

Schwartz, C. E., Snidman, N., & Kagan, J. (1999). Adolescent social anxiety as an outcome of inhibited
temperament in childhood. Journal of the American Academy of Child & Adolescent Psychiatry, 38,
1008–​1015. doi: 10.1097/​00004583-​199908000-​00017
Settles, R. E., Fischer, S., Cyders, M. A., Combs, J. L., Gunn, R. L., & Smith, G. T. (2012). Negative urgency:
A personality predictor of externalizing behavior characterized by neuroticism, low conscientiousness,
and disagreeableness. Journal of Abnormal Psychology, 121, 160–​172. doi: 10.1037/​a0024948
Snyder, H. R. (2013). Major depressive disorder is associated with broad impairments on
neuropsychological measures of executive function: A meta-​analysis and review. Psychological Bulletin,
139, 81–​132. doi:10.1037/​a0028727
Spinrad, T. L., Eisenberg, N., Gaertner, B., Popp, T., Smith, C. L., Kupfer, A., … Hofer, C. (2007).
Relations of maternal socialization and toddlers’ effortful control to children’s adjustment and social
competence. Developmental Psychology, 43, 1170–​1186. doi:10.1037/​0012-​1649.43.5.1170
Spinrad, T. L., Eisenberg, N., Silva, K. M., Eggum, N. D., Reiser, M., Edwards, A., … Gaertner, B. M.
(2012). Longitudinal relations among maternal behaviors, effortful control and young children’s
committed compliance. Developmental Psychology, 48, 552–​566. doi:10.1037/​a0025898
Sportel, B. E., Nauta, M. H., de Hullu, E., de Jong, P. J., & Hartman, C. A. (2011). Behavioral inhibition
and attentional control in adolescents: Robust relationships with anxiety and depression. Journal of
Child and Family Studies, 20, 149–​156. doi:10.1007/​s10826-​010-​9435-​y
Stieben, J., Lewis, M. D., Granic, I., Zelazo, P. D., Segalowitz, S., & Pepler, D. (2007). Neurophysiological
mechanisms of emotion regulation for subtypes of externalizing children. Development and
Psychopathology, 19, 455–​480. doi:10.1017/​S0954579407070228
Stifter, C. A., Putnam, S., & Jahromi, L. (2008). Exuberant and inhibited toddlers: Stability of temperament
and risk for problem behavior. Development and Psychopathology, 20, 401–​421. doi:10.1017/​
S0954579408000199
Sulik, M. J., Blair, C., Mills-​Koonce, R., Berry, D., Greenberg, M., & The Family Life Project Investigators.
(2015). Early parenting and the development of externalizing behavior problems: Longitudinal
mediation through children’s executive function. Child Development, 86, 1588–​1603. doi:10.111/​
cdev.12386.
Telzer, E. H., Masten, C. L., Berkman, E. T., Lieberman, M. D., & Fuligni, A. J. (2011). Neural regions
associated with self control and mentalizing are recruited during prosocial behaviors towards the
family. Neuroimage, 58, 242–​249. doi:10.1016/​j.neuroimage.2011.06.013
Valiente, C., Eisenberg, N., Smith, C. L., Reiser, M., Fabes, R. A., Losoya, S., … Murphy, B. C. (2003). The
relations of effortful control and reactive control to children’s externalizing problems: A longitudinal
assessment. Journal of Personality, 71, 1171–​1196. doi:10.1111/​1467-​6494.7106011
Valiente, C., Eisenberg, N., Spinrad, T. L., Reiser, M., A., Cumberland, Losoya, S., & Liew, J. (2006).
Relations among mothers’ expressivity, children’s effortful control, and their problem behaviors: A four-​
year longitudinal study. Emotion, 6, 459–​472. doi:10.1037/​1528-​3542.6.3.459
Verstraeten, K., Vasey, M. W., Raes, F., & Bijttebier, P. (2009). Temperament and risk for depressive
symptoms in adolescence: Mediation by rumination and moderation by effortful control. Journal of
Abnormal Child Psychology, 37, 349–​361. doi:10.1007/​s10802-​008-​9293-​x
Wang, F. L., Chassin, L., Eisenberg, N., & Spinrad, T. L. (2015). Effortful control predicts adolescent
antisocial-​aggressive behaviors and depressive symptoms: Co-​occurrence and moderation by
impulsivity. Child Development, 86, 1812–​1829. doi:10.111/​cdev.12406.
Wang, F. L., Chassin, L., Geiser, C., & Lemery-​Chalfant, K. (2016). Mechanisms in the relation between
GABRA2 and adolescent externalizing problems. European Child and Adolescent Psychiatry, 25, 67–​80.
doi:10.1007/​s00787-​015-​0703-​7.
Wang, M. T., Brinkworth, M., & Eccles, J. (2013). Moderating effects of teacher-​student relationship in
adolescent trajectories of emotional and behavioral adjustment. Developmental Psychology, 49, 690–​705.
doi:10.1037/​a0027916
24

42 The Relation of Self-Regulation to Children's Externalizing and Internalizing Problems

White, B. A., Jarrett, M. A., & Ollendick, T. H. (2012). Self-​regulation deficits explain the link between
reactive aggression and internalizing and externalizing behavior problems in children. Journal of
Psychopathology and Behavioral Assessment, 35, 1–​9. doi:10.1007/​s10862-​012-​9310-​9
White, L. K., McDermott, J. M., Degnan, K. A., Henderson, H. A., & Fox, N. A. (2011). Behavioral
inhibition and anxiety: The moderating roles of inhibitory control and attention shifting. Journal of
Abnormal Child Psychology, 39, 735–​747. doi:10.1007/​s10802-​011-​9490-​x
Xu, Y., Farver, J. M., & Zhang, Z. (2009). Temperament, harsh and indulgent parenting, and Chinese
children’s proactive and reactive aggression. Child Development, 80, 244–​258. doi:10.1111/​
j.1467-​8624.2008.01257.x
Yap, M. B. H., Allen, N. B., O’Shea, M., Di Parsia, P., Simons, J. G., & Sheeber, L. (2011). Early adolescents’
temperament, emotion regulation during mother–​child interactions, and depressive symptomatology.
Development and Psychopathology, 23, 267–​282. doi:10.1017/​S0954579410000787
Zalewski, M., Lengua, L. J., Wilson, A. C., Trancik, A., & Bazinet, A. (2011). Emotion regulation
profiles, temperament, and adjustment problems in preadolescents. Child Development, 82, 951–​966.
doi:10.1111/​j.1467-​8624.2011.01575.x
Zhou, Q., Eisenberg, N., Wang, Y., & Reiser, M. (2004). Chinese children’s effortful control and
dispositional anger/​frustration: Relations to parenting styles and children’s social functioning.
Developmental Psychology, 40, 352–​366. doi:10.1037/​0012-​1649.40.3.352
Zhou, Q., Lengua, L. J., & Wang, Y. (2009). The relations of temperament reactivity and effortful control to
children’s adjustment problems in China and the United States. Developmental Psychology, 45, 724–​739.
doi:10.1037/​a0013776
Zhou, Q., Main, A., & Wang, Y. (2010). The relations of temperamental effortful control and anger/​
frustration to Chinese children’s academic achievement and social adjustment: A longitudinal study.
Journal of Educational Psychology, 102, 180–​196. doi:10.1037/​a0015908
34

Chapter 3

Biological and Physiological Aspects


of Emotion Regulation
Kateri McRae & Michelle Shiota

Adaptive emotion responses


Theory and research on emotions widely recognize that emotional responses are adaptive, and
often useful (Keltner & Gross, 1999; Levenson, 1994). Even negative emotions, though unpleas-
ant in the moment, evolved to prepare our minds and bodies to handle important situations
quickly and efficiently (Öhman & Mineka, 2001; Parrott, 1993). However, we sometimes produce
emotional responses that are problematic for the situation that we are in making us needlessly
unhappy or causing more harm than good. Emotions that are mismatched to our needs in terms
of magnitude, type, or duration therefore, benefit from being regulated (Gross, 2015).
Research on emotion regulation has focused primarily on describing and manipulating the
various strategies people employ to change their emotions, and measuring the outcomes associ-
ated with the use of these strategies. Biological measures have played an important role in this
work, both documenting the relative effectiveness and strategy-╉specific consequences of various
emotion regulation strategies in a more objective manner, and helping to elucidate the psycho-
logical mechanisms supporting emotion regulation as a process. In this chapter we emphasize the
importance of differentiating these two applications of biological measures in emotion regulation
science, and review the empirical literature relevant to each. We then consider the contributions
of each approach to the important task of comparing and contrasting different emotion regulation
strategies.

Emotion regulation: definition and research methods


Emotion regulation consists of “the processes by which individuals influence which emotions they
have, when they have them, and how they experience and express these emotions” (Gross, 1998a,
p. 275). Emotion regulation can be used to up-╉or down-╉regulate positive or negative emotion, or
to switch from one kind of emotional response to another (Gross, 1998b). However, the bulk of
the empirical literature has focused upon emotion regulation for the purpose of down-╉regulating
negative emotion, as this work provides the most obvious bridge to clinical application. Although
co-╉regulation of emotion between two interacting individuals has long been of interest in devel-
opmental psychology (Cole, Martin, & Dennis, 2004; Evans & Porter, 2009), and is now a cutting-╉
edge interest in research on adult close relationships as well (Butler & Randall, 2013), the great
majority of research has examined intra-individual emotion regulation processes and outcomes.
Much research on emotion regulation has focused upon delineating the different emotion regu-
lation strategies that individuals commonly use (Gross, 1998; Gross, 2015, McRae 2016). These
strategies are then compared and contrasted to determine which are associated with positive and
negative short-╉term and long-╉term emotional outcomes. Using observational methods, studies
have often examined how frequently individuals use different strategies (Aldao, Nolen-╉Hoeksema,
4

44 Biological and Physiological Aspects of Emotion Regulation

& Schweizer, 2010; Gross & John, 2003). In experimental designs, individuals are taught to use
different strategies in the laboratory, and the success of each strategy is determined based on the
degree to which emotion measures are influenced by its use (Gross, 1998; Gross & Levenson,
1993; Jackson, Malmstadt, Larson, & Davidson, 2000). Emotion outcome measures commonly
include self-╉reported affect (how emotional people report feeling at any point in time), facial
expressions of emotion, measures of peripheral psychophysiology that reflect bodily responses to
emotion, and functional signals from brain regions thought to be involved in emotion, such as the
amygdala, insula and nucleus accumbens (McRae, 2016). The changes in these different measures
are not always coordinated (Mauss, Levenson, McCarter, Wilhelm, & Gross, 2005), but concur-
rent changes in multiple measures of emotion are taken as convergent evidence that successful
emotion regulation has occurred.
Emotion regulation strategies can be organized according to the point in the emotion genera-
tion process in which they are enlisted. Referred to as the process model of emotion regulation
(Gross, 1998; Gross, 2015), this organizational framework highlights five broad categories of emo-
tion regulation strategies, those that implement regulation by: 1) situation selection, 2) situation
modification, 3) attentional deployment, 4) cognitive restructuring, and 5) response modulation.
Comparison of specific strategies across these categories has proven to be extremely fruitful, but
studies have contrasted strategies within categories as well. Because much of the empirical litera-
ture has focused upon cognitive reappraisal, it is discussed in the greatest detail below.
Cognitive reappraisal refers to attempts to reconsider, reframe, or gain new perspective on
an emotional situation in a way that changes its emotional meaning (Giuliani & Gross, 2009).
Cognitive interventions, including cognitive therapy and cognitive behavioral therapy, refer to
cognitive reappraisal as cognitive reframing (Beck & Dozois, 2011). Reappraisal itself is a nod to
the fact that our emotions are a downstream consequences of interpreting what a given stimulus
means for our goals and well-╉being (Arnold, 1960; Frijda, 1988; Lazarus & Folkman 1984; Scherer,
1997). The same objective situation or stimulus can thus evoke different emotions, depending on
whether it is appraised as consistent with, or inconsistent with, our goals. Reappraisal involves
replacing or supplementing one’s initial (often negative) interpretation of a situation’s meaning
with another meaning (often, though not always, neutral or even positive; McRae 2016). For
example, if one initially appraises the end of a romantic relationship as a loss, evoking sadness,
one might regulate this emotion in part by thinking of the benefits associated with being single
(not having to consult another person about meals, regaining activities given up during the rela-
tionship), or even looking forward to the possibility of finding a new partner. Reappraisal involves
still thinking about the emotion eliciting event, but reframing the event’s perceived meaning and
implications, and therefore the emotions that follow.

Methods in emotion regulation research


A rich observational literature has documented correlations between the use of dispositional emo-
tion regulation strategies and various outcome measures, such as depressive symptomatology and
well-╉being (Aldao et al., 2010; Gross & John, 2003). This work relies primarily on self-╉report mea-
sures of how often participants use a variety of strategies. These studies have been valuable in sug-
gesting that, at the level of individual differences, the implications of all regulation strategies are
not alike. However, correlational designs and trait-╉level self-╉report measures are not well-╉suited to
identifying the more immediate consequences of using these strategies, or the mechanisms by which
strategies have their effects. Experimental methods are better suited to establishing how successful
different strategies are at achieving their emotional goals (e.g., Gross, 1998a; McRae et al, 2010;
Shiota & Levenson, 2012), what cognitive and social “side effects” each strategy might have (e.g.,
54

Emotion regulation outcomes: Measures of regulation success 45

Butler et al., 2003), and what psychological mechanisms are needed to employ each strategy (e.g.,
Kanske, Heissler, Schönfelder, Bongers, & Wessa, 2011; McRae et al., 2010; Ochsner et al., 2004).
One design commonly used to investigate the effects of different emotion regulation strategies
has been to instruct participants to use a particular emotion regulation strategy, and measure one
or more dependent variables while that strategy is being used during some emotion-​eliciting task.
This can be done in a between-​subjects design, randomly assigning some participants to one strat-
egy (for example, expressive suppression) and others to a no-​regulation control condition, and/​or
another regulation condition (for example, cognitive reappraisal). Where feasible, within-​subjects
designs in which participants are asked to alternate between multiple strategies, or between regu-
lation and control instructions, provide additional statistical power.
With these designs, biological measures can be used in two ways. First, they can be used to doc-
ument the extent to which some emotion regulation strategy is effective in producing the desired
emotional change. Strong negative emotions, in particular, commonly include “fight-​flight” sym-
pathetic nervous system responses that can be detected peripherally with non-​invasive sensors
(Cacioppo, Berntson, Larsen, Poehlmann, & Ito, 2000; Kreibig, 2010). Stressors—​especially social
stressors—​lasting longer than 15–​20 minutes evoke a rise in cortisol that can be detected in saliva
(Kirschbaum, Pirke, & Hellhammer, 1993). Emotional stimuli also elicit increased activity in neu-
ral structures such as the amygdala (Costafreda, Brammer, David, & Fu, 2008; Phelps & LeDoux,
2005), insula (Kurth, Zilles, Fox, Laird, & Eickhoff, 2010; Phan, Wager, Taylor, & Liberzon, 2002),
and anterior cingulate cortex (Etkin, Egner, & Kalisch, 2011; Phan et al., 2002). In a typical study,
where the goal of emotion regulation is to reduce the intensity of emotional distress, such mea-
sures provide a more objective index of regulation success than subjective self-​reports of affect.
These uses of biological measures are outcome-​focused—​measures of emotion regulation success.
Biological measures can also be used to better characterize the process of regulating emotion.
For example, a growing body of research documents the activation of brain structures known to
mediate cognitive control, such as the prefrontal cortex, while people are engaged in instructed
cognitive reappraisal (Buhle et al., 2013). Rather than documenting success in altering emotional
experience, these studies indicate that effortful cognitive control is required to implement the
instructed strategy. Although somewhat more controversial, measures of parasympathetic ner-
vous system influence on the heart (e.g., respiratory sinus arrhythmia, high-​frequency heart rate
variability) have also been linked to emotion regulation effort (Butler, Wilhelm, & Gross, 2006).
These uses of biological measures are process-​focused, capturing emotion regulation as a psycho-
logical experience in its own right. In the sections below, we offer more detailed explanations of
the biological measures used in each type of question, as well as offering examples from the exist-
ing literature of how these approaches have been applied.

Emotion regulation outcomes: Measures of regulation


success
Studies asking whether a given emotion regulation strategy is successful in reducing participants’
distress require outcome measures capturing intensity of the emotional experience. Most com-
monly, self-​reports of emotion are used to examine changes in subjective affect. Many studies use
a single item asking participants to rate their feelings either on a single valence: (How negative
do you feel?) or a bivalent scale (Bradley & Lang, 1994) or rating dial (Butler, Wilhelm, & Gross,
2006) ranging from negative to positive, while others use separate questions to ask about positive
and negative emotion (McRae, Ciesielski, & Gross, 2012; Shiota & Levenson, 2012) or a larger bat-
tery of questions to examine effects on specific emotions (Troy & Mauss, 2011). On the one hand,
an emotion regulation strategy should certainly be expected to improve one’s subjective feelings
64

46 Biological and Physiological Aspects of Emotion Regulation

if it is effective. However, self-╉report scales are used by participants in highly idiosyncratic ways
(Clark & Watson, 1995), and can be susceptible to bias associated with demand effects (Watson &
Vaidya, 2003). For these reasons, convergent evidence from more objective measures of emotional
reactivity is often desirable.

Measuring expressive behavior


In some studies of emotion regulation, behavioral measures of emotion are used. Most commonly,
facial expression data are used to measure emotional responding (Gross, 1998; Gross & Levenson,
1993), especially when studying strategies that target the facial expression of emotion. Facial
expression data can be collected via video cameras in the lab, and coded according to the facial
action coding system (FACS; Ekamn & Friesen, 1978), or more global coding of facial expressive
behaviors (Gross & Levenson, 1993). Electromyography can also be used to measure subtle acti-
vation of muscles of facial expression that may not be perceived by the naked eye (Lang, 2009 &
Bradley,), or eyetracking can be used to precisely track the parts of emotional images that cap-
ture and keep visual attention (Kellough, Beevers, Ellis, & Wells, 2008; Urry, 2010). In addition,
some researchers have used body posture as indicative of emotional responding (Riskind & Gotay,
1982; Shafir, Taylor, Atkinson, Langenecker, & Zubieta, 2013), and documented how this might
change in response to the use of emotion regulation strategies (Woodward et al., 2015).

Peripheral measures of sympathetic nervous system activity


From William James’ (James, 1884)  early proposal that emotional feelings reflect awareness of
visceral responses to environmental stimuli, to laboratory research on emotion in the present
day, physiological changes such as increased heart rate, peripheral vasoconstriction, and increase
in the skin’s electrical conductivity have been used as indicators of emotion. Walter Cannon’s
(Cannon, 1929) research offered a major advance in understanding the role of the sympathetic
branch of the autonomic nervous system (ANS) in mediating mammalian behavior in fear and
rage. Sympathetic activation causes increased heart rate; increased cardiac contractility (i.e., pre-╉
ejection period) and output (i.e., volume of blood pumped per minute); increased overall arte-
rial resistance to blood flow (resulting in higher blood pressure); peripheral vasoconstriction in
particular, leading to chilly hands and feet; increased respiration rate and depth; a burst of sweat
gland activity leading to an increase in skin electrical conductivity; piloerection; and pupil dila-
tion, among other effects.
The term “sympathetic” or “same pathway” refers to a long-╉held assumption that, due to the
structure of this branch of the ANS, all of these effects should tend to happen together (Cannon,
1929). The prototypical, full-╉blown “fight-╉flight” response presumably serves to facilitate intense
physical movement in a high-╉stress situation, increasing oxygen intake and carbon dioxide
release, circulating blood more quickly and effectively, and preferentially delivering resources to
the large skeletal muscles and brain. Reflecting this assumption, a large number of studies have
examined one of these measures, or a composite of several of these measures, as an index of
emotion regulation success while people view negative emotion stimuli, re-╉live a distressing per-
sonal experience, or complete a stressful task. If an emotion regulation strategy is successful in
reducing distress during said task, then this should manifest in a reduced sympathetic nervous
system response. A large body of research finds that cognitive reappraisal is effective at reduc-
ing physiological reactivity during negative emotion tasks (Gross, 1998; Jackson et al., 2000; Ray,
McRae, Ochsner, & Gross, 2010), especially in comparison to suppressing emotional expression
(e.g., Butler et al., 2003). Although the great majority of this research has focused on alleviating
negative emotion, at least one study has demonstrated comparable effects of reappraisal aimed at
74

Emotion regulation outcomes: Measures of regulation success 47

up-╉and down-╉regulating amusement (Giuliani, McRae, & Gross, 2008), an emotional state that is
also characterized by heightened physiological arousal (Kreibig, 2010). Developments in psycho-
physiology increasingly demonstrate, however, that the sympathetic nervous system is not quite
that simple. Although the neurons carrying sympathetic messages to the various visceral organs
do tend to travel together, they use at least two different neurotransmitters, and there is differen-
tiation of receptors as well. At a minimum, it has proved important to distinguish among effects
mediated by alpha-╉adrenergic receptors, including vasoconstriction and piloerection and among
effects mediated by beta-╉adrenergic receptors, including cardiac effects; and effects mediated by
cholinergic receptors, including sweat gland activity (Stemmler, 2003). Each major receptor type
has subtypes as well, with consequences that are still under investigation. The physiological effects
mediated by these various mechanisms often diverge, and in patterns that correspond to different
psychological states (Kreibig, 2010; Shiota, Neufeld, Yeung, Moser, & Perea, 2011).
Recognition of this diversity and fine-╉tuning of sympathetic responses is recharging research
on autonomic specificity, or the extent to which different emotions show different physiological
profiles, and has proved important for emotion regulation research as well. For example, beta-╉
adrenergic receptor-╉mediated cardiac effects (i.e., increased heart rate and cardiac output, short-
ened pre-╉ejection period) are seen when people engage in a difficult task, regardless of whether
they’ve been instructed to appraise that task as a threat or a challenge (Tomaka, Blascovich, Kibler,
& Ernst, 1997). With threat appraisal (wherein instructions emphasize task difficulty), increased
cardiac activity is accompanied by an increase in total vasoconstriction, consistent with a com-
prehensive sympathetic response. With challenge appraisal (wherein task instructions encourage
participants to “think of the task as a challenge” and of “yourself as someone capable of meeting
that challenge”), the increase in cardiac activity is intensified, yet accompanied by a reduction
in peripheral vasoconstriction, suggesting withdrawal of sympathetic influence on the arteries.
This discovery has paved the way for a rich body of research documenting threat vs. challenge
responses in a variety of important situations, as well as the benefits of challenge appraisal as
a strategy for regulating emotion in the face of stress (e.g., (Mendes, Reis, Seery, & Blascovich,
2003). As we learn more about the diversity of ANS responses that can accompany emotions
and their regulation, research may uncover similar kinds of distinctions in the effects of specific
regulation strategies.

Cortisol response to stress


Effects of emotion on measures of ANS activity can be detected within seconds. Longer-╉term
stressors, especially social stressors, may often evoke an increase in the hormone cortisol as well
(Dickerson & Kemeny, 2004; Kirschbaum, Pirke, & Hellhammer, 1993). Cortisol produces some
of the same peripheral effects as sympathetic nervous system activation, as well as a variety of
additional effects, but is much slower to onset and can last much longer as well. Cortisol is easily
measured through saliva, which can be assayed by a number of external labs. A number of stud-
ies have examined the implications of trait (e.g., Lam, Dickerson, Zoccola, & Zaldivar, 2009) and
experimentally instructed (Stansbury & Gunnar, 1994) use of various emotion regulation strate-
gies for cortisol reactivity during a stressful task.
However, some caveats are in order regarding use of cortisol as a measure of emotion regula-
tion success. One is that cortisol shows a diurnal (daily) rhythm, typically peaking around the
time that we wake up and then dropping over the course of the day and night (Stone et al., 2001).
Because this diurnal rhythm can confound the effects of laboratory manipulations, it is important
to run all study participants at around the same time of day. Also, people who have experienced
chronic stress may show blunted cortisol reactivity, observable in both flattened diurnal slope
84

48 Biological and Physiological Aspects of Emotion Regulation

(with blunted morning peaks but higher levels throughout the day) and reduced reactivity to
stress tasks (Elzinga et al., 2008; Miller, Chen, & Zhou, 2007). As a result, it can be difficult to
interpret individual differences in people’s cortisol responses to some tasks—╉do smaller responses
indicate successful regulation, or an individual who is so depleted that they no longer show a
healthy response to a challenging situation? Given these developments, caution is needed in inter-
preting cortisol reactivity data while these issues are being worked out.

Neural measures of emotional reactivity


Finally, many studies have examined the effects of emotion regulation on neural measures of
emotion. In these studies, the most commonly studied region is the amygdala. The amygdala is
a subcortical, bilateral structure known to be necessary for the formation of new, emotionally-╉
enhanced memories, such as those that are formed during fear conditioning (LaBar, Gatenby,
Gore, LeDoux, & Phelps, 1998; Phelps, 2006). It is frequently activated during neuroimaging stud-
ies of negative affect, induced by pictures, videos, words or autobiographical memory prompts.
Although the amygdala also responds to positive stimuli (Anderson et al., 2003; Cunningham,
Van Bavel, & Johnsen, 2008), in the context of regulation of negative stimuli, its deactivation is
usually thought to index successful down-╉regulation of negative emotion, such as during cogni-
tive reappraisal (Buhle et al., 2013).
The amygdala is not the only neural region thought to respond to negative emotion in the brain,
and to be down-╉regulated by cognitive reappraisal. Activation of the insula, which is thought to
relate to interoception and bodily responses during emotional experience, and the visual cortex
are also successfully down-╉regulated by successful reappraisal (Buhle et al., 2013). In addition,
studies using scalp surface electrodes to measure event-╉related potentials (ERPs) to emotional
stimuli have examined the effect of emotion regulation on the late positive potential (LPP),
which is thought to index the intensity or personal salience of an emotional stimulus. Studies of
instructed reappraisal have documented diminished LPPs in accordance with successful down-╉
regulation via reappraisal (Dennis & Hajcak, 2009). Finally, activation in reward-╉related regions
such as the nucleus accumbens can be regulated using reappraisal (Kim & Hamann, 2007), for
example, activation here can be down-╉regulated by using reappraisal to re-╉think cravings for sub-
stances of abuse (Kober, Kross, Mischel, Hart, & Ochsner, 2010).

Emotion regulation process: regulation effort


We have focused so far on the use of biological measures to index changes in emotion state that
are in accordance with the regulatory goal. In addition, however, bodily and neural changes can
be used to index regulatory processes that are brought online during attempts to change emo-
tions. In this work, biological measures capture the process, rather than the result, of emotion
regulation. For the most part, these measures serve as indices of effortful control rather than
processes specific to emotion regulation. Subjective measures of difficulty or effort indicate that
most individuals find following instructions to regulate more difficult than responding naturally
to an emotional stimulus (McRae et al., 2010). In addition, several studies have documented the
cognitive costs of ongoing emotion regulation effort by measuring performance on a concurrent
task (e.g., Anderson, 2013), with impaired performance on that task indicating competition with
emotion regulation for cognitive resources. Similarly, studies of ego depletion effects have shown
that engaging in effortful emotion regulation can lead to disruption of performance on subse-
quent self-╉regulatory tasks (Gailliot et al., 2007).
94

Emotion regulation process: regulation effort 49

Parasympathetic influence on the heart


One biological variable that has been used to index emotion regulation effort is vagal parasympa-
thetic influence on the heart (Porges, 2007). In an effect opposite to that of sympathetic influence,
parasympathetic nervous system (PNS) influence slows heart rate down. Because this effect is
blocked by respiratory intake, the slowing is absent while people inhale, and present while they
exhale; the resulting change in heart rate over the course of the respiratory cycle can be used as
an operational measure of parasympathetic influence on the heart. The first step in data process-
ing is to calculate the intervals in milliseconds between each successive heartbeat, over some
epoch of interest (there should be at least 30 seconds, preferably 60, to provide a stable estimate).
Most studies then use a method such as Fourier transform to decompose the interbeat inter-
val time-╉series into different frequency ranges, and calculate the amount or proportion of heart
rate variability that can be accounted for by frequencies typical of breathing (typically .12–╉.40
Hz). Commonly used measures include RSA, HF-╉HRV, and the ratio of high-╉frequency to low-╉
frequency heart rate variability.
Early studies identified a correlation between parasympathetic influence on the heart at rest, or
“vagal tone,” and a variety of indices of dispositional emotion regulation ability. This effect is now
well established by a rich body of research (Appelhans & Luecken, 2006). The effect is present
in children as well as adults; for example, studies with middle-╉school and high-╉school students
both found that those with conduct disorder showed lower resting baseline RSA than controls
(Beauchaine, Gatzke-╉Kopp, & Mead, 2007). Higher vagal tone has also been found to predict
spontaneous use of reappraisal while participants watched a distressing film clip (Volokhov &
Demaree, 2010). Another study found that the relationship between resting vagal tone and subjec-
tive well-╉being was mediated by habitual use of emotion regulation strategies requiring executive
cognitive control, but not by less effortful strategies (Geisler, Vennewald, Kubiak, & Weber, 2010).
Documenting an in-╉the-╉moment effect of emotion regulation on parasympathetic influence on
the heart has proved trickier; vagal withdrawal is commonly observed during strong negative
emotion (Kreibig, 2010), so it can be difficult to tease effects linked to emotional reactivity and
regulation apart. However, a growing body of research suggests that baseline-╉to-╉trial increases in
RSA during a negative emotion task are likely to indicate effortful emotion regulation in progress.
For example, Butler and colleagues (Butler et al., 2006) observed increases in participants’ RSA
during instructed reappraisal and expressive suppression, consistent with proposals that both
strategies require cognitive control.
More recently, studies have shown that greater increases in RSA during laboratory negative
emotion tasks are associated with lower levels of trait-╉level emotional dysfunction consistent
with deficits in regulation. For example, Austin and colleagues (Austin, Riniolo, & Porges, 2007)
found that RSA increased over the course of several emotion-╉eliciting films in healthy controls,
but dropped in patients with borderline personality disorder. Di Simplicio and colleagues (Di
Simplicio et al., 2012) showed that neuroticism interacted with image type in predicting RSA
trajectories while participants viewed sets of images; those low on neuroticism showed greater
increases in RSA during unpleasant pictures relative to neutral pictures, consistent with greater
regulatory effort in the former condition, whereas those high on neuroticism showed the opposite
pattern. These effects can be detected as early as preschool; Hastings and colleagues (Hastings et
al., 2008) found that children’s RSA increases during a laboratory-╉based social challenge signifi-
cantly predicted both internalizing and externalizing symptoms. In each case, normative baseline-╉
to-╉trial increases in RSA during some aversive task were less pronounced, or even absent, among
individuals who also showed trait-╉level signs of difficulty regulating emotions.
05

50 Biological and Physiological Aspects of Emotion Regulation

Two important caveats are needed in conducting and interpreting research using RSA and HRV
as operational measures of emotion regulation as a process. First, as noted earlier, these mea-
sures serve as markers for effortful control processes that generalize to a variety of tasks requiring
executive function or cognitive control (Hansen, Johnsen, & Thayer, 2009; Segerstrom & Nes,
2007; Thayer, Hansen, Saus-╉Rose, & Johnsen, 2009). This means not only that these measures are
not specific to emotion regulation, but also that they are only likely to capture emotion regulation
processes that require effortful control. Second, there is ongoing debate over whether it is nec-
essary and appropriate to control for respiration rate and depth when analyzing RSA/╉HF-╉HRV
data, as changes in breathing have their own effects on vagal activation (Porges, 2007). Although
this step is commonly recommended, it should be done with an eye to theory, and the extent to
which breathing might itself reflect or covary with a psychological process of interest. Consider
that “taking a deep breath” is, in fact, a common emotion regulation strategy; by controlling for
this behavior one may actually wipe out covariation linking regulatory effort to RSA or HF-╉HRV.
Our recommendation is to report analyses both with and without controlling for respiration, and
to consider these issues carefully when interpreting findings.

Associations with neural activation


The same complexities that impact the interpretation of RSA/╉HRV are also true of responding in
neural regions thought to index negative emotion and support regulatory processes. Early reports
of decreased amygdala responding during cognitive reappraisal, for example, were questioned
with regard to whether the decreased activation was due to the completion of any effortful sec-
ondary task, rather than reflecting the affective consequences of regulation (Ochsner, Bunge,
Gross, & Gabrieli, 2002; Schaefer et  al., 2002). However, these concerns can be addressed by
instructing individuals to use the same regulation strategy to achieve opposing emotional goals
(Kim & Hamann, 2007; Ochsner et al., 2004). Any measure of regulatory effort should be elevated
in all regulation conditions compared to a non-╉regulation baseline. By contrast, regions reflecting
the successful regulation of emotion should change in accordance with the regulation goal. To
address the question of changes in amygdala responding during reappraisal, this was resolved by
instructing participants to use reappraisal to decrease or increase negative affect (Ochsner et al.,
2004). This instruction resulted in decreased and increased amygdala activation respectively. This
pattern was consistent with amygdala modulation reflecting the successful regulation of emotion
rather than the effort exerted during regulatory processing.
In addition, in these same early studies, a network of prefrontal regions was activated dur-
ing the use of reappraisal to decrease and increase negative affect. This is strong evidence that
these are the neural correlates of the regulatory processes involved in reappraisal, rather than an
emotional consequence of reappraisal (Ochsner et al., 2004). These regions are highly overlap-
ping with those engaged during cognitive control more broadly, and include: Dorsolateral regions
thought to implement higher-╉order goals, ventrolateral regions thought to be involved in keep-
ing instructions in mind as well as generating verbal narratives as part of the re-╉interpretation,
midline regions implicated in attention and awareness of one’s own emotional experience, and
posterior parietal regions involved in the allocation of attention in space (Buhle et al., 2013).
These prefrontal and parietal regions, implicated in cognitive control, are thought to do the
“heavy lifting” involved in cognitive reappraisal. The overlap between these regions and those
engaged during other types of cognitive control indicates that individual differences in emo-
tion regulation ability should be related to other cognitive skills, a prediction that is borne out
by behavioral evidence (McRae, Jacobs, Ray, John, & Gross, 2012; Schmeichel, Volokhov, &
Demaree, 2008).
15

Putting an emotion regulation framework to work 51

Putting an emotion regulation framework to work


This framework has been fruitful in better understanding different emotion regulation strate-
gies, as well as providing a consistent structure for studies that compare strategies. A  grow-
ing number of studies have compared two or more strategies, first in terms of their outcomes
(emotional, social, cognitive, or all three), and then in terms of the cognitive processes engaged
during their implementation. Based on these comparisons, researchers are able to form new
predictions about the circumstances under which, and people for whom, certain strategies
might be most adaptive.

Reappraisal vs. suppression


Comparing effects of regulation
One emotion regulation strategy to which reappraisal is frequently compared is expressive sup-
pression, or the outward inhibition of an internally felt emotion (Gross & Levenson, 1993).
Suppression is sometimes necessary as a last-╉ditch effort to regulate emotional expression in social
contexts (e.g., When a strong feeling of sadness about the end of a romantic relationship swells
during a business meeting, when crying would be seen as inappropriate). In Western contexts,
however, the use of suppression is generally thought to be less effective, helpful, and adaptive than
reappraisal. For example, individuals instructed to reappraise a disgusting film report successfully
decreasing subjective negative emotion, whereas those using expressive suppression report little
or no reduction in subjective distress (Gross, 1998). Those using reappraisal also show dimin-
ished physiological responding, as outlined above, in terms of reduced startle response (Jackson
et al., 2000; Ray et al., 2010), and responding in neural regions such as the amygdala (Ochsner
et al., 2002; Schaefer et al., 2002). By contrast, there is some evidence that suppression can “back-
fire” and actually increase physiological reactivity to negative stimuli, as measured by increases
in peripheral physiology and amygdala activation (Goldin, McRae, Ramel, & Gross, 2009; Gross,
1998; Roberts, Levenson, & Gross, 2008).
Beyond biology, there is evidence that, at least in Western contexts, suppression involves other
costs as well. Some costs are social, in the form of lower levels of social responsiveness and more
negative perceptions by social partners (Butler, Lee, & Gross, 2007). In addition, there are cogni-
tive costs to suppression relative to reappraisal. Multiple studies have demonstrated that indi-
viduals instructed to use suppression show impaired memory for the emotion-╉eliciting stimulus,
whereas those instructed to use reappraisal show improved memory (Dillon, Ritchey, Johnson, &
LaBar, 2007; Hayes et al., 2010; Richards & Gross, 2000).

Comparing processes recruited to regulate


As discussed above, reappraisal has been shown to recruit a wide network of cognitive-╉control
related brain regions, including the dorosolateral prefrontal cortex, ventrolateral prefrontal cor-
tex, medial prefrontal cortex, and bilateral parietal regions. Together, these regions support the
re-╉consideration of emotional meaning by representing the current emotional meaning (medial
PFC), re-╉directing attention (bilateral parietal corticies) and generating new narratives (ventro-
lateral PFC) that change in accordance with the broader emotion regulation goal (dorosolateral
PFC). Suppression, by contrast, is thought to be a more targeted cognitive control process, related
more specifically to the inhibition of facial expression. Studies comparing reappraisal and sup-
pression demonstrate greater prefrontal recruitment for reappraisal than suppression, especially
early in the emotion regulation process (the first five seconds of a 15-╉second film; (Goldin et al.,
2009). By contrast, regions engaged during suppression continue to ramp up over the regulation
25

52 Biological and Physiological Aspects of Emotion Regulation

period (Goldin et al., 2009). Finally, suppression also recruits the right inferior gyrus to a greater
extent than reappraisal, a region that is thought to be engaged during the inhibition of prepo-
tent responses more broadly (Garavan, Hester, Murphy, Fassbender, & Kelly, 2006; Rubia, Smith,
Brammer, & Taylor, 2003).

Reappraisal vs. distraction


Comparing effects of regulation
Reappraisal has also been compared to strategies involving attentional reallocation, such as dis-
traction. Behavioral research indicates that both internally-╉directed distraction (telling people to
think of something else; e.g., Sheppes & Meiran, 2007) and externally-╉directed distraction (giv-
ing people a concurrent task to complete; e.g., McRae et  al., 2010)  result in successful down-╉
regulation of subjective negative emotion experience. Within the brain, externally-╉directed
distraction reduces amygdala activation to a greater extent than reappraisal (McRae et al., 2010).
However, it is likely that specific contextual factors moderate the extent to which distraction vs.
reappraisal is most effective; this is a promising area for future research.

Comparing processes recruited to regulate


By and large, many of the cognitive control regions that are recruited during reappraisal are also
recruited during distraction. However, most of these regions demonstrate significantly stronger
activation during reappraisal than distraction (Kanske, Heissler, Schönfelder, Bongers, & Wessa,
2011; McRae et al., 2010). This suggests that reappraisal may require more “executive” cognitive
resources than distraction. Distraction more strongly recruits regions of a posterior attentional
system (bilateral parietal regions, amongst others) than reappraisal (Kanske et al., 2011; McRae
et al., 2010). Because attentional resources might be more quickly and less effortfully deployed
than a coordinated response across several cognitive control systems, distraction may be easier
and more effective in some ways than reappraisal. However, shifts in attention might be shorter-╉
lived than the shifts in meaning generated during reappraisal, requiring greater investment of
effort over time.
Behavioral research has confirmed the first of these predictions, showing that there are contexts
in which distraction may be more effective than reappraisal. Specifically, distraction might be more
effective if the time available for reappraisal is severely limited (Sheppes & Gross, 2011; Sheppes &
Meiran, 2007). Reappraisal requires time to generate possible reappraisals, select amongst them,
implement the selected reappraisal, and monitor for success (Ochsner & Gross, 2005). There is
also some evidence that distraction may be a better choice when the emotional stimulus is of a
very high intensity (Thiruchselvam et al., 2011). In this case, individuals may feel overwhelmed
by their initial negative response, and unable to engage the cognitive resources required to gener-
ate, select, and/╉or implement a high quality reappraisal. Attempting to reappraise while emotional
intensity runs high is more effortful (Silvers, Weber, Wager, & Ochsner, 2014), and less likely to
be successful, which might lead to frustration, and perhaps decreases an individual’s self-╉efficacy
beliefs regarding emotion regulation.
Despite these advantages, studies have also spoken to the second prediction above—╉that
the effects of reappraisal are longer lasting than those of distraction, carrying over into any
re-╉exposure to the reappraised stimulus rather than requiring new investment of effort. Thus,
distraction might be a good short-╉term strategy, but individuals who reappraise a situation are
better equipped if that situation is ongoing or arises again (Denny, Inhoff, Zerubavel, Davachi, &
Ochsner, 2015; Thiruchselvam et al., 2011). Taken together, these findings suggest that the relative
merits of reappraisal and distraction depend on context, and may change over time. When time is
35

Putting an emotion regulation framework to work 53

tight, or an emotional situation is of high intensity, distraction may be a great short-╉term strategy.
Once the individual has more time, more cognitive resources at his or her disposal, and/╉or the
initial intensity of the experience has diminished, reappraisal might be a good choice to ensure
more lasting effects.

Variety among reappraisal strategies


Comparing effects of regulation
Although reappraisal is often contrasted with other families of emotion regulation strategies, there
are also qualitatively different varieties of reappraisal. A number of recent studies have contrasted
detached reappraisal, or concentrating on non-╉emotion-╉eliciting aspects and implications of the
target situation, with positive reappraisal, concentrating on the situation’s positive aspects and
implications. For example, if one had just had a car accident, the thought “these things happen,
it’s not a big deal” would reflect detached reappraisal, whereas the thought “I wanted to get my
bumper replaced anyway, and now insurance will pay for it!” would reflect positive reappraisal.
Some studies have explicitly instructed participants in the use of these strategies (e.g., Shiota &
Levenson, 2012), whereas others used instructions emphasizing the goals of increasing positive
versus decreasing negative emotion, and measured the resulting strategy use (McRae, Ciesielski,
& Gross, 2012).
The handful of studies comparing these strategies suggest that positive reappraisal might be
more desirable than detached reappraisal, or reappraising with the intent to return to a neutral
point (McRae et al., 2012; Shiota & Levenson, 2012; Troy & Mauss, 2010). Both strategies are effec-
tive in reducing the intensity of subjective distress, but when positive and negative affect are mea-
sured separately, positive reappraisal is more likely to increase positive feelings (e.g., McRae et al.,
2012; Shiota & Levenson, 2012). Positive reappraisal also promotes greater smiling—╉not only an
expression of positive emotion, but also a key affiliative signal (Papa & Bonanno, 2008)—╉than
detached reappraisal (Shiota & Levenson, 2012). From a biological standpoint, multiple studies
have found that positive reappraisal leads to a higher level of autonomically-╉mediated physiologi-
cal arousal (e.g., cardiac reactivity, skin conductance) than detached reappraisal (McRae et al.,
Shiota & Levenson, 2012). Prior research on physiological activity associated with positive emo-
tion suggests that, rather than indicating a failure to regulate, this maintained arousal indicates
positive, approach-╉oriented engagement with the stimulus (Kreibig, 2010; Shiota et al., 2011).
Less data are available contrasting the effects of detached and positive reappraisal on neural
measures of emotional responding. One study instructed participants to increase their positive
emotion in response to a positive picture, finding that this increases amygdala responding (Kim
& Hamann, 2007). This effect is consistent with the proposal that amygdala activation can reflect
whichever valence of emotion is most salient at the time a stimulus is processed, not just distress
(Cunningham et al., 2008). This does not, however, speak to the effects of using positive reap-
praisal for the purpose of regulating one’s response to a negative stimulus. However, prelimi-
nary evidence suggests that although detached and positive reappraisal have comparable effects in
shortening the duration of amygdala responding, the intensity of amygdala activation is reduced
only by detached reappraisal (Waugh et al., in press 2016). These results parallel the peripheral
physiology findings above, with detached reappraisal resulting in greater down-╉regulation of bio-
logical markers of emotional engagement than positive reappraisal.

Comparing processes recruited to regulate


To date, most studies have compared the neural regions recruited during down-╉regulation of
negative responses to negative stimuli with those recruited in up-╉regulation of positive emotion
45

54 Biological and Physiological Aspects of Emotion Regulation

responses to positive stimuli, rather than contrasting the regions recruited in detached versus
positive reappraisal of negative stimuli. These studies have been helpful in demonstrating that up-╉
regulation of positive emotion and down-╉regulation of negative emotion activate similar neural
regions: Left-╉lateralized dorso-╉and ventro-╉lateral prefrontal cortex, midline prefrontal regions
and bilateral parietal regions. However, some subtle differences may exist. Specifically, in one
study the rostral medial prefrontal cortex (mPFC), associated with focusing on current affec-
tive experience, was recruited more during up-╉regulation of positive emotion than during down-╉
regulation of negative emotion (Waugh et al., in press 2016).
However, one study has directly compared implications of detached versus situational reap-
praisal of negative stimuli for neural activity (Ochsner et al., 2004). A region of the left lateral PFC
was recruited more strongly during instructed situational reappraisal, whereas an area of the right
medial PFC region was recruited more strongly during detached reappraisal. This distinction may
reflect the greater emphasis on reducing perceived self-╉relevance of the stimulus in detached reap-
praisal, in contrast with the greater emphasis on manipulating information about the outside
world in situational reappraisal (Ochsner et al., 2004). Much more information is needed about
the differential cognitive mechanisms recruited by detached versus situational reappraisal, and
neuroimaging studies are likely to prove valuable in this endeavor.

Conclusion
Studies of emotion regulation have compared and contrasted several emotion regulation strategies,
using multiple measures. Measures of self-╉reported affect, sympathetic responding, and neural
activation from regions thought to index emotional intensity have been used to evaluate the suc-
cess of various emotion regulation strategies. In addition, measures of subjective effort, RSA/╉HRV,
and engagement of neural regions have been used to characterize the effortful processes engaged
during regulation. This framework has allowed for the rich characterization of a number of emo-
tion regulation strategies, but has been particularly useful for documenting the success of cognitive
reappraisal, as well, inspiring new hypotheses about the contexts in which it works best. Compared
with other strategies, cognitive reappraisal is thought to be a relatively effective way to decrease
negative emotion, although it does require intact cognitive control resources. The importance of
cognitive reappraisal in various disorders can be seen in the various chapters in this volume.

References
Aldao, A., Nolen-╉Hoeksema, S., & Schweizer, S. (2010). Emotion-╉regulation strategies across
psychopathology: A meta-╉analytic review. Clinical Psychology Review, 30(2), 217–╉237. doi:10.1016/╉
j.cpr.2009.11.004
Anderson, A. K., Christoff, K., Stappen, I., Panitz, D., Ghahremani, D. G., Glover, G., … Sobel, N. (2003).
Dissociated neural representations of intensity and valence in human olfaction. Nature Neuroscience, 6,
196–╉202. doi:10.1038/╉nn1001
Appelhans, B. M., & Luecken, L. J. (2006). Heart rate variability as an index of regulated emotional
responding. Review of General Psychology, 10(3), 229.
Arnold, M. B. (1960). Emotion and personality. New York, NY: Columbia University Press.
Austin, M. A., Riniolo, T. C., & Porges, S. W. (2007). Borderline personality disorder and emotion
regulation: Insights from the polyvagal theory. Brain and Cognition, 65(1), 69–╉76.
Beck, A. T., & Dozois, D. J. (2011). Cognitive therapy: Current status and future directions. Annual Review
of Medicine, 62, 397–╉409.
Bradley, M. M., & Lang, P. J. (1994). Measuring emotion: The self-╉assessment manikin and the semantic
differential. Journal of Behavior Therapy: Experimental Psychiatry, 25(1), 49–╉59.
5

Conclusion 55

Buhle, J. T., Silvers, J. A., Wager, T. D., Lopez, R., Onyemekwu, C., Kober, H., … Ochsner, K. N. (2013).
Cognitive reappraisal of emotion: A meta-​analysis of human neuroimaging studies. Cerebral Cortex,
24(11), 2981–​2990.doi:10.1093/​cercor/​bht154
Butler, E. A., Egloff, B., Wlhelm, F. H., Smith, N. C., Erickson, E. A., & Gross, J. J. (2003). The social
consequences of expressive suppression. Emotion, 3(1), 48.
Butler, E. A., Lee, T. L., & Gross, J. J. (2007). Emotion regulation and culture: Are the social consequences
of emotion suppression culture-​specific? Emotion, 7(1), 30–​48.
Butler, E. A., & Randall, A. K. (2013). Emotional coregulation in close relationships. Emotion Review, 5(2),
202–​210.
Butler, E. A., Wilhelm, F. H., & Gross, J. J. (2006). Respiratory sinus arrhythmia, emotion, and emotion
regulation during social interaction. Psychophysiology, 43(6), 612–​622.
Cacioppo, J. T., Berntson, G. G., Larsen, J. T., Poehlmann, K. M., & Ito, T. A. (2000). The
psychophysiology of emotion. In M. Lewis & J. M. Haviland-​Jones (Eds.), The handbook of emotion.
New York: Guildford Press.
Cannon, W. B. (1929). Bodily changes in pain, hunger, fear and rage. Oxford, England: Appleton.
Clark, L. A., & Watson, D. (1995). Constructing validity: Basic issues in objective scale development.
Psychological assessment, 7(3), 309.
Cole, P. M., Martin, S. E., & Dennis, T. A. (2004). Emotion regulation as a scientific
construct: Methodological challenges and directions for child development research. Child
Development, 75(2), 317–​333.
Costafreda, S. G., Brammer, M. J., David, A. S., & Fu, C. H. (2008). Predictors of amygdala activation
during the processing of emotional stimuli: A meta-​analysis of 385 pet and fmri studies. Brain Research
Reviews, 58(1), 57–​70.
Cunningham, W. A., Van Bavel, J. J., & Johnsen, I. R. (2008). Affective flexibility: Evaluative
processing goals shape amygdala activity. Psychological Science, 19(2), 152–​160. doi:10.1111/​
j.1467-​9280.2008.02061.x
Dennis, T. A., & Hajcak, G. (2009). The late positive potential: A neurophysiological marker for emotion
regulation in children. J Child Psychol Psychiatry, 50(11), 1373–​1383. doi:JCPP2168 [pii] 10.1111/​
j.1469-​7610.2009.02168.x
Denny, B. T., Inhoff, M. C., Zerubavel, N., Davachi, L., & Ochsner, K. N. (2015). Getting over it long-​
lasting effects of emotion regulation on amygdala response. Psychological science, 26(9), 1377–​1388.
Di Simplicio, M., Costoloni, G., Western, D., Hanson, B., Taggart, P., & Harmer, C. (2012). Decreased
heart rate variability during emotion regulation in subjects at risk for psychopathology. Psychological
Medicine, 42(08), 1775–​1783.
Dickerson, S. S., & Kemeny, M. E. (2004). Acute stressors and cortisol responses: A theoretical integration
and synthesis of laboratory research. Psychological bulletin, 30(3), 355–​391.
Dillon, D. G., Ritchey, M., Johnson, B. D., & LaBar, K. S. (2007). Dissociable effects of conscious emotion
regulation strategies on explicit and implicit memory. Emotion, 7(2), 354–​365.
Ekamn, P., & Friesen, W. (1978). Facial action coding system (facs): Manual: Consulting
Psychologists Press.
Elzinga, B. M., Roelofs, K., Tollenaar, M. S., Bakvis, P., van Pelt, J., & Spinhoven, P. (2008). Diminished
cortisol responses to psychosocial stress associated with lifetime adverse events: A study among healthy
young subjects. Psychoneuroendocrinology, 33(2), 227–​237.
Etkin, A., Egner, T., & Kalisch, R. (2011). Emotional processing in anterior cingulate and medial prefrontal
cortex. Trends Cogn Sci, 15(2), 85–​93.
Evans, C. A., & Porter, C. L. (2009). The emergence of mother–​infant co-​regulation during the first
year: Links to infants’ developmental status and attachment. Infant Behavior and Development, 32(2),
147–​158.
Frijda, N. H. (1988). The laws of emotion. American Psychologist, 43(5), 349–​358.
65

56 Biological and Physiological Aspects of Emotion Regulation

Gailliot, M. T., Baumeister, R. F., DeWall, C. N., Maner, J. K., Plant, E. A., Tice, D. M., … Schmeichel, B.
J. (2007). Self-​control relies on glucose as a limited energy source: Willpower is more than a metaphor.
Journal of Personality and Social Psychology, 92(2), 325.
Garavan, H., Hester, R., Murphy, K., Fassbender, C., & Kelly, C. (2006). Individual differences in the
functional neuroanatomy of inhibitory control. Brain Research, 1105(1), 130–​142.
Geisler, F. C., Vennewald, N., Kubiak, T., & Weber, H. (2010). The impact of heart rate variability on
subjective well-​being is mediated by emotion regulation. Personality and Individual Differences, 49(7),
723–​728.
Giuliani, N. R., & Gross, J. J. (2009). Reappraisal. In D. Sander & K. R. Scherer (Eds.), Oxford companion to
the affective sciences (pp. 329–​330). New York: Oxford University Press.
Giuliani, N. R., McRae, K., & Gross, J. J. (2008). The up-​and down-​regulation of amusement: Experiential,
behavioral, and autonomic consequences. Emotion, 8(5), 714–​719. doi:10.1037/​a0013236
Goldin, P. R., McRae, K., Ramel, W., & Gross, J. J. (2009). The neural bases of emotion
regulation: Reappraisal and suppression of negative emotion. Biological Psychiatry, 65, 170–​180.
Gross, J. J. (1998). Antecedent-​and response-​focused emotion regulation: Divergent consequences for
experience, expression, and physiology. Journal of Personality and Social Psychology, 74(1), 224–​237.
doi:10.1037//​0022-​3514.74.1.224
Gross, J. J. (1998). The emerging field of emotion regulation: An integrative review. Review of General
Psychology, 2(3), 271–​299. doi:10.1037//​1089-​2680.2.3.271
Gross, J. J. (2015). Emotion regulation: Current status and future prospects. Psychological Inquiry, 26, 1–​26.
Gross, J. J., & John, O. P. (2003). Individual differences in two emotion regulation processes: Implications
for affect, relationships, and well-​being. Journal of Personality and Social Psychology, 85(2), 348–​362.
doi: 10.1037/​0022-​3514.85.2.348
Gross, J. J., & Levenson, R. W. (1993). Emotional suppression: Physiology, self-​report, and expressive
behavior. Journal of Personality and Social Psychology, 64(6), 970–​986.
Hansen, A. L., Johnsen, B. H., & Thayer, J. F. (2009). Relationship between heart rate variability and
cognitive function during threat of shock. Anxiety, Stress, & Coping, 22(1), 77–​89.
Hastings, P. D., Nuselovici, J. N., Utendale, W. T., Coutya, J., McShane, K. E., & Sullivan, C. (2008).
Applying the polyvagal theory to children’s emotion regulation: Social context, socialization, and
adjustment. Biological Psychology, 79(3), 299–​306.
Hayes, J. P., Morey, R. A., Petty, C. M., Seth, S., Smoski, M. J., McCarthy, G., & LaBar, K. S. (2010). Staying
cool when things get hot: Emotion regulation modulates neural mechanisms of memory encoding.
Frontiers in Human Neuroscience, 4, Article 230, 230–​240.
Jackson, D. C., Malmstadt, J. R., Larson, C. L., & Davidson, R. J. (2000). Suppression and enhancement
of emotional responses to unpleasant pictures. Psychophysiology, 37(4), 515–​522. doi:10.1111/​
1469-​8986.3740515
James, W. (1884). What is an emotion? Mind, 9, 188–​205.
Kanske, P., Heissler, J., Schönfelder, S., Bongers, A., & Wessa, M. (2011). How to regulate emotion? Neural
networks for reappraisal and distraction. Cerebral Cortex, 21(6), 1379–​1388. doi:10.1093/​cercor/​bhq216
Kellough, J. L., Beevers, C. G., Ellis, A. J., & Wells, T. T. (2008). Time course of selective attention in
clinically depressed young adults: An eye tracking study. Behaviour research and therapy, 46(11), 1238–​
1243. doi: http://​dx.doi.org/​10.1016/​j.brat.2008.07.004
Keltner, D., & Gross, J. J. (1999). Functional accounts of emotions. Cognition and Emotion, 13(5), 467–​480.
doi:10.1080/​026999399379140
Kim, S. H., & Hamann, S. (2007). Neural correlates of positive and negative emotion regulation. Journal of
Cognitive Neuroscience, 19(5), 776–​798. doi:10.1162/​jocn.2007.19.5.776
Kirschbaum, C., Pirke, K.-​M., & Hellhammer, D. H. (1993). The “trier social stress test”: A tool
for investigating psychobiological stress responses in a laboratory setting. Neuropsychobiology,
28(1–​2), 76–​81.
75

Conclusion 57

Kober, H., Kross, E. F., Mischel, W., Hart, C. L., & Ochsner, K. N. (2010). Regulation of craving by
cognitive strategies in cigarette smokers. Drug & Alcohol Dependence, 106(1), 52–​55. doi:S0376-​
8716(09)00309-​3 [pii] 10.1016/​j.drugalcdep.2009.07.017
Kreibig, S. D. (2010). Autonomic nervous system activity in emotion: A review. Biological Psychology, 84(3),
394–​421.
Kurth, F., Zilles, K., Fox, P. T., Laird, A. R., & Eickhoff, S. B. (2010). A link between the
systems: Functional differentiation and integration within the human insula revealed by meta-​analysis.
Brain Structure and Function, 214(5–​6), 519–​534.
LaBar, K. S., Gatenby, J. C., Gore, J. C., LeDoux, J. E., & Phelps, E. A. (1998). Human amygdala
activation during conditioned fear acquisition and extinction: A mixed-​trial fmri study. Neuron,
20(5), 937–​945. Retrieved from http://​www.sciencedirect.com/​science/​article/​B6WSS-​418PT65-​G/​2/​
6db719d8fc4eafdeb7c27f57d9c4cd48
Lam, S., Dickerson, S. S., Zoccola, P. M., & Zaldivar, F. (2009). Emotion regulation and cortisol reactivity
to a social-​evaluative speech task. Psychoneuroendocrinology, 34(9), 1355–​1362. Retrieved from http://​
www.sciencedirect.com/​science/​article/​B6TBX-​4WBR6MR-​1/​2/​aca407e23f5603e5dc44c5e721bf27ee
Lang, P., & Bradley, M. M. (2009). The international affective picture system (IAPS) in the study of emotion
and attention. In J. A. Coan & J. B. Allen (Eds.), Handbook of emotion elicitation and assessment,.
New York, NY: Oxford University Press.
Lazarus, R. S., & Folkman, S. (1984). Stress, appraisal, and coping. New York: Springer.
Levenson, R. W. (1994). Human emotion: A functional view. In P. Ekman & R. J. Davidson (Eds.), The
nature of emotion: Fundamental questions (pp. 123–​126). New York: Oxford University Press.
Mauss, I. B., Levenson, R. W., McCarter, L., Wilhelm, F. H., & Gross, J. J. (2005). The tie that binds?
Coherence among emotion experience, behavior, and physiology. Emotion, 5(2), 175–​190. Retrieved
from http://​www.apa.org http://​www.apa.org/​journals/​emo.html
McRae, K., Ciesielski, B., & Gross, J. J. (2012). Unpacking cognitive reappraisal: Goals, tactics, and
outcomes. Emotion, 12(2), 250–​255. doi:10.1037/​A0026351
McRae, K., Hughes, B., Chopra, S., Gabrieli, J. D., Gross, J. J., & Ochsner, K. N. (2010). The neural bases of
distraction and reappraisal. Journal of Cognitive Neuroscience, 22(2), 248–​262.
McRae, K., Hughes, B., Chopra, S., Gabrieli, J. D. E., Gross, J. J., & Ochsner, K. N. (2010). The neural
bases of distraction and reappraisal. Journal of Cognitive Neuroscience, 22(2), 248–​262. doi:10.1162/​
jocn.2009.21243
McRae, K., Jacobs, S. E., Ray, R. D., John, O. P., & Gross, J. J. (2012). Individual differences in reappraisal
ability: Links to reappraisal frequency, well-​being, and cognitive control. Journal of Research in
Personality, 46(1), 2–​7. doi:10.1016/​J.Jrp.2011.10.003
Mendes, W. B., Reis, H. T., Seery, M. D., & Blascovich, J. (2003). Cardiovascular correlates of
emotional expression and suppression: Do content and gender context matter? J Pers Soc Psychol, 84(4),
771–​792.
Miller, G. E., Chen, E., & Zhou, E. S. (2007). If it goes up, must it come down? Chronic stress and the
hypothalamic-​pituitary-​adrenocortical axis in humans. Psychological Bulletin, 133(1), 25.
Ochsner, K. N., Bunge, S. A., Gross, J. J., & Gabrieli, J. D. E. (2002). Rethinking feelings: An fmri study of
the cognitive regulation of emotion. Journal of Cognitive Neuroscience, 14(8), 1215–​1229. doi:10.1162/​
089892902760807212
Ochsner, K. N., & Gross, J. J. (2005). The cognitive control of emotion. Trends in Cognitive Sciences, 9(5),
242–​249. doi:10.1016/​j.tics.2005.03.010
Ochsner, K. N., Ray, R. D., Cooper, J. C., Robertson, E. R., Chopra, S., Gabrieli, J. D. E., & Gross, J. J.
(2004). For better or for worse: Neural systems supporting the cognitive down-​and up-​regulation of
negative emotion. Neuroimage, 23, 483–​499. doi:10.1016/​j.neuroimage.2004.06.030
Öhman, A., & Mineka, S. (2001). Fears, phobias, and preparedness: Toward an evolved module of fear and
fear learning. Psychological Review, 108(3), 483–​522.
85

58 Biological and Physiological Aspects of Emotion Regulation

Papa, A., & Bonanno, G. A. (2008). Smiling in the face of adversity: The interpersonal and intrapersonal
functions of smiling. Emotion, 8(1), 1.
Parrott, W. G. (1993). Beyond hedonism: Motives for inhibiting good moods and for maintaining bad
moods. In D. M. Wegner & J. W. Pennebaker (Eds.), Handbook of mental control (pp. 278–​305). Upper
Saddle River, NJ: Prentice Hall.
Phan, K. L., Wager, T. D., Taylor, S. F., & Liberzon, I. (2002). Functional neuroanatomy of
emotion: A meta-​analysis of emotion activation studies in pet and fmri. NeuoroImage, 16, 331–​348.
Phelps, E. A. (2006). Emotion and cognition: Insights from studies of the human amygdala. Annual Review
of Psychology, 57(1), 27–​53. doi:10.1146/​annurev.psych.56.091103.070234
Phelps, E. A., & LeDoux, J. E. (2005). Contributions of the amygdala to emotion processing: From animal
models to human behavior. Neuron, 48(2), 175–​187. Retrieved from http://​www.sciencedirect.com/​
science/​article/​B6WSS-​4HC6GJV-​8/​2/​1331df121450dfc81ccc3f30e982c49d
Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–​143.
Ray, R. D., McRae, K., Ochsner, K. N., & Gross, J. J. (2010). Cognitive reappraisal of negative
affect: Converging evidence from emg and self-​report. Emotion, 10(4), 587–​592. doi:10.1037/​a0019015
Richards, J. M., & Gross, J. J. (2000). Emotion regulation and memory: The cognitive costs of keeping one’s
cool. Journal of Personality and Social Psychology, 79(3), 410–​424.
Riskind, J. H., & Gotay, C. C. (1982). Physical posture: Could it have regulatory or feedback effects on
motivation and emotion? Motivation and Emotion, 6(3), 273–​298.
Roberts, N. A., Levenson, R. W., & Gross, J. J. (2008). Cardiovascular costs of emotion suppression cross
ethnic lines. International Journal of Psychophysiology, 70(1), 82–​87. Retrieved from http://​www.
sciencedirect.com/​science/​article/​B6T3M-​4STGRRV-​1/​2/​f27ee02477f8a4035febf7e602ba5a00
Rubia, K., Smith, A. B., Brammer, M. J., & Taylor, E. (2003). Right inferior prefrontal cortex mediates
response inhibition while mesial prefrontal cortex is responsible for error detection. Neuroimage, 20(1),
351–​358.
Schaefer, S. M., Jackson, D. C., Davidson, R. J., Aguirr, G. K., Kimberg, D. Y., & Thompson-​Schill, S. L.
(2002). Modulation of amygdalar activity by the conscious regulation of negative emotion. Journal of
Cognitive Neuroscience, 14, 913–​921.
Scherer, K. R. (1997). Profiles of emotion-​antecedent appraisal: Testing theoretical predictions across
cultures. Cognition & Emotion Vol 11(2) (Mar 1997): 113–​150, 11(2), 113–​150. Retrieved from http://​
firstsearch.oclc.org/​WebZ/​DARead?key=0269-​9931%28199703%2911%3A2%3C113%3APOEATT%3E
%26sp06sw02-​2100-​dym7j20i-​sabo1o%264588e776ee29fbb7264bb0c2bfe47a13ceb40a60d8eca260f8c56
399f336deb3&sessionid=0&db=ECO_​FT&format=URL
Schmeichel, B. J., Volokhov, R. N., & Demaree, H. A. (2008). Working memory capacity and the self-​
regulation of emotional expression and experience. Journal of Personality and Social Psychology, 95(6),
1526–​1540. Retrieved from 10.1037/​a0013345 http://​0-​search.ebscohost.com.bianca.penlib.du.edu/​
login.aspx?direct=true&db=psyh&AN=2008-​16429-​020&login.asp&site=ehost-​live
Segerstrom, S. C., & Nes, L. S. (2007). Heart rate variability reflects self-​regulatory strength, effort, and
fatigue. Psychological Science, 18(3), 275–​281.
Shafir, T., Taylor, S. F., Atkinson, A. P., Langenecker, S. A., & Zubieta, J.-​K. (2013). Emotion regulation
through execution, observation, and imagery of emotional movements. Brain and Cognition, 82(2),
219–​227. doi: http://​doi.org/​10.1016/​j.bandc.2013.03.001
Sheppes, G., & Gross, J. J. (2011). Is timing everything? Temporal considerations in emotion regulation.
Pers Soc Psychol Rev., 15(4), 319–​331. doi:1088868310395778 [pii] 10.1177/​1088868310395778
Sheppes, G., & Meiran, N. (2007). Better late than never? On the dynamics of online regulation of
sadness using distraction and cognitive reappraisal. Personality and Social Psychology Bulletin, 33(11),
1518–​1532.
Shiota, M. N., & Levenson, R. W. (2012). Turn down the volume or change the channel? Emotional effects
of detached versus positive reappraisal. Journal of Personality and Social Psychology, 103(3), 416.
95

Conclusion 59

Shiota, M. N., Neufeld, S. L., Yeung, W. H., Moser, S. E., & Perea, E. F. (2011). Feeling good: Autonomic
nervous system responding in five positive emotions. Emotion, 11(6), 1368.
Silvers, J. A., Weber, J., Wager, T. D., & Ochsner, K. N. (2014). Bad and worse: Neural systems underlying
reappraisal of high and low intensity negative emotions. Social Cognitive and Affective Neuroscience,
nsu043.
Stansbury, K., & Gunnar, M. R. (1994). Adrenocortical activity and emotion regulation. Monographs of the
Society for Research in Child Development, 59(2–​3), 108–​134.
Stemmler, G. (2003). Methodological considerations in the psychophysiological study of emotion.
Handbook of affective sciences, 225–​255.
Stone, A. A., Schwartz, J. E., Smyth, J., Kirschbaum, C., Cohen, S., Hellhammer, D., & Grossman, S.
(2001). Individual differences in the diurnal cycle of salivary free cortisol: A replication of flattened
cycles for some individuals. Psychoneuroendocrinology, 26(3), 295–​306.
Thayer, J. F., Hansen, A. L., Saus-​Rose, E., & Johnsen, B. H. (2009). Heart rate variability, prefrontal neural
function, and cognitive performance: The neurovisceral integration perspective on self-​regulation,
adaptation, and health. Annals of Behavioral Medicine, 37(2), 141–​153.
Thiruchselvam, R., Blechert, J., Sheppes, G., Rydstrom, A., & Gross, J. J. (2011). The temporal dynamics
of emotion regulation: An eeg study of distraction and reappraisal. Biological Psychology, 87, 2. doi:
10.1016/​j.biopsycho.2011.02.009
Tomaka, J., Blascovich, J., Kibler, J., & Ernst, J. M. (1997). Cognitive and physiological antecedents of
threat and challenge appraisal. J Pers Soc Psychol, 73(1), 63–​72.
Troy, A. S., & Mauss, I. B. (2010). The up-​regulation of positive emotions in negative situations predicts
well-​being. Paper presented at the Poster Presented: Emotion pre-​conference, Society for Social and
Personality Psychology, Las Vegas, NV.
Troy, A. S., & Mauss, I. B. (2011). Resilience in the face of stress: Emotion regulation ability as a
protective factor. In D. C. S. Southwick, M. Friedman, & B. Litz (Ed.), Resilience to stress (pp. 30–​44).
Cambridge: Cambridge University Press.
Urry, H. L. (2010). Seeing, thinking, and feeling: Emotion-​regulating effects of gaze-​directed cognitive
reappraisal. Emotion, 10(1), 125.
Volokhov, R. N., & Demaree, H. A. (2010). Spontaneous emotion regulation to positive and negative
stimuli. Brain and Cognition, 73(1), 1–​6.
Watson, D., & Vaidya, J. (2003). Mood measurement: Current status and future directions. Handbook of
psychology, New York: Wiley.
Waugh, C. E., Zarolia, P., Mauss, I., Lumian, D., Ford, B., Davis, T., … McRae, K. (2016). Emotion
regulation changes the duration of the bold response to emotional stimuli. Social, Cognitive & Affective
Neuroscience, 11(10), 1550–​1559.
Woodward, S. H., Shurick, A. A., Alvarez, J., Kuo, J., Nonyieva, Y., Blechert, J., … Gross, J. J. (2015).
Seated movement indexes emotion and its regulation in posttraumatic stress disorder. Psychophysiology,
52(5), 679–​686. doi:10.1111/​psyp.12386
06

Chapter 4

Cultural and Social Aspects of Emotion


Regulation
Selda Koydemir & Cecilia A. Essau

Managing emotions
In the last two decades, the field of psychology has witnessed a substantial interest in human emo-
tionality, including how emotions are experienced, expressed, and managed. Emotion regulation
is critical for well-╉being given that studies have demonstrated the positive relationship between
healthy emotion regulation and well-╉being domains such as adjustment, mental health, and posi-
tive social relationships (Gross, 2007). Research also shows that failure to regulate emotions is
linked to a wide range of psychopathology as well as interpersonal, social and cognitive impair-
ments (Aldao, Nolen-╉Hoeksema, & Schweizer, 2010; Rottenberg & Gross, 2003; 2007).
Despite the growing literature on emotional processes, researchers continue to debate whether
emotions are inborn, evolutionary reactions to the outside world or if they are a result of social
and cultural practices (Ekman & Friesen, 1971; Lutz, 1988). Early research on emotion sug-
gested emotions are universal and are accompanied by distinct bodily reactions (Ekman, 1965;
1984; Mead, 1975). For instance, although Ekman discussed the ability of individuals to regulate
their emotions based on what cultures determine is the appropriate emotion expression, he also
proposed that emotions should be perceived as cross-╉culturally invariant. However, contempo-
rary research using a culture-╉specific perspective of emotions suggests that emotions are social
constructions and can be best understood on the social level (Kitayama, Markus, Matsumoto, &
Norasakkunkit, 1997; Markus & Kitayama, 1991; Matsumoto, 1990). In fact, today, the universal
nature of emotions is widely accepted, but the important role socialization processes and cultural
values play on emotional expressions and processes is also considered. As such, emotions are
perceived to be the product of cultural and social processes by which their physiological, neuro-
logical, and psychological components are elicited (Cole, Tamag, & Shrestha, 2006; Kitayama &
Markus, 1994).
Cultural theories propose that the self and emotion are shaped by cultural meanings and
practices (Bruner, 1990; Markus & Kitayama, 1991; Miller, 1999). The self is closely linked to
regulation since regulatory processes have an effect on the way emotions are experienced as well
as how those emotions are expressed in social situations (Srivastava, Tamir, McGonigal, John, &
Gross, 2009). Besides, culture functions to maintain social order, and describes certain norms
regarding emotion regulation (Keltner, Ekman, Gonzaga, & Beer, 2003). In line with these
assumptions, past research has shown that there are cultural differences with respect to many
aspects of emotion regulation including emotion-╉related appraisals (Mauro, Sato, & Tucker,
1992; Roseman, Dhawan, Rettek, Naidu, & Thapa, 1995), coping (Taylor, Sherman, Kim, Jarcho,
Takagi, & Dunagan 2004; Yeh & Inose, 2002), and suppression (Matsumoto, Yoo, Hirayama, &
Petrova, 2005).
16

SOCIALIZATION AND EMOTION REGULATION 61

This chapter focuses on the social and cultural aspects of emotion regulation by examining
cultural explanations of emotional regulation differences, and documenting empirical evidence
garnered from cross-╉cultural research.

Understanding cross-╉cultural research


Culture is one of the most commonly used concepts in contemporary psychological research.
This popularity comes from the recognition that culture is a powerful tool in guiding our percep-
tion, attention, behavior, and emotion, while also determining how we establish and maintain
social relationships (D’Andrade & Strauss, 1992; Markus & Kitayama, 1991). Although culture
has long been regarded as being restricted to nations, it has become important among researchers
to emphasize not only geographical differences but also different aspects of culture itself includ-
ing age, gender, ethnicity, religious affiliation, race, and traditions. Approaches in cross-╉cultural
psychology, thus, conceptualize culture as dynamic systems. Early on, Triandis (1972) defined cul-
ture by referring to “the shared attitudes, beliefs, categorizations, expectations, norms, roles, self-╉
definitions, values, and other such elements of subjective culture found among individuals whose
interactions were facilitated by shared language, historical period, and geographical region.” (p. 3).
Cross-╉cultural psychology deals with the study of relationships between behaviors and the cul-
tural context. In essence, it compares behavior of interest across two or more cultures (Matsumoto,
1996). Cross-╉cultural psychology investigates both differences and similarities of constructs com-
mon to a range of cultural contexts. In understanding behaviors and emotions, cross-╉cultural
psychology provides important sources to researchers by examining whether the psychological
knowledge of one culture is applicable to another. Besides, in the field of clinical psychology,
cross-╉cultural studies enable researchers to examine the universality of psychiatric disorders and
symptoms; the cultural differences of experience of emotions and related emotional and behav-
ioral problems; the meaning of psychiatric disorders and symptoms in different cultures; and the
information that will facilitate culturally sensitive treatment options for disorders. Cross-╉cultural
psychology also helps researchers to examine the variations in emotional display and functioning
in different cultures, as well as the socialization practices that influence these variations in emo-
tions (Ellsworth, 1994; Wang & Fivush, 2005).
Cultural psychology, which is distinguishable from cross-╉cultural psychology, studies the rela-
tionships within a given culture and certain psychological constructs in relation to individuals liv-
ing in that particular region (Shiraev & Levy, 2010). Despite the differences in understanding and
explaining the relationship between culture and psychological constructs, both cross-╉cultural psy-
chology and cultural psychology contribute valuable information with regard to human behavior.

Socialization and emotion regulation


Cultural norms influence emotional development in the context of early parent-╉child relation-
ships (Thompson, 1994) and by prescribing which, when, and how emotions should be displayed
(Matsumoto & Juang, 2013). There are considerable cultural as well as intra-╉societal variability
in the experience and expressions of emotions (Mesquita & Karasawa, 2004). It is reasonable to
assume that although the underlying biological system of emotions is universal, one should make
certain changes in order to adapt to a variety of socio-╉cultural contexts in which they are embed-
ded or exposed to (Mesquita & Albert, 2007; Ochsner & Gross, 2007). It is essential to understand
the social environments surrounding the emotion regulation and related behavior so that the pro-
cess and outcomes can be fully understood (Zeman, Cassano, Perry-╉Parrish, & Stegall, 2006). As
argued by Averill (1980), “emotions are not just remnants of our phylogenetic past, nor can they
26

62 Cultural and Social Aspects of Emotion Regulation

be explained in strictly physiological terms. Rather, they are social constructions, and they can be
fully understood only on a social level of analysis” (p. 309).
The way parents respond to children’s emotions is crucial to children’s self-​and emotion-​
regulation capabilities. Parents can either directly affect children’s emotion regulation by coach-
ing self-​regulation of children or indirectly by managing the emotional demands in the family
(Thompson & Meyer, 2007). One study revealed that mothers’ problem solving responses to their
children’s negative emotions were correlated with children’s constructive coping with problems
(Eisenberg, et al., 1996). In other studies it was revealed that children of mothers who valued guid-
ing emotion development had a better emotional understanding, emotional competence and psy-
chosocial adjustment (Dunsmore & Karn, 2001; Katz, Maliken, & Stettler, 2012). Cunningham,
Kliwer, and Garner (2009) showed that mothers’ emotion coaching is negatively associated with
later internalizing and externalizing behavior. In a more recent study by Meyer et al. (2014), it
was found that children of parents who attended to and accepted emotional experiences, and
maintained more positive emotion socialization had children who had more constructive self-​
regulatory strategies.
Parents’ socialization of emotions in their children seems to be important for the well-​being of
children. However, cultural differences are also present in relation to the socialization of emotions.
Cross-​cultural research provides us with an understanding of the similarities and differences in
the way emotions are experienced across different cultures as well as the socialization practices
that play a role in the variations of emotions (Wang & Fivush, 2005). Both the experience and
the expressions of emotion are culturally constructed and shaped by a given context (Lutz, 1988;
Russell, 1991). People are socialized in a way that teaches them which emotions are appropriate
and inappropriate in varying contexts (Elfenbein & Ambady, 2002; Parkinson, 1996). A variety
of emotional processes show cultural differences given that different beliefs and values shape our
affective life (Matsumoto, Kudoh, Scherer, & Walbott, 1988; Parkinson, 1996). Therefore, it is dif-
ficult to understand emotion regulation without an understanding of the context of the practices
that the sociocultural world creates (Cheung & Park, 2010).
Gross and Thompson (2006) in their modal model theory of emotion regulation asserted that
cognitive appraisals related to emotions are constructed by significant others’ and environment
reactions to the specific behaviors in the concept of reasons and results. They concluded that chil-
dren’s emotion regulation processes are highly influenced by their own culture in terms of their
overt versus covert behavior, the level at which it is deemed socially acceptable to express personal
goals and the degree of coping with problems. Besides, one’s goals, which are not only shaped
by internal processes but also the socio-​cultural environment, are important in deciding how to
manage a particular emotion. Prosocial goals, such as the desire to avoid negatively affecting oth-
ers with one’s emotional expressions, can be an example of interpersonal goals shaped by others.
Among the emotion processes, emotion regulation is crucial for physical and mental health
and well-​being. Emotion regulation consists of processes through which individuals modulate
their emotions either consciously and nonconsciously so that they appropriately respond to the
environment (Rottenberg & Gross, 2003; Thompson, 1994). Matsumoto (2006) defined emotion
regulation as “the ability to manage and modify one’s emotional reactions to achieve goal-​directed
outcomes” (Matsumoto, 2006, p. 421). The concept of emotion regulation is based on the idea that
individuals are active agents in their emotional processes, and that they can control their emo-
tions by using different regulation processes (Gross, 2007). It consists of selections of and changes
in the duration, intensity, and balance of emotion-​related behaviors (Cole, Martin, & Dennis,
2004; Thompson, 1990). Thus, emotion regulation is a dynamic construct and does not only imply
the suppression or control of emotion, but it also includes emotional substitution and the altera-
tion of one’s emotions depending on the purposes.
36

SOCIALIZATION AND EMOTION REGULATION 63

Knowing one’s feelings, what emotions should be expressed when, and what to do with
emotions are skills that are essential for adaptive social interactions and behavioral develop-
ment (Halberstadt, Denham, & Dunsmore, 2001; Hubbard & Coie, 1994). In the development
of these skills (i.e., emotion regulation), socialization plays a key role. To some extent, emo-
tion regulation is learned through observation of others and the teachings of parents as to the
accepted ways of expressing emotions (Denham, 1998). However, it is important to note that
the norms and beliefs of the appropriateness of emotional experience and expression change
among cultures. As Saarni (1999) states, cultural expectations from an individual are funda-
mental for emotion regulation. Learning and understanding which group of emotions are
appropriate to express, the way of expressing them, the best time to express them and selecting
the appropriate person to express them to are all constructed depending on the familial values
(Southam-​Gerow, 2013).
There have been a limited number of studies that have examined the links between emotion
socialization processes and children’s emotive functioning in different cultures. For example, in a
study that examined parenting practices of mothers of pre-​school-​age children living in Mainland
China and those living in the Unites States, Chinese mothers’ tendency to encourage modesty in
their children was found more often as compared to American mothers (Wu et al., 2002). Chinese
mothers also considered shaming and love withdrawal to be more acceptable in terms of emotion
parenting styles in comparison with American mothers. Suveg and collaegues (2014) compared
families from the United States and China regarding family emotional expressiveness, children’s
emotional experiences and regulation. Children and families from the United States were found to
have greater emotional expressiveness than their Chinese counterparts. Furthermore, American
children reported greater under-​controlled emotion that comprised externalizing types of manag-
ing emotional experiences, such as slamming doors when angry and fussing/​whining when sad
when compared to the Chinese children cohort. This study also showed that family expression of
positive emotion was related to effortful emotion regulation among American children, whereas
family expression of negative emotion was associated with under-​controlled emotion for both
United States and Chinese children.
It is known that typical emotion socialization of European American parents is supportive
(Warren & Stifter, 2008). Western parents also prefer to talk about the causes and consequences
of emotions (Wang, 2006). On the other hand, East Asian mothers use minimization as much
as expressive encouragement (Tao, Zhou, & Wang, 2010), and do not support children’s emo-
tion expression (Wang, 2006). In one study, to their child’s aggression toward peers, European
American mothers reported non-​supportive responses such as punishment as compared to
Chinese mothers who used discussion and education to a greater extent (Cheah & Rubin, 2004).
Additionally, European American mothers reported that they would be disappointed by child
aggression whereas Chinese mothers thought they would be angry. Culture, thus, affects how
parents use socialization of emotions.
In sum, parents’ beliefs about emotion socialization and their child rearing practices are impor-
tant in children’s emotion regulation. In fact, caregivers’ emotion socialization is key to emo-
tion regulation development throughout childhood. Children take these beliefs as “… a meaning
system for constructing the self, others, and social relations.” (Trommsdorff & Heikamp, 2013,
p. 69). In general, supportive and constructive responses of parents (e.g., encouragement of emo-
tion expression) facilitate the development of competent emotion regulation skills in children
(Thompson & Meyer, 2007), while non-​supportive responses (e.g., punitive responses) are associ-
ated with children’s poorer emotional competence (Denham & Grout, 1993). Besides, the experi-
ence and expression of emotion is culturally constructed, and emotions are socialized in line with
socially and culturally appropriate norms and expectations in different contexts. Therefore, it is
46

64 Cultural and Social Aspects of Emotion Regulation

important to understand the socialization processes in different cultures in order to make sense
of emotion regulation differences.

Cultural models of self and emotion regulation


Kitayama and Markus (1994, p. 4) argued that emotion is “fully encultured” and should be under-
stood in terms of a cultural frame. Regarding the relationships between culture and human behav-
ior, cultural models are known to involve certain beliefs and social practices which determine
what is appropriate, moral, and desirable in terms of self and relationships. Following Hofstede’s
(1980) important and popular work on cross-╉cultural differences in values, researchers became
increasingly interested in the constructs of individualism and collectivism, and used this dimen-
sion of culture as a theoretical model in their studies. Most of the cross-╉cultural studies in emo-
tion regulation also made use of this dimension (e.g., Elfenbein & Ambady, 2003; Kwon, Yoon,
Joormann, & Kwon, 2013). Generally speaking, individualism was described as a cultural pattern
emphasizing an individual’s goal attainment and personal well-╉being, whereas collectivism was
perceived to place more emphasis on the collective such as the family or group.
Researchers use this dimension and attribute individual traits to people from Western cultures,
and attribute collective characteristics to people from non-╉Western cultures. Individualism is
strongly associated with autonomy, competence, and personal achievement; whereas collectiv-
ism is closely related with duty toward one’s group, interdependence, and maintaining harmony
in relationships. Therefore, in most of the individualistic cultures such as the United States, the
self is construed in independent terms as a unique entity, and differentiated from other people.
On the other hand, in collectivist cultures such as Japan, the self is construed in interdependent
terms that emphasize connectedness with other people, which become meaningful in the large
context of social relationships (Heine, Lehman, Markus, & Kitayama, 1999; Hofstede, 2001). Not
surprisingly, parents in collectivistic countries (e.g., China) tend to encourage the suppression of
ego-╉focused emotions (e.g., anger) in order to maintain interpersonal harmony, and encourage
their children to express group-╉oriented emotions such as gratitude, whereas in individualistic
cultures individual-╉oriented emotions such as happiness are promoted to a greater extent (Saw &
Okazaki, 2010).
However, as it is well-╉known, cultures are not homogenous, and depending on the situational
cues, both individualistic and collectivist orientations can be manifested in individuals of very
different cultures. In order to capture intra-╉and inter-╉cultural variance, one needs to move from
a simple dichotomy such as individualism versus collectivism (Bond & Van de Vijver, 2011;
Oyserman, Coon, & Kemmelmeler, 2002). As argued by Ford and Mauss (2015), it is not a mem-
bership to a particular culture that is important in regulating emotion, but rather their motivation
to do so. Independence and interdependence are present in all cultures; yet some cultures place
more emphasis on one or the other. In this respect, when emotions and behaviors are contextual-
ized in terms of cultural models of the self, it requires the acknowledgment that emotions and
behaviors can be best understood at different levels.
Triandis (1989) made an association between the macro-╉level cultural differences in individual-
ism and collectivism and the micro-╉level self. He proposed that depending on a particular cul-
ture’s emphasis on individualism and collectivism, different self-╉conceptualizations become more
prevalent in a society. In this respect, in individualistic cultures, people are more concerned about
themselves or individuals, and their thoughts are openly expressed. In collectivist cultures, people
are more concerned with the in-╉groups that they belong to. Markus and Kitayama (1991) extended
this model to create a distinction between the independent and interdependent self. The main
differences between independent and interdependent models of the self is that while the former
56

Cross-cultural differences in emotion regulation 65

is concerned with individual autonomy and self-╉achievement, the latter is concerned primarily
with social goals and maintaining harmony. This kind of a difference helps us to understand not
only the differences among different cultural groups, such as Western and Eastern cultures, but
also individuals who dominantly operate at either of these levels. To what extent relationships
are valued and the ways they are evaluated differ as one’s independent and interdependent selves
function. For example, in those cases in which independence is valued, relationships are evaluated
in terms of meeting one’s personal needs (Rothbaum, Weisz, Pott, Miyake, & Morelli, 2000). On
the other hand, in interdependent cultures, people want to fit in the social relationships, and one’s
self is evaluated in terms of meeting the expectations of others (Oishi & Diener, 2003; Mesquita &
Markus, 2004; Rothbaum et al, 2000).
These arguments and empirical findings have several implications for the experience of human
emotionality. The majority of the cross-╉cultural studies discuss cultural differences in emotion
regulation in relation to the American notion of emotion regulation versus non-╉Western notions.
These studies were based on the discussed cultural differences in individualism and collectiv-
ism and the micro-╉level self. The following is a review of studies that examined the relationship
between culture and emotion regulation.

Cross-╉cultural differences in emotion regulation


Contemporary research on emotion regulation has taken a social-╉oriented approach and sug-
gested that people can use emotion regulation with the goal of acting in accordance with others’
expectations (Rothbaum & Wang, 2010; Trommsdorff, 2009). The distinction between self-╉
oriented and social-╉oriented emotion regulation is consistent with the view of the self as inde-
pendent versus interdependent (Markus & Kitayama, 1991). The individual-╉oriented approach
observed in many Western cultures emphasizes authentic expression of emotions motivated by
autonomy goals, whereas the social-╉oriented approach observed in many non-╉Western societies
is concerned with the goal of interdependence and relatedness (Kitayama, Mesquita, & Karasawa,
2006; Kitayama et al., 2000).
In line with these assumptions, in cross-╉cultural research, a distinction has been made between
socially engaging and disengaging emotions (Kitayama, Markus, & Kurokawa, 2000). It has been
proposed that in cultures where the self is constructed with respect to interpersonal relationships
and group cohesion, engaging emotions such as respect and shame become more salient. On the
other hand, in cultures where the self is defined in individual and independent terms, disengag-
ing emotions such as pride and anger are more salient (Kitayama, Mesquita, & Karasawa, 2006;
Markus & Kitayama, 1991; 1994). Depending on which goals—╉independent or interdependent—╉
are thwarted, certain emotions will arise which will eventually facilitate either social engagement
or disengagement of the self. For example, caregivers in individualistic cultures tend to foster hap-
piness as a goal which fits with the goal of autonomy (Heine at al., 1999; Mesquita & Albert, 2007),
whereas in collectivistic cultures emotional harmony rather than intense happiness is fostered by
caregivers since emotional harmony is associated with harmonious relationships and social order
(Mesquita & Albert, 2007; Uchida & Kitayama, 2009).
Another cultural difference manifested is in East Asian collectivistic cultures where negative emo-
tions can be tolerated to a larger extent, down regulation of intense positive emotions is common,
and, in some of these cultures, the balance between and moderation of positive and negative emo-
tions are seen as dominant cultural scripts (Eid & Diener, 2001; Kitayama et al., 2000; Peng & Nisbett,
1999; Uchida & Kitayama, 2009). Morover, Americans are more likely than Japanese to experience
positive emotions more frequently while Japanese are more likely than Americans to experience
both positive and negative emotions in moderation more frequently (Miyamoto & Ryff, 2011).
6

66 Cultural and Social Aspects of Emotion Regulation

One fundamental aspect of emotion regulation, which is also particularly relevant for under-
standing the role of culture, is expressive suppression. Suppression is concerned with the inhibi-
tion of the expressive, behavioral component of emotion, such as gestures or verbal expressions. It
is well known that collectivist cultures are more likely than individualistic cultures to emphasize
adjusting the self and behavior in order to maintain relationship harmony and social cohesion.
This characteristic of the culture suggests that when the expression of an emotion is possibly det-
rimental for one’s relationships, people tend to use emotional suppression which is in line with
the interdependent model of the self. Thus, suppression of emotions may be more encouraged in
collectivist cultures, with a motive to fulfill prosocial goals (e.g., suppression of anger to preserve
group harmony). On the other hand, given that suppression may create a discrepancy between
one’s inner experience of emotion and observable expressive behavior, the use of suppression may
interfere with one’s self-​concept which is a characteristic of the independent self. Therefore, indi-
viduals from Western cultures could be expected to be less likely than individuals from East Asian
cultures to use emotional suppression as a regulatory process.
In line with these assumptions, Matsumoto and colleagues (2008) determined that cultures
emphasizing social order and hierarchy scored higher on emotion suppression. Additionally,
they reported that a positive correlation exists between emotion suppression and reappraisal for
cultures emphasizing social order and hierarchy. On the other hand, for cultures emphasizing
autonomy and egalitarianism, suppression scores were lower, and there was a negative correlation
between suppression and reappraisal. Other studies support this finding, as they have demon-
strated that suppression is not only more frequently applied in collectivistic cultures (Gross &
John, 2003), but it is also associated with less negative social consequences and lower negative
affect (Butler, Lee, & Gross, 2007; Soto, Levenson, & Eberling, 2005). An earlier study (Scherer,
Matsumoto, Wallbott, & Kudoh, 1988) noted that Japanese individuals reported fewer gestures
and body movement than Americans in situations of fear, anger, and sadness, as well as happiness.
This can be explained by the argument that emotional expressions and behaviors are exhibited in
a consistent manner that fits with cultural models.
Cultures also differ with respect to the promotion of events that are associated with particular
emotions. In other words, the extent to which certain events are created or facilitated varies in
accordance with cultural goals. For people in Western/​individualistic cultures, the dominant cul-
tural pattern is to promote or create events that will maximize the experience of positive emotions
and minimize negative emotions (Kitayama, Markus, & Kurakawa, 2000; Tsai & Levenson, 1997).
In the typical North American culture, for instance, happiness activation is valued to a great
extent, people are encouraged and reinforced to feel happy, many contexts in which happiness is
likely to occur are created or promoted, and happiness is perceived as a result of fulfilling one’s
personal goals (Hochschild, 1995; Mesquita & Walker, 2003; Wierzbicka, 1994). Besides, indi-
viduals themselves tend to select situations in which they would engage in activities that promote
happiness (Diener & Suh, 1999). In contrast, in Japan, happiness is not one of the most important
goals of life. Pursuit of one’s individual happiness is not encouraged in the context of the society.
Instead, pursuit of the happiness of the groups or the society is the main focus. Furthermore, indi-
viduals in collectivist cultures are more likely to approach situations that foster contribution to
others and emotions are cultivated as the means to harmonious relationships (Heine et al., 1999).
Cultural differences in emotion regulation are also evident in terms of situation modification
which is an important component of emotion regulation. Individuals with a dominant indepen-
dent self emphasize personal well-​being, and their own preferences; hence they tend to change the
situations to fit their needs. On the contrary, since individuals with an interdependent self are pri-
marily focused on the expectations and needs of others, they tend to accommodate others in dif-
ficult situations (Kitayama, Duffy, & Uchida, 2007; Rothbaum & Trommsdorf, 2007). Given that
76

Cultural scripts and emotional experience 67

individuals with an independent self-╉concept aim for happiness and autonomy, they use strategies
to increase self-╉confidence. For individuals with an interdependent self-╉concept, harmony is a
common goal and thus they are more willing to accept other people. Research shows that children
who are motivated by autonomy tend to change stress-╉eliciting situations to ones that facilitates
happiness (Heine et al., 2001). On the other hand, in stress-╉eliciting environments, children who
are motivated by social harmony tend to restore calmness rather than seek happiness.
Differences in emotion regulation styles within cultures were studied more than a decade ago by
Weisz, Suwanlert, Chaiyasit, Weiss, Achenbach, and Eastman (1993) within Thai and American
adolescents. In this study, parents’ reports were used to measure the differences. Researchers
reported that Thai adolescents showed more control over (e.g., shyness, compulsivity) problems
than Americans and their emotion regulation strategies were different from each other such that
Americans were more direct, open and controlled aggressive towards others under controlled
situations whereas Thai adolescents showed introversion behaviors.
Morelen and colleagues (2012) compared the way in which children in Ghana, Kenya and
America manage their anger in times of sadness. Children in Ghana were found to report display-
ing their anger in more overt, under controlled ways than Kenyan and American children. Kenyan
children on the other hand reported suppressing their anger more than children in Ghana and in
the US. These findings suggested that children in Ghana were more expressive with wider fluctua-
tions in their emotionality than children from the US and Kenya. In terms of sadness, American
children were found to exert more control over this emotion than the two groups of African
children; however, Kenyan children responded calmly to their sadness more than Ghanaian and
American children. The authors argued that these differences may be related to socialization expe-
riences, in that emotional expressivity is shaped by the expectations and responses of others to
anger expression. Specifically, most of the children who lived in the village often received harsh
repercussions for their overt emotional displays by family, whereas similar responses were not
observed in the suburban areas. Speculatively, the village children might have learned to control
their anger in response to the expectation that they would receive a punitive response to emo-
tional displays.
In another study, Zhou and Bishop (2012) examined experiential and cardiovascular out-
comes of three anger regulation strategies (expression, suppression and reappraisal) in Chinese
and Caucasian undergraduate students during a role-╉play that was used to induce anger. Results
indicated that Chinese students reported using reappraisal more frequently in anger situations
than did Caucasians; whereas, no differences were obtained for suppression. Their findings also
showed that cultural background moderated the effects of regulation strategy on cardiovas-
cular reactivity (CVR) following anger provocation. Specifically, when asked to suppress their
emotions, Caucasians showed stronger CVR, whereas, Chinese students showed stronger CVR
when instructed to express their anger. The Chinese students’ greater use of reappraisal com-
pared with Caucasians is interpreted as being consistent with the “other orientation” among the
Chinese, indicating that Chinese people are attuned to others on psychological and behavioral
levels (Yang, 1995). “Other orientation” is also related to a tendency to conform to others, strong
concern about social norms, and an attempt to create a better impression on others through self-╉
monitoring (Zhou & Bishop, 2012). Furthermore, the ability to control the impulse to express
anger is regarded by the Chinese as a good quality and is pursued as an achievement (Yang, 1995).

Cultural scripts and emotional experience


It is evident from cross-╉cultural research that Americans are more likely to appraise emotional
situations as more pleasant when compared to Asians. For instance, Mesquita and Karasawa
86

68 Cultural and Social Aspects of Emotion Regulation

(2002) found that Americans appraised emotional situations as positively different from neutral;
however, Japanese and Taiwanese perceived situations in their lives as neither positive nor nega-
tive. Other studies (e.g., Kitayama at al., 2000) also evidenced that Americans were more likely to
report a higher frequency of positive than negative emotions than Japanese. These findings are in
line with the cultural models that account for the differences between independent and interde-
pendent orientations of the self.
The general assumption is that people want to feel positive emotions (Larsen, 2000), however,
the extent to which people want to regulate hedonically (i.e., to dampen their positive emotions
or to not savor them) differs across cultures. For example, Americans have been found to mainly
focus on the positive aspects of happiness; whereas, Japanese are more likely to indicate negative
aspects of happiness more so than positive ones. Research has also shown that Easterners when
compared to Westerners are more likely to experience positive and negative emotions in pre-
dominantly pleasant situations while no differences are observed in the experience of emotions
in predominantly unpleasant situations (Miyamoto, Kumagai, Lang, & Nunn, 2010). According
to Gross (1998), these cultural differences in emotional experiences are determined by cultural
scripts. The dominant cultural script in Western culture is to maximize positive emotions and
minimize negative emotions (Kitayama et al., 2000). On the other hand, the cultural script that
is dominant in Eastern culture is characterized by a tendency to seek a middle way by balancing
positive and negative emotions.
In many Eastern cultures, emotion moderation which refers to balancing positive emotions is a
more preferred emotion regulation strategy (Miyamoto & Ma, 2011). It is also possible to under-
stand the emphasis on positive-​negative balance in Eastern cultures by looking at the relationship
between positive and negative emotions. Studies have revealed no significant correlation between
positive and negative affects among Western samples; whereas, in individualistic cultures there
is a negative, though small, correlation (Schimmack, Osihi, & Diener, 2002). This is in line with
the findings of other studies that observed that for Americans, positive and negative emotions are
opposites, whereas in Eastern cultures all emotions are accepted more readily (Heine et al., 1999;
Miyamoto et al., 2010).
Some emotions, such as guilt, are more valued in collectivistic cultures than individualistic cul-
tures; whereas, emotions such as pride are perceived as more positive in individualistic cultures
(Eid & Diener, 2001). The intensity and level of arousal of emotions also has a cultural component.
For instance, low arousal, pleasant emotions such as relaxation are valued to a greater extent in
collectivist cultures given that these emotions promote adjustment to others, while high arousal,
pleasant emotions such as excitement, are more valued in individualistic cultures since these emo-
tions promote influencing others (Tsai, Knutson, & Fung, 2006; Tsai et al., 2007). Westerners are
also known to be more likely to think that positive emotions are desirable and appropriate and
negative emotions are undesirable and inappropriate (Eid & Diener, 2001).
It should be noted, however, that in some cultures, collectivism versus individualism may not
be a relevant cultural dimension in terms of emotion regulation. For example, although Mexican
culture is a relatively collectivist culture, in this culture there is a culture script called simpatia
which basically emphasizes promotion of group harmony through the expression of positive emo-
tion (Triandis, Marin, Lisansky, & Betancourt, 1984). Individuals in this culture tend to prefer
high arousal emotions such as enthusiasm over low arousal emotions such as relaxation (Ruby et
al., 2012) which contradicts with other studies (e.g., Tsai et al., 2006). Therefore, it is important to
recognize that two different collectivist cultures may hold different cultural scripts; thus, research-
ers should be careful about generalizing their findings.
Cultural differences have also been reported in the prevalence and appreciation of anger, such
that anger tends to be less prevalent in interdependent than in independent cultures (Markus &
96

Cultural scripts and emotional experience 69

Kitayama, 1991). In a study by Miyake et al. (1986), infants from relatively interdependent cul-
tures were found to react stronger to their mother’s vocal expression of anger (but not joy or fear).
This finding was explained in terms of the low frequency of anger in interdependent cultures.
Moreover, the control of anger is related to high social functioning among Chinese school chil-
dren (Zhou et al., 2004). Anger not only occurred in a low frequency in interdependent culture,
but individuals from interdependent cultures also tolerated less anger. When anger was expressed
in simulated negotiations Asians and Asian Americans made smaller concessions, whereas
European Americans made larger concessions (Adam et al., 2010).
Zahn-​Waxler et al. (1996) investigated how compared Japanese and American preschool
children by investigating how preschoolers reacted to hypothetical interpersonal dilemmas.
American, compared to Japanese children were reported to show more anger and undercon-
trolled emotions such as disorganized, unusual, or incoherent displays of emotion. American
mothers also encouraged their children to express emotions more than Japanese mothers.
Japanese mothers, on the other hand, used more guilt and anxiety induction strategies and
showed disappointment in the child if they failed to meet parental expectations when com-
pared with American mothers. In a study by Lewis and colleagues (2010), white American,
black American, and Japanese pre-​schoolers were compared on how they reacted to success
and failure on a sticker matching task. Results showed that during the failure manipulation
condition, American children expressed more sadness than Japanese children. During the
success condition, American compared to Japanese children showed more pride; Japanese
children, on the other hand, expressed more embarrassment than American children. In dis-
cussing this finding, Lewis and colleagues argued that the Japanese children’s greater display
of embarrassment across conditions is most likely related to cultural differences in response to
being the object of another’s attention.
In attempting to understand the above findings, it is important to note that children are
socialized to regulate their emotion in accordance to their cultural script. A study by Miller,
Wang, Sandel, and Cho (2002) indicated that American mothers considered it important to
highlight their children’s success; Chinese mothers on the other hand considered it important
to discipline children. Children’s emotional responses are closely tied to the differences in
their parent’s response patterns to an event. For instance, while American parents empha-
size their children’s academic success, for Chinese children the case is the opposite (Ng,
Pomerantz, & Lam, 2007). In this study it was found that American mothers were more likely
to provide positive comments (e.g., “You are so smart!”) than Chinese mothers; Chinese
mothers on the other hand were more likely than American mothers to provide neutral and
task-​relevant statements (e.g., “Did you understand what the questions were asking or did
you just randomly guess?”). Furthermore, Chinese children were reported to experience
fewer positive emotions after success, and more negative emotions after a failure as com-
pared to American children. Thus, cultural differences in parenting may affect children’s
emotional expression towards certain events. Considering the cultural models of the self, it
can be argued that being raised in an interdependent context can make it possible to be more
sensitive to negative information.
In a recent study by Miyamoto and Ma (2011), Easterners (i.e., East Asian Undergraduates)
were found to recall engaging in hedonic emotion regulation less than Westerners (i.e., European
American undergraduates) did. They also found cultural differences in emotion regulation to be
mediated by dialectical beliefs about positive emotions. Furthermore, cultural differences in emo-
tion that changed over time were partly explained by dialectical beliefs about positive emotions.
These findings were interpreted in terms of the role that cultural scripts have in shaping emotion
regulation and emotional experiences.
07

70 Cultural and Social Aspects of Emotion Regulation

Emotion regulation and psychopathology


An accumulating number of studies have shown a significant association between emotional sup-
pression and psychopathology as well as negative health and social outcomes (Butler et al., 2007;
Gross & John, 2003; Srivastava, Tamir, McGonigal, John, & Gross, 2009). For instance anxiety, dis-
tress, and depression have been found to be linked with difficulties in emotion regulation (Gross
& Munoz, 1995; Mennin et al., 2005). Studies also show that poor emotion regulation predicts eat-
ing problems and alcohol abuse (Polivy & Herman, 2002; Tice, Bratslavsky, & Baumeister, 2001).
Furthermore, difficulty in emotion regulation is a risk factor for such psychopathologies as social
phobia, major depressive disorder, and bipolar disorder (Johnson, 2005; Kashdan & Breen, 2008;
Rottenberg, Gross, and Gotlib, 2005). In a recent meta-╉analysis, Aldao et al. (2010) documented
that maladaptive emotion regulation strategies such as avoidance and suppression were associ-
ated with psychopathology, whereas adaptive strategies such as reappraisal and acceptance were
associated with less psychopathology.
Emotional suppression has been one of the most widely studied emotion regulation strategies.
Many studies documented that suppression is a positive predictor of depression and anxiety, and
a negative predictor of life satisfaction (Gross & John, 2003; Kashdan & Breen, 2008; Wenzlaff &
Luxton, 2003). In particular, anger suppression has been reported to be positively associated with
high levels of guilt, irritability, and depression (Martin & Dahlen, 2005). However, most of these
studies have been conducted in Western countries which emphasize independent cultural values.
Questions have been raised whether this conclusion is universal.
More recent studies have reported that the consequences of suppression are dependent on
cultural context. As shown by Butler, Lee, and Gross (2007) female undergraduates with higher
Asian values tended to suppress their emotion more often in their daily activities compared to
those with European American values. Furthermore, cultural values were found to moderate the
relation between emotion suppression and negative social outcomes; specifically, suppression
seemed to serve prosocial goals among those with Asian values while among those with Western
values, suppression seemed to serve a self-╉protective function. In interpreting this finding, anger
suppression was considered as a form of emotion regulation that promotes social engagement
and psychological well-╉being for interdependent individuals (Markus & Kitayama, 1991). In
a study by Cheung and Park (2010) among college students, anger suppression mediated the
effects of trait anger and family processes on depression. However, the link between anger sup-
pression and depression was attenuated by an Asian American status.
In one study Soto, Perez, Kim, Lee, and Minnick (2011) found that expressive suppression was
linked with poor psychological functioning for European Americans but not for Chinese. Other
studies also documented evidence for the positive effect of habitual suppression on negative affec-
tivity in Western cultures (e.g., Butler et al., 2007). These findings are in line with the arguments
that since Eastern cultures emphasize interdependence and harmony in social relationships,
expressive suppression is more encouraged than in Western cultures in which personal values
and independence are more important (Markus & Kitayama, 1991). It is also known that indi-
viduals in Eastern cultures emphasize moderation of emotions more than European Americans
(Matsumoto, 1993). Many studies showed a positive relationship between the interdependent self
and depression (Mak, Law, & Teng, 2011).
Cultural scripts, in relation to the extent to which positive emotions are valued, also have clini-
cal importance. Leu, Wang, and Koo (2011) compared college students from different cultural
contexts and found that perceived stress affected depression by means of intensity of positive
emotions among European Americans, whereas for immigrant Asians, no such relationship was
17

Conclusion 71

observed. Additionally, for European Americans, but not immigrant Asians, positive emotions
were associated with decreased depression. Furthermore, pure positive emotions predict better
health outcomes among Western samples, whereas mixed emotions predict better physical health
outcomes among Japanese (Miyamoto & Ryff, 2011).

Conclusion
Despite the general understanding that emotion regulation has a biological basis, the important
role of cultural context in emotion regulation has been recognized both theoretically and empiri-
cally in recent years (Cheung & Park, 2010). The process of emotion regulation takes place in
socio-╉cultural contexts and thus is affected by the environment in which it occurs. As such, how
individuals regulate their emotions is imperative in order to successfully live in social contexts
(Keltner & Haidt, 2001; Lazarus, 1991). In this chapter we reviewed research to document to what
extent an individual desires to start, intensify, or terminate emotions depends on cultural factors
as well as cultural scripts and cultural models of the self.
Several aspects of emotion regulation are influenced by cultural differences such as emotional
expression (Matsumoto & Kupperbusch, 2001), cognitive reappraisal (Yeh & Inose, 2002), and
emotional suppression (Matsumoto, Yoo, Hirayama, & Petrova, 2005). The larger social context
offers standards for what is appropriate to feel and express, and how frequently an emotion regu-
lation strategy is to be used. These standards provide expectations about the ways emotions are
regulated (Kitayama et al., 2000; Mesquita, 2001). In Western cultures, promoting one’s autonomy
and maintaining a positive self-╉view serves as important goals for emotion regulation. On the
other hand, in many non-╉Western cultures the most important goal for emotion regulation is
meeting the expectations of others and maintaining harmonious relationships.
Most of the cross-╉cultural research in this field has compared samples selected from Western
and non-╉Western samples and arrived at similarities as well as differences in emotion regulation.
For example, studies showed that emotion suppression is quite common among Asians (Gross
& John, 1998; Matsumoto et al., 2008) while emotion expression is more common in Western
cultures (Kim & Sherman, 2007). Empirical findings also suggest that the same emotion regu-
lation strategies have different effects in different cultures. Although many studies have shown
that emotional suppression is associated with psychopathology in Western countries, suppressing
emotions leads to positive outcomes for East Asians (Matsumoto et al., 2008). Therefore, behav-
iors are more likely to appear and feel right when it fits the individual’s goals. Since different
socializing practices put differing emphases on autonomy versus harmony, emotion regulation
strategies need to be tailored to these goals whilst considering the cultural microcosm in which
the person resides.
Despite the increasing studies on the social and cultural differences in emotion regulation, most
studies use Japan or China as representations of collectivist cultures and the American as repre-
sentative of individualistic cultures. More studies are needed in other Western and non-╉Western
cultures in order to reach more reliable and generalizable conclusions. For instance Middle
Eastern cultures or cultures balancing both individualism and collectivism may be interesting
to compare. Besides, since cultures are not homogenous and often support both autonomy and
harmony, within-╉culture differences should be investigated more closely. Although cross-╉national
comparisons involving cultural variables are common methods of cross-╉cultural research, coun-
tries cannot be considered cultures. The findings obtained from studies in social and cultural
aspects of emotion regulation should be examined to determine the degree to which they replicate
in other cultural groups.
27

72 Cultural and Social Aspects of Emotion Regulation

References
Adam, H., Shirako, A., & Maddux, W. W. (2010). Cultural variance in the interpersonal effects of anger in
negotiations. Psychological Science, 21(6), 882–╉889.
Aldao, A., Nolen-╉Hoeksema, S., & Schweizer, S. (2010). Emotion-╉regulation strategies across
psychopathology: A meta-╉analytic review. Clinical Psychology Review, 30(2), 217–╉237.
Averill, J. R. (1980). A constructivist view of emotion. In R. Plutchik & H. Kellerman (Eds.),
Emotion: Theory, research and experience (Vol. 1, pp. 305–╉339). New York, NY: Academic Press.
Bond, M. H., & Van de Vijver, F. J. R. (2011). Making scientific sense of cultural differences in
psychological outcomes: Unpackaging the magnum mysterium. In M. H. Bond (Eds), Cross-╉cultural
research methods in psychology, pp. 75–╉100. Cambridge: Cambridge University Press.
Bruner, J. (1990). Culture and human development: A new look. Human Development, 33(6), 344–╉355.
Butler, E. A., Lee, T. L., & Gross, J. J. (2007). Emotion regulation and culture: are the social consequences of
emotion suppression culture-╉specific? Emotion, 7(1), 30–╉48.
Cheah, C. L., & Rubin, K. H. (2004). European American and Mainland Chinese mothers’ responses to
aggression and social withdrawal in preschoolers. International Journal of Behavioral Development,
28, 83–╉94.
Cheung, R. Y. M., & Park, I. J. K. (2010). Anger suppression, interdependent self-╉construal, and depression
among Asian American and European American college students. Cultural Diversity & Ethnic Minority,
16 (4), 517–╉525.
Cole, P. M., Martin, S. E., & Dennis, T. A. (2004). Emotion regulation as a scientific construct:
Methodological challenges and directions for child development research. Child Development, 75,
317–╉333.
Cole, P., Tamang, B. L., & Shrestha, S. (2006). Cultural variations in the socialization of young children’s
anger and shame. Child Development, 77, 1237–╉1251.
Cunningham, J. N., Kliewer, W., & Garner, P. W. (2009). Emotion socialization, child emotion
understanding and regulation, and adjustment in urban African American families: Differential
associations across child gender. Development and Psychopathology, 21, 261–╉283.
D’Andrade, R. G., & Strauss, C. (1992). Human motives and cultural models (Vol. 1). New York: Cambridge
University Press.
Denham, S. A. (1998). Emotional development in young children. New York: Guilford Press.
Denham, S. A., & Grout, L. (1993). Socialization of emotion: Pathway to preschoolers’ affect regulation.
Journal of Nonverbal Behavior, 17, 215–╉227.
Diener, E., & Suh, E. (1999). National differences in subjective well-╉being. In D. Kahneman, E. Diener, &
N. Schwarz (Eds.).Well-╉being: The foundations of hedonic psychology, (pp. 434–╉452). New York: Russell
Sage Foundation.
Dunsmore, J. C., & Karn, M. A. (2001). Mothers’ beliefs about feelings and children’s emotional
understanding. Early Education and Development, 12, 117–╉138.
Eid, M., & Diener, E. (2001). Norms for experiencing emotions in different cultures: inter-╉and intranational
differences. Journal of Personality and Social Psychology, 81(5), 869–╉885.
Eisenberg, N., Fabes, R. A., Guthrie, I. K., Murphy, B. C., Maszk, P., Holmgren, R., & Suh, K. (1996).
The relations of regulation and emotionality to problem behavior in elementary school children.
Development and Psychopathology, 8, 141–╉162.
Ekman, P. (1965). Differential communication of affect by head and body cues. Journal of Personality and
Social Psychology, 2, 725–╉735.
Ekman, P. (1984). Expression and the nature of emotion. In K. R. Scherer & E Ekman (Eds.). Approaches to
emotion, (pp. 319–╉344). Hillsdale, NJ: Erlbaum.
Ekman, P., & Friesen, W. V. (1971). Constants across cultures in the face and emotion. Journal of Personality
and Social Psychology, 17, 124–╉129.
37

Conclusion 73

Elfenbein, H. A., & Ambady, N. (2002). On the universality and cultural specificity of emotion
recognition: a meta-​analysis. Psychological Bulletin, 128(2), 203.
Elfenbein, H. A., & Ambady, N. (2003). Cultural similarity’s consequences: A distance perspective on cross-​
cultural differences in emotion recognition. Journal of Cross Cultural Psychology, 34, 92–​110.
Ellsworth, P. C. (1994). Sense, culture, and sensibility. In S. Kitayama, & H. R. Markus (Eds.), Emotion and
culture: Empirical studies of mutual influence, (pp. 23–​50). Washington, DC: American Psychological
Association.
Ford, B. Q., & Mauss, I. B. (2015). Culture and emotion regulation. Current Opinion in Psychology, 3, 1–​5.
Gross, J. J. (1998). Antecedent-​and response-​focused emotion regulation: Divergent consequences for
experience, expression, and physiology. Journal of Personality and Social Psychology, 74(1), 224–​237.
Gross, J. J., & Muñoz, R. F. (1995). Emotion regulation and mental health.Clinical Psychology: Science and
Practice, 2(2), 151–​164.
Gross, J. J., & John, O. P. (1998). Mapping the domain of expressivity: multimethod evidence for a
hierarchical model. Journal of Personality and Social psychology, 74(1), 170–​191.
Gross, J. J., & John, O. P. (2003). Individual differences in two emotion regulation processes: implications
for affect, relationships, and well-​being. Journal of Personality and Social Psychology, 85(2), 348–​362.
Gross. J. (Ed.). (2007). Handbook of emotion regulation. New York: Guilford Press.
Gross, J. J., & Thompson, R.A. (2006). Emotion regulation: Conceptual foundations. In J. J. Gross (Ed.),
Handbook of Emotion Regulation, (pp. 3–​26). New York: Guildford Press.
Halberstadt, A. G., Denham, S. A., & Dunsmore, J. C. (2001). Affective social competence. Social
Development, 10(1), 79–​119.
Heine, S. J., Kitayama, S., Lehman, D. R., Takata, T., Ide, E., Leung, C., & Matsumoto, H. (2001).
Divergent consequences of success and failure in Japan and North America: An investigation of
self improving motivations and malleable selves. Journal of Personality and Social Psychology, 81(4),
599–​615.
Heine, S. J., Lehman, D. R., Markus, H. R., & Kitayama, S. (1999). Is there a universal need for positive
self-​regard? Psychological Review, 106(4), 766–​794.
Hochschild, A. R. (1995). The culture of politics: traditional, postmodern, cold-​modern, and warm modern
ideals of care. Social Politics: International Studies in Gender, State & Society, 2(3), 331–​346.
Hofstede, G. (1980). Culture’s consequences: International differences in work related values. Beverly Hills,
CA: Sage.
Hofstede, G. H. (2001). Culture’s consequences: Comparing values, behaviors, institutions and
organizations across nations. Sage, Thousand Oaks, CA.
Hubbard, J. A., & Cole, J. D. (1994). Emotional correlates of social competence in children’s peer
relationships. Merrill-​Palmer Quarterly, 40, 1–​20.
Johnson, S. L. (2005). Mania and dysregulation in goal pursuit: A review. Clinical Psychology Review, 25,
241–​262.
Kashdan, T. B., & Breen, W. E. (2008). Social anxiety and positive emotions: A prospective examination
of a self-​regulatory model with tendencies to suppress or express emotions as a moderating variable.
Behavior Therapy, 39, 1−12.
Katz, L. F., Maliken, A. C., & Stettler, N. M. (2012). Parental meta-​emotion philosophy: A review of
research and theoretical framework. Child Development Perspectives, 6(4), 417–​422.
Keltner, D., & Haidt, J. (2001). Social functions of emotions. In T. J. Mayne & G. A. Bonanno (Eds.),
Emotions: Current issues and future directions. Emotions and social behavior, (pp. 192–​213).
New York: Guilford.
Keltner, D., Ekman, P., Gonzaga, G. C., & Beer, J. (2003). Facial expression of emotion. In Davidson,
R. J., Scherer, K. R., & Goldsmith, H. H. (Eds.), Handbook of affective sciences, (pp. 415–​432).
New York: Oxford University Press.
47

74 Cultural and Social Aspects of Emotion Regulation

Kim, H. S., & Sherman, D. K. (2007). “Express yourself ”: Culture and the effect of self-​expression on
choice. Journal of Personality and Social psychology, 92(1), 1–​11.
Kitayama, S. E., & Markus, H. R. E. (1994). Emotion and culture: Empirical studies of mutual influence.
Washington, DC, American Psychological Association.
Kitayama, S., Duffy, S., & Uchida, Y. (2006). Self as cultural mode of being. In. S. Kitayama & D. Cohen
(Eds.), Handbook of cultural psychology, (pp. 136–​173). New York: Guilford Press.
Kitayama, S., Markus, H. R., & Kurokawa, M. (2000). Culture, emotion, and well-​being: Good feelings in
Japan and the United States. Cognition and Emotion, 14(1), 93–​124.
Kitayama, S., Markus, H.R., Matsumoto, H., & Norasakkunkit, V. (1997). Individual and collective
processes in the construction of the self: Self-​enhancement in the United States and self-​criticism in
Japan. Journal of Personality and Social Psychology, 72, 1245–​1267.
Kitayama, S., Mesquita, B., & Karasawa, M. (2006). Cultural affordances and emotional experience: socially
engaging and disengaging emotions in Japan and the United States. Journal of Personality and Social
Psychology, 91(5), 890–​903.
Kwon, H., Yoon, K. L., Joormann, J., & Kwon, J. H. (2013). Cultural and gender differences in emotion
regulation: Relation to depression. Cognition and Emotion, 27, 769–​782.
Larsen, R. J. (2000). Toward a science of mood regulation. Psychological Inquiry, 11(3), 129–​141.
Lazarus, R. S. (1991). Emotion and adaptation. Oxford, UK: Oxford University Press.
Leu, J., Wang, J., & Koo, K. (2011). Are positive emotions just as “positive” across cultures? Emotion, 11(4),
994–​999.
Lewis, M., Takai-​Kawakami, K., Kawakami, K., & Sullivan, M. (2010). Cultural differences in emotional
responses to success and failure. International Journal of Behavioral Development, 34, 53–​61.
Lutz, C. A. (1988). Unnatural emotions: Everyday sentiments on a Micronesian atoll and their challenge to
Western theory. Chicago: University of Chicago Press.
Mak, W. S., Law, R. W., & Teng, Y. (2011). Cultural model of vulnerability to distress: The role of self-​
construal and sociotropy on anxiety and depression among Asian Americans and European Americans.
Journal of Cross-​Cultural Psychology, 42(1), 75–​88.
Markus, H. R., & Kitayama, S. (1991). Culture and the self: Implications for cognition, emotion, and
motivation. Psychological Review, 98(2), 224–​253.
Martin, R. C., & Dahlen, E. R. (2005). Cognitive emotion regulation in the prediction of depression,
anxiety, stress, and anger. Personality and Individual Differences, 39, 1249−1260.
Matsumoto, D. (1990). Cultural similarities and differences in display rules. Motivation and Emotion, 14(3),
195–​214.
Matsumoto, D. (1993). Ethnic differences in affect intensity, emotion judgments, display rule attitudes, and
self-​reported emotional expression in an American sample. Motivation & Emotion, 17, 107–​123.
Matsumoto, D., Kudoh, T., Scherer, K., & Wallbott, H. (1988). Antecedents of and reactions to emotions in
the United States and Japan. Journal of Cross-​Cultural Psychology, 19(3), 267–​286.
Matsumoto, D., & Kupperbusch, C. (2001). Idiocentric and allocentric differences in emotional expression
and experience. Asian Journal of Social Psychology, 4, 113–​131.
Matsumoto, D., Yoo, S. H., Hirayama, S., & Petrova, G. (2005). Validation of an individual-​level measure
of display rules: The display rule assessment inventory (DRAI). Emotion, 5(1), 23–​40.
Matsumoto, D. (1996). Culture and psychology. Pacific Grove, CA: Brooks Cole.
Matsumoto, D. (2006). Are cultural traits in emotion regulation mediated by personality traits? Journal of
Cross-​Cultural Psychology, 37(4), 421–​437.
Matsumoto, D., & Juang, L. (2013). Culture and psychology (5th ed.). Belmont, CA: Wadsworth/​Cengage
Learning.
Matsumoto, D., Yoo, S. H., Fontaine, J., Anguas-​Wong, A. M., Arriola, M., Ataca, B., Grossi, E. (2008).
Mapping expressive differences around the world: The relationship between emotional display rules and
individualism versus collectivism. Journal of Cross-​Cultural Psychology, 39, 55–​74.
57

Conclusion 75

Mauro, R., Sato, K., & Tucker, J. (1992). The role of appraisal in human emotions: a cross-​cultural study.
Journal of Personality and Social Psychology, 62(2), 301–​317.
Mead, D. J. (1975). Review of Darwin and facial expression. Journal of Communication, 25, 209–​213.
Mennin, D. S., Heimberg, R. G., Turk, C. L., & Fresco, D. M. (2005). Preliminary evidence for an emotion
dysregulation theory of generalized anxiety disorder. Behavior Research and Therapy, 43, 1281–​1310.
Mesquita, B. (2001). Emotions in collectivist and individualist contexts. Journal of Personality and Social
Psychology, 80, 68–​74.
Mesquita, B., & Albert, D. (2007). The cultural regulation of emotions. In J. J. Gross (Ed.), The handbook of
emotion regulation, (pp. 486–​503). New York: Guilford Press.
Mesquita, B., & Karasawa, M. (2002). Different emotional lives. Cognition and Emotion, 16(1), 127–​141.
Mesquita, B., & Karasawa, M. (2004). Self-​conscious emotions as dynamic cultural processes. Psychological
Inquiry, 161–​166.
Mesquita, B., & Markus, H. R. (2004). Culture and emotion: Models of agency as sources of cultural
variation in emotion. In N. H. Frijda, A. S. R. Manstead, & A. H. Fischer (Eds.), Feelings and
emotions: The Amsterdam Symposium, (pp. 341–​358). Cambridge: Cambridge University Press.
Mesquita, B., & Walker, R. (2003). Cultural differences in emotions: A context for interpreting emotional
experiences. Behaviour Research and Therapy, 41(7), 777–​793.
Meyer, S., Raikes, H. A., Virmani, E. A., Waters, S., & Thompson, R. A. (2014). Parent emotion
representations and the socialization of emotion regulation in the family. International Journal of
Behavioral Development, 38(2), 164–​173.
Miyake, K., Campos, J. J., Kagan, J., & Bradshaw, D. L. (1986). Issues in socioemotional development. In
H. Stevenson, H. Azuma & K. Hakuta (Eds.), Child development and education in Japan, (pp. 239–​261).
New York, NY: W.H. Freeman.
Miller, D. T. (1999). The norm of self-​interest. American Psychologist, 54(12), 1053–​1060.
Miller, P. J., Wang, S. H., Sandel, T., & Cho, G. E. (2002). Self-​esteem as folk theory: A comparison of
European American and Taiwanese mothers’ beliefs. Parenting: Science and Practice, 2(3), 209–​239.
Miyamoto, T., Kumagai, T., Lang, M.S. & Nunn, M.E. (2010) Compliance as a prognostic indicator.
II. Impact of patient’s compliance to the individual tooth survival. Journal of Periodontology, 81,
1280–​1288.
Miyamoto, Y., & Ryff, C. D. (2011). Cultural differences in the dialectical and non-​dialectical emotional
styles and their implications for health. Cognition and Emotion, 25(1), 22–​39.
Miyamoto, Y., & Ma, X. (2011). Dampening or savoring positive emotions: A dialectical cultural script
guides emotion regulation. Emotion, 11, 1346–​1357.
Morelen,D., Zeman,J., Perry-​Parrish, C., & Anderson, E. (2012). Children’s emotion regulation across
and within nations: A comparison of Ghanaian, Kenyan, and American youth. British Journal of
Developmental Psychology, 30, 415–​431.
Ng, F. F. Y., Pomerantz, E. M., & Lam, S. F. (2007). European American and Chinese parents’ responses to
children’s success and failure: implications for children’s responses. Developmental Psychology, 43(5),
1239–​1255.
Ochsner, K. N., & Gross, J. J. (2007). The neural architecture of emotion regulation. In J. J. Gross (Ed.),
Handbook of emotion regulation, (pp. 87–​109). New York: The Guilford Press.
Oishi, S., & Diener, E. (2003). Culture and well-​being: The cycle of action, evaluation, and decision.
Personality and Social Psychology Bulletin, 29(8), 939–​949.
Oyserman, D., Coon, H. M., & Kemmelmeier, M. (2002). Rethinking individualism and
collectivism: evaluation of theoretical assumptions and meta-​analyses. Psychological Bulletin, 128(1), 3.
Parkinson, B. (1996). Emotions are social. British Journal of Psychology, 87(4), 663–​684.
Peng, K., & Nisbett, R. E. (1999). Culture, dialectics, and reasoning about contradiction. American
Psychologist, 54(9), 741–​754.
Polivy, J., & Herman, C. P. (2002). Causes of eating disorders. Annual Review of Psychology, 53(1), 187–​213.
67

76 Cultural and Social Aspects of Emotion Regulation

Roseman, I. J., Dhawan, N., Rettek, S. I., Naidu, R. K., & Thapa, K. (1995). Cultural differences and
cross-​cultural similarities in appraisals and emotional responses. Journal of Cross-​Cultural Psychology,
26(1), 23–​48.
Rothbaum, F., & Trommsdorff, G. (2007). Do roots and wings complement or oppose one another: The
socialization of relatedness and autonomy in cultural context. In J. Grusec, & P. Hastings (Eds.),
Handbook of socialization, (pp. 461–​489). New York: Guilford Press.
Rothbaum, F., & Wang, Y. Z. (2010). Fostering the child’s malleable views of the self and the
world: Caregiving practices in East Asian and European-​American communities. Psychologie–​Kultur
Gesellschaft, Springer, Berlin (pp. 101–​120).
Rothbaum, F., Weisz, J., Pott, M., Miyake, K., & Morelli, G. (2000). Attachment and culture: Security in the
United States and Japan. American Psychologist, 55(10), 1093–​1104.
Rottenberg, J., & Gross, J. J. (2003). When emotion goes wrong: Realizing the promise of affective science.
Clinical Psychology: Science and Practice, 10, 227–​232.
Rottenberg, J., & Gross, J. J. (2007). Emotion and emotion regulation: a map for psychotherapy researchers.
Clinical Psychology: Science and Practice, 14, 323–​328.
Rottenberg, J., Gross, J. J., & Gotlib, I. H. (2005). Emotion context insensitivity in major depressive
disorder. Journal of Abnormal Psychology, 114, 627−639.
Ruby, M. B., Falk, C. F., Heine, S. J., Villa, C., & Silberstein, O. (2012). Not all collectivisms are
equal: Opposing preferences for ideal affect between East Asians and Mexicans. Emotion, 12, 1206–​
1209. doi:10.1037/​a0029118.
Russell, J. A. (1991). Culture and the categorization of emotions. Psychological Bulletin, 3, 426–​450.
Saarni, C. (1999). The development of emotional competence. New York: Guilford Press.
Saw, A., & Okazaki, S. (2010). Family emotion socialization and affective distress in Asian American and
White American college students. Asian American. Journal of Psychology, 1, 81–​92.
Scherer, K. R., Matsumoto, D., Wallbott, H. G., & Kudoh, T. (1988). Emotional experience in cultural
context: a comparison between Europe, Japan, and the US. In K. R. Scherer (Ed.), Facets of emotion,
(pp. 5–​39). Hillsdale, NJ: Lawrence Erlbaum.
Schimmack, U., Diener, E., & Oishi, S. (2002). Life‐satisfaction is a momentary judgment and a stable
personality characteristic: The use of chronically accessible and stable sources. Journal of Personality,
70(3), 345–​384.
Shiraev, E. B., & Levy, D. A. (2010). Cross-​cultural psychology: Critical thinking and contemporary
applications (4th ed.). Boston, MA: Allyn & Bacon.
Suveg, C., Raley, J. N., Morelen, D., Wang, W., Han, R. Z., & Campion, S. (2014). Child and Family
Emotional Functioning: A Cross-​National Examination of Families from China and the United States.
Journal of Child and Family Studies, 23, 1444–​1454.
Soto, J. A., Levenson, R. W., & Ebling, R. (2005). Cultures of moderation and expression: emotional
experience, behavior, and physiology in Chinese Americans and Mexican Americans. Emotion, 5(2),
154–​165.
Soto, J. A., Perez, C. R., Kim, Y. H., Lee, E. A., & Minnick, M. R. (2011). Is expressive suppression always
associated with poorer psychological functioning? A cross-​cultural comparison between European
Americans and Hong Kong Chinese. Emotion, 11(6), 1450–​1455.
Southam-​Gerow, M. A. (2013). Emotion regulation in children and adolescents: A practitioner’s guide.
New York: Guilford Press.
Srivastava, S., Tamir, M., McGonigal, K. M., John, O. P., & Gross, J. J. (2009). The social costs of emotional
suppression: a prospective study of the transition to college. Journal of Personality and Social Psychology,
96(4), 883–​897.
Tao, A., Zhou, Q., & Wang, Y. (2010). Parental reactions to children’s negative emotions: Prospective
relations to Chinese children’s psychological adjustment. Journal of Family Psychology, 24, 135–​144.
7

Conclusion 77

Taylor, S. E., Sherman, D. K., Kim, H. S., Jarcho, J., Takagi, K., & Dunagan, M. S. (2004). Culture and
social support: who seeks it and why? Journal of Personality and Social Psychology, 87(3), 354–​562.
Thompson, R. A. (1990). Emotion and self-​regulation. In R. A. Thompson (Ed.), Socioemotional
development. Nebraska Symposium on Motivation, (pp. 367–​467). Lincoln: University of Nebraska Press.
Thompson, R. (1994). Emotion regulation: A theme in search of a definition. Emotion regulation biological
and behavioral considerations. Monographs of the Society for Research in Child Development, 59, 25–​52.
Thompson, R. A., & Meyer, S. (2007). Socialization of emotion regulation in the family. In J. J. Gross (Ed.),
Handbook of emotion regulation, 249–​268. New York, NY: Guilford Press.
Tice, D. M., Bratslavsky, E., & Baumeister, R. F. (2001). Emotional distress regulation takes precedence
over impulse control: If you feel bad, do it! Journal of Personality and Social Psychology, 80, 53−67.
Triandis, H. C. (1972). The analysis of subjective culture. New York: Wiley.
Triandis, H. C. (1989). The self and social behavior in differing cultural contexts. Psychological Review,
96(3), 506–​520.
Triandis, H. C., Marin, G., Lisansky, J., & Betancourt, H. (1984). Simpatia as a cultural script of Hispanics.
Journal of Personality and Social Psychology, 47(6), 1363–​1374.
Trommsdorff, G. (2009). Intergenerational relations and cultural transmission. In A. Gari & K. Mylonas
(Eds.), Quod erat demonstrandum: From herodotus’ ethnographic journeys to cross-​cultural research.
Book of selected chapters of the 18th international congress of the International Association for Cross-​
Cultural Psychology (pp. 221–​230). Athens, Greece: Pedio Books Publishing.
Trommsdorff, G., & Heikamp, T. (2013). Socialization of emotions and emotion regulation in cultural
context. In S. Barnow, & N. Balkir (Eds.), Cultural variations in psychopathology, (pp. 67–​92).
Göttingen, Germany: Hogrefe.
Tsai, J. L., & Levenson, R. W. (1997). Cultural influences on emotional responding Chinese American and
European American dating couples during interpersonal conflict. Journal of Cross-​Cultural Psychology,
28(5), 600–​625.
Tsai, J. L., Miao, F. F., Seppala, E., Fung, H. H., & Yeung, D. Y. (2007). Influence and adjustment
goals: Sources of cultural differences in ideal affect. Journal of Personality and Social Psychology, 92,
1102–​1117.
Tsai, J. L., Knutson, B., & Fung, H. H. (2006). Cultural variation in affect valuation. Journal of Personality
and Social Psychology, 90(2), 288–​307.
Uchida, Y., & Kitayama, S. (2009). Happiness and unhappiness in east and west: themes and variations.
Emotion, 9(4), 441–​456.
Warren, H. K., & Stifter, C. A. (2008). Maternal Emotion‐related Socialization and Preschoolers’
Developing Emotion Self‐awareness. Social Development, 17(2), 239–​258.
Wang, Q. (2006). Relations of maternal style and child self-​concept to autobiographical memories in
Chinese, Chinese immigrant, and European American 3-​year-​olds. Child Development, 77 (6),
1794–​1809.
Wang, Q., & Fivush, R. (2005). Mother–​child conversations of emotionally salient events: Exploring the
functions of emotional reminiscing in European‐American and Chinese families. Social Development,
14(3), 473–​495.
Weisz, J. R., Suwanlert, S., Chaiyasit, W., Weiss, B., Achenbach, T. M., & Eastman, K. L. (1993). Behavioral
and emotional problems among Thai and American adolescents: Parent reports for ages 12–​16. Journal
of Abnormal Psychology, 102(3), 395–​403.
Wenzlaff, R. M., & Luxton, D. D. (2003). The role of thought suppression in depressive rumination.
Cognitive Therapy and Research, 27, 293−308.
Wierzbicka, A. (1994). Emotion, language, and cultural scripts. In S. Kitayama & H. R. Markus (Eds.)
Emotion and culture: Empirical studies of mutual influence, (pp. 130–​196). Washington, D.C.: American
Psychological Association.
87

78 Cultural and Social Aspects of Emotion Regulation

Wu, P., Robinson, C. C., Yang, C., Hart, C. H., Olsen, S. F., Porter, C. L., Jin, S., Wo, J., & Wu, X. (2002).
Similarities and differences in mothers’ parenting of preschoolers in China and the United States.
International Journal of Behavioral Development, 26, 481–​491.
Yang, K. S. (1995). Chinese Social Orientation: an Integrated Analysis. In Lin, T. Y., Tseng, W. S., Yeh, E. K.
(Eds.), Chinese societies and mental health, (pp. 19–​39). Oxford University Press, New York.
Yeh, C., & Inose, M. (2002). Difficulties and coping strategies of Chinese, Japanese, and Korean immigrant
students. Adolescence, 37(145), 69–​82.
Zahn‐Waxler, C., Friedman, R. J., Cole, P. M., Mizuta, I., & Hiruma, N. (1996). Japanese and United States
preschool children’s responses to conflict and distress. Child Development, 67(5), 2462–​2477.
Zeman, J., Cassano, M., Perry-​Parrish, C., & Stegall, S. (2006). Emotion regulation in children and
adolescents. Journal of Developmental & Behavioral Pediatrics, 27(2), 155–​168.
Zhou, T. & Bishop, G. D. (2012). Culture moderates the cardiovascular consequences of anger regulation
strategy. International Journal of Psychophysiology, 86, 291–​298.
Zhou, Q., Eisenberg, N., Wang, Y., & Reiser, M. (2004). Chinese children’s effortful control and
dispositional anger/​frustration: relations to parenting styles and children’s social functioning.
Developmental Psychology, 40(3), 352–​366.
97

Chapter 5

Research Domain Criteria (RDoC)


and Emotion Regulation
Michael Sun, Meghan Vinograd, Gregory A. Miller,
& Michelle G. Craske

Emotion control
The ability to control one’s emotional experiences can lead to richer, more productive, and
healthier lives. For this reason, the study of emotion regulation has played a growing role in
cross-╉disciplinary psychological research over the last two decades. As a result, we know that
appropriate emotion regulation in childhood is associated with better mental (Gross & Muñoz,
1995) and physical health (Gross & Levenson, 1993, 1997; John & Gross, 2004), reduced stress
(Martin & Dahlen, 2005), improved relationships (John & Gross, 2004; Lopes et al., 2011; Lopes,
Salovey, Côté, Beers, & Petty, 2005), increased resistance to temptation (Casey et al., 2011), and
more efficient workplace organization (Coté, 2005; Grandey, 2000). On the other hand, poor emo-
tion down-╉regulation can lead to slower reaction times to emotional pictures (Ortner, Zelazo, &
Anderson, 2013) and poorer long-╉term memory (Richards & Gross, 1999, 2000). Emotion regula-
tion has established itself as a valuable domain of inquiry in psychological science given its associ-
ations with a variety of important outcomes. Although emotion regulation has often been invoked
as a concept of relevance to the etiology, pathophysiology, and presentation of mental disorders
(e.g., Aldao, Nolen-╉Hoeksema, & Schweizer, 2010; Campbell-╉Sills & Barlow, 2007; Cicchetti,
Ackerman, & Izard, 1995; Gross, 1998; Gross & Muñoz, 1995), the concept itself has been met
with confusion (e.g., Bridges, Denham, & Ganiban, 2004; Gross & Barrett, 2011; Thompson,
1994). A theoretical understanding of emotion generation and regulation that incorporates rel-
evant findings emerging from the neuroscience literature is essential for scientific advancement
and new avenues for clinical treatment.

Mental health and disorder: A brief history


The sixth United States Census, held in 1840, was the first national attempt to account for individ-
uals with mental illness. There was but a single category: “idiocy/╉insanity” (American Psychiatric
Association, 2016). It was not until 40 years later that attempts were made to refine this broad
category, and seven categories emerged: mania, melancholia, monomania, paresis, dementia,
dipsomania, and epilepsy (American Psychiatric Association, 2016). There was little need for
increased conceptual separation at the time as there was little verifiable understanding of etiology
or pathophysiology. Neither were there effective treatments to address mental disorders differen-
tially once identified. The motivation for increased clinical utility did not emerge until 1917 in a
collaborative effort between the American Medico-╉Psychological Association (now the American
Psychiatric Association) and the National Commission on Mental Hygiene (American Psychiatric
Association, 2016). In 1949, the World Health Organization published the highly influential, sixth
08

80 Research Domain Criteria (RDoC) and Emotion Regulation

edition of the International Classification of Diseases (ICD-​6), which included 17 categories for
mental health and psychological traits for the first time (American Psychiatric Association, 2016).
The American Psychiatric Association (APA) Committee on Nomenclature and Statistics com-
missioned the first edition of the Diagnostic and Statistical Manual of Mental Disorders (DSM-​I)
in 1952. This initial version was poorly accepted and, when utilized, exhibited poor reliability.
Clinicians could not agree which individuals belonged in which category, and it was lambasted by
Dr. Erwin Stengel who was commissioned by the World Health Organization to provide a com-
prehensive review of diagnostic issues facing both DSM-​I and ICD-​6 and 7 (American Psychiatric
Association, 2016). Stengel recommended progress be made toward explicitly defining mental
disorders for the purposes of reliable clinical diagnosis. It was not until DSM-​III in 1980 that these
suggestions were taken fully into consideration through the construction of several important
conceptual innovations: 1) An effort toward explicit diagnostic criteria with more emphasis on
overt behavioral manifestations and 2) a multiaxial diagnostic assessment system that included
acute symptoms, trait abnormalities, medical concerns, and other psychosocial considerations
(American Psychiatric Association, 2016). Its 494 pages provided 265 diagnostic categories
(American Psychiatric Association, 1980), but concerns regarding diagnostic reliability and valid-
ity continued.
Efforts were made in subsequent revisions of the DSM to improve reliability. The DSM-​5, pub-
lished in 2013, features 157 diagnostic categories. Despite attempts to improve its reliability as a
diagnostic tool, concerns remained. The traditional problem of under-​accounting for dysfunction
(i.e., a disorder yet to be discovered or separated from another disorder) persisted, and another
problem emerged: The over-​accounting and over-​separation of mental disorder concepts which
required scientific and clinical reconceptualization (e.g., Asperger’s and autism being moved to a
single diagnostic spectrum; American Psychiatric Association, 2013). Moreover, the progression
of the DSM has historically given more attention to reliability over validity, although the DSM-​5
did make significant advances in terms of validity (e.g., Clarke et al., 2013; Freedman et al., 2013;
Narrow et al., 2013; Regier et al., 2013). As such, diagnostic nosologies such as the DSM and the
ICD are continual works in progress, but each is a cultural artifact that roots itself in the language
of clinical research, clinical practice, and broader traditions and trends extant at the time.
The invocation of diagnostic categories reifies conceptualizations of mental function and dys-
function. While generally reliable across health care providers, the literature often indicates that
conventional diagnostic categories fail to achieve validity through convergent evidence with allied
disciplines such as neuroscience (e.g., Casey et al., 2013; Cuthbert, 2014; Insel et al., 2010). It is
anticipated that rapid developments in new methodologies such as genetics and neuroscience will
enhance the granularity of analysis of mental disorders. Diagnostic categories describe subgroups
with wide-​ranging symptom profiles, leading to problematically heterogeneous patient groups
(e.g., Chen, Eaton, Gallo, & Nestadt, 2000; Litten et al., 2015; Wåhlstedt, Thorell, & Bohlin, 2009).
The categories limit the capacity to identify new concepts that may advance our understanding of
etiology, pathophysiology, and treatment for mental disorders. Solutions to these problems may
be found by interweaving behavioral sciences with genetics, molecular biology, cellular biology,
and neuroscience.
In response to these issues with the DSM, the National Institute of Mental Health (NIMH)
began an initiative in 2008 known as the Research Domain Criteria (RDoC; National Institutes of
Mental Health, 2008). The RDoC calls for the “development … of new ways of classifying psycho-
pathology based on dimensions of observable behaviors and neurobiological measures” (National
Institutes of Health, 2014). The RDoC is a framework in-​progress, mapping current psychological
concepts onto seven units of analysis: genes, molecules, cells, brain circuits, physiology, behav-
ior, and self-​report. The RDoC also offers space to accommodate current and to-​be-​developed
18

Diverse perspectives of emotion 81

paradigms designed to elicit data relevant to putative psychological concepts. The ultimate goal
of the RDoC is to foster research that identifies relationships between the psychological and
biological phenomena central to mental illness and, to the extent possible, develop new, hybrid
constructs (Kozak & Cuthbert, 2016) and conceptualizations (detailed below in “Importing New
Concepts”) to better understand and improve mental health.
The research and developing perspectives on emotion and emotion regulation will be used in
this chapter to illustrate the application of the RDoC methods and models. The integration of
emotion regulation within the RDoC framework fosters intellectual cross-╉fertilization and inter-╉
disciplinary progress in understanding psychopathology. Furthermore, the controversy between
contemporary diagnostic systems and the RDoC mimics the controversy between the basic and
the dimensional conceptualization of emotion (e.g., Barrett, 1998; Barrett et al., 2007; Hamann,
2012; Izard, 2010, 2011; Panksepp & Watt, 2011). By extension, the field of emotion regulation
is also routinely criticized for issues of reliable conceptualization (e.g., Gross & Barrett, 2011;
Thompson, 1994). The RDoC, a unifying framework that is substantially agnostic theoretically,
can be a useful approach for organizing the current state of research on emotion and emotion
regulation. In doing so, it can identify key gaps on which to focus future inquiry. The aim of
this chapter is to describe the RDoC’s relevance to the study of emotions, their regulation, and
dysregulation. We outline how the tension between emphasizing DSM/╉ICD categories versus
the RDoC dimensions parallels historic disputes between views that emphasize discrete ver-
sus dimensional accounts of emotions. The two common views described here—╉the Bradley-╉
Lang view and the Gross-╉Ochsner view—╉are dimensional accounts of emotion that complement
rather than oppose common-╉sense discrete emotion (e.g., anger, sadness, fear, disgust, surprise,
happiness) views. The RDoC readily accommodates each of these views of emotion in the service
of understanding psychological disorders. At the same time, it preserves the option for viewing
them categorically.

Diverse perspectives of emotion


When William James stated “everyone knows what it is” with regard to the patterns of panic,
grief, anger, and sadness (James, 1884, p. 197), he would have never surmised that by 2016 the
field would fail to achieve consensus with regard to emotion. Bradley and Lang (2006) claimed
that “there are almost as many definitions [of emotion] as there are investigators.” Naturally,
defining emotion regulation presupposes an understanding of emotion, making the lack of an
agreed-╉upon definition of emotion a major impediment. Despite these conceptual obstacles, the
construct of emotion dysregulation is widely recognized as a common factor across psychiatric
disorders (Hofmann, Sawyer, Fang, & Asnaani, 2012; Kring, 2008). At its most basic level, emo-
tion dysregulation can be seen as simply problems with emotion regulation. Given this defini-
tion, possible classifications of emotion dysregulation can be proposed based on contemporary
theories of emotion and emotion regulation. Below we briefly outline some of the major views
of emotion and emotion regulation to provide a comparative context in which the RDoC is then
placed. A more thorough map of theories of emotions is presented by others (e.g., Gendron &
Barrett, 2009).

Emotion as action preparation, regulation as action change


As suggested by Bradley and Lang (2006, p. 589), one way to think of emotions is in terms of
their behavioral functions: “Emotion is, inherently, a disposition to act.” This follows from the
William James tradition of searching for ways to index the “organic reverberation” of emotion
(James, 1884)  that results in overt behavior and preparation for it by the autonomic nervous
28

82 Research Domain Criteria (RDoC) and Emotion Regulation

system. Two major ideas have emerged from the important work of Bradley and Lang. First is the
biphasic model of emotion, with the notion that emotions are organized along two motivational
systems: Approach and avoidance systems. The second is the bioinformational model of emotion,
the notion that emotion-​related information is linked to neural activity via propositions. This
second model posits that emotions are linked to neural activity that specifically represents the
organism’s response propositions (i.e., actions and action tendencies) distinct from its stimulus or
meaning propositions. In this section, we review insights derived from these models, explore what
they mean for emotion regulation, and evaluate them from the RDoC perspective.
The biphasic model of emotion is historically rooted in James’s hypothesis that physiology is
the basis for every emotional response—​that the peripheral physiology is the emotion. Based on
the James-​Lange theory of emotion, Bradley and Lang evaluated whether discrete emotions (e.g.,
fear, joy, sadness) have unique physiological profiles. Results from this line of inquiry have thus far
been largely disappointing (see Cacioppo, Berntson, & Hatfield, 1993). Instead, the physiological
reactions to emotion-​laden stimuli relative to non-​emotional stimuli support the idea that there
are several robust physiological indicators of emotion in general, such as electrodermal, skeletal
muscle, pupillary, and cardiac activity as well as numerous non-​invasive electromagnetic, opti-
cal, and hemodynamic metrics of central nervous system activity. These indicators are loosely
correlated with one another and form valenced motivational sets (or phases) of approach and
avoidance behaviors that clearly diverge at increasing levels of emotion arousal. In other words,
increases in the coupling of central and peripheral psychophysiological indicators seems to be
determined by whether the organism is motivated to approach or avoid based on whether the
focal stimulus in question (real or imagined) is appetitive or aversive. The regulation of emotion
from this standpoint is a stimulus-​induced change in one or more of these action indicators. Its
dysregulation therefore can be viewed as some inappropriate deviation from adaptive elicitation
of an approach or avoidance action set.
The biphasic motivational view complements learning and memory phenomena that play the-
oretically important roles in the etiology and pathophysiology of mental disorders and behav-
ioral dysfunction. For example, the conditioning equations described by Rescorla-​Wagner and
their theoretical progeny (Mackintosh, 1975; Pearce & Hall, 1980; Rescorla, Wagner, & others,
1972)  feature formulations that include summations of the intensity of both non-​emotional
conditional stimuli and emotional unconditional stimuli, as well as the organism’s learning rate.
In studies of conditioned fear or reward learning, typical physiological and behavioral patterns
related to avoidance and approach to unconditional stimuli are robustly elicited. As the associa-
tion between conditional stimuli and unconditional stimuli increases over repeated trials, the pat-
terns elicited in response to conditional stimuli increasingly resemble the response elicited by the
unconditional one, demonstrating the phenomenon of emotional learning—​one route of emotion
regulation to a stimulus.
Of course, emotions are elicited by imagined as well as environmental stimuli. Recognizing
this, Lang and Bradley posited that emotions are elicited through mentalization conceptualized
in a bioinformational theory of emotion (Lang, 1979), which they study through imagery para-
digms. This theory conceives of emotion as based on sensory, conceptual, and action informa-
tion in a network of propositions that includes biology as encoded neural action units. Sensory
information is coded as stimulus propositions (e.g., the spider is black), and conceptual informa-
tion is coded as meaning propositions (e.g., the spider is dangerous). Response propositions (not
stimulus or meaning propositions) are directly linked to emotions, since emotions are viewed as
actions and action tendencies: In this model, physiological activity that accompanies an emotion
is viewed as being part of, not a response to, the emotion (Miller & Kozak, 1993; Miller, 1996).
When imagining stimuli, stimulus-​induced neural activation that corresponds to features and
38

Diverse perspectives of emotion 83

meanings that do not result in action (e.g., “a black circle that is a shape” leads to no action)
is unemotional. Nonetheless, such activation may influence response proposition-╉related neural
activation that is emotional (e.g., “a black spider that is dangerous and should be avoided”, leads to
flinching upon presentation of a spider). Imagined stimuli may prompt a different physiological
profile than in vivo stimuli, in that the mental generation of such stimuli typically induces a non-╉
valence-╉specific cardiac acceleration (thus for both positive and negative imaginal stimuli), which
is viewed as reflecting emotional action engagement (e.g., Cuthbert, Vrana, & Bradley, 1991). In
contrast, in vivo stimuli such as pictures and sounds seem to elicit differential patterns of heart
rate and corrugator electromyography (EMG) based on affective valence, with negative and posi-
tive stimuli eliciting increases and decreases, respectively (Bradley, Codispoti, Cuthbert, & Lang,
2001; Bradley, Moulder, & Lang, 2005).
In the theories outlined above, relying on the three-╉systems view (Lang, 1968), emotions are
equated with action, manifested in language expression, central and peripheral physiology, and/╉
or overt behavior. Whereas non-╉human animals commonly show fairly close correspondence
between preparation and behavior (e.g., Panksepp, 1998), the link between motivational engage-
ment and overt action is weaker in humans, putatively because of the development of cortical
control mechanisms that implement a regulatory function. Emotion regulation can thus be under-
stood as a differential correspondence between central and peripheral physiological preparation
and behavior: A smile without approach, a shudder without avoidance. The three-╉systems view
foregrounds publicly observable data and intervening variables derived from them without pro-
posing hypothetical constructs that characterize their relationship to emotion as a concept (for a
review of the relationship between intervening variables and hypothetical constructs, see Kozak &
Miller, 1982; MacCorquodale & Meehl, 1948). A theory of emotion would articulate hypothetical
constructs as well as the bridge principles that connect observable variables with the constructs.
How a theory constructs these relationships sets the stage for how we understand emotion regula-
tion and its dysregulation. Some examples are delineated in Table 5.1.
Although such bridge principles can be specified, and the overall psychophysics involved can be
further refined, this view provides little with which to construct necessary and sufficient criteria
for understanding emotion dysregulation or psychopathology. What makes a changed action set
adaptive as opposed to maladaptive? How much deviation from normality can confer functional
impairment via emotion dysregulation? The Bradley and Lang view does not have much to say
about emotion regulation, and its extension to subsume that construct is not obvious. Thus, the
“emotion is action” perspective and its vast empirical database occupy a critical though incom-
plete place in the RDoC framework.

Emotion and emotion regulation—╉two steps of valuation


The view that motivated action is necessary and sufficient for emotion is not accepted by some
other theorists, who develop more elaborate sets of hypothetical constructs that are connected to
observable data by more elaborate and less explicated sets of bridge principles. For example, Barrett
and colleagues (e.g., Barrett, 2013; Barrett, Wilson-╉Mendenhall, & Barsalou, 2013; Harré, 1986)
consider mental construction as a core aspect of emotion. Appraisal processes also often loom large
(e.g., Scherer, 2009). A particularly popular view comes from a Gross-╉Ochsner understanding of
emotion regulation, which was initially derived from peripheral psychophysiological data (Gross
& Levenson, 1993) and has been elaborated more recently in a subdomain of psychophysiology
now known as social cognitive affective neuroscience that puts more emphasis on phenomena of
the central nervous system (Ochsner & Gross, 2005). These theories tend to involve much more
diverse cognitive machinery to separate cognitive, emotional, and observable phenomena much
more than do the Bradley-╉Lang models, with emotion being framed as a feeling state that may
48

84 Research Domain Criteria (RDoC) and Emotion Regulation

Table 5.1  Left-​most column: bridge principles that can be derived from the three-​systems view.
Examples of emotion regulation and dysregulation in neighboring columns. “→” symbolizes “is” or
“which means that”, “¬” symbolizes negation.

Emotion Bridge Principle Emotion Regulation Examples Emotion Dysregulation Examples


Physiology, behavior, or self-​ Regulated breathing, on-​ Maladaptive breathing, saccadic
reported experience is (loosely) target gaze, clear speech, eye movement not supportive
coordinated → and reported confidence in of task performance, tremors
a motivated action set; → the context of a job interview in voice, and reported desire to
a motivated action set occurred → reflects a lack of an avoidance leave the room in the context of
an emotion motivation that is functionally a job interview reflects avoidance
occurred appropriate. motivation that is not functionally
appropriate.
Stimulus propositions → Changing networked response Changing networked response
neurally encoded propositions to be adaptive, propositions to be maladaptive,
information¬→ i.e., “an interview is not e.g., “an interview is something
emotion; something to run away from.” that will make my heart race and is
Meaning propositions → something I will stammer and want
neurally encoded to run away from.”
information¬→
emotion;
Response propositions →
neurally encoded
information →
action →
emotion
a motivated action set can be The set of avoidance actions The set of avoidance actions can be
changed → can be reduced to only heightened or feature an additional
emotion can be changed hand sweating, reducing response of sweating hands,
(i.e. regulated) self-​reported feelings of increasing self-​reported feelings of
anxiousness. anxiousness.

prompt actions rather than as action dispositions and observable actions. In Gross’s view, “emo-
tion regulation requires the activation of a goal to up-​or down-​regulate either the magnitude or
duration of the emotional response” (Gross, Sheppes, & Urry, 2011, as cited in Gross, 2013, p. 359).
Here, goal activation holds the logical status of an unobservable psychological construct that serves
as the target of regulation, instead of observable psychophysiological action (or action preparation)
that is to be changed through regulatory behavior typical of the Bradley-​Lang models. Gross’s view
of emotion, which combines four major views of mentalized phenomena contributing to emo-
tion (basic, appraisal, psychological constructionism, and social constructionism), emphasizes the
commonalities among the four views, that 1) emotions elicit loosely coordinated changes in feel-
ing states, physiology, and behavior (a problematic triad that mixes inferred and observed phe-
nomena), 2) emotions unfold over time, and 3) emotions are context-​dependent (Gross, 2015a).
Gross’s initial conceptualization, a major contribution to this area of research, was the delineation
of emotion regulation as behavior classified along a time-​unfolding emotion-​generation process,
dubbed the process model (Gross, 1998; Kalisch, Wiech, Herrmann, & Dolan, 2006; Webb, Miles,
& Sheeran, 2012). The process model demarcates emotion generation as one that begins with a
situation, followed by the attendance to a stimulus, then to a cognition, and ends in an emotional
response that provides input and changes the situation (beginning the cycle over again). This model
provided a way to classify differing emotion regulatory behaviors by the time at which they were
58

Diverse perspectives of emotion 85

elicited in the emotion-​generation cycle. Specifically, emotions can be regulated by selecting one-
self into or out of a situation, known as situation selection strategies (e.g., when one chooses to
leave a noisy room while studying), changing the situation in some way, known as situation modi-
fication strategies (e.g., telling others to “stop all the commotion”), shifting one’s attention away
from an emotional stimulus, known as attentional allocation or attentional redeployment strate-
gies (e.g., putting on headphones or distracting oneself), changing one’s interpretations towards
an emotional stimulus to reduce the emotional impact, known as cognitive change or reappraisal
strategies (e.g., interpreting the midnight noise of studying as an integral part of a memorable
university experience), and changing one’s behaviors, known as response modulation (e.g., biting
one’s lip to avoid expressing contempt at the noisemakers). Further, emotion-​regulatory behaviors
can be directed at the self, known as intrinsic emotion regulation, or directed at others, known as
extrinsic emotion regulation.
In what are now classic studies, Gross demonstrated that regulatory strategies that occur early
in such an emotion-​generation process are more effective at downregulating indices of negative
emotion than are response-​modulation strategies, a dichotomization he termed antecedent-​
focused versus response-​focused emotion regulation (Gross & Levenson, 1993). Critically, he also
defined successful emotion regulation by relating the outcome to the individual’s goals. He pos-
ited that regulation is successful if the resulting emotion meets the regulator’s goals, regardless
of social norms or long-​term adaptive value (Gross, 1998; Gross & Thompson, 2007; Thompson
& Calkins, 1996). Gross was able to demonstrate that instructed cognitive reappraisal, an unob-
servable process, causes robust reductions in observable indicators of emotion (those featured
in the three-​systems model) relative to conditions where participants were instructed to emote
as normal or to use response-​focused expressive suppression (e.g., the resistance of expressing
visible emotion on the face; Gross & Levenson, 1993). Controlled comparisons have shown that
an instruction to change one’s interpretation reduces physiological responding and self-​reported
negative affect and alters facial behavior toward an emotional stimulus. At the same time, it was
found that expressive suppression had no effect on psychophysiological or self-​reported indica-
tors of negative emotional responding and some studies even showed prolongation of respond-
ing. Researchers then concluded that these strategies were maladaptive because it was presumed
that the individual’s goal was to reduce negative emotional responding. Following this reasoning,
empirical links have been found indicating negative predictive effects on important indices of
cognitive performance such as memory (Richards & Gross, 1999, 2000), social function (Butler
et  al., 2003; Gross & John, 2003), mental health (Gross & Muñoz, 1995), and physical health
(DeSteno, Gross, & Kubzansky, 2013). Later work sought to validate this understanding of emo-
tion regulation neurobiologically, to demonstrate both its temporality and to identify the top-​
down sources of cortical control (Goldin, McRae, Ramel, & Gross, 2008; Ochsner & Gross, 2005;
Thiruchselvam, Blechert, Sheppes, Rydstrom, & Gross, 2011). Findings generally indicate that
regions such as prefrontal cortex, orbital frontal cortex, and cingulate are active in paradigms of
negative emotion downregulation via instructed cognitive control and that cognitive reappraisal
activates these areas earlier than do expressive suppression strategies.
The process model view has since been extended by borrowing concepts from cybernetic control
(Carver & Scheier, 1982; Wiener, 1961). This updated view, dubbed the extended process model
(EPM; Gross, 2015a), achieved several things. Conceptually, it separated emotion generation from
emotion regulation, it defined emotion and emotion regulation in terms of valuations, it posited
and defined three explicit stages of the emotion regulation process (identification, selection, and
implementation), and it proposed two ways in which emotions may be dysregulated at each step
of the process (regulatory failure and misregulation). By using concepts from cybernetics, the
processes of both emotion generation and emotion regulation could be defined strictly through a
68

86 Research Domain Criteria (RDoC) and Emotion Regulation

goal-​instantiated information-​to-​action feedback loop. First, the information from the situational
context of the world (W; World) is perceived in a sensory array (P; Perception). Next, a discrepancy
calculation is made between the percept and the goal state. If this discrepancy is large enough that
it crosses some threshold, this information will be evaluated in the form of a “good for me/​bad
for me” proposition (V; Valuation). This valuation will generate allostatic emotional actions, both
mental and physical, in an effort to reduce this discrepancy. Information is finally fed back to pro-
cess the need for continued emotive action or the termination of emotive action (A; Action) if the
goal state(s) is met. Taken together, these stages are called the World, Perception, Valuation, Action
(WPVA) cycle (Gross, 2015a). The conceptual separation between emotion generation and emo-
tion regulation here is one of first-​order (a control system on a passive percept) vs. second-​order
cybernetics (a control system imposed on the control system that is emotion; Von Foerster, 2003).
In other words, emotion generation involves a calculated valuation of percepts made relative to
goal states, and these valuations can be generated in a multitude of ways. Indeed, various models of
emotion explain the generation of valuations in different ways, for example as an “affect program”
from the basic view, an appraisal, or as a result of a psychological or social construction. Emotion
regulation, on the other hand, is a valuation of emotions themselves. In summary, emotion regula-
tion begins with the process of evaluating a valuation (Gross, 2014).
Emotion regulation involving an evaluation of a valuation is described not as one but as three
sequential WPVA cycles (see Figure 5.1). Specifically, these are 1) the emotion identification cycle,
2) the emotion regulation strategy selection cycle, and finally 3) the emotion regulation strategy
implementation cycle (Gross, 2014, 2015a, 2015b). These cycles and their components imply a
host of ways emotion regulation can go wrong. Gross posits two classes of such dysregulation.
The first class is emotion-​regulation failure, which refers to the lack of appropriate emotion regu-
lation when it would serve one’s goals. The second class is emotion misregulation, the enacting
of regulatory behavior that is counterproductive to one’s goals (Gross, 2015a, 2015b). We further
observe that emotion-​regulation failure or emotion misregulation can occur at either perception
or valuation stages in each cycle. For example, in the emotion identification cycle, perception fail-
ure is a failure to detect a counterproductive emotional display (e.g., Michael impulsively screams
in a fit of rage at a cashier while trying to return an item, leading him to be kicked out of the store
without receiving a refund), whereas valuation failure is a failure to understand that one’s emo-
tional display is counterproductive to one’s goals (e.g., Meghan decides to scream at a cashier, not
understanding that this behavior will not make the cashier more likely to help her). Misregulation
can be exemplified in the emotion identification cycle as well, where perception misregulation
might involve thinking that one has or will have panic symptoms (ostensibly and ironically to

Emotion Generation

W P V A
Emotion Regulation
W P V A W P V A W P V A
Identification Selection Implementation
Figure 5.1  The extended process model of emotion generation and regulation
Reproduced from James J. Gross, The Extended Process Model of Emotion Regulation: Elaborations, Applications,
and Future Directions, Psychological Inquiry, 26 (1), pp. 130–​137, doi.org/​10.1080/​1047840X.2015.989751,
Copyright © 2015 Routledge, with permission.
78

The takeaway 87

avoid feeling panic) and valuation misregulation might involve a valuation that one’s heart rate
represents anxious feelings (a “bad-╉for-╉me” valuation) in a social evaluative situation, leading to
greater attention to one’s heart rate, which serves to further increase one’s anxiety. As the example
illustrates, the distinctions between perception and valuation, as well as regulation failure and
misregulation, will likely have mechanistic and conceptual overlap, as each will implicate the
other. Nonetheless, it is a useful nuance that fosters the construction of hypotheses regarding the
specific mechanistic implementations of emotion regulation.
Ochsner and Gross have worked to elaborate the neural implementations of emotion regulation
from the EPM viewpoint (Ochsner & Gross, 2014). This research has demonstrated that regions
including dorsal medial prefrontal cortex, rostral medial prefrontal cortex, ventral medial pre-
frontal cortex, ventral striatum, amygdala, and insula provide neural implementation of various
valuation systems. Regions such as dorsal anterior cingulate, dorsal posterior medial prefrontal
cortex, dorsolateral prefrontal cortex, ventrolateral prefrontal cortex, and inferior parietal lobe
implement various regulatory control systems (Ochsner & Gross, 2014). Acknowledging the mul-
tifaceted nature of valuation itself, they propose that valuation can range from stimulus-╉generated
core valuations (e.g., amygdala-╉implemented threat or ventral striatum-╉implemented reward),
context-╉derived valuations (e.g., ventral medial prefrontal cortex-╉implemented contextual under-
standing) and conceptual valuation (e.g., dorso-╉medial prefrontal cortex-╉implemented concep-
tual/╉categorical understanding). The insula notably plays a role in all valuation processes in that it
is posited to account for interoceptive awareness when calculating values in a way that resembles
the James-╉Lange view (Ochsner & Gross, 2014).
Summarily, in the EPM view, emotions and emotion regulation are conceptually separate and
maintain their logical status as unobserved hypothetical constructs. The valuation component of
the EPM incorporates private data. Therefore, this viewpoint does not lend itself to the construc-
tion of bridge principles between emotion or emotion regulation and empirical indices. Brain
regions important for emotion regulation are conceptualized as either control areas or valua-
tion areas. The valuation areas can themselves be sorted into regions central to core evaluation
(important for emotion generation) or central to contextual evaluation (important for appraisal
and meaning creation). Both psychological and neuroanatomical conceptualizations of emotion
regulation based on the EPM imply a taxonomy of dysregulation and possible sources of dys-
regulation in the brain (see section “Scientific Advancement of Emotion Regulation with RDoC-╉
Consistent Thinking” below).

The takeaway
The Bradley and Lang view, supported by a wealth of empirical data, allows for a bridging between
the hypothetical construct of emotion (or specific, more granular concepts demonstrating clear
relationships with emotion) and hallmark psychophysiological observables (e.g., hemodynamic
and electromagnetic neuroimaging, skin conductance, heart rate). It also readily complements
research in other areas of psychology such as motivation, attention, learning, and memory. While
this view allows for private data such as subjective awareness, it does not describe a role for them,
which some other theorists believe should be in a comprehensive theory of emotion and psycho-
pathology. The Gross and Ochsner EPM offers an elaborated, multiple-╉process model that includes
a taxonomy of emotion dysregulation, testable across multiple implementing neurobiological
mechanisms, yet remains highly speculative. They do not equate emotion with its indicators, but
they make many assumptions about private data that cannot yet be tested given limitations in the
articulation of the model to date. The EPM appears to assume a parity between unobservables and
observables (e.g., valuation causes mental and physical action) in a way that the Bradley and Lang
8

88 Research Domain Criteria (RDoC) and Emotion Regulation

view does not. Arguably, the virtue of each of the two views is its weakness: Bradley and Lang are
more circumspect, so data and construct map well over but cover more limited terrain, whereas
Gross and Ochsner reach further conceptually and further from observable phenomena.

The research domain criteria


It is possible to study emotion without terms such as joy, fear, or sadness, or any other classifica-
tion of discrete emotion. As reviewed by Tracy and Randles (2011), contemporary theories of dis-
crete emotions vary in the criteria used to define emotion. This results in lists that can vary from
four—╉happiness, sadness, fear, and anger (Ekman & Cordaro (2011); Izard (2011), happiness is
labelled enjoyment by Levenson (2011), and PLAY by Panksepp and Watt (2011)), while sadness
and anger are respectively labelled PANIC/╉GRIEF and RAGE by Panksepp and Watt (2011)—╉to
as many as 11 emotions (along with the aforementioned ones, disgust (Ekman & Cordaro, 2011;
Izard, 2011; Levenson, 2011), interest (Izard, 2011; Levenson, 2011) or SEEKING (Panksepp &
Watt, 2011), contempt (Ekman & Cordaro, 2011; Izard, 2011), love (Levenson, 2011) or LUST
(Panksepp & Watt, 2011), CARE (Panksepp & Watt, 2011), relief (Levenson, 2011), and surprise
(Ekman & Cordaro, 2011)). Both the Bradley-╉Lang view and the Gross-╉Ochsner EPM comple-
ment discrete emotion classification. In the same way, the RDoC presents a complementary
dimensional framework for diagnostic categorization (Hyman, 2010). Thus, as with the study
of emotion, one need not invoke diagnostic labels of depression, anxiety, or schizophrenia (or
any other label) to study psychopathology. The dimensional approach invites investigators from
multiple disciplines to study the constructs that underlie mental health and mental illness with a
different emphasis than that of the DSM tradition.
The metaphorical workspace provided by the RDoC is represented visually as a matrix of
rows and columns. The rows represent psychological constructs of interest organized in broad
domains of neurobiological organization. (A few are arguably both psychological and biologi-
cal; more on such a merger below.) These domains include: Negative Valence Systems, Positive
Valence Systems, Cognitive Systems, Social Systems, and Arousal/╉Regulatory Systems. Within
each domain are more specific psychological constructs, selected because they are concepts with
a relatively well-╉developed supporting neuroscience literature. A  current list of the constructs
and sub-╉constructs that have been agreed upon can be found on the RDoC website (http://╉www.
nimh.nih.gov/╉research-╉priorities/╉rdoc/╉constructs/╉rdoc-╉matrix.shtml; (National Institutes of
Mental Health, 2016). They represent specified functional dimensions of behavior that can be
characterized across units of analysis. The columns represent these units: genes, molecules, cells,
brain circuits, physiology, behavior, and self-╉report. A final column includes the paradigms used
to elicit the putative constructs of interest. This array of measurable psychological and biologi-
cal phenomena is intended to eventually inform interventions (Kozak & Cuthbert, 2016; Miller,
Rockstroh, Hamilton, & Yee, 2016; Yee, Javitt, & Miller, 2015).
Table 5.2 displays the matrix domains and units, with example targets and citations of research
that could reside in each cell. Of course, the vast, multidisciplinary breadth of this work precludes
a complete presentation of the extant literature in such a matrix, even within cells, let alone rela-
tionships between cells.
The matrix is a research framework that represents the consensus of experts attending a series of
NIMH workgroups, one for each domain. The goals of these workgroups were to form and clarify
formal definitions of each construct within each domain. The matrix is incomplete and is meant
to be a work in progress (Cuthbert, 2014). A common misunderstanding is that the matrix is the
RDoC, when in fact it just an initial proposal, an exemplar (Miller et al., 2016). Further work is
expected to refine the matrix. Researchers are encouraged to evaluate, modify, and further define
98
Table 5.2  The RDoC Matrix with relevant exemplar literature citations

Units of Analysis
Domains Genes Molecules Cells Circuits Physiology Behavior Self -​Report Paradigms
Negative Valence Systems

Constructs: e.g. 5-​HTT (Hariri & e.g. Serotonin e.g. GABAergic cells e.g. anterior cingulate e.g. fear e.g. avoidance e.g. Behavioral e.g. Fear
e.g. Fear (Hartley & Phelps, Holmes, 2006) (Krakowski, 2003) (Capogna, 2014) cortex-​amygdala potentiated startle (Schmader & Lickel, Inhibition Schedule Conditioning
2010) (Hung, Smith, & Taylor, (Lissek et al., 2008) 2006) (Carver & White, (Hermans, Craske,
2012) 1994) Mineka, & Lovibond,
2006)

Positive Valence Systems

Constructs: e.g. DRD2 (Peciña e.g. Dopamine e.g. dopaminergic e.g. nucleus e.g. postauricular e.g. approach e.g. Behavioral e.g. Appetitive
e.g. Approach Motivation et al., 2013) (Salgado-​Pineda, neurons (Chinta & accumbens—​ventral reflex (Benning, (Schmader & Lickel, Activation Conditioning (Shabel
(Wacker, Mueller, Pizzagalli, Delaveau, Blin, & Andersen, 2005) tegmentum Patrick, & Lang, 2006) Schedule (Carver & & Janak, 2009)
Hennig, & Stemmler, 2013) Nieoullon, 2005) (Der-​Avakian & 2004) White, 1994)
Markou, 2012)

Cognitive Systems

Constructs: e.g. COMT (Drabant e.g. Glutamate, e.g. Pyramidal e.g. dorsolateral e.g. pupillometry e.g. distraction e.g. Emotion e.g. Reappraise/​
e.g. Cognitive Control et al., 2006) GABA (Stan et al., cells (Helmeke, prefrontal cortex-​ (Silk et al., 2009) behaviors Regulation Watch—​Negative vs.
(Ochsner, Silvers, & Buhle, 2014) Ovtscharoff, Poeggel, amygdala (Goldin (Thiruchselvam Questionnaire Neutral (Goldin et al.,
2012) & Braun, 2001) et al., 2008) et al., 2011) (Gross & John, 2008)
2003)

Systems for Social Processes

Constructs: e.g. OXTR (Kim e.g. Oxytocin e.g. Fusiform gyrus e.g. fusiform e.g. eye contact e.g. emotion e.g. Affect e.g. Dyadic
e.g. Social Communication et al., 2011) (Quirin, Kuhl, & neurons (Pizzagalli gyrus—​amygdala (Adams Jr & Kleck, recognition Valuation Inventory conversation tasks
(Laurent & Powers, 2007) Düsing, 2011) et al., 2002) (Faivre, Charron, Roux, 2005) performance (Tsai, Knutson, & with one member of
Lehéricy, & Kouider, (Szanto et al., 2012) Fung, 2006) the dyad instructed
2012) to regulate (Richards,
Butler, & Gross,
2003)

Arousal and Regulatory Systems

Constructs: e.g. Apoe4 e.g. Amyloid e.g. Pineal cells e.g. hypothalamic-​ e.g. cortisol (Lam, e.g. sleep (Mauss, e.g. Self-​ e.g. Actigraphy
e.g. Arousal (Cuthbert, (Delano-​Wood plaque deposition (Waider, Araragi, pituitary-​adrenal Dickerson, Zoccola, Troy, & LeBourgeois, Assessment (Baum et al., 2014)
Schupp, Bradley, Birbaumer, et al., 2008) (Sturm et al., Gutknecht, & Lesch, gland axis (Laurent & & Zaldivar, 2009) 2013) Manikin Arousal
& Lang, 2000) 2013) 2011) Powers, 2007) Scale (Bradley &
Lang, 1994)
09

90 Research Domain Criteria (RDoC) and Emotion Regulation

constructs, as well as consider new constructs as warranted. Indeed, NIMH leadership advocates
development of “hybrid” psychological-​biological constructs (Kozak & Cuthbert, 2016). Perceived
gaps or paucities in certain cells could represent potential avenues of valuable research (Cuthbert
& Insel, 2013; National Institutes of Mental Health, 2014).
The RDoC matrix represents three efforts toward validity. The first is to provide descriptive
validity by answering questions about how largely psychological constructs are implemented bio-
logically (mechanistic). This is represented in the columns of genes, molecules, cells, and brain
circuitry and will involve research in genetics, endocrinology, immunology, neuroscience, etc.
The second is to increase our ability to validly explain neurobiological functions through a pre-
liminary consensus on the questions pertaining to why a biological system works the way it does
or for what reason a neurobiological structure is as it is. This consensus is represented by the rows
of domains and constructs, which will involve the efforts of psychology working in tandem with
neuroscience and other disciplines. The third effort addresses predictive validity by providing sci-
entific evidence that may answer what might happen should changes in a certain unit of analysis
occur. In the clinical care of mental illness, such answers may be most important for quality of
life, indexed most prominently in the columns of physiology, overt behavior, and self-​report. This
research will involve a cross-​fertilization of research from the fields of psychology, epidemiology,
and sociology.
As it pertains to emotion, emotion regulation, and emotion dysregulation, the RDoC matrix
comfortably accommodates the concepts presented by Bradley and Lang. Arousal exists as a con-
struct within the Arousal and Regulatory Systems domain, and motivational valence is repre-
sented in two domains (Negative Valence System and Positive Valence Systems). Related concepts
of attention and perception reside in the Cognitive Systems domain. The three systems of prepa-
ratory psychophysiology, overt behaviors related to approach and avoidance, and self-​report are
represented in the physiology, behavior, and self-​report columns respectively. The matrix also
accommodates the concepts of Gross’s process model:
• Situation selection and modification belongs in Positive Valence System Domain > Approach
Motivation > Action Selection/​Preferential Decision Making.
• Attentional deployment belongs in Cognitive Systems > Attention and Perception.
• Cognitive appraisal/​reappraisal belongs in Cognitive Systems > Perception, Language, and
Cognitive Control.
• Response modulation/​suppression belongs in Cognitive Systems > Cognitive Control >
Response Selection; Inhibition/​Suppression.
It is abundantly clear that even these concepts can be, and perhaps need to be, unpacked.
Cognitive reappraisal, for example, involves perception, language, and cognitive control (see
McRae, Ciesielski, & Gross, 2012, for such an analysis of cognitive reappraisal).
One of the characteristics that differentiates the RDoC from the DSM is that the RDoC constructs
are selected to be firmly grounded in research on neuroscience phenomena (e.g., Infantolino,
Crocker, Heller, Yee, & Miller, in press). To date, hemodynamic neuroimaging studies focused on
emotion-​regulation strategies such as those developed by Gross (1998) have primarily focused on
fear conditioning (e.g., Milad, Rosenbaum, & Simon, 2014; Phelps, Delgado, Nearing, & LeDoux,
2004) and cognitive reappraisal (Banks, Eddy, Angstadt, Nathan, & Phan, 2007; Ochsner & Gross,
2008; Ochsner et al., 2012), although some work has examined emotional suppression (Goldin
et  al., 2008). Fear conditioning paradigms commonly include phases of fear habituation, fear
acquisition, fear extinction, and fear extinction recall. These components of fear learning involve
both unique and shared neural circuitry. Subdivisions and subnuclei of the amygdala, as well as
the dorsal anterior cingulate, are considered central to fear acquisition (Milad et al., 2014). The
19

The takeaway 91

basolateral complex of the amygdala, hippocampus, and infralimbic region of the medial prefron-
tal cortex are thought to interact during fear extinction (Milad & Quirk, 2012). In terms of cogni-
tive reappraisal paradigms, a meta-​analysis of 48 hemodynamic neuroimaging studies revealed
involvement of cognitive control regions, including dorsomedial prefrontal cortex, dorsolateral
prefrontal cortex, ventrolateral prefontal cortex, and posterior parietal lobe, as well as bilateral
amygdala (Buhle et al., 2014). Interestingly, this meta-​analysis did not find evidence of differential
ventromedial prefrontal cortex activation during cognitive reappraisal, a region that had been
thought to be important for this process in previous studies (e.g. Diekhof, Geier, Falkai, & Gruber,
2011). Functional connectivity analyses using EEG, MEG, fMRI, PET, and optical imaging meth-
ods also provide valuable information regarding the relationships between brain regions during
emotion-​regulation paradigms. Functional connectivity is assessed by examining the coactiva-
tion of neural regions; regions are said to be functionally connected if their activity increases or
decreases in tandem. These analyses have demonstrated significant positive correlations of ven-
tromedial prefrontal cortex and hippocampus with amygdala during fear extinction recall (Milad
et al., 2007), and dorsolateral, dorsal medial, anterior cingulate, and orbital cortices correlate with
amygdala during cognitive reappraisal (Banks et al., 2007).
For an investigation to be aligned with the RDoC framework, it should assess specific
psychopathology-​related constructs (whether basic or clinical), preferably across multiple units
of analysis. For an investigation to make progress pertinent to the RDoC aim of constructing and
clarifying conceptual granulanda, the “anchor” unit of interest might be brain circuitry (Insel
et al., 2010; Insel & Cuthbert, 2015). This focus on brain circuits, rather than brain regions in iso-
lation, reflects an important shift in neuroscience toward viewing the brain more often as a highly
interconnected and dynamic organ.
The matrix is not yet structured to foreground the effects of the environment (e.g., parenting,
culture), development (e.g., stages, such as infancy, childhood, adolescence, adulthood, and old
age, relevant biological processes, such as puberty and senescence), and learning history. These
could be conceived as dimensions orthogonal to the matrix, thus related to all of its cells, and
often must be taken into consideration when making a determination about what is healthy versus
abnormal (National Institutes of Mental Health, 2012). One could imagine a two-​dimensional
matrix for each cultural context and for each stage of development, both powerful forces that
shape one’s learning history. For example, developmentally, outward exuberance may not be dys-
regulatory for children too young to understand contexts where such displays are inappropriate
(e.g., a three-​year-​old laughing at a solemn funeral). Culturally speaking, exuberance may not
necessarily be a positively construed experience (Tsai et al., 2006) or even functional for the indi-
vidual (Tsai, Sun, Wang, & Lau, 2016) if one’s culture imposes limits upon positive emotional dis-
plays (Matsumoto, 1990). Therefore, one of the RDoC-​relevant goals will be to reliably and validly
identify for whom, where, and at what stage to predict the development of emotion dysregulation
difficulties across the lifespan to help guide intervention and prevention.
Problematic functional configurations in the Negative Valence System might be labeled anxi-
ety, whereas problematic configurations in the Positive Valence System domain might be labeled
depression, and so on. The RDoC matrix itself presents no obstacle to this kind of categorical
classification. Indeed, complementation would be preferred, as it would be unwise to disregard
decades of DSM-​based clinical research. However, the bulk of the research used to support diag-
nostic classification systems is limited to one unit of analysis: self (or clinician) report. As emo-
tion scientists know well, self-​reports are ultimately unsatisfactory due to their dependence on
cultural norms and individual differences in disclosure (Bradley & Lang, 2006). Some would
even consider them flawed hypotheses about one’s own functioning (Kozak & Miller, 1982; Miller
& Kozak, 1993). By encouraging study and integration of multiple units of analysis, the RDoC
29

92 Research Domain Criteria (RDoC) and Emotion Regulation

framework overcomes limitations inherent to sole reliance on self-╉report without dismissing


them. Furthermore, even though the RDoC can complement existing nosology, it offers starting
points for scientific advancement by highlighting heterogeneity within these existing categories
(e.g., anhedonia within the broader category of depression).
RDoC investigators can choose to advance the field in several ways. They may take an empiri-
cally expansive approach, such as clarifying emotion concepts by taking into account more units of
analysis from the RDoC framework. They may take a conceptually granular approach. This might
involve subdividing existing constructs and subconstructs into components that more closely
match the resolution of the anchoring unit of analysis, introducing new theoretical concepts,
or investigating endophenotypes that result in smaller causal chains (more on this in “Scientific
Advancement of Emotion Regulation with RDoC-╉Consistent Thinking” below). They can choose
to advance the field by developing bridge principles, such as by ascertaining cutoffs or a profile
of multiple units of analysis that is sensitive to dysfunction. They can develop robust laboratory
paradigms that reliably elicit constructs of interest. Scientific progress is guided and sometimes
constrained by the availability of experimental paradigms and stimuli, and these paradigms often
need to be as natural as possible to reflect naturalistic and ecologically valid elicitation of emotion,
and its regulation, and dysregulation.
The RDoC matrix provides space for intellectual flexibility. Instead of appealing to unob-
servable constructs derived from folk theories of psychology that are latent and hypotheti-
cal in nature (Bentler, 1980)  (e.g., stress), we can examine observable principal components
(Jolliffe, 2002) (e.g., HPA-╉axis release of corticotropin releasing hormone) from multiple units
of analysis. These principal components may even serve as a bridge to laymen’s folk concepts.
To be sure, the RDoC is not yet that bridge, and RDoC users must avoid the unhelpful concept
reification that came to characterize traditional classification systems. Metaphorically speak-
ing, the clarification of constructs in each row reinforces the constitution of the bridge toward
a well mapped understanding of psychological function and biological function. Developing
rows representing more fine-╉grained constructs will tell us where to take the next step. Both
clarification and development will be primary tasks for the discipline of psychology at large,
with input from investigators in the areas of emotion generation, regulation, and dysregulation,
broadly conceived.

Scientific advancement with RDoC-consistent thinking


Empirical advancement: Genetic, molecular, and cellular units
of analyses
Since its introduction in 2008, the RDoC has spurred a great deal of research on emotion regula-
tion along multiple units of analysis. There has been substantial growth in the literature contain-
ing the words “emotion,” “emotion regulation,” and “emotion dysregulation” in combination with
keywords at each unit of analysis. RDoQ displays a cross-╉tabulated index and visual literature map
(see Table 5.3 and Figures 5.2a, 5.2b, 5.2c) of PubMed articles of combined search terms, which,
although crude, can provide some preliminary insight in future areas of growth in the study of
emotion regulation (RDoQ, 2016; http://╉datamining.cs.ucla.edu/╉rdoq/╉).
Given the relative citation numbers, it is clear that connections between genes, molecules, cells,
and emotion warrant far more study. Studying genes as they relate to emotion dysregulation using
the RDoC approach may be particularly fruitful. For example, the serotonin transporter gene
(5HTTLPR) has been identified as a candidate gene for populations diagnosed with DSM-╉IV
Major Depressive Disorder, but this has not been replicated in genome-╉wide association studies,
39

Table 5.3  Cross-​tabulation of new publications containing the keywords “Emotion,” “Emotion


Regulation,” and “Emotion Dysregulation” with important keywords pertaining to each unit of analysis
for the years 1976, 1995, 2000, 2008, 2010, and 2015.

1976 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms


Emotion 5 26 45 142 6
Emotion 1 4 7 4
Regulation
Emotion
Dysregulation
1995 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms
Emotion 2 33 45 106 1619 98
Emotion 4 8 13 49 4
Regulation
Emotion 1 4 1
Dysregulation
2000 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms
Emotion 3 102 46 117 2380 186
Emotion 12 8 15 83 8
Regulation
Emotion 3 1 8 2
Dysregulation
2008 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms
Emotion 75 5 490 60 153 4861 564
Emotion 16 81 8 19 265 44
Regulation
Emotion 15 2 29 4
Dysregulation
2010 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms
Emotion 115 5 660 78 187 5853 706
Emotion 28 112 9 21 354 65
Regulation
Emotion 24 1 3 46 9
Dysregulation
2015 Genes Molecules Cells Circuits Physiology Behaviors Self-​Reports Paradigms
Emotion 733 240 440 1326 2808 19848 10124 1212
Emotion 215 59 80 252 408 2413 911 141
Regulation
Emotion 15 2 7 52 66 375 175 21
Dysregulation

Data from RDoQ search retrieved February 25, 2016 (http://​datamining.cs.ucla.edu/​rdoq/​).


49
(a) 14000
Since RDoC

7000
"Emotion"

0
1976 1979 1982 1985 1988 1991 1994 1997 2000 2003 2006 2009 2012 2015

300 300 300 300

+ ”Genes” + ”Psychophysiology”
150 150 + ”Cells” 150 150 + ”Self-Report”

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012
300 300 300 300

+ ”Behavior” + ”Paradigms”
150 + ”Molecules” 150 + ”Circuits” 150 150

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012
59
(b) 2000

Since RDoC

"Emotion Regulation"
1000

0
1976 1979 1982 1985 1988 1991 1994 1997 2000 2003 2006 2009 2012 2015

2000 2000 2000 2000

1000 1000 1000 1000


+ ”Genes” + ”Cells” + ”Physiology” + ”Self-Report”

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012
2000 2000 2000 2000

1000 1000 1000 1000


+ ”Molecules” + ”Circuits” + ”Behavior” + ”Paradigms”

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012

Figure 5.2  (Continued)
69
(c) 300
Since RDoC

150 "Emotion
Dysregulation"

0
1976 1979 1982 1985 1988 1991 1994 1997 2000 2003 2006 2009 2012 2015

300 300 300 300

150 150 150 150


+ ”Genes” + ”Cells” + ”Psychophysiology” + ”Self-Report”

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012
300 300 300 300

150 150 150 150


+ ”Molecules” + ”Circuits” + ”Behavior” + ”Paradigms”

0 0 0 0
1976 1994 2012 1976 1994 2012 1976 1994 2012 1976 1994 2012

Figure 5.2 New publications per year containing the keywords (a) “Emotion,” (b) “Emotion Regulation,” and (c) “Emotion Dysregulation” in combination
with words pertaining to each unit of analysis, curated by the NIH-​funded UCLA RDoQ Tool as a predefined term set. Shaded area represents the years that
the RDoC initiative was implemented. Note that year-​by-​year numbers and figures do not represent cumulative publications.
© Michael Sun, 2016.
79

Scientific advancement with RDoC-consistent thinking 97

likely in part due to the heterogeneity inherent in this diagnosis (Bosker et  al., 2011). Results
of candidate-​gene research and genome-​wide association studies may converge should the field
move toward defining emotion regulatory phenotypes from the RDoC perspective. For example,
research that selects individuals based on their scores on measures of positive valence may reveal
new genetic relationships as they relate to specifically defined constructs of dysregulation, such as
failure in reward learning. Future study of molecular and cellular units of analysis in the context
of the RDoC, not only diagnostic categories, will also surely build upon existing research. The role
of dopamine receptors in reward processing is one example of existing research that fits within the
RDoC framework because this process is not unique to a specific diagnostic category (e.g., Peciña
et al., 2013; Pizzagalli et al., 2008; Vrieze et al., 2013). The RDoC is well positioned to foster and
benefit from research on potential endophenotypes, which bridge gene and disease expression
(such as major depression). Research on endophenotypes pursues narrower relationships and
shorter, intermediate causal chains (Miller & Rockstroh, 2013; Miller et al., 2016) and can readily
accommodate specific roles for emotion dysregulation in mental illness. Endophenotypes can be
used to parse the genetics of emotion dysregulation into smaller, more tractable components. It
is presumed that such components would have simpler genetic architectures than a disorder itself
(Goldstein & Klein, 2014).
In investigations that seek to clarify psychopathology, the emphasis on microbiological or
behavioral units of analysis need not and should not distract from considerations of clinical
practice and understanding. A vision for the future of clinical research must be complemented
with an understanding of present needs in the field. We recommend that researchers embark-
ing on RDoC-​congruent investigations of emotion dysregulation consider questions of applica-
tion, whether that be in patient conceptualization or clinical decision making. Clinical science
has traditionally focused on impaired function and subjective distress (Antony et al., 1994), and
replacing this with a reductionistic focus on a “gene/​molecule/​cell/​circuit for psychopathology”
impoverishes rather than enriches the phenomenological understanding of psychiatric disorders
(Miller, 2010). Research on hybrid psychology-​biology concepts with diverse degrees of granular-
ity may improve assessment in such a way that prediction of dysfunction and disease course could
be made earlier. Furthermore, empirical understanding at multiple levels can provide better cut
points and more finely tuned bridge principles that can help delineate function from dysfunction,
including healthy regulation from dysregulation. Holistic assessment of an individual can result
in precision clinical care on a personal level (Insel & Cuthbert, 2015). Finally, new treatments may
be developed through a combined use of pharmacology, behavior, and mediating technological
devices (Craske, Meuret, Ritz, Treanor, & Dour, in press; Cuthbert, 2014).
How has RDoC-​inspired thinking already advanced clinical understanding and clinical prac-
tice? One example comes from the study of fear-​based disorders. From an observational stand-
point both in the laboratory and in clinical practice (the same standpoint that brought us the
DSM), it was surmised that repeated exposure to conditional fear stimuli reduced fearful respond-
ing (Foa & Kozak, 1986). The extinction learning laboratory paradigm and its clinical analogue,
exposure therapy, were birthed from this observation, and the latter is regularly employed for
individuals with anxiety disorders (in which excessive fear responding is a central dysregulated
mechanism). It was believed that reduced fearful responding evidenced the erasure of fearful
associations between conditional and unconditional stimuli, and efforts were made to ensure
that a reduction in fear responding was observed in-​session (van Minnen & Hagenaars, 2002).
However, efforts to understand emotional learning and memory across multiple disciplines led
investigators to conclude that the reduced response was actually due to the formation of new,
non-​threat-​associated memories of the conditional stimulus, contrary to the widely held view of
exposure therapy. These new inhibitory memories need to be consolidated over time (typically
89

98 Research Domain Criteria (RDoC) and Emotion Regulation

over 24 hours). Evidence of short-╉term reduction in fearful responding (e.g., habituation) during
exposure therapy sessions is no longer viewed as essential to clinical progress. Indeed, it has been
found that short-╉term extinction does not predict long-╉term extinction learning (e.g., Brown,
LeBeau, Chat, & Craske, in press; Peters, Dieppa-╉Perea, Melendez, & Quirk, 2010). These findings
have prompted a revised Emotional Processing Theory (Rauch & Foa, 2006). This new theory has
the effect of clarifying how fear regulation occurs in nature (inhibitory learning as opposed to
habituation), as well as increasing therapeutic efficiency by saving a great deal of clinician and cli-
ent effort. Although these advances predate the RDoC proper, it is this type of empirical advance-
ment that the RDoC seeks to systematically reproduce more broadly for emotion regulation.

Conceptual advancement: Hybrid matrices


Researchers coming from other perspectives of emotion, emotion regulation, or emotion dys-
regulation may find the transition to the RDoC approach challenging. Indeed, as illustrated in
Table 5.4, concepts from the EPM perspective may be difficult to place cleanly within the RDoC
matrix. It should be acknowledged that functional conceptualization can be evaluated on dimen-
sions according to the degree of heuristic value (likely a subjective evaluation), as well as its dis-
tance from a reliably observable unit of analysis (an objective one). These concerns are addressed
by the EPM through the invocation of a single valuation unit: The WPVA cycle, which scales flex-
ibly and accommodates real-╉life biological information-╉processing such as DNA (e.g., Santini,
Bath, Turberfield, & Tyrrell, 2012)  or neural systems implementing memory (e.g. Hampson,
Hedberg, & Deadwyler, 2000). They are also apparent in the RDoC’s matrix organization and
inclusion of granular subconstructs within constructs.
Table 5.5 is an example of designing an RDoC-╉type matrix to accommodate a Gross-╉Ochsner
perspective on emotion generation, regulation and dysregulation. With such a strategy, more
granular concepts than those already included in the RDoC matrix can eventually be imported
when sufficient evidence accumulates.
The example hybrid-╉matrix provided would, like the RDoC, be an imperfect framework subject
to revision. For example, the construct of extrinsic regulation may very well be subsumed under
intrinsic regulation, as the motivation to extrinsically regulate could arise as a way of regulating
one’s intrinsic emotions (e.g., a mother shushing her baby as it cries in public, serving to regulate
the mother’s own anxiety and embarrassment, as well as the child’s). For the sake of organizational
and visual clarity, intrinsic and extrinsic regulation are presented separately. Many of the subcon-
structs, such as those involving valuation, may be mediated by the same mechanism. Indeed, most
of these constructs have yet to be well-╉defined neurally. The hybrid framework serves as an illus-
tration for how evolving conceptualizations that are independent of the RDoC may give rise to
alternative concepts, spurring more research to validate or reject these notions. Having multiple
imperfect frameworks may be valuable to pit strong hypotheses against each other. Core dysregu-
latory constructs, such as invoked depressogenic schemas in major depressive disorder targeted
by Cognitive Behavioral Therapy, might be more analogous to the construct of [cognitive] genera-
tive action (Beck, 1976). The hypergeneration followed by a slow return to baseline in borderline
personality disorder conceptualized using Dialectical Behavioral Therapy (Linehan, 1993) may be
more analogous to generative action (for the hypergeneration) and intrinsic emotion regulatory
stopping action (for the slow return to baseline). The experiential avoidance explanation respon-
sible for psychopathology transdiagnostically in Acceptance and Commitment Therapy (Hayes,
Strosahl, & Wilson, 1999) may be more analogous to the action of misregulated emotion identi-
fication. These associations appear to have better associability than what may be offered from the
9
Table 5.4  A rough mapping between Gross-​Ochsner emotion regulatory concepts and the RDoC constructs.

Domains Constructs
Negative Valence Acute Threat Potential Harm Responses Sustained Threat Frustrative Non-​Reward Loss
Systems Responses Responses

Emotion generation to Emotion generation to Emotion regulatory Emotion generation to Emotion generation to loss
acute threat potential threat maintenance, stop, and frustration, Non-​reward
switch to sustained threat Valuation

Positive Valence Approach Initial responsiveness to Sustained/​Longer-​term Reward Learning Habit


Systems Motivation reward attainment responsiveness to
reward attainment

Trait emotion Emotion generation to reward Emotion regulatory Reward-​Valuation Emotion generation and
generation toward acquisition maintenance, stop, switch regulation traits, effortfulness and
rewards to sustained reward automaticity of emotion
generation and emotion regulation

Cognitive Systems Attention Perception Declarative Memory Language Cognitive Control Working Memory

External sensory-​ Internal perception in all Stored representations that Appraisal and Reappraisal Emotion Regulation Selection, Holding
array level perception WPVA cycles, appraisal and influence perception Situation Selection, Situation representations
in all WPVA reappraisal Modification, Reappraisal, of perception and
cycles, attentional Suppression referent for valuation
redeployment processing

Systems for Social Affiliation and Social Communication Perception and Perception and
Processes Attachment Understanding of the Understanding of Others
Self

Social emotion Social emotion generation, Social emotion generation, Social emotion generation,
generation, Extrinsic Extrinsic emotion regulation Extrinsic emotion Extrinsic emotion regulation
emotion regulation regulation

Arousal and Arousal Circadian Rhythms Sleep and Wakefulness Consider: Default Mode
Regulatory Network
Systems

Magnitude of action in Goal engagement necessary


all WPVA cycles for emotion regulation

Note: Default Mode Network was a considered construct within the Arousal and Regulatory Systems but was determined to not yet be ready for inclusion due to a lack of sufficient evidence. Nonetheless, it
putatively serves an important role in the goal engagement required for emotion regulation, which is not adequately captured by other constructs.
01
Table 5.5  An example of a hybrid EPM-​RDoC view of emotion regulation to exemplify RDoC-​type thinking.

Rows—​Domain: Emotion Columns—​
Units of Analysis
Constructs Sub-​Constructs
Generation Initial Initial Valuation Generative Action—​ Initial Perception Initial Valuation Generative Genes
Perception Cognition and Action—​Cognition
Behavior and Behavior
Regulation Intrinsic Regulation Extrinsic Regulation
Identification Perception of Valuation of own Intrinsic Action –​ Perception of other’s Valuation of other’s Extrinsic Action –​ Molecules
own emotive emotive states Cognition and emotive states emotive states Cognition and
states Behavior to Identify Behavior to Identify
own emotion other’s emotion
Selection Intrinsic Intrinsic Valuation of Selective Action on Knowhow (Perception) Valuation of Selective Action on Circuits
Perception Situation Selection, the emotive self of using Situation using Situation the emotive other
of Situation Situation Modification, Selection, Situation Selection, Situation
Selection, Attentional Allocation, Modification, Modification,
Situation Cognitive Reappraisal, Attentional Allocation, Attentional Allocation,
Modification, and Response Cognitive Reappraisal, Cognitive Reappraisal, Cells
Attentional Modulation and Response and Response
Allocation, Modulation on others Modulation on
Cognitive others
Reappraisal,
and Response
Modulation
Implementation Contextual Contextual valuation Implemented Contextual knowhow Contextual valuation Implemented Physiology
knowhow of of emotion regulatory regulatory action on of implementing a of implementation regulatory action
implementing a appropriateness the self in context selected strategy on others in
selected strategy context
1
0
Maintenance Perceiving one’s Valuation of Action to Maintain Perceiving one’s online Valuation of Action to Maintain Behavior
online intrinsic maintaining one’s emotion self-​ extrinsic regulation as maintaining one’s extrinsic regulation
regulation as online intrinsic regulatory behavior one to maintain extrinsic regulation
one to maintain regulation
Stopping Perceiving one’s Valuation of stopping Action to Stop Perceiving one’s online Valuation of stopping Action to Stop the Self-​Report
online intrinsic one’s online intrinsic emotion self-​ extrinsic regulation as one’s extrinsic extrinsic regulation
regulation as regulation regulatory behavior one to stop regulation
one to stop
Switching Perceiving one’s Valuation of switching Action to Switch to Perceiving one’s online Valuation of switching Action to Switch Paradigms
online intrinsic one’s online intrinsic another emotion self-​ extrinsic regulation as one’s extrinsic to another extrinsic
regulation as regulation regulatory behavior one to switch regulation regulatory behavior
one to switch
2
0
1

102 Research Domain Criteria (RDoC) and Emotion Regulation

current RDoC framework. Time will tell which functionalism serves as a better explanation across
units of analysis and in their associations with one another.
It might be argued that emotion regulation, given its centrality in psychopathology, deserves
its own domain in the RDoC matrix. Indeed, this discussion did occur in the Cognitive Systems
domain workgroup (National Institute of Mental Health, 2011). As it is viewed currently, research
relevant to emotion regulation is scattered across the matrix. Research on maladaptively up-╉
regulated fear and anxiety falls primarily within the Negative Valence Systems domain, whereas
the lack of generated reward valuation in anhedonia and adaptive habit formation involves the
Positive Valence System domain. Top-╉down regulatory strategies involve the Cognitive Systems
domain, and their adaptive and maladaptive use in social contexts involves the Social Systems
domain. Finally goal engagement, a critical referent in emotion regulation, involves the Arousal
and Regulatory Systems domain. This distribution across the matrix of constructs related to
emotion regulation may fuel more confusion than seems necessary. It should be noted that, as
improvements in conceptual clarity accrue, we are likely to see a blurring of boundaries between
traditional psychological concepts such as cognition and emotion (Miller, 2010). One may wish
to adopt a hybrid conceptualization for this reason. Whether one decides to adopt the current
RDoC matrix or a hybrid matrix yet to be drafted, it should be noted that “the usefulness of any
approach will ultimately rest upon the degree to which it promotes an empirical understanding of
the emotions and their role in behavior” (Panksepp, 1982, p. 421).

Summary and conclusion


The primary aim of this chapter was to explicate the RDoC framework and its development, and to
describe its relevance to the study of emotions and their regulation and dysregulation. The tension
between classical nosology and the development of the RDoC parallels historic disputes between
views of discrete and dimensional emotions. Both the Bradley-╉Lang view and the Gross-╉Ochsner
EPM view are dimensional accounts of emotion that complement discrete emotion views. The
RDoC matrix accommodates each of these theories while complementing the clinically useful
categorical views of psychological disorders represented in the DSM.
The tension between the RDoC and traditional classification systems mimics the tension in
emotion science between discrete and dimensional theory. Just as most emotion researchers uti-
lize both discrete and dimensional conceptualizations (Ekman, 2016) clinical research can and
indeed must move forward with discrete and dimensional conceptualizations of psychopathology.
As emotion researchers have long known, many of these dimensions lie in units of analysis out-
side of the traditional tools in psychology. Instead of rejecting this reality, a wiser policy will be to
embrace it for the sake of advancing the understanding of emotion regulation and increasing the
precision of clinical care. The RDoC is a promising platform for all of these aims.

Acknowledgements
We would like to thank Richard LeBeau, Anna S. Lau and members of the Anxiety and Depression
Research Center (ADRC) as well as the Culture Attention Emotion Science and Research
(CAESAR) lab for their helpful comments during the preparation of this chapter. Michael Sun
gratefully acknowledges the generous financial support of the Graduate Research Mentorship
award (with Michelle Craske) from the UCLA Graduate Division as well as a training fellowship
with the National Institute of Mental Health (NIMH) of the National Institutes of Health (NIH)
under award number T32-╉MH015750.
3
0
1

Acknowledgements 103

Meghan Vinograd gratefully acknowledges the generous financial support of the Graduate
Research Mentorship award (with Michelle Craske) from the UCLA Graduate Division as well as
the UCLA Depression Grand Challenge Research Fellowship.

References
Adams, R. B. Jr, & Kleck, R. E. (2005). Effects of direct and averted gaze on the perception of facially
communicated emotion. Emotion, 5(1), 3.
Aldao, A., Nolen-╉Hoeksema, S., & Schweizer, S. (2010). Emotion-╉regulation strategies across
psychopathology: A meta-╉analytic review. Clinical Psychology Review, 30(2), 217–╉237.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (DSM-╉5®).
American Psychiatric Pub.
American Psychiatric Association. (2016). DSM History. Retrieved February 24, 2016, from http://╉www.
psychiatry.org/╉psychiatrists/╉practice/╉dsm/╉history-╉of-╉the-╉dsm
Association, A. P., & others. (1980). DSM-╉III-╉R: Diagnostic and statistical manual of mental disorders.
American Psychiatric Association.
Banks, S. J., Eddy, K. T., Angstadt, M., Nathan, P. J., & Phan, K. L. (2007). Amygdala–╉frontal connectivity
during emotion regulation. Social Cognitive and Affective Neuroscience, 2(4), 303–╉312.
Barrett, L. F. (1998). Discrete emotions or dimensions? The role of valence focus and arousal focus.
Cognition & Emotion, 12(4), 579–╉599.
Barrett, L. F. (2013). Psychological construction: The Darwinian approach to the science of emotion.
Emotion Review, 5(4), 379–╉389.
Barrett, L. F., Lindquist, K. A., Bliss-╉Moreau, E., Duncan, S., Gendron, M., Mize, J., & Brennan, L. (2007).
Of mice and men: Natural kinds of emotions in the mammalian brain? A response to Panksepp and
Izard. Perspectives on Psychological Science, 2(3), 297–╉312.
Barrett, L. F., Wilson-╉Mendenhall, C. D., & Barsalou, L. W. (2013). A psychological construction account
of emotion regulation and dysregulation: The role of situated conceptualizations.
Baum, K. T., Desai, A., Field, J., Miller, L. E., Rausch, J., & Beebe, D. W. (2014). Sleep restriction worsens
mood and emotion regulation in adolescents. Journal of Child Psychology and Psychiatry, 55(2),
180–╉190.
Beck, A. T. (1976). Cognitive therapy and the emotional disorders. New York: The New American
Library. Inc.
Benning, S. D., Patrick, C. J., & Lang, A. R. (2004). Emotional modulation of the post-╉auricular reflex.
Psychophysiology, 41(3), 426–╉432.
Bentler, P. M. (1980). Multivariate analysis with latent variables: Causal modeling. Annual Review of
Psychology, 31(1), 419–╉456.
Bosker, F. J., Hartman, C. A., Nolte, I. M., Prins, B. P., Terpstra, P., Posthuma, D., … others. (2011). Poor
replication of candidate genes for major depressive disorder using genome-╉wide association data.
Molecular Psychiatry, 16(5), 516–╉532.
Bradley, M. M., Codispoti, M., Cuthbert, B. N., & Lang, P. J. (2001). Emotion and motivation I: defensive
and appetitive reactions in picture processing. Emotion, 1(3), 276.
Bradley, M. M., & Lang, P. J. (1994). Measuring emotion: the self-╉assessment manikin and the semantic
differential. Journal of Behavior Therapy and Experimental Psychiatry, 25(1), 49–╉59.
Bradley, M. M., & Lang, P. J. (2006). Emotion and motivation. Handbook of Psychophysiology, 2, 602–╉642.
Bradley, M. M., Moulder, B., & Lang, P. J. (2005). When good things go bad the reflex physiology of
defense. Psychological Science, 16(6), 468–╉473.
Bridges, L. J., Denham, S. A., & Ganiban, J. M. (2004). Definitional issues in emotion regulation research.
Child Development, 75(2), 340–╉345.
4
0
1

104 Research Domain Criteria (RDoC) and Emotion Regulation

Brown, L. A., LeBeau, R. T., Chat, K. Y., & Craske, M. G. (in press). Associative learning versus fear
habituation as predictors of long-╉term extinction retention.
Buhle, J. T., Silvers, J. A., Wager, T. D., Lopez, R., Onyemekwu, C., Kober, H., … Ochsner, K. N. (2014).
Cognitive reappraisal of emotion: a meta-╉analysis of human neuroimaging studies. Cerebral Cortex,
24(11), 2981–╉2990.
Butler, E. A., Egloff, B., Wlhelm, F. H., Smith, N. C., Erickson, E. A., & Gross, J. J. (2003). The
social consequences of expressive suppression. Emotion, 3(1), 48–╉67. http://╉dx.doi.org/╉10.1037/╉
1528-╉3542.3.1.48
Cacioppo, J. T., Berntson, G. C., & Hatfield, E. (1993). The psychophysiology of emotion. Handbook of
Emotions, 119–╉142.
Campbell-╉Sills, L., & Barlow, D. H. (2007). Incorporating emotion regulation into conceptualizations and
treatments of anxiety and mood disorders. Handbook of Emotion Regulation, 2, 542–╉559.
Capogna, M. (2014). GABAergic cell type diversity in the basolateral amygdala. Current Opinion in
Neurobiology, 26, 110–╉116.
Carver, C. S., & Scheier, M. F. (1982). Control theory: A useful conceptual framework for personality–╉
social, clinical, and health psychology. Psychological Bulletin, 92(1), 111.
Carver, C. S., & White, T. L. (1994). Behavioral inhibition, behavioral activation, and affective responses to
impending reward and punishment: the BIS/╉BAS scales. Journal of Personality and Social Psychology,
67(2), 319.
Casey, B. J., Craddock, N., Cuthbert, B. N., Hyman, S. E., Lee, F. S., & Ressler, K. J. (2013). DSM-╉5 and
RDoC: progress in psychiatry research? Nature Reviews Neuroscience, 14(11), 810–╉814.
Casey, B. J., Somerville, L. H., Gotlib, I. H., Ayduk, O., Franklin, N. T., Askren, M. K., … others. (2011).
Behavioral and neural correlates of delay of gratification 40 years later. Proceedings of the National
Academy of Sciences, 108(36), 14998–╉15003.
Chen, L.-╉S., Eaton, W. W., Gallo, J. J., & Nestadt, G. (2000). Understanding the heterogeneity of depression
through the triad of symptoms, course and risk factors: a longitudinal, population-╉based study. Journal
of Affective Disorders, 59(1), 1–╉11.
Chinta, S. J., & Andersen, J. K. (2005). Dopaminergic neurons. The International Journal of Biochemistry &
Cell Biology, 37(5), 942–╉946.
Cicchetti, D., Ackerman, B. P., & Izard, C. E. (1995). Emotions and emotion regulation in developmental
psychopathology. Development and Psychopathology, 7(01), 1–╉10.
Clarke, D. E., Narrow, W. E., Regier, D. A., Kuramoto, S. J., Kupfer, D. J., Kuhl, E. A., … Kraemer, H. C.
(2013). DSM-╉5 field trials in the United States and Canada, part I: study design, sampling strategy,
implementation, and analytic approaches. American Journal of Psychiatry, 170(1), 43–╉58. Retrieved
from http://╉ajp.psychiatryonline.org/╉doi/╉pdf/╉10.1176/╉appi.ajp.2012.12070998
Coté, S. (2005). A social interaction model of the effects of emotion regulation on work strain. Academy of
Management Review, 30(3), 509–╉530.
Craske, M. G., Meuret, A. E., Ritz, T., Treanor, M., & Dour, H. J. (in press). Treatment for Anhedonia.
Cuthbert, B. N. (2014). The RDoC framework: facilitating transition from ICD/╉DSM to dimensional
approaches that integrate neuroscience and psychopathology. World Psychiatry, 13(1), 28–╉35.
Cuthbert, B. N., & Insel, T. R. (2013). Toward the future of psychiatric diagnosis: the seven pillars of RDoC.
BMC Medicine, 11(1), 126.
Cuthbert, B. N., Schupp, H. T., Bradley, M. M., Birbaumer, N., & Lang, P. J. (2000). Brain potentials
in affective picture processing: covariation with autonomic arousal and affective report. Biological
Psychology, 52(2), 95–╉111. http://╉doi.org/╉10.1016/╉S0301-╉0511(99)00044-╉7
Cuthbert, B. N., Vrana, S. R., & Bradley, M. M. (1991). Imagery: Function and physiology. Advances in
Psychophysiology, 4, 1–╉42.
Delano-╉Wood, L., Houston, W. S., Emond, J. A., Marchant, N. L., Salmon, D. P., Jeste, D. V., … Bondi,
M. W. (2008). APOE genotype predicts depression in women with Alzheimer’s disease: a retrospective
study. International Journal of Geriatric Psychiatry, 23(6), 632–╉636. http://╉doi.org/╉10.1002/╉gps.1953
5
0
1

Acknowledgements 105

Der-​Avakian, A., & Markou, A. (2012). The neurobiology of anhedonia and other reward-​related deficits.
Trends in Neurosciences, 35(1), 68–​77. http://​doi.org/​10.1016/​j.tins.2011.11.005
DeSteno, D., Gross, J. J., & Kubzansky, L. (2013). Affective science and health: The importance of emotion
and emotion regulation. Health Psychology, 32(5), 474.
Diekhof, E. K., Geier, K., Falkai, P., & Gruber, O. (2011). Fear is only as deep as the mind allows: a
coordinate-​based meta-​analysis of neuroimaging studies on the regulation of negative affect.
Neuroimage, 58(1), 275–​285.
Drabant, E. M., Hariri, A. R., Meyer-​Lindenberg, A., Munoz, K. E., Mattay, V. S., Kolachana, B. S., …
Weinberger, D. R. (2006). Catechol O-​methyltransferase val158met genotype and neural mechanisms
related to affective arousal and regulation. Archives of General Psychiatry, 63(12), 1396–​1406.
Ekman, P., & Cordaro, D. (2011). What is meant by calling emotions basic. Emotion Review, 3, 364–​370.
Ekman, P. (2016). What Scientists Who Study Emotion Agree About. Perspectives on Psychological Science,
11(1), 31–​34.
Faivre, N., Charron, S., Roux, P., Lehéricy, S., & Kouider, S. (2012). Nonconscious emotional processing
involves distinct neural pathways for pictures and videos. Neuropsychologia, 50(14), 3736–​3744. http://​
doi.org/​10.1016/​j.neuropsychologia.2012.10.025
Foa, E. B., & Kozak, M. J. (1986). Emotional processing of fear: exposure to corrective information.
Psychological Bulletin, 99(1), 20.
Freedman, R., Lewis, D. A., Michels, R., Pine, D. S., Schultz, S. K., Tamminga, C. A., … others. (2013).
The initial field trials of DSM-​5: new blooms and old thorns. American Journal of Psychiatry, 170(1),
1–​5. Retrieved from http://​ajp.psychiatryonline.org/​doi/​pdf/​10.1176/​appi.ajp.2012.12091189
Gendron, M., & Barrett, L. F. (2009). Reconstructing the past: A century of ideas about emotion in
psychology. Emotion Review, 1(4), 316–​339.
Goldin, P. R., McRae, K., Ramel, W., & Gross, J. J. (2008). The neural bases of emotion
regulation: reappraisal and suppression of negative emotion. Biological Psychiatry, 63(6), 577–​586.
Goldstein, B. L., & Klein, D. N. (2014). A review of selected candidate endophenotypes for depression.
Clinical Psychology Review, 34(5), 417–​427.
Grandey, A. A. (2000). Emotional regulation in the workplace: A new way to conceptualize emotional labor.
Journal of Occupational Health Psychology, 5(1), 95.
Gross, J. J. (1998). The emerging field of emotion regulation: An integrative review. Review of General
Psychology, 2(3), 271–​299. http://​dx.doi.org/​10.1037/​1089-​2680.2.3.271
Gross, J. J. (2013). Emotion regulation: taking stock and moving forward. Emotion, 13(3), 359.
Gross, J. J. (2014). Emotion regulation: Conceptual and empirical foundations. Handbook of Emotion
Regulation, 2, 3–​20.
Gross, J. J. (2015a). Emotion regulation: Current status and future prospects. Psychological Inquiry,
26(1), 1–​26.
Gross, J. J. (2015b). The Extended Process Model of Emotion Regulation: Elaborations, Applications, and
Future Directions. Psychological Inquiry, 26(1), 130–​137.
Gross, J. J., & Barrett, L. F. (2011). Emotion generation and emotion regulation: One or two depends on
your point of view. Emotion Review, 3(1), 8–​16.
Gross, J. J., & John, O. P. (2003). Individual differences in two emotion regulation processes: Implications
for affect, relationships, and well-​being. Journal of Personality and Social Psychology, 85(2), 348–​362.
http://​dx.doi.org/​10.1037/​0022-​3514.85.2.348
Gross, J. J., & Levenson, R. W. (1993). Emotional suppression: physiology, self-​report, and expressive
behavior. Journal of Personality and Social Psychology, 64(6), 970.
Gross, J. J., & Levenson, R. W. (1997). Hiding feelings: the acute effects of inhibiting negative and positive
emotion. Journal of Abnormal Psychology, 106(1), 95.
Gross, J. J., & Muñoz, R. F. (1995). Emotion regulation and mental health. Clinical Psychology: Science and
Practice, 2(2), 151–​164.
6
0
1

106 Research Domain Criteria (RDoC) and Emotion Regulation

Gross, J. J., Sheppes, G., & Urry, H. L. (2011). Emotion generation and emotion regulation: A distinction
we should make (carefully). Cognition and Emotion, 25(5), 765–​781.
Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. In J. J. Gross (Ed).
(2007). Handbook of emotion regulation, (pp. 3–​24). New York, NY, US: Guilford Press, xvii, 654 pp.
Hamann, S. (2012). Mapping discrete and dimensional emotions onto the brain: controversies and
consensus. Trends in Cognitive Sciences, 16(9), 458–​466.
Hampson, R. E., Hedberg, T., & Deadwyler, S. A. (2000). Differential information processing by
hippocampal and subicular neurons. Annals of the New York Academy of Sciences, 911(1), 151–​165.
Hariri, A. R., & Holmes, A. (2006). Genetics of emotional regulation: the role of the serotonin transporter
in neural function. Trends in Cognitive Sciences, 10(4), 182–​191.
Harré, R. (1986). The Social Construction of Emotions. London: Basil Blackwell.
Hartley, C. A., & Phelps, E. A. (2010). Changing fear: the neurocircuitry of emotion regulation.
Neuropsychopharmacology, 35(1), 136–​146.
Hayes, S. C., Strosahl, K. D., & Wilson, K. G. (1999). Acceptance and commitment therapy.
New York: Guilford Press.
Helmeke, C., Ovtscharoff, W., Poeggel, G., & Braun, K. (2001). Juvenile emotional experience alters
synaptic inputs on pyramidal neurons in the anterior cingulate cortex. Cerebral Cortex, 11(8), 717–​727.
Hermans, D., Craske, M. G., Mineka, S., & Lovibond, P. F. (2006). Extinction in human fear conditioning.
Biological Psychiatry, 60(4), 361–​368.
Hofmann, S. G., Sawyer, A. T., Fang, A., & Asnaani, A. (2012). Emotion dysregulation model of mood and
anxiety disorders. Depression and Anxiety, 29(5), 409–​416.
Hung, Y., Smith, M. L., & Taylor, M. J. (2012). Development of ACC–​amygdala activations in processing
unattended fear. Neuroimage, 60(1), 545–​552.
Hyman, S. E. (2010). The diagnosis of mental disorders: the problem of reification. Annual Review of
Clinical Psychology, 6, 155–​179.
Infantolino, Z. P., Crocker, L. D., Heller, W., Yee, C. M., & Miller, G. A. (In press). Psychophysiology in
pursuit of psychopathology. In J. T. Cacioppo, L. G. Tassinary, & G. G. Berntson (Eds.), Handbook of
Psychophysiology ( 4th Ed.). Cambridge, UK: Cambridge University Press.
Insel, T., Cuthbert, B., Garvey, M., Heinssen, R., Pine, D. S., Quinn, K., … 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.
Insel, T. R., & Cuthbert, B. N. (2015). Brain disorders? Precisely. Science, 348(6234), 499–​500.
Izard, C. E. (2010). The many meanings/​aspects of emotion: Definitions, functions, activation, and
regulation. Emotion Review, 2(4), 363–​370.
Izard, C. E. (2011). Forms and functions of emotions: Matters of emotion–​cognition interactions. Emotion
Review, 3(4), 371–​378.
James, W. (1884). II.—​What is an emotion? Mind, (34), 188–​205.
John, O. P., & Gross, J. J. (2004). Healthy and unhealthy emotion regulation: Personality processes,
individual differences, and life span development. Journal of Personality, 72(6), 1301–​1334.
Jolliffe, I. (2002). Principal component analysis. Wiley Online Library. Retrieved from http://​onlinelibrary.
wiley.com/​doi/​10.1002/​9781118445112.stat06472/​full
Kalisch, R., Wiech, K., Herrmann, K., & Dolan, R. J. (2006). Neural correlates of self-​distraction from
anxiety and a process model of cognitive emotion regulation. Journal of Cognitive Neuroscience, 18(8),
1266–​1276.
Kim, H. S., Sherman, D. K., Mojaverian, T., Sasaki, J. Y., Park, J., Suh, E. M., & Taylor, S. E. (2011).
Gene–​culture interaction oxytocin receptor polymorphism (OXTR) and emotion regulation. Social
Psychological and Personality Science, 2(6), 665–​672.
Kozak, M. J., & Cuthbert, B. N. (2016). The NIMH Research Domain Criteria Initiative: Background,
issues, and pragmatics. Psychophysiology, 53(3), 286–​297. http://​doi.org/​10.1111/​psyp.12518
7
0
1

Acknowledgements 107

Kozak, M. J., & Miller, G. A. (1982). Hypothetical constructs versus intervening variables: A re-​appraisal of
the three-​systems model of anxiety assessment. Behavioral Assessment, 4(3), 347–​358.
Krakowski, M. (2003). Violence and serotonin: influence of impulse control, affect regulation, and social
functioning. The Journal of neuropsychiatry and clinical neurosciences, 15(3), 294–​305.
Kring, A. M. (2008). Emotion disturbances as transdiagnostic processes in psychopathology. Handbook of
Emotion, 3, 691–​705.
Lam, S., Dickerson, S. S., Zoccola, P. M., & Zaldivar, F. (2009). Emotion regulation and cortisol reactivity
to a social-​evaluative speech task. Psychoneuroendocrinology, 34(9), 1355–​1362.
Lang, P. J. (1979). A bio-​informational theory of emotional imagery. Psychophysiology, 16(6), 495–​512.
Laurent, H., & Powers, S. (2007). Emotion regulation in emerging adult couples: Temperament,
attachment, and HPA response to conflict. Biological Psychology, 76(1), 61–​71.
Levenson, R. W. (2011). Basic emotion questions. Emotion Review, 3, 379–​386.
Linehan, M. (1993). Cognitive-​behavioral treatment of borderline personality disorder.
New York: Guilford press.
Lissek, S., Biggs, A. L., Rabin, S. J., Cornwell, B. R., Alvarez, R. P., Pine, D. S., & Grillon, C. (2008).
Generalization of conditioned fear-​potentiated startle in humans: experimental validation and clinical
relevance. Behaviour Research and Therapy, 46(5), 678–​687.
Litten, R. Z., Ryan, M. L., Falk, D. E., Reilly, M., Fertig, J. B., & Koob, G. F. (2015). Heterogeneity
of alcohol use disorder: understanding mechanisms to advance personalized treatment.
Alcoholism: Clinical and Experimental Research, 39(4), 579–​584.
Lopes, P. N., Nezlek, J. B., Extremera, N., Hertel, J., Fernández-​Berrocal, P., Schütz, A., & Salovey, P.
(2011). Emotion regulation and the quality of social interaction: Does the ability to evaluate emotional
situations and identify effective responses matter? Journal of Personality, 79(2), 429–​467.
Lopes, P. N., Salovey, P., Côté, S., Beers, M., & Petty, R. E. (2005). Emotion regulation abilities and the
quality of social interaction. Emotion, 5(1), 113.
MacCorquodale, K., & Meehl, P. E. (1948). On a distinction between hypothetical constructs and
intervening variables. Psychological Review, 55(2), 95.
Mackintosh, N. J. (1975). A theory of attention: variations in the associability of stimuli with reinforcement.
Psychological Review, 82(4), 276.
Martin, R. C., & Dahlen, E. R. (2005). Cognitive emotion regulation in the prediction of depression, anxiety, stress,
and anger. Personality and Individual Differences, 39(7), 1249–​1260. http://​doi.org/​10.1016/​j.paid.2005.06.004
Matsumoto, D. (1990). Cultural similarities and differences in display rules. Motivation and Emotion,
14(3), 195–​214.
Mauss, I. B., Troy, A. S., & LeBourgeois, M. K. (2013). Poorer sleep quality is associated with lower
emotion-​regulation ability in a laboratory paradigm. Cognition & Emotion, 27(3), 567–​576.
McRae, K., Ciesielski, B., & Gross, J. J. (2012). Unpacking cognitive reappraisal: goals, tactics, and
outcomes. Emotion, 12(2), 250.
Milad, M. R., & Quirk, G. J. (2012). Fear extinction as a model for translational neuroscience: ten years of
progress. Annual Review of Psychology, 63, 129–​151.
Milad, M. R., Rosenbaum, B. L., & Simon, N. M. (2014). Neuroscience of fear extinction: implications for
assessment and treatment of fear-​based and anxiety related disorders. Behaviour Research and Therapy,
62, 17–​23.
Milad, M. R., Wright, C. I., Orr, S. P., Pitman, R. K., Quirk, G. J., & Rauch, S. L. (2007). Recall of fear
extinction in humans activates the ventromedial prefrontal cortex and hippocampus in concert.
Biological Psychiatry, 62(5), 446–​454.
Miller, G. A. (1996). How we think about cognition, emotion, and biology in psychopathology.
Psychophysiology, 33, 615–​628.
Miller, G. A. (2010). Mistreating psychology in the decades of the brain. Perspectives on Psychological
Science, 5(6), 716–​743.
8
0
1

108 Research Domain Criteria (RDoC) and Emotion Regulation

Miller, G. A., & Kozak, M. J. (1993). A philosophy for the study of emotion: Three-​systems theory. In N.
Birbaumer & A. Öhman (Eds.), The Structure of Emotion: Physiological, Cognitive and Clinical Aspects
(pp. 31–​47). Seattle: Hogrefe & Huber.
Miller, G. A., & Rockstroh, B. (2013). Endophenotypes in psychopathology research: where do we stand?
Annual Review of Clinical Psychology, 9, 177–​213.
Miller, G. A., Rockstroh, B. S., Hamilton, H. K., & Yee, C. M. (2016). Psychophysiology as a core strategy
in RDoC. Psychophysiology, 53(3), 410–​414.
Narrow, W. E., Clarke, D. E., Kuramoto, S. J., Kraemer, H. C., Kupfer, D. J., Greiner, L., & Regier, D. A.
(2013). DSM-​5 field trials in the United States and Canada, Part III: development and reliability testing
of a cross-​cutting symptom assessment for DSM-​5. American Journal of Psychiatry, 170(1), 71–​82.
Retrieved from http://​ajp.psychiatryonline.org/​doi/​pdf/​10.1176/​appi.ajp.2012.12071000
National Institutes of Health. (2014). Research Domain Criteria Database. Retrieved February 24, 2016,
from http://​rdocdb.nimh.nih.gov/​about/​
National Institutes of Mental Health. (2008). The National Institute of Mental Health Strategic Plan.
Retrieved March 4, 2016, from http://​www.nimh.nih.gov/​about/​strategic-​planning-​reports/​
index.shtml
National Institutes of Mental Health. (2011). NIMH » Cognitive Systems: Workshop Proceedings.
Retrieved February 14, 2016, from http://​www.nimh.nih.gov/​research-​priorities/​rdoc/​cognitive-​
systems-​workshop-​proceedings.shtml
National Institutes of Mental Health. (2012). Social Processes: Workshop Proceedings. Retrieved March 4,
2016, from http://​www.nimh.nih.gov/​research-​priorities/​rdoc/​social-​processes-​workshop-​proceedings.
shtml
National Institutes of Mental Health. (2014). NIMH » RDoC Frequently Asked Questions (FAQ).
Retrieved February 24, 2016, from http://​www.nimh.nih.gov/​research-​priorities/​rdoc/​rdoc-​frequently-​
asked-​questions-​faq.shtml
National Institutes of Mental Health. (2016). NIMH  RDoC Matrix. Retrieved February 24, 2016, from
http://​www.nimh.nih.gov/​research-​priorities/​rdoc/​constructs/​rdoc-​matrix.shtml
Ochsner, K. N., & Gross, J. J. (2005). The cognitive control of emotion. Trends in Cognitive Sciences, 9(5),
242–​249.
Ochsner, K. N., & Gross, J. J. (2008). Cognitive emotion regulation insights from social cognitive and
affective neuroscience. Current Directions in Psychological Science, 17(2), 153–​158.
Ochsner, K. N., & Gross, J. J. (2014). The neural bases of emotion and emotion regulation: A valuation
perspective. Handbook of Emotion Regulation, 2, 23–​42.
Ochsner, K. N., Silvers, J. A., & Buhle, J. T. (2012). Functional imaging studies of emotion regulation: a
synthetic review and evolving model of the cognitive control of emotion. Annals of the New York
Academy of Sciences, 1251(1), E1–​E24.
Ortner, C. N., Zelazo, P. D., & Anderson, A. K. (2013). Effects of emotion regulation on concurrent
attentional performance. Motivation and Emotion, 37(2), 346–​354.
Panksepp, J. (1982). Toward a general psychobiological theory of emotions. Behavioral and Brain Sciences,
5(03), 407–​422.
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. Oxford
University Press.
Panksepp, J., & Watt, D. (2011). What is basic about basic emotions? Lasting lessons from affective
neuroscience. Emotion Review, 3(4), 387–​396.
Pearce, J. M., & Hall, G. (1980). A model for Pavlovian learning: variations in the effectiveness of
conditioned but not of unconditioned stimuli. Psychological Review, 87(6), 532.
Peciña, M., Mickey, B. J., Love, T., Wang, H., Langenecker, S. A., Hodgkinson, C., … others. (2013). DRD2
polymorphisms modulate reward and emotion processing, dopamine neurotransmission and openness
to experience. Cortex, 49(3), 877–​890.
9
0
1

Acknowledgements 109

Peters, J., Dieppa-​Perea, L. M., Melendez, L. M., & Quirk, G. J. (2010). Induction of fear extinction with
hippocampal-​infralimbic BDNF. Science, 328(5983), 1288–​1290.
Phelps, E. A., Delgado, M. R., Nearing, K. I., & LeDoux, J. E. (2004). Extinction learning in humans: role
of the amygdala and vmPFC. Neuron, 43(6), 897–​905.
Pizzagalli, D. A., Evins, A. E., Schetter, E. C., Frank, M. J., Pajtas, P. E., Santesso, D. L., & Culhane, M.
(2008). Single dose of a dopamine agonist impairs reinforcement learning in humans: behavioral
evidence from a laboratory-​based measure of reward responsiveness. Psychopharmacology, 196(2),
221–​232.
Pizzagalli, D. A., Lehmann, D., Hendrick, A. M., Regard, M., Pascual-​Marqui, R. D., & Davidson, R.
J. (2002). Affective judgments of faces modulate early activity (160 ms) within the fusiform gyri.
Neuroimage, 16(3), 663–​677.
Quirin, M., Kuhl, J., & Düsing, R. (2011). Oxytocin buffers cortisol responses to stress in individuals with
impaired emotion regulation abilities. Psychoneuroendocrinology, 36(6), 898–​904.
Rauch, S., & Foa, E. (2006). Emotional processing theory (EPT) and exposure therapy for PTSD. Journal of
Contemporary Psychotherapy, 36(2), 61–​65.
RDoQ. (2016). Retrieved February 24, 2016, from http://​datamining.cs.ucla.edu/​rdoq/​
Regier, D. A., Narrow, W. E., Clarke, D. E., Kraemer, H. C., Kuramoto, S. J., Kuhl, E. A., & Kupfer, D.
J. (2013). DSM-​5 field trials in the United States and Canada, Part II: test-​retest reliability of selected
categorical diagnoses. American Journal of Psychiatry, 170(1), 59–​70. Retrieved from http://​ajp.
psychiatryonline.org/​doi/​abs/​10.1176/​appi.ajp.2012.12070999
Rescorla, R. A., Wagner, A. R., & others. (1972). A theory of Pavlovian conditioning: Variations in the
effectiveness of reinforcement and nonreinforcement. Classical Conditioning II: Current Research and
Theory, 2, 64–​99.
Richards, J. M., Butler, E. A., & Gross, J. J. (2003). Emotion Regulation in Romantic Relationships: The
Cognitive Consequences of Concealing Feelings. Journal of Social and Personal Relationships, 20(5),
599–​620. http://​doi.org/​10.1177/​02654075030205002
Richards, J. M., & Gross, J. J. (1999). Composure at Any Cost? The Cognitive Consequences of Emotion
Suppression. Personality and Social Psychology Bulletin, 25(8), 1033–​1044. http://​doi.org/​10.1177/​
01461672992511010
Richards, J. M., & Gross, J. J. (2000). Emotion regulation and memory: the cognitive costs of keeping one’s
cool. Journal of Personality and Social Psychology, 79(3), 410.
Salgado-​Pineda, P., Delaveau, P., Blin, O., & Nieoullon, A. (2005). Dopaminergic contribution to the
regulation of emotional perception. Clinical Neuropharmacology, 28(5), 228–​237.
Santini, C. C., Bath, J., Turberfield, A. J., & Tyrrell, A. M. (2012). A DNA network as an information
processing system. International Journal of Molecular Sciences, 13(4), 5125–​5137.
Scherer, K. R. (2009). The dynamic architecture of emotion: Evidence for the component process model.
Cognition and Emotion, 23(7), 1307–​1351.
Schmader, T., & Lickel, B. (2006). The approach and avoidance function of guilt and shame
emotions: Comparing reactions to self-​caused and other-​caused wrongdoing. Motivation and Emotion,
30(1), 42–​55.
Shabel, S. J., & Janak, P. H. (2009). Substantial similarity in amygdala neuronal activity during conditioned
appetitive and aversive emotional arousal. Proceedings of the National Academy of Sciences, 106(35),
15031–​15036.
Silk, J. S., Siegle, G. J., Whalen, D. J., Ostapenko, L. J., Ladouceur, C. D., & Dahl, R. E. (2009). Pubertal
changes in emotional information processing: Pupillary, behavioral, and subjective evidence during
emotional word identification. Development and Psychopathology, 21(01), 7–​26.
Stan, A. D., Schirda, C. V., Bertocci, M. A., Bebko, G. M., Kronhaus, D. M., Aslam, H. A., …
others. (2014). Glutamate and GABA contributions to medial prefrontal cortical activity to
emotion: Implications for mood disorders. Psychiatry Research: Neuroimaging, 223(3), 253–​260.
0
1

110 Research Domain Criteria (RDoC) and Emotion Regulation

Sturm, V. E., Yokoyama, J. S., Seeley, W. W., Kramer, J. H., Miller, B. L., & Rankin, K. P. (2013).
Heightened emotional contagion in mild cognitive impairment and Alzheimer’s disease is associated
with temporal lobe degeneration. Proceedings of the National Academy of Sciences, 110(24), 9944–​9949.
Szanto, K., Dombrovski, A. Y., Sahakian, B. J., Mulsant, B. H., Houck, P. R., Reynolds, C. F. III, & Clark,
L. (2012). Social Emotion Recognition, Social Functioning, and Attempted Suicide in Late-​Life
Depression. The American Journal of Geriatric Psychiatry, 20(3), 257–​265. http://​doi.org/​10.1097/​
JGP.0b013e31820eea0c
Thiruchselvam, R., Blechert, J., Sheppes, G., Rydstrom, A., & Gross, J. J. (2011). The temporal dynamics of
emotion regulation: an EEG study of distraction and reappraisal. Biological Psychology, 87(1), 84–​92.
Thompson, R. A. (1994). Emotion regulation: A theme in search of definition. Monographs of the Society for
Research in Child Development, 59(2-​3), 25–​52.
Thompson, R. A., & Calkins, S. D. (1996). The double-​edged sword: Emotional regulation for children at
risk. Development and Psychopathology, 8(01), 163–​182.
Tracy, J. L., & Randles, D. (2011). Four models of basic emotions: a review of Ekman and Cordaro, Izard,
Levenson, and Panksepp and Watt. Emotion Review, 3(4), 397–​405.
Tsai, J. L., Knutson, B., & Fung, H. H. (2006). Cultural variation in affect valuation. Journal of Personality
and Social Psychology, 90(2), 288.
Tsai, W., Sun, M., Wang, S., & Lau, A. S. (2016). Implications of Emotion Expressivity for Daily and
Trait Interpersonal and Intrapersonal Functioning Across Ethnic Groups. Asian American Journal of
Psychology, 7(1), 52–​63. http://​doi.org/​10.1037/​aap0000043
van Minnen, A., & Hagenaars, M. (2002). Fear activation and habituation patterns as early process
predictors of response to prolonged exposure treatment in PTSD. Journal of Traumatic Stress, 15(5),
359–​367.
Von Foerster, H. (2003). Cybernetics of cybernetics. Understanding Understanding: Essays on Cybernetics
and Cognition, (pp. 283–​286). Springer New York.
Vrieze, E., Pizzagalli, D. A., Demyttenaere, K., Hompes, T., Sienaert, P., de Boer, P., … Claes, S. (2013).
Reduced reward learning predicts outcome in major depressive disorder. Biological Psychiatry, 73(7),
639–​645.
Wacker, J., Mueller, E. M., Pizzagalli, D. A., Hennig, J., & Stemmler, G. (2013). Dopamine-​D2-​Receptor
Blockade Reverses the Association Between Trait Approach Motivation and Frontal Asymmetry
in an Approach-​Motivation Context. Psychological Science, 24(4), 489–​497. http://​doi.org/​10.1177/​
0956797612458935
Wåhlstedt, C., Thorell, L. B., & Bohlin, G. (2009). Heterogeneity in ADHD: neuropsychological pathways,
comorbidity and symptom domains. Journal of Abnormal Child Psychology, 37(4), 551–​564.
Waider, J., Araragi, N., Gutknecht, L., & Lesch, K.-​P. (2011). Tryptophan hydroxylase-​2 (TPH2) in
disorders of cognitive control and emotion regulation: A perspective. Psychoneuroendocrinology, 36(3),
393–​405. http://​doi.org/​10.1016/​j.psyneuen.2010.12.012
Webb, T. L., Miles, E., & Sheeran, P. (2012). Dealing with feeling: a meta-​analysis of the effectiveness of
strategies derived from the process model of emotion regulation. Psychological Bulletin, 138(4), 775.
Wiener, N. (1961). Cybernetics or control and communication in the animal and the machine (Vol. 25). MIT
Press Cambridge, Massachusettes.
Yee, C. M., Javitt, D. C., & Miller, G. A. (2015). Replacing DSM categorical analyses with dimensional
analyses in psychiatry research: The Research Domain Criteria Initiative. JAMA Psychiatry, 72(12),
1159–​1160.
Part II

Emotion Regulation
and Child and Adolescent
Psychopathology
2
1
3
1

Chapter 6

Emotion Regulation and Attention


Deficit Hyperactivity Disorder
Blossom Fernandes, Roseann Tan-╉Mansukhani,
& Cecilia A. Essau

Diagnostic criteria
Attention Deficit Hyperactivity Disorder (ADHD) is considered one of the most pervasive disor-
ders of childhood (Castellanos & Tannock, 2002). ADHD frequently persists into adolescence and
adulthood and is consistently associated with a range of negative outcomes. The term ADHD was
first coined as part of the restructure and enhancement of the Diagnostic and Statistical Manual
of Mental Disorders (DSM) by the American Psychiatric Association (American Psychiatric
Association, 1980); prior to this, children with ADHD were diagnosed with brain dysfunction or
brain damage with hyperkinesia (Barkley, 1990).
The diagnostic criteria for ADHD has undergone several changes, but most notably, the DSM-╉
IV (American Psychiatric Association, 2000)  has classified behavioural symptoms comprising
three major subtypes:  Inattention (I), hyperactivity-╉impulsivity (HI), and Combined. Children
presenting with the inattentive subtype have difficulty with tasks that require sustained mental
effort, are more disorganized and are easily distracted and forgetful when compared to peers of
a similar age (Sergeant, Oosterlaan, & van der Meere, 1999). Children with the HI subtype were
characterized as more fidgety, restless and “squirmy” when compared to typically developing chil-
dren. The combined subtype is the most commonly diagnosed sub-╉type and involves six or more
symptoms of each of the inattention and hyperactivity subtypes (APA, 2000). Research has shown
children with ADHD, as compared to other children, also have difficulty inhibiting on-╉going
behavior (Oosterlaan, Logan, & Sergeant, 1998) and difficulty inhibiting immediate gratification
(Douglas & Parry, 1983). As with attention problems, these difficulties lead to serious problems in
home and school functioning.
The fifth revision of the DSM (DSM-╉5; American Psychiatric Association, 2013) characterized
ADHD as a neurodevelopmental disorder consisting of a pattern of inattention and/╉or hyperac-
tivity-╉impulsivity that affects every day functioning. It further specifies the severity of ADHD as
mild (involving minor impairments with few, if any, symptoms in excess of the six required for
diagnosis), moderate (impairment between mild and severe), and severe (marked impairment
and with several symptoms in excess of those necessary for a diagnosis). The DSM-╉5 suggests that
these symptoms should have persisted for at least six months, directly impacting social and aca-
demic/╉occupational activities and be present before the age of 12 years. Additionally, the DSM-╉5
allows for diagnosis in adolescents and adults, including types of behaviour and examples of how
the disorder could manifest itself in different age groups. The symptoms are expected to persist for
at least five months in individual above the age of 17, unlike six months for children and adoles-
cents; moreover, those aged 17 and older need to exhibit only five of the requisite symptoms, not
the six required for younger children.
4
1

114 Emotion Regulation and Attention Deficit Hyperactivity Disorder

Prevalence
There is growing consensus that ADHD occurs in approximately 5% of school-╉aged children, a
prevalence rate that is seen across cultures (Esser, Schmidt, & Woerner, 1990; Polanczyk et al.
2007; Polanczyk et al. 2014). However, ADHD prevalence rates diverge widely as a result of age,
definition of disorder and assessment method. Distinctions based on definitions (e.g., clinical
classifications of ADHD versus scores on a behavior checklist) and more rigorous assessment cri-
teria lead to fewer cases of ADHD. For example, a review of 86 studies using the DSM-╉IV criteria
(American Psychiatric Association, 1994) indicated the prevalence of ADHD ranged from 5.9%
to 7.1% (Willcutt, 2012). ADHD diagnosis also seems to be influenced by parental practices and
beliefs. As reported by Asherson and colleagues (2012), in Asian countries such as Hong Kong
and Taiwan parental monitoring of child behavior is considered essential in reducing disruptive
behaviors and poor habits, thereby influencing diagnosis and access to treatment. Variations in
the prevalence rates of ADHD in different countries has raised a question to whether ADHD is
a universal syndrome affecting children worldwide regardless of race and society (Bauermeister
et al., 2010; Goetz et al., 2010; Polanczyk et al., 2007) or a cultural construct generally based on a
Western conceptualization (Asherson et al., 2012; Faraone et al., 2003; Jacobsen, 2002; Timimi &
Taylor, 2003).

Gender differences
ADHD is observed more often in boys than girls, with a male to female ratio approximating
three to one (Skounti, Philalithis, & Galanakis, 2007). Relatedly, girls have been found to have
lower levels of inattention, hyperactivity, and oppositional/╉defiant behaviour compared to boys.
Research shows that boys under the age of 13 years tend to be overt and display severe disruptive
behaviours in the classroom; whereas, girls appear to exhibit more cognitive and academic prob-
lems (Gaub & Carlson, 1997). In addition, females are less likely to be identified in samples due to
the manifestation of the disorder, as they are less likely to exhibit disruptive behaviors compared
to ADHD males, and are more likely to go unnoticed if they present inattentive behaviours (Gaub
& Carlson, 1997). Importantly, a large number of referral for ADHD males who exhibit disruptive
behaviours occur in school settings; therefore, females who similarly display disruptive behaviors,
may be ignored (Gaub & Carlson, 1997; Gershon & Gershon, 2002).

Psychosocial impairment
As mentioned previously, children with ADHD often suffer from academic and social impair-
ments. Academic deficits, school-╉related problems, and peer neglect tend to be most associated
with elevated symptoms of inattention; whereas, peer rejection and, to a lesser extent, accidental
injury are frequently linked with symptoms of hyperactivity or impulsivity (Willcutt et al., 2012).
Compounding the stress for a person with ADHD, family relationships are consistently strained
and lead to discord and negative interactions. In addition, attentional problems frequently have a
significant impact on rates of mother–╉child rejection; however for fathers, rejection seems to sig-
nificantly impact their children’s attention problems (Lifford, Harold, & Thapar, 2008). Moreover,
peer relationships are affected by peer rejection, neglect, or teasing of the individual with ADHD.
In its severe presentation, ADHD is markedly impairing, due to its deleterious impact on social,
familial, and scholastic/╉occupational functioning (Hinshaw & Melnick, 1995; Hoza et al., 2005).
Individuals with ADHD have significant difficulty regulating their initial thoughts, behav-
iors, and emotions during a given task, thereby impacting their ability to successfully manage
tasks and achieve their desired outcomes (Barkley, 2006). A main feature of ADHD is difficulty
5
1

Comorbidity 115

with behavioral inhibition and self-╉regulation, with several models supporting this (Cleary &
Zimmerman, 2004). One such model primarily conceptualizes ADHD as an issue of behavioral
inhibition, which in turn, leads to a flawed concept of time awareness culminating in ineffective
time management (Barkley, 2006). This model closely links ADHD and its constructs to executive
function—╉a system that underlies the capacity for self-╉organisation and goal-╉directed actions;
thus, impairments in executive functioning result in behavioral disinhibition. Barkley claims that
the foundation for the key symptoms of ADHD (i.e., impulsivity, inattention, and hyperactivity) is
the result of the initial inability to diminish pre-╉potent responses to a given situation. Behavioral
inhibition allows individuals to halt an on-╉going response or response pattern; thus, creating a
delay and permitting self-╉directed action (Barkley, 2006). These self-╉directed actions are outlined
by core executive function processes, such as planning and working memory (Elliott, 2003); this
delay in time and executive functioning during normative functioning is what leads to effectual
and appropriate actions in addition to appropriate expression of emotions in relation to a task.
In contrast, for individuals with ADHD, difficulties inhibiting behavior and creating this delay
indicates they are often unable to prevent immediate responses to situations, such as answering
or talking out of turn, moderating emotional responses, controlling movements, or maintaining
attention and focusing on tasks with little immediate reward or positive consequence (Travell &
Visser, 2006).
Emotional impairments in children and adolescents with ADHD involve poor self-╉regulation
of emotion, excessive emotional expression, problems with anger and aggression, and greater
problems coping with frustration and empathy. Studies show that children with emotional and
behavioral difficulties are impulsively emotional and lack the ability to regulate their behavioral
responses to emotionally provoking events when compared to children without emotional and
behavioral difficulties (Cross, 2011).

Comorbidity
ADHD is highly comorbid with externalising disorders such as conduct disorder and opposi-
tional defiant disorder (ODD) (with comorbidity rates ranging from 43% to 93%) and internalis-
ing disorders (with comorbidity rates ranging from 13% to 51%) including anxiety and depression
(Jarrett & Ollendick, 2008). Moreover, children with ADHD are highly likely to develop ODD,
which involves difficulties with expressions of anger, hostility, frustration, and aggression toward
others, especially towards authority figures such as parents, alongside problems such as dis-
obedience. Boys with ADHD and comorbid ODD or conduct disorder in particular, have been
found to suffer from the impaired regulation of negative emotions (Melnick & Hinshaw, 2000).
Concordantly, approximately 45% of children with ADHD may also develop conduct disorder.
Furthermore within a subset of those with ADHD and conduct disorder the likelihood of child-
hood psychopathy such as callousness, lack of emotion and low empathy for others is increased
(Waschbusch, 2002).
Importantly, callous unemotional traits have been found to be prevalent in ADHD even after
controlling for conduct disorder (Musser et al., 2013). Marsh et al. (2013) compared ten to 17 year
olds with and without psychopathic traits on the subjective experiences of emotion during five
recent emotionally evocative life events. Their findings revealed that fewer children with psycho-
pathic traits reported the subjective experience of fear relative to other emotions. These results
suggest that comorbid psychopathy impairs fear learning, physiological responses to threats, and
the recognition of fear in others, as these children have difficulties expressing and displaying pro-╉
social emotions and behaviors, which is characterized by lower levels of empathy, a lack of a sense
of guilt or remorse, shallow or blunted affect, in conjunction with physiological under arousal
6
1

116 Emotion Regulation and Attention Deficit Hyperactivity Disorder

(Kimonis et al., 2008). These callous unemotional traits are therefore, important when consider-
ing emotional arousal and regulation in ADHD. As but one example, Musser et al. (2013) tested
ADHD children with age appropriate levels of pro-╉social behaviors and those with low levels
of pro-╉social behaviors on affect based tasks measuring emotional suppression and arousal. The
results from this study showed that children with ADHD and low pro-╉social behaviours displayed
a reduced level of arousal and elevated emotion dysregulation, which highlights the significance
of physiological responses in ADHD and emotion regulation.
ADHD also has a negative effect on the emotional wellbeing of the affected child or adolescent,
including those at risk for major depression (Edbom et al., 2006). Research additionally shows
that 75% of children diagnosed with ADHD are likely to have mood disorders and are therefore,
at an increased risk of developing depression (Biederman et al. 2008). Furthermore youths with
ADHD show greater levels of depressive symptoms, compared to those without ADHD (Lee et al.,
2008). A recent study by Seymour et al. (2014) found that emotion regulation mediated symptoms
of depression in ADHD youth, such that young people with ADHD and comorbid depression
exhibited poor emotion regulation strategies. Seymour et al. argue that this could be as a result of
executive function deficits, in particular working memory. Specifically, those with impairments
in working memory and inhibition experience and express heightened emotions in response to
emotionally laden stimuli when compared to individuals with intact working memory; as work-
ing memory affects the ability to effectively appraise emotional stimuli and supress negative and
positive emotions.

Executive dysfunctions
Executive functions are a set of inter-╉related cognitive processes that allow for effective prob-
lem solving, and facilitate goal directed activities; these processes are comprized of inhibition,
working memory, attention shifting, planning, initiating tasks, detecting and correcting errors
(Willcutt, Doyle, Nigg, Faraone, & Pennington, 2005). Researchers indicate that self-╉regulatory
processes underlie cognitive, behavioral and emotional regulation (Berger, Kofman, Livneh, &
Henik, 2007; Posner & Rothbart, 1998). This suggests that executive functions are involved in the
self-╉regulation of emotions in goal directed situations (Zelazo & Cunningham, 2007).
Emotion regulation has been found to be consistently linked to inhibitory processes; for exam-
ple, a study of typically developing preschool children’s performance on an emotion regulation
task (i.e., responses to a disappointing gift) significantly correlated with responses on tasks inves-
tigating inhibitory processes (i.e., Simon Says) and suppression or slowing of responses (e.g., not
pulling a lever or drawing a line very slowly) (Carlson & Wang, 2007). Moreover, a study measur-
ing the performance of young adults during a Stroop task revealed that this measure of inhibitory
functions and conflict monitoring was linked with the ability to successfully manage negative
responses to unfamiliar and visually unappetizing food (Kieras, Tobin, Graziano, & Rothbart,
2005). Additionally, when asked to divide their attention by remembering an eight-╉digit number
during a task to challenge executive function processing capacity, individuals were increasingly
found to have difficulties modulating their negative responses. This is supported by Walcott and
Landau’s (2004) findings, in that emotion regulation was strongly associated with the speed of the
inhibition process using tasks such as the Stop Signal Reaction Time Task (SSRT).
Hoeksma, Oosterlan, and Schipper (2004) found that in children aged between ten and 13
anger variability over a number of days was strongly associated with outcomes on SSRT, which
measures the time needed to stop an inappropriate response. This is a further indication of behav-
ioral inhibition, as studies show that deficits in SSRT also reflect impairments in attentional and
cognitive processes (Alderson, Rapport, & Kofler, 2007). Rich et al. (2008) showed that children
7
1

Executive dysfunctions 117

with severe mood disorder had problems with attentional orienting and initial attentional pro-
cessing; moreover, approximately 80% of their participants had comorbid ADHD. This suggests
that the underlying processes involved in ADHD are also related to attentional processes and
emotion regulation, supporting the view that executive function task difficulties are closely linked
to ADHD and emotion regulation (Skirrow, McLoughlin, Kuntsi, & Asherson, 2009).
As outlined earlier, children with ADHD consistently display deficits in most areas of executive
functions (Barkley, 2006). Research suggests that behavioral disinhibition is an important charac-
teristic of ADHD; Nigg (2001) suggests there are two distinct forms of impairments in inhibition
that can be applied to ADHD. Firstly, motivational inhibition automatically ceases an on-​going
response that is usually caused by fear or anxiety as a result of a novel event. Secondly, execu-
tive inhibition involves processing of the deliberate suppression of a response for goal-​directed
purposes. As proposed by Barkley’s (1997) behavioral disinhibition theory, children with ADHD
do not effectively respond to social circumstances, but rather display rules detached from the
emotional context of the situation. These children therefore, appear more dysregulated as they fail
to consider social cues and rules, thereby appearing more socially dysregulated. According to this
theory, the successful regulation of emotions would therefore, depend on successful behavioral
inhibition. This is supported by a study investigating behavioral disinhibition and its associations
to emotion regulation using a frustration-​inducing task (Walcott & Landau, 2004). In this study,
boys with and without ADHD were explicitly given instruction to hide their emotional display
in the presence of a peer. Results determined that boys with ADHD failed to succeed on this
task, whereas, non-​symptomatic boys were more effective at regulating their emotional displays
in response to contextual demands. Importantly, disinhibition scores were higher for boys with
ADHD than those without ADHD (Walcott & Landau, 2004).
The findings discussed above are generally supported by imaging data investigating executive
processes and ADHD, whereby the frontal regions of the brain are associated with inhibitory con-
trol and emotional processing (Posner et al., 2011). Thus, children with ADHD are shown to have
increased activation in prefrontal regions, relative to healthy controls on an emotional process-
ing task; these findings were specific to emotional processing even after controlling for cognitive
processes. Essentially, this indicates that normal function in the prefrontal regions are impaired
in ADHD, however, they may also mediate or facilitate affective responses i.e., negatively valenced
words such as “kill” drew increased attention than neutral words such as “month;” similarly posi-
tive words could induce self-​reflection to a greater extent than neutral words (Posner et al., 2011).
In relation to emotion regulation, Shaw et  al. (2014) proposed a top-​down regulatory pro-
cess and bottom-​up mechanistic theory to explain the processes affecting emotion regulation in
ADHD. According to a bottom-​up psychological mechanism, the attention systems identify emo-
tionally significant stimuli and exert control—​an aspect that is thought to be impaired in ADHD.
In contrast, in healthy individuals, affectively salient stimuli receive appropriate sensory coding
and early detection, whilst this effect is significantly reduced in ADHD as a result of heightened
emotions (i.e., the over perception of negative stimuli). Concordantly, the accurate identification
of emotions in human faces is associated with well-​regulated behavior; thus, misperception could
be caused as a result of emotion dysregulation. Furthermore, aversion to delayed rewards is an
indication of impulsivity; this is mediated in the limbic regions of the brain, which are also respon-
sible for emotion processing; thus, it is probable that these brain regions may also be involved in
emotion regulation (Musser et al., 2013; Shaw et al., 2014).
In relation to top-​down regulatory processes, the importance of the autonomic nervous system
is paramount, as it recognizes emotional valence and task demands, particularly when the stimuli
are negative rather than positive. This is difficult for those with ADHD because they lack physi-
ological indicators of regulation. Thus, the inability to focus on a goal or allocate appropriate levels
8
1

118 Emotion Regulation and Attention Deficit Hyperactivity Disorder

of attention to a task means that individuals with ADHD have difficulties managing emotions or
focusing on emotional stimuli. For example, when completing an emotional Stroop task, the per-
formance of adolescents with ADHD is severely impaired when compared to healthy counterparts
(Posner et al., 2011).

Neural mechanisms
ADHD is a neurodegenerative disorder with most models highlighting deficits in the frontal lobe
networks. In particular the prefrontal cortex (PFC) region has been consistently found to mediate
cognitive control processes, including decision-╉making and emotion regulation, in particular the
orbitofrontal cortex, dorsomedial prefrontal cortex, anterior cingulate gyrus, dorsolateral pre-
frontal cortex and ventrolateral prefrontal cortex (Phillips, Ladouceur & Drevets, 2008). Shaw
et al. (2014) claim that for individuals with ADHD, the prefrontal regions, including the ventro-
lateral, orbitofrontal and medial prefrontal cortices are impaired. Plessen and colleagues (2009)
suggest that deficits in the connections between the amygdala and orbitofrontal cortex may lead to
behavioral disinhibition. The orbitofrontal cortex is strongly connected with the amygdala, thala-
mus and multiple cortical regions, thus, it is an important region involved in emotion regulation
processes. In addition, the amygdala plays a crucial role as it is involved in processing emotion
and emotional behavior.
The majority of studies have shown amygdala hyperactivation in ADHD, during both the
subliminal perception of fearful expressions and while subjects rated their fear of neutral faces
(Malisza et al., 2011). These findings are similar to behavioral measures of delay aversion, dur-
ing which amygdala hyperactivation was observed for the processing of delayed rewards (Plichta
et al., 2009). The anticipation (and receipt) of rewards causes reduced ventral striatum responsive-
ness in ADHD, thus contributing to aversion delay. This is supported by dysfunction in a neural
network composed of the amygdala, ventral striatum, and orbitofrontal cortex, which mediates
emotional stimuli, and is implicated in emotion regulation. Therefore, Shaw et  al. (2014) have
argued that emotion dysregulation in ADHD implicates dysfunction in the amygdala, ventral
striatum and orbitofrontal cortex. Relatedly, lesion studies have shown that the orbitofrontal
region, in particular, is important for the generation of emotional states and emotion regulation
(Ochsner & Gross, 2004). Thus, neural theory predicts (Shaw et al., 2014) deficits in these regions
are strongly associated with symptoms of both ADHD and emotion dysregulation.

Emotion dysregulation
One of the earlier models for emotional dysregulation (ED) (Cicchetti, Ackerman, & Izard,
1995) posited that regulating emotion requires certain control mechanisms involving structure
or a strategy that will allow for co-╉ordination and actions. Cicchetti and colleagues (1995) out-
lined four of these aspects: Firstly, control concerns the cause of felt emotion, involving cogni-
tive and affective mechanisms. Secondly, control structures mediate the output of this emotional
system, whereby earlier mechanisms of cognitive and affective processes are reflected in expres-
sion. Thirdly, control structures coordinate expression and inhibit responses based on context.
Finally, this control structure deviates for those with externalising and internalising problems, as
these individuals tend to suffer from weak or absent control structures. As such, individuals with
ADHD are thought to have problems moderating or suppressing the emotional reactions they
experience, leading to impulsive and severe emotional reactions toward events when compared
to non-╉ADHD individuals of a similar age. Emotion dysregulation therefore, results from a lack
of knowledge concerning affective behavior or difficulty in modulating emotional responses to
social situations or environmental demands (Saarni, 1999).
9
1

Development 119

Impairments in emotional control are closely associated with hyperactive and impulsive symp-
toms, and likely arise from the poor inhibitory capacity involved in ADHD (Barkley, Murphy,
& Fisher, 2008). Observational studies show that children with ADHD display heightened emo-
tional reaction and frustrations compared to their non-╉ADHD peers; this is further supported by
parent reports of increased levels of sadness, anger and guilt. Importantly, these youth have dif-
ficulty self-╉regulating these negative emotions (Berlin, Bohlin, Nyberg, & Janols, 2004; Braaten &
Rosen, 2000; Melnick & Hinshaw, 2000). Moreover, as irritability is an aspect of reactive aggres-
sion and emotional outbursts, it is considered one of the main outcomes of emotion dysregula-
tion in ADHD (Leibenluft, 2011). In fact, a study examining ADHD children with and without
irritability found increased rates of ODD and depression/╉dysthymia in children with irritable
mood and ADHD (Ambrosini, Bennet, & Elia, 2013).
A recent meta-╉analysis by Shaw et  al. (2014) revealed a consistent increase in aggressive
behavior in ADHD compared to non-╉ADHD samples. Their results suggest a strong association
between aggression and hyperactivity-╉impulsivity rather than between aggression and inatten-
tion. Emotion dysregulation was further reflected in frustration inducing situations in ADHD.
In addition, children with ADHD were more likely to express negative affect and have emotional
outbursts when compared with non-╉ADHD participants during challenging tasks. Based on their
meta-╉analysis Shaw et al. (2014) described three distinct features of ADHD and emotion dysregu-
lation. The first feature suggests that at its core, emotion dysregulation is a main characteristic
of ADHD and its symptoms of hyperactivity, impulsivity and inattention, are reflective of defi-
cits in executive functions. The second feature considers ADHD and emotion dysregulation as
a unique entity, formed as a result of distinct neurocognitive features and the clinical outcomes
for those with the combination of ADHD and emotion dysregulation. The third feature refers
to the fact that symptoms of ADHD and emotion dysregulation overlap and are underlined by
dissociable neurocognitive deficits such as impairments in executive function, which impacts
decision-╉making and emotional control. This model is supported by correlations observed for
deficits in emotional processes, for example in emotion recognition and frustration tolerance
(Banaschewski et al., 2012); however it is important to note that not all those with ADHD display
impaired levels of emotion dysregulation
Results from longitudinal studies reveal that ADHD symptoms and emotion dysregulation
difficulties emerge in early childhood and continue into adulthood (Biederman et al., 2012).
Skirrow, McLoughlin, Kuntsi, and Asherson (2009) argue that these symptoms of emotional
dysregulation significantly differ from mood instability, as mood instability is used to describe
volatile, irritable and changeable mood with a hot temper and low frustration tolerance in
the absence of underlying deficits. Emotion dysregulation however, is believed to be an active
modification or alteration of on-╉going emotional responses. These responses are associated to
emotions linked with the environment and therefore, part of emotional patterns. Therefore,
those with emotion dysregulation do not usually suffer from mood instability, as mood insta-
bility arises from existing processes that lead to deviant emotional responses independent of
regulatory processes.

Development
Research shows a strong association between ADHD and emotion dysregulation (Sjöwall, Roth,
Lindqvist, & Thorell, 2012). Stringaris and Goodman’s (2009) study examining 5,326 youth
found mood lability (i.e., poorly controlled shifts in emotion) in 38% of children with ADHD.
Parent reports of the Child Behaviour Checklist revealed that adolescents with mood and aggres-
sion problems also tended to suffer from attention difficulties and were more likely to suffer from
0
2
1

120 Emotion Regulation and Attention Deficit Hyperactivity Disorder

emotion dysregulation among those likely to have ADHD (Althoff et al., 2006). Shaw et al. (2014)
noted that clinic-╉based studies in young people with ADHD conveyed similar levels of emotion
dysregulation, ranging between 24% and 50%.
Longitudinal research of children with ADHD spanning into adulthood has rarely consid-
ered emotion dysregulation, but rather has focused on outcomes from the DSM-╉IV disruptive
and antisocial disorders (Klein et al., 2012). Stringaris, Maughan, and Goodman (2010) con-
ducted a longitudinal study of 7,140 children and found that temperamental emotionality in
three-╉year-╉olds predicted co-╉morbid ADHD with internalising disorders by the age of seven.
Another longitudinal study by Sanson, Smart, Prior, and Oberklaid (1993) showed that infants
who developed hyperactive symptoms alone did not differ in their temperament from typical
infants; whereas, children who developed ADHD and aggressive traits were prominently unco-
operative and irritable from infancy. Therefore, a difficult temperament with significant nega-
tive emotionality has been linked with later ADHD combined with emotion dysregulation.
Nonetheless, environmental factors such as parental criticism and hostility were associated
with the development of conduct problems in children with ADHD, and with the development
of childhood ADHD in pre-╉schoolers with behavioral problems. Shaw et al. (2014) claims that
poor parental emotion regulation is reflected in high levels of hostility, thereby contributing to
the development of emotional dysregulation in children with ADHD.

Transition from adolescence to adulthood


Transitions during the adolescent years has been associated with numerous issues affecting
social interaction and emotional outcomes. Resultantly, low self esteem may manifest due to
poor relationships with peers and the inability to effectively participate in social exchanges
such as sharing, cooperating, and turn taking. These issues are adversely affected for those with
ADHD as a result of delayed self-╉regulation (Barkley, 2006). These outcomes are further linked
to impairments in self-╉esteem and sociability (Hoy et al., 1978). Research shows that the symp-
toms of ADHD are continuously changing during the adolescent years and into adulthood
(Wolraich et al., 2006). In particular, hyperactivity becomes less prominent during this age,
compared to inattention, which tends to remain persistent during adulthood (Barkley, 2006).
Unlike hyperactivity, symptoms of inattention and executive function difficulties greatly affect
academic achievement more so than symptoms of hyperactivity and impulsivity. Nonetheless,
deficits in certain aspects of executive functioning (e.g., working memory) may prevent ado-
lescents from reading, listening and comprehending and therefore, planning, which results in
future rewards being less valued (Barkley, 2006). Furthermore, adolescents with ADHD are
more likely to display poor delay of gratification and are less likely to persevere with set goals
and have poor emotion regulation competencies as a result of deficits in anger and frustration
control (Barkley, 2006).
Emotional dysregulation is considered to be an important feature of adult ADHD with 34–╉70%
of adults in clinical samples of ADHD reporting impaired emotion regulation (Able, Johnston,
Adler, & Swindle, 2007). Aggressive behaviors also continue to be persistent, as indicated in a
study of 950 adults diagnosed with ADHD. As reported by Able et al. (2007), those with ADHD
scored themselves higher in interpersonal conflict and reported negative, conflicted social ties.
Somewhat similarly, a longitudinal study investigating the outcomes of ADHD children found
higher rates of emotion dysregulation in adults with persistent ADHD when compared with
adults with remitted ADHD. This suggests that as symptoms of ADHD improve, so too does emo-
tion dysregulation (Shaw et al., 2014).
1
2

TREATMENT 121

Clinical implications
Empirical findings demonstrate physiological and observable behaviors consistent with ED in
children with ADHD (Musser et al., 2011; Musser, Galloway-╉Long, Frick, & Nigg, 2013; Seymour
et al., 2012; Walcott & Landau, 2004). These include demoralization, learned helplessness, low
self-╉esteem, fear and anxiety, increased frustration and occupational challenges. Previous stud-
ies have found ADHD boys to be socially inflexible, emotionally intense with poor attention and
concentration levels (Sanson, Smart, Prior, & Oberklaid, 1993). Using an unsolvable puzzle task
to elicit aggression, boys with ADHD who were considered highly aggressive were further found
to be more emotionally reactive and less effective at emotion regulation than boys with low lev-
els of aggression and without ADHD (Hinshaw & Melnick, 1995). This suggests that aggression
rather than ADHD is responsible for this level of emotional response. Additional manifestations
of emotion dysregulation involve over-╉reactivity to positive and negative emotions (Martel &
Nigg, 2006), lack of emotional control (Erhardt & Hinshaw, 1994; Saunders & Chambers, 1996);
and impatience which most likely leads to peer perceptions of youth with ADHD as easily excited,
disruptive, or intrusive in their social interactions (Landau & Moore, 1991). Children with inat-
tentive presentations of ADHD show emotion dysregulation enhanced by emotional intensity and
display heightened emotions (Wheeler, Maedgen & Carlson, 2000).
Negative emotionality however is also a characteristic of ODD which includes loss of temper, as
the child gets easily angry and resentful (Barkley et al., 2010). Negative emotionality is similar to
emotional dysregulation, however it is a risk-╉factor for ED (Belsky, Friedman, & Hsieh, 2001) and
is considered a risk factor for developing ODD in children with ADHD (Martel & Nigg, 2006).
Children and adolescents with ADHD are therefore more likely to experience impairments in
social relationships, as they exhibit aggressive behaviors and consistent rule breaking unlike typi-
cally developing peers (Buhrmester, Whalen, Henker, MacDonald, & Hinshaw, 1992).

Treatment
Treatments for ADHD involve a broad range of options including behavioral therapy, psycho-
therapeutic approaches and pharmacotherapy; the aim of treatment is to treat the disorder as early
and as effectively as possible. When considering non-╉pharmacological treatments, studies show
that parent and family education is important, along with effective parent training in behavioral
management involving teachers to improve classroom behaviors. These treatments indicate that
with appropriate behavior modification training and special education placement, outcomes for
children with ADHD can be greatly improved (Thompson et al., 2004). In addition, treatment
programs have shown that the management of adolescents with ADHD can be effective; these
include parent and teacher training in behavioral management, particularly contingency man-
agement methods applied in classrooms and similar settings, such as summer camp (Antshel &
Barkley, 2008). However, Barkley (2006) argues that interventions for behavioral management in
children with ADHD are most effective when inappropriate behaviors are targeted in the child’s
natural environment, as it occurs. Subsequently, Barkley suggests it is important to assist the
child/╉individual in understanding suitable behavior which is contextually expected.
Most psychosocial treatment programs involve a multimodal treatment plan part of which
includes medication (Jensen et  al., 2001). The Multimodal Treatment Study of Children with
ADHD investigated long term outcomes of interventions, including medication and behavior
modification in combination and alone. The results showed that medication alone and medica-
tion with behavior modification was superior to behavior modification alone or standard com-
munity care (MTA Cooperative Group, 1999). Apart from decreasing levels of ADHD symptoms,
21

122 Emotion Regulation and Attention Deficit Hyperactivity Disorder

these two intervention strategies improved aggressive behavior, social skills, academic achieve-
ment, and parent-╉child relationships. Stimulant medications such as methylphenidate have been
found to be effective in improving academic outcomes and emotional wellbeing. In addition, a
study evaluating the effectiveness of multimodal psychosocial treatment of children with ADHD
being treated with methylphenidate reported a consistent pattern of improvement in academic
achievement and emotional status, particularly self-╉esteem and ratings of depression (Hechtman
et al., 2004). Non-╉stimulant medications, such as atomoxetine have also been found to reduce core
ADHD symptoms, improve social interactions and quality of life in children and adolescents with
ADHD (Cheng et al., 2007; Wilens et al., 2006).
In relation to psychological intervention, Cognitive Behavioural Therapy (CBT) has been
shown to benefit individuals with ADHD by helping them to understand and categorize the emo-
tions they experience accurately. Importantly, CBT has been found to help with labeling emotions
correctly and coping with intense negative reactions (Mongia & Hechtman, 2012). Moreover,
these skills can be developed alongside mindfulness training (Mongia & Hechtman, 2012), which
promotes present centered focused awareness of emotions (Farb et al., 2007). Additionally, inter-
ventions aiming to treat avoidance behavior and mood disturbances in ADHD may also improve
emotion regulation by enhancing motivation and providing individuals with strategies to cope
with daily life (Mongia & Hechtman, 2012).
Considering emotion dysregulation in ADHD treatment has been challenging, primarily
this is due to the fact that studies have measured emotional changes as a secondary outcome
(Shaw et al., 2014). However, one of the few studies measuring the attributes of stimulants on
emotional expression found improvement in emotional dysregulation, parallel to improve-
ments observed in hyperactivity and impulsivity (Mccracken et al., 2003). According to Manos
et al.’s (2011) literature review, emotional lability and irritability reduced by 3% in ADHD as
a result of medication alone. Stimulants have also been found to improve emotion recogni-
tion, whilst concurrently improving performance (Conzelmann et al., 2011). These findings
are supported by neural activities, as medicated adolescents have been found to have reduced
activity in the prefrontal regions, similar to healthy controls, contrasted by increased reactiv-
ity found in ADHD participants not taking medication. ADHD adolescents taking medica-
tion were found to have better performance on emotional processing tasks when compared to
ADHD adolescents without medication. Shaw et al. (2014) suggest that stimulant treatment
of the core symptoms of ADHD also leads towards improvement in emotion dysregulation.
Additionally, behaviour modification combined with medication is effective at reducing exter-
nalising and internalising symptoms, which are linked with emotion dysregulation (Stringaris
& Goodman, 2009).

Conclusions
ADHD is one of the most commonly occurring psychiatric disorders of childhood (Spencer,
Biederman, & Mick, 2007). Moreover, it frequently persists into adolescence and adulthood and is
associated with multiple functional impairments. Research has revealed that externalizing behav-
ioral problems and social impairment are associated with emotion dysregulation in children
with ADHD (Wheeler, Maedgen & Carlson, 2000; Melnick & Hinshaw, 2000; Parker, Majeski, &
Collin, 2004). Emotion dysregulation is strongly linked to inhibitory deficit, which may manifest
into socially inappropriate behavioral responses to extreme emotional expression and the inability
to self-╉regulate (Barkley, 2006). This then suggests the individual finds it difficult to self-╉soothe
during enhanced emotional experiences, focus on the task at hand, and to organize thoughts
to achieve goal driven behavior (Lynn, Carroll, Houghton, & Cobham, 2013). The association
3
2
1

Conclusions 123

between emotion dysregulation and ADHD has been mainly explored in children; therefore, gen-
eralizability across developmental stages remains largely unaddressed. Moreover, ADHD is highly
comorbid with the internalising/╉externalizing disorders which significantly impact emotion dys-
regulation, yet very few studies have considered the effect of subtype or comorbidity on emotion
dysregulation (Wheeler, Maedgen & Carlson, 2000; Melnick & Hinshaw, 2000).
In summary, emotion dysregulation affects approximately 25–╉╉45% of children and between
30–╉70% of adults with ADHD. It represents a major source of impairment and presages a poor
clinical outcome (Shaw et al., 2014). Emotion dysregulation in ADHD may be caused through
deficits at multiple levels, ranging from abnormal early orientation to emotional stimuli to deficits
in cognitive processes, in particular working memory and response inhibition. Although these
deficits may contribute to emotion dysregulation they alone do not explain its presence in ADHD,
as the underlying mechanism is likely complex, and is influenced by impairments in neural net-
works in the prefrontal cortex and executive functioning processes.

References
Able, S. L., Johnston, J. A., Adler, L. A., & Swindle, R. W. (2007). Functional and psychosocial impairment
in adults with undiagnosed ADHD. Psychological Medicine, 37(1), 97–╉107.
Alderson, R. M., Rapport, M. D., & Kofler, M. J. (2007). Attention-╉deficit/╉hyperactivity disorder and
behavioral inhibition: a meta-╉analytic review of the stop-╉signal paradigm. Journal of Abnormal Child
Psychology, 35(5), 745–╉758.
Althoff, R. R., Copeland, W. E., Stanger, C., Derks, E. M., Todd, R. D., Neuman, R. J., Van Beijsterveldt,
T. C., Boomsma, D. I. and Hudziak, J. J. (2006). The latent class structure of ADHD is stable across
informants. Twin Research and Human Genetics, 9(04), 507–╉522.
Ambrosini, P. J., Bennett, D. S., & Elia, J. (2013). Attention deficit hyperactivity disorder characteristics: II.
Clinical correlates of irritable mood. Journal of Affective Disorders, 145(1), 70–╉76.
American Psychiatric Association. (1980). Diagnostic and statistical manual of mental disorders (3rd ed.).
Washington, DC: American Psychiatric Association.
American Psychiatric Association. (1994). Diagnostic and statistical manual of mental disorders (4th ed.).
Washington, DC: American Psychiatric Association.
American Psychiatric Association. (2000). Diagnostic and statistical manual of mental disorders (4th text
revision ed.). Washington, DC: American Psychiatric Association.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders. (5th ed.).
Washington, DC: American Psychiatric Association.
Antshel, K. M., & Barkley, R. (2008). Psychosocial interventions in attention deficit hyperactivity disorder.
Child and Adolescent Psychiatric Clinics of North America, 17(2), 421–╉437.
Asherson, P., Akehurst, R., Kooij, J., Huss, M., Beusterien, K., Sasane, R., … Hodgkins, P. (2012). Under
Diagnosis of Adult ADHD: Cultural Influences and Societal Burden. Journal of Attention Disorders, 16,
20S–╉38S.
Banaschewski, T., Jennen-╉Steinmetz, C., Brandeis, D., Buitelaar, J. K., Kuntsi, J., Poustka, L., … Asherson,
P. (2012). Neuropsychological correlates of emotional lability in children with ADHD. Journal of Child
Psychology and Psychiatry, 53(11), 1139–╉1148.
Barkley, R. A. (1990). Attention deficit hyperactivity disorder: A handbook for diagnosis and treatment.
New York, NY, US: Guilford.
Barkley, R. A. (1997). Behavioral inhibition, sustained attention, and executive functions: constructing a
unifying theory of ADHD. Psychological Bulletin, 121(1), 65.
Barkley, R. A. (2006). Primary symptoms, diagnostic criteria, prevalence, and gender differences. Attention-╉
deficit hyperactivity disorder: A handbook for diagnosis and treatment, 3, 76–╉121.
Barkley, R. A., Murphy, K. R., & Fischer, M. (2010). ADHD in adults: What the science says. New York, NY,
US: Guilford Press.
4
2
1

124 Emotion Regulation and Attention Deficit Hyperactivity Disorder

Bauermeister, J. J., Canino, G., Polanczyk, G., & Rohde, L. A. (2010). ADHD across cultures: Is there
evidence for a bidimensional organization of symptoms? Journal of Clinical Child & Adolescent
Psychology, 39(3), 362–​372.
Belsky, J., Friedman, S. L., & Hsieh, K. H. (2001). Testing a Core Emotion‐Regulation Prediction: Does
Early Attentional Persistence Moderate the Effect of Infant Negative Emotionality on Later
Development? Child Development, 72(1), 123–​133.
Berger, A., Kofman, O., Livneh, U., & Henik, A. (2007). Multidisciplinary perspectives on attention and the
development of self-​regulation. Progress in Neurobiology, 82(5), 256–​286.
Berlin, L., Bohlin, G., Nyberg, L., & Janols, L. O. (2004). How well do measures of inhibition and other
executive functions discriminate between children with ADHD and controls? Child Neuropsychology,
10(1), 1–​13.
Biederman, J., Melmed, R. D., Patel, A., McBurnett, K., Konow, J., Lyne, A., & Scherer, N. (2008). A
randomized, double-​blind, placebo-​controlled study of guanfacine extended release in children and
adolescents with attention-​deficit/​hyperactivity disorder. Pediatrics, 121(1), e73–​e84.
Biederman, J., Petty, C. R., O’Connor, K. B., Hyder, L. L., & Faraone, S. V. (2012). Predictors of persistence
in girls with attention deficit hyperactivity disorder: results from an 11‐year controlled follow‐up study.
Acta Psychiatrica Scandinavica, 125(2), 147–​156.
Braaten, E. B., & Rosen, L. A. (2000). Self-​regulation of affect in attention deficit-​hyperactivity disorder
(ADHD) and non-​ADHD boys: differences in empathic responding. Journal of Consulting and Clinical
Psychology, 68(2), 313.
Buhrmester, D., Whalen, C. K., Henker, B., MacDonald, V., & Hinshaw, S. P. (1992). Prosocial behavior
in hyperactive boys: Effects of stimulant medication and comparison with normal boys. Journal of
Abnormal Child Psychology, 20(1), 103–​121.
Castellanos, F. X., & Tannock, R. (2002). Neuroscience of attention-​deficit/​hyperactivity disorder: the
search for endophenotypes. Nature Reviews Neuroscience, 3(8), 617–​628.
Carlson, S. M., & Wang, T. S. (2007). Inhibitory control and emotion regulation in preschool children.
Cognitive Development, 22(4), 489–​510.
Cheng, J. Y., Chen, R. Y., Ko, J. S., & Ng, E. M. (2007). Efficacy and safety of atomoxetine for attention-​
deficit/​hyperactivity disorder in children and adolescents—​meta-​analysis and meta-​regression analysis.
Psychopharmacology, 194(2), 197–​209.
Cicchetti, D., Ackerman, B. P., & Izard, C. E. (1995). Emotions and emotion regulation in developmental
psychopathology. Development and Psychopathology, 7(01), 1–​10.
Conzelmann, A., Woidich, E., Mucha, R. F., Weyers, P., Jacob, C. P., Lesch, K. P., & Pauli, P. (2011).
Methylphenidate normalizes emotional processing in adult patients with attention-​deficit/​hyperactivity
disorder: preliminary findings. Brain Research, 1381, 159–​166.
Cleary, T. J., & Zimmerman, B. J. (2004). Self‐regulation empowerment program: A school‐based program
to enhance self‐regulated and self‐motivated cycles of student learning. Psychology in the Schools, 41(5),
537–​550.
Cross, M. (2011). Children with social, emotional and behavioural difficulties and communication
problems: There is always a reason. London, UK: Jessica Kingsley Publishers.
Douglas, V. I., & Parry, P. A. (1983). Effects of reward on delayed reaction time task performance of
hyperactive children. Journal of Abnormal Child Psychology, 11, 313–​326
Elliott, R. (2003). Executive functions and their disorders Imaging in clinical neuroscience. British Medical
Bulletin, 65(1), 49–​59.
Edbom, T., Lichtenstein, P., Granlund, M., & Larsson, J. O. (2006). Long‐term relationships between
symptoms of Attention Deficit Hyperactivity Disorder and self‐esteem in a prospective longitudinal
study of twins. Acta Pædiatrica, 95(6), 650–​657.
Esser, G., Schmidt, M. H., & Woerner, W. (1990). Epidemiology of course of psychiatric disorders in
school-​age children—​Results of a longitudinal study. Journal of Child and Psychology and Psychiatry
and Allied Disciplines, 31, 243–​263.
5
2
1

Conclusions 125

Erhardt, D., & Hinshaw, S. P. (1994). Initial sociometric impressions of attention-​deficit hyperactivity
disorder and comparison boys: Predictions from social behaviors and from nonbehavioral variables.
Journal of Consulting and Clinical Psychology, 62(4), 833.
Faraone, S. V., Sergeant, J., Gillberg, C., & Biederman, J. (2003). The worldwide prevalence of ADHD: is it
an American condition? World Psychiatry, 2(2), 104–​113.
Farb, N. A., Segal, Z. V., Mayberg, H., Bean, J., McKeon, D., Fatima, Z., & Anderson, A. K. (2007).
Attending to the present: mindfulness meditation reveals distinct neural modes of self-​reference. Social
Cognitive and Affective Neuroscience, 2(4), 313–​322.
Gaub, M., & Carlson, C. L. (1997). Gender differences in ADHD: a meta-​analysis and critical review.
Journal of the American Academy of Child & Adolescent Psychiatry, 36(8), 1036–​1045.
Gershon, J. (2002). Gender Differences in ADHD. The ADHD Report, 10(4), 8–​16.
Gershon, J., & Gershon, J. (2002). A meta-​analytic review of gender differences in ADHD. Journal of
Attention Disorders, 5(3), 143–​154.
Goetz, M., Yeh, C.-​B., Ondrejka, I., Akay, A., Herczeg, I., Dobrescu, I., … Treuer, T. (2010). A 12-​month
prospective, observational study of treatment regimen and Quality of Life associated with ADHD in
Central and Eastern Europe and Eastern Asia. Journal of Attention Disorders, 20(10), 1–​16.
Hechtman, L., Abikoff, H., Klein, R. G., Weiss, G., Respitz, C., Kouri, J., & Pollack, S. (2004). Academic
achievement and emotional status of children with ADHD treated with long-​term methylphenidate and
multimodal psychosocial treatment. Journal of the American Academy of Child & Adolescent Psychiatry,
43(7), 812–​819.
Hinshaw, S. P., & Melnick, S. M. (1995). Peer relationships in boys with attention-​deficit hyperactivity
disorder with and without comorbid aggression. Development and Psychopathology, 7(04), 627–​647.
Hoeksma, J. B., Oosterlaan, J., & Schipper, E. M. (2004). Emotion regulation and the dynamics of
feelings: A conceptual and methodological framework. Child Development, 75(2), 354–​360.
Hoy, E., Weiss, G., Minde, K., & Cohen, N. (1978). The hyperactive child at adolescence: Cognitive,
emotional, and social functioning. Journal of Abnormal Child Psychology, 6(3), 311–​324.
Hoza, B., Mrug, S., Gerdes, A. C., Hinshaw, S. P., Bukowski, W. M., Gold, J. A., … & Arnold, L. E. (2005).
What aspects of peer relationships are impaired in children with attention-​deficit/​hyperactivity
disorder?. Journal of Consulting and Clinical Psychology, 73(3), 411.
Jacobson, K. (2002). ADHD in Cross-​Cultural Perspective: Some Empirical Results. American
Anthropologist, 104(1), 283–​287.
Jarrett, M. A., & Ollendick, T. H. (2008). A conceptual review of the comorbidity of attention-​deficit/​
hyperactivity disorder and anxiety: Implications for future research and practice. Clinical Psychology
Review, 28(7), 1266–​1280.
Jensen, P. S., Hinshaw, S. P., Kraemer, H. C., Lenora, N., Newcorn, J. H., Abikoff, H. B., … & Vitiello, B.
(2001). ADHD comorbidity findings from the MTA study: comparing comorbid subgroups. Journal of
the American Academy of Child & Adolescent Psychiatry, 40(2), 147–​158.
Kieras, J. E., Tobin, R. M., Graziano, W. G., & Rothbart, M. K. (2005). You Can’t Always Get What You
Want: Effortful Control and Children’s Responses to Undesirable Gifts. Psychological Science, 16(5),
391–​396.
Kimonis, E. R., Frick, P. J., Skeem, J. L., Marsee, M. A., Cruise, K., Munoz, L. C., … & Morris, A. S. (2008).
Assessing callous–​unemotional traits in adolescent offenders: Validation of the Inventory of Callous–​
Unemotional Traits. International Journal of Law and Psychiatry, 31(3), 241–​252.
Klein, R. G., Mannuzza, S., Olazagasti, M. A. R., Roizen, E., Hutchison, J. A., Lashua, E. C., &
Castellanos, F. X. (2012). Clinical and functional outcome of childhood attention-​deficit/​hyperactivity
disorder 33 years later. Archives of General Psychiatry, 69(12), 1295–​1303.
Landau, S., & Moore, L. A. (1991). Social skill deficits in children with attention-​deficit hyperactivity
disorder. School Psychology Review, 20(2), 235–​251.
Lee, S. S., Lahey, B. B., Owens, E. B., & Hinshaw, S. P. (2008). Few preschool boys and girls with ADHD are
well-​adjusted during adolescence. Journal of Abnormal Child psychology, 36(3), 373–​383.
6
2
1

126 Emotion Regulation and Attention Deficit Hyperactivity Disorder

Leibenluft, E. (2011). Severe mood dysregulation, irritability, and the diagnostic boundaries of bipolar
disorder in youths. American Journal of Psychiatry, 168 (2), 129–​142.
Lifford, K. J., Harold, G. T., & Thapar, A. (2008). Parent–​child relationships and ADHD symptoms: a
longitudinal analysis. Journal of Abnormal Child Psychology, 36(2), 285–​296.
Lynn, S., Carroll, A., Houghton, S., & Cobham, V. (2013). Peer relations and emotion regulation of
children with emotional and behavioural difficulties with and without a developmental disorder.
Emotional and Behavioural Difficulties, 18(3), 297–​309.
Malisza, K. L., Clancy, C., Shiloff, D., Holden, J., Jones, C., Paulson, K., … & Chudley, A. E. (2011).
Functional magnetic resonance imaging of facial information processing in children with autistic
disorder, attention deficit hyperactivity disorder and typically developing controls. International Journal
of Adolescent Medicine and Health, 23(3), 269–​277.
Manos, M. J., Brams, M., Childress, A. C., Findling, R. L., López, F. A., & Jensen, P. S. (2011). Changes
in emotions related to medication used to treat ADHD. Part I: literature review. Journal of Attention
Disorders, 15(2), 101–​112.
Marsh, A. A., Finger, E. C., Fowler, K. A., Adalio, C. J., Jurkowitz, I. T., Schechter, J. C., … & Blair, R. J. R.
(2013). Empathic responsiveness in amygdala and anterior cingulate cortex in youths with psychopathic
traits. Journal of Child Psychology and Psychiatry, 54(8), 900–​910.
Martel, M. M., & Nigg, J. T. (2006). Child ADHD and personality/​temperament traits of reactive and
effortful control, resiliency, and emotionality. Journal of Child Psychology and Psychiatry, 47(11),
1175–​1183.
Mccracken, J. T., Biederman, J., Greenhill, L. L., Swanson, J. M., Mcgough, J. J., Spencer, T. J., Posner,
K., Wigal, S., Pataki, C., Zhang, Y,. & Tulloch, S. (2003). Analog classroom assessment of a once-​
daily mixed amphetamine formulation, SLI381 (Adderall XR), in children with ADHD. Journal of the
American Academy of Child & Adolescent Psychiatry, 42(6), 673–​683.
Melnick, S. M., & Hinshaw, S. P. (2000). Emotion regulation and parenting in AD/​HD and comparison
boys: Linkages with social behaviors and peer preference. Journal of Abnormal Child Psychology,
28(1), 73–​86.
Mongia, M., & Hechtman, L. (2012). Cognitive behavior therapy for adults with attention-​deficit/​
hyperactivity disorder: a review of recent randomized controlled trials. Current Psychiatry Reports,
14(5), 561–​567.
MTA Cooperative Group. (1999). A 14-​month randomized clinical trial of treatment strategies for
attention-​deficit/​hyperactivity disorder. Archives of General Psychiatry, 56(12), 1073.
Musser, E. D., Galloway-​Long, H. S., Frick, P. J., & Nigg, J. T. (2013). Emotion regulation and heterogeneity
in attention-​deficit/​hyperactivity disorder. Journal of the American Academy of Child & Adolescent
Psychiatry, 52(2), 163–​171.
Musser, E. D., Backs, R. W., Schmitt, C. F., Ablow, J. C., Measelle, J. R., & Nigg, J. T. (2011). Emotion
Regulation via the Autonomic Nervous System in Children with Attention-​Deficit/​Hyperactivity
Disorder (ADHD). Journal of Abnormal Child Psychology, 39(6), 841–​852.
Nigg, J. T. (2001). Is ADHD a disinhibitory disorder?. Psychological bulletin, 127(5), 571.
Ochsner, K. N., & Gross, J. J. (2004). Thinking makes it so: A social cognitive neuroscience approach
to emotion regulation. In Vohs, K. D., & Baumeister, R. F. (Eds.). (2011). Handbook of self-​
regulation: Research, theory, and applications. (pp. 229–​255). New York, NY, US: Guilford Press.
Oosterlaan, J., Logan, G. D., & Sergeant, J. A. (1998). Response inhibition in AD/​HD, CD, comorbid AD/​
HD + CD, anxious, and control children: A meta-​analysis of studies with the stop task. Journal of Child
Psychology and Psychiatry, 39, 411–​425.
Parker, J. D., Majeski, S. A., & Collin, V. T. (2004). ADHD symptoms and personality: Relationships with
the five-​factor model. Personality and Individual Differences, 36(4), 977–​987.
Phillips, M. L., Ladouceur, C. D., & Drevets, W. C. (2008). A neural model of voluntary and automatic
emotion regulation: implications for understanding the pathophysiology and neurodevelopment of
bipolar disorder. Molecular Psychiatry, 13(9), 833–​857.
7
2
1

Conclusions 127

Plessen, K. J., Bansal, R., & Peterson, B. S. (2009). Imaging evidence for anatomical disturbances and
neuroplastic compensation in persons with Tourette syndrome. Journal of Psychosomatic Research,
67(6), 559–​573.
Plichta, M. M., Vasic, N., Wolf, R. C., Lesch, K. P., Brummer, D., Jacob, C., … & Grön, G. (2009). Neural
hyporesponsiveness and hyperresponsiveness during immediate and delayed reward processing in adult
attention-​deficit/​hyperactivity disorder. Biological Psychiatry, 65(1), 7–​14.
Posner, J., Maia, T. V., Fair, D., Peterson, B. S., Sonuga-​Barke, E. J., & Nagel, B. J. (2011). The attenuation
of dysfunctional emotional processing with stimulant medication: an fMRI study of adolescents with
ADHD. Psychiatry Research: Neuroimaging, 193(3), 151–​160.
Posner, M. I., & Rothbart, M. K. (1998). Attention, self–​regulation and consciousness. Philosophical
Transactions of the Royal Society of London B: Biological Sciences, 353(1377), 1915–​1927.
Polanczyk, G., de Lima, M. S., Horta, B. L., Biederman, J., & Rohde, L. A. (2007). The worldwide
prevalence of ADHD: a systematic review and metaregression analysis. The American Journal of
Psychiatry, 164(6), 942–​948.
Polanczyk, G. V., Willcutt, E. G., Salum, G. A., Kieling, C., & Rohde, L. A. (2014). ADHD prevalence
estimates across three decades: an updated systematic review and meta-​regression analysis.
International Journal of Epidemiology, 43(2), 434–​442.
Rich, B. A., Grimley, M. E., Schmajuk, M., Blair, K. S., Blair, R. J. R., & Leibenluft, E. (2008). Face emotion
labeling deficits in children with bipolar disorder and severe mood dysregulation. Development and
Psychopathology, 20(02), 529–​546.
Saunders, B., & Chambers, S. M. (1996). A review of the literature on Attention‐Deficit Hyperactivity
Disorder children: Peer interactions and collaborative learning. Psychology in the Schools, 33(4),
333–​340.
Saarni, C. (1999). The development of emotional competence. New York, NY, US: Guilford Press.
Sanson, A., Smart, D., Prior, M., & Oberklaid, F. (1993). Precursors of hyperactivity and aggression.
Journal of the American Academy of Child & Adolescent Psychiatry, 32(6), 1207–​1216.
Sergeant, J. A., Oosterlaan, J., & van der Meere, J. (1999). Information processing and energetic factors in
attention-​deficit/​hyperactivity disorder. In H. C. Quay & A. E. Hogan (Eds.), Handbook of disruptive
behavior disorders (pp. 75–​104). New York: Kluwer Academic/​Plenum Publishers.
Seymour, K. E., Chronis-​Tuscano, A., Halldorsdottir, T., Stupica, B., Owens, K., & Sacks, T. (2012).
Emotion regulation mediates the relationship between ADHD and depressive symptoms in youth.
Journal of Abnormal Child Psychology, 40(4), 595–​606.
Seymour, K. E., Chronis-​Tuscano, A., Iwamoto, D. K., Kurdziel, G., & MacPherson, L. (2014). Emotion
Regulation Mediates the Association Between ADHD and Depressive Symptoms in a Community
Sample of Youth. Journal of Abnormal Child Psychology, 42(4), 611–​621.
Shaw, P., Stringaris, A., Nigg, J., & Leibenluft, E. (2014). Emotion dysregulation in attention deficit
hyperactivity disorder. American Journal of Psychiatry, 171, 276–​293.
Skirrow, C., McLoughlin, G., Kuntsi, J., & Asherson, P. (2009). Behavioral, neurocognitive and treatment
overlap between attention-​deficit/​hyperactivity disorder and mood instability. Expert review of
neurotherapeutics, 9(4), 489–​503.
Skounti, M., Philalithis, A., & Galanakis, E. (2007). Variations in prevalence of attention deficit
hyperactivity disorder worldwide. European Journal of Pediatrics, 166(2), 117–​123.
Spencer, T. J., Biederman, J., & Mick, E. (2007). Attention-​deficit/​hyperactivity disorder: diagnosis,
lifespan, comorbidities, and neurobiology. Journal of Pediatric Psychology, 32(6), 631–​642.
Stringaris, A., & Goodman, R. (2009). Longitudinal outcome of youth oppositionality: irritable,
headstrong, and hurtful behaviors have distinctive predictions. Journal of the American Academy of
Child & Adolescent Psychiatry, 48(4), 404–​412.
Stringaris, A., Maughan, B., & Goodman, R. (2010). What’s in a disruptive disorder? Temperamental
antecedents of oppositional defiant disorder: findings from the Avon longitudinal study. Journal of the
American Academy of Child & Adolescent Psychiatry, 49(5), 474–​483.
8
2
1

128 Emotion Regulation and Attention Deficit Hyperactivity Disorder

Timimi, S., & Taylor, E. (2003). ADHD is best understood as a cultural construct. The British Journal of
Psychiatry, 184(1), 8–​9.
Travell, C., & Visser, J. (2006). “ADHD does bad stuff to you”: young people’s and parents’ experiences
and perceptions of Attention Deficit Hyperactivity Disorder (ADHD). Emotional and Behavioural
Difficulties, 11(3), 205–​216.
Thompson, M. J., Brooke, X. M., West, C. A., Johnson, H. R., Bumby, E. J., Brodrick, P., … & Scott, N.
(2004). Profiles, co-​morbidity and their relationship to treatment of 191 children with AD/​HD and
their families. European Child & Adolescent Psychiatry, 13(4), 234–​242.
Walcott, C. M., & Landau, S. (2004). The relation between disinhibition and emotion regulation in boys
with attention deficit hyperactivity disorder. Journal of Clinical Child and Adolescent Psychology, 33(4),
772–​782.
Waschbusch, D. A. (2002). A meta-​analytic examination of comorbid hyperactive-​impulsive-​attention
problems and conduct problems. Psychological Bulletin, 128(1), 118.
Wheeler Maedgen, J., & Carlson, C. L. (2000). Social functioning and emotional regulation in the attention
deficit hyperactivity disorder subtypes. Journal of Clinical Psychology, 29(1), 30–​42.
Wilens, T. E., Gignac, M., Swezey, A., Monuteaux, M. C., & Biederman, J. (2006). Characteristics of
adolescents and young adults with ADHD who divert or misuse their prescribed medications. Journal
of the American Academy of Child & Adolescent Psychiatry, 45(4), 408–​414.
Willcutt, E. G., Doyle, A. E., Nigg, J. T., Faraone, S. V., & Pennington, B. F. (2005). Validity of the executive
function theory of attention-​deficit/​hyperactivity disorder: a meta-​analytic review. Biological Psychiatry,
57(11), 1336–​1346.
Willcutt, E. G. (2012). The Prevalence of DSM-​IV Attention-​Deficit/​Hyperactivity Disorder: A Meta-​
Analytic Review. Neurotherapeutics, 9(3), 490–​499.
Wolraich, M. L., Wibbelsman, C. J., Brown, T. E., Evans, S. W., Gotlieb, E. M., Knight, J. R., … & Wilens,
T. (2005). Attention-​deficit/​hyperactivity disorder among adolescents: a review of the diagnosis,
treatment, and clinical implications. Pediatrics, 115(6), 1734–​1746.
Zelazo, P. D., & Cunningham, W. A. (2007). Executive Function: Mechanisms Underlying Emotion
Regulation. In James J. Gross (Ed.), Handbook of emotion regulation (pp. 135−158). New York, NY,
US: Guilford Press.
9
2
1

Chapter 7

Emotion Regulation
and Conduct Disorder: The Role
of Callous-╉Unemotional Traits
Nicholas D. Thomson, Luna C. M. Centifanti,
& Elizabeth A. Lemerise

Conduct disorder
Although all children disobey adults at times, children with conduct disorder (CD) persistently
break the rules, engage in norm-╉breaking behavior, defy adults and authority figures across situa-
tions, and repeatedly and seriously violate the rights of others (American Psychological Association
[APA], 2013). CD was first introduced as a psychiatric diagnosis in the second edition of the
American Psychiatric Association’s Diagnostic & Statistical Manual of Mental Disorders (DSM).
Since this time the diagnosis of CD has become more refined (Kimonis, Frick, & McMahon, 2014).
There are four types of symptoms that define CD: 1) aggression towards people and animals (e.g.,
fighting, bullying); 2)  destruction of property (e.g., fire setting, vandalism); 3)  deceitfulness, or
theft (e.g., conning, shoplifting); and 4)  serious violations of rules (e.g., truancy, running away
from home [APA, 2013]). CD is one of the most prevalent mental health concerns for children and
adolescents and is considered one of the most challenging childhood disorders to treat (Dadds &
Fraser, 2003). To further complicate matters, children with CD are often viewed as “bad” rather
than having a mental illness because their symptoms result in the violation of the rights of oth-
ers (e.g., hostility, aggression, cruelty). Further, conduct problems represent a large cost to society
(Welsh et al., 2008). Although CD is considered a behavioral disorder, differences in emotion (dys)
regulation might identify subgroups of youth with CD. Some children with CD may exhibit irrita-
bility and mood swings resulting in aggressive responding; whereas, other children with CD may be
emotionally disconnected from others so they callously hurt others. This emotional heterogeneity
may explain why some children with CD fail to be bothered by the effects of their behavior on other
people, whereas others experience anxiety over their negative behavior (Pardini & Frick, 2013).
In this chapter, we will discuss the evidence for considering how children with CD manage their
emotions because subgroups of children with CD may show different developmental trajectories
based on having strong or poor emotion regulation abilities. We will also discuss the implications
for clinical practice in managing CD based on this heterogeneity.
Based on the severity and number of symptoms displayed, CD can be classified from “mild,”
such that the youth displays few symptoms and/╉or causes minor harm (e.g., lying, truancy), to
“severe,” such that the youth displays many more symptoms than required for a diagnosis and
considerable harm to others is caused (e.g., forced sex, use of a weapon). Severity of CD has
been found to affect the persistence of the disorder, with youths in the moderate to severe scale
of CD more likely to retain CD symptoms into their adolescence (Cohen, Cohen, & Brook,
1993) and suffers from educational problems (Kim-╉Cohen et al., 2005). Although the number of
0
3
1

130 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

children who display early and pervasive antisocial behavior is small in number (5% [Hinshaw
& Lee, 2003]), they account for almost half the crime in the United States (Loeber, Burke, Lahey,
Winters, & Zera, 2000).
In the DSM-╉5 (APA, 2013), heterogeneity in CD diagnosis is recognized, such that persis-
tence of antisocial behavior beyond childhood is characteristic of a subgroup of those with CD.
Currently, the diagnosis takes into account the age at which the symptoms onset, because early
onset of behavioral problems typically relates to lifetime-╉persistence of these behaviors (Moffitt,
1993). Childhood-╉onset, which is defined as onset before the age of ten years, has been associated
with greater cognitive impairment, mental health concerns, and more harmful, violent behavior
than adolescent-╉onset CD (onset after the age of ten years) (Johnson, Kemp, Heard, Lennings,
& Hickie, 2015). Heterogeneity in CD is in line with the dual taxonomy of offending posited by
Moffitt (1993). Moffitt theorized that those who have an onset of offending during adolescence
(adolescent-╉limited offenders) tend to cease their delinquency by early adulthood, whereas those
with childhood-╉onset (lifecourse-╉persistent offenders) continue their antisocial behavior into
adulthood (Moffitt, 1993).
A test of Moffitt’s (1993) taxonomic predictions revealed four antisocial behavior trajectories
roughly in line with Moffitt: 1) lifecourse-╉persistent, 2) adolescence-╉limited, 3) childhood-╉limited,
4) and low (Odgers et al., 2008). There was empirical evidence for the developmental trajectories
of antisocial behavior which coincided most with the designations of CD: lifecourse-╉persistent
and adolescent-╉onset antisocial trajectories coincided with childhood-╉and adolescent-╉onset
CD. However, the outcomes related to the trajectories differed for the adolescent-╉onset group
(Odgers et al., 2008). Some individuals in the adolescent-╉onset trajectory continued to show anti-
social behavior into adulthood. Also, a trajectory not originally posited by Moffitt was identi-
fied: A childhood-╉limited trajectory. People on this trajectory of antisocial behavior desisted past
childhood. However, of importance to psychopathology, they only showed minor problems with
smoking, managing finances, and internalizing behavior problems (e.g., anxiety and depression).
Thus, there are some who present with early conduct problems but who grow out of them, only
seeming to be left with the remains of their poor behavior management choices. Thus, it could be
that other factors like emotion or behavior management could be useful in delineating heteroge-
neity within CD.
Yet, in line with childhood-╉onset CD diagnoses, research finds that children with early-╉onset of
CD typically have a prior diagnosis of Oppositional Defiant Disorder (ODD). ODD is considered
a precursor to and milder variant of CD (Loney & Lima, 2003). Longitudinal samples have shown
that 80% (Loeber, Green, Keenan, & Lahey, 1995) of children with CD had a former diagnosis of
ODD, and about 90% of clinically referred children with CD diagnosis meet the criteria for ODD
(Faraone, Biederman, Keenan, & Tsuang, 1991). CD and ODD can co-╉occur, and a comorbid
diagnosis of CD with ODD can be given (APA, 2013). In research, the term “conduct problems”
(CP) is often used to jointly describe children with severe behavioral problems, or a diagnosis of
CD or ODD (Kimonis, Frick et al., 2014).

Prevalence and course
Conduct disorder is one of the most prevalent disorders for children and adolescents (Kessler
et  al., 2012; Lindhiem, Bennett, Hipwell, & Pardini, 2015), but the negative impact is not
limited to these early years and is associated with lifelong adjustment, mental health, legal,
social, occupational, and physical health problems (Jones, 2013; Odgers et  al., 2008). Based
on study samples, the prevalence of CD is estimated to be between 2% and 15% (APA, 2013;
Egger & Angold; Kim-╉Cohen et al., 2005) with more cases evident of adolescent-╉onset than
1
3

Prevalence and course 131

childhood-​onset (Nock, Kazdin, Hiripi, & Kessler, 2006; Perou et al., 2013). Overall, boys are
twice as likely as girls to receive a CD diagnosis (4.6% versus 2.2% with current CD diagno-
sis [Perou et  al.,  2013]). However, gender differences are greater early in childhood. At the
age of five years, boys are three to five times more likely to be diagnosed with CD than girls
(Kim-​C ohen et al., 2005). Emerging into mid-​adolescence, gender differences tend to reduce
significantly with female CD prevalence peaking at the age of 16  years (Esser, Schmidt, &
Woerner, 1990; McGee et  al., 1990). For adolescent girls, CD is the second most common
psychiatric diagnosis with a prevalence rate of about 10% in community samples (Dalwani
et al., 2015; Pajer et al., 2008) and 36% in detention center samples (Washburn et al., 2007).
Although research on CD tends to focus on male samples, there is evidence supporting simi-
larities between males and females in biological (Fairchild et al., 2014) and psychosocial vul-
nerabilities (Bardone et al., 1998; Pajer, 1998).
Moffit’s (1993, 2006) dual taxonomy of conduct problems identifies two groups of youth based
on the timing of onset of behavioral problems. The developmental typology suggests that con-
duct problems developing in early childhood lead to “life-​course-​persistent” antisocial behavior,
whereas antisocial behavior that begins in adolescence is limited to the teenage years (Moffitt &
Caspi, 2001). Prior research suggests that children who develop CD during childhood differ in
the underlying mechanisms compared to youths who develop CD during adolescence. Evidence
suggests that children with childhood-​onset are exposed to family, social, and inherited neurode-
velopmental risk factors more than youths with adolescent-​onset of CD (Moffitt & Caspi, 2001;
Odgers et al., 2008). By the age of 32, adults with a history of childhood-​onset of CD show greater
perpetration of violence, and more mental and physical health problems (Odgers et al., 2008). The
different trajectories based on age of onset are attributed to causal mechanisms in the child’s envi-
ronment as well as biological factors that seem to distinguish the two groups. Childhood-​onset
has been associated with poorer neurological functioning (e.g., self-​control, memory and verbal
abilities), which in turn, negatively impacts the successful navigation of social relationships, man-
agement of emotions, and the ability to control behaviors (Johnson et  al., 2015; Moffitt, 2006;
Pardini & Frick, 2013). The child is more likely to experience childhood maltreatment (Johnson
et  al., 2015), poorer parenting strategies (i.e., harsh and inconsistent discipline), and greater
family-​level conflict, poverty, mental health problems (Odgers et al., 2008) and parental history
of antisocial behavior (McCabe, Hough, Wood, & Yeh, 2001). Whereas, youths with adolescent-​
onset CD are less likely to have a childhood history of ADHD or ODD, have neurological deficits,
and have less severe family dysfunction and aggression and a greater remission rate of antisocial
behavior into adulthood (Moffitt, Caspi, Dickson, Silva, & Stanton, 1996). Prior research supports
the dual taxonomic trajectory of antisocial behavior; thus, the DSM-​5 categorization of age of
onset is an important factor. However, there is considerable evidence showing etiological hetero-
geneity within the childhood-​onset group based on emotionality (Frick & Viding, 2009; Pardini
& Frick, 2013).
Further heterogeneity in CD has recently been identified based on callous-​unemotional (CU)
traits. Research has identified a subgroup of youth with CD and callous unemotional (CU)
traits, although the term used in the DSM is Limited Prosocial Emotions (LPE; APA, 2013). To
meet diagnostic criteria for LPE, the youth must display two of the following four character-
istics: a lack of remorse or guilt, a callous lack of empathy, shallow or deficient affect, or lack
of concern about performance (Blair, Leibenluft, & Pine, 2014). Thus, children with LPE are
emotionally cold and experience little concern over the effects that their problem behaviors
may cause (see Munoz & Frick, 2012). Children with CU traits have a lack of concern for the
welfare of others; they often act cruelly to others with the intention to cause physical or emo-
tional harm in order to achieve a goal (e.g., exerting dominance [Pardini & Byrd, 2012]). This
2
3
1

132 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

group of children has a lack of emotionality (Essau, Sasagawa, & Frick, 2006) making them
fearless perpetrators of antisocial behavior without consideration of the consequences of their
actions (Fanti, Panayiotou, Lazarou, Michael, & Georgiou, 2015). Emotional deficits such as low
empathy and guilt are suggested to play an integral role in the atypical development of moral
values for children with CU traits (Frick, Ray, Thornton, & Kahn, 2014). Although children
with conduct problems continue to show behavioral problems throughout childhood, children
with conduct problems (CP) and CU traits continue to show the greatest levels of conduct prob-
lems, delinquency, and police contacts (Frick, Stickle, Dandreaux, Farrell, & Kimonis, 2005).
Compared to children with CD-╉only, children with CD + CU have been described as having
different etiological mechanisms (i.e., social, genetic, behavioral, and cognitive vulnerabilities
[Kimonis, Centifanti, Allen, & Frick, 2014; Sebastian et al., 2015]), and are hypoactive in their
emotional responses (Frick & Viding, 2009; Sebastian et al., 2015). Evidence from twin stud-
ies suggests that children with CU traits and conduct problems are more likely to have inher-
ited contributing factors, whereas conduct problems in children with low levels of CU traits
are explained mostly by environmental vulnerabilities (Viding, Blair, Moffitt, & Plomin, 2005;
Viding, Jones, Frick, Moffitt, & Plomin, 2008).
Therefore, prior research with community samples and clinic-╉referred samples has found that
children and adolescents with CD and LPE (CD + LPE) are characteristically different from youth
with only CD (CD-╉only) in terms of long-╉term outcomes. When compared to youth with CD-╉
only, youth with CD + LPE are more likely to display severe and persistent psychopathology (Rowe
et al., 2010), get involved in criminal activities at a younger age (Pechorro, Jiménez, Hidalgo, &
Nunes, 2015), have greater levels of externalizing behaviors (e.g., aggression, delinquency, psy-
chopathic traits [Colins & Andershed, 2015]), and are more likely to develop antisocial personal-
ity disorder symptoms in adulthood (McMahon, Witkiewitz, & Kotler, 2010).

Development of emotion regulation


Emotion is a complex and subjective phenomenon, which is composed of physiological arousal
and expressive behavior (Calkins, 1994; Thompson & Calkins, 1996). Functionalists believe the
purpose of emotions in infants is to serve as effective communication to caregivers (see Lemerise
& Arsenio, 2000). Emotions regulate behavior and communicate the infant’s immediate needs
and internal states to caregivers (Marshall, Fox, & Henderson, 2000). The caregiver’s consistent
and sensitive responses to their infant’s emotions greatly influences the development of attach-
ment security and contributes to the infant’s developing ability to regulate emotion (Lemerise
& Dodge, 2008; Lemerise & Harper, 2014). Emotional competence is multifaceted and includes
emotion awareness, identification, and understanding in self and others, as well as regulation of
emotion/╉arousal in service of adaptive coping during an emotionally arousing event (Bohnert,
Crnic, & Lim, 2003; Halberstadt, Denham, & Dunsmore, 2001; Lemerise & Harper, 2010, 2014;
Saarni, 1999).
Although biologically based temperament plays a role in conduct problems, emotional com-
petence is important and develops in a relational context, first with parents and later with peers.
From the earliest days of infancy, mothers respond differentially to infants’ emotions, with the
net result that children’s positive and neutral expressions increase, whereas negative expres-
sions decrease (e.g., Malatesta, Culver, Tesman, & Shepard, 1989). Mothers continue to social-
ize emotions indirectly, and as children develop language, more direct socialization methods are
employed (Lemerise & Dodge, 2008; Lemerise & Harper, 2014).
Socialization practices that act to modulate children’s emotional arousal scaffold children’s
learning about emotions and how to regulate them. In particular, parents who employ the
31

Temperament and emotionality 133

practice of correctly labeling children’s emotions and coach their children on coping strategies
and problem-╉solving have children who display better emotional competence (emotion aware-
ness/╉understanding and emotion regulation [Gottman, Katz, & Hooven, 1997; Laible & Panfile,
2009; Thompson, 2006]). For example, mothers who remark appropriately about their child’s
mental states appear to “scaffold” a richness in children’s understanding of emotions (Centifanti,
Meins, & Fernyhough, 2015). Further, mothers’ appropriate verbal comments about their infant’s
mental states and desires led to lower levels of CU traits at age ten years through a greater emotion
understanding at age four years (Centifanti, et al., 2016). However, caregivers’ hostile responses
to their children’s emotions, including anger, tend to increase children’s arousal, interfering with
learning about emotions and with the regulation of emotions, raising risks for children’s aggres-
sive and other problem behaviors (Lemerise & Dodge, 2008). Moreover, the stresses associated
with poverty (a well-╉known risk factor for aggressive behavior and CD) interfere with children’s
regulatory development as well as with the supportive parenting that might buffer children from
these stressors (Blair & Raver, 2015).
With language development, children soon learn that they can communicate their needs more
effectively using their words. However, Thompson (1994) theorized that emotions continue to
serve in social signaling, defensive motivations, and in communication of one’s needs, but also
serve to maintain affiliational ties. Indeed, nonhuman primates rely on expressions of threat to
inhibit agonistic behavior from other primates (see Izard, 1991). Thus, emotions can communi-
cate in addition to other forms of communication. Human infants respond to pain with crying,
seemingly to summon their caregiver’s attention, as infants lack the ability to defend themselves.
By 19  months, however, children respond to pain with anger (after a short period of crying),
which might serve to inhibit any perpetrator. This developmental change in their emotional
expression may, arguably, be due to many other factors related to the regulation of emotion and
the emerging theory of mind (which both depend on cognitive processes [Lemerise & Dodge,
1993; Meltzoff, 2002]).

Temperament and emotionality


Emotional reactivity or emotionality, which has been variously defined as reflecting onset, dura-
tion, and intensity of emotion, is thought to be central to theories of temperament (Eisenberg
et  al., 1997; Rothbart & Bates, 1998)  and figure prominently in emotion regulation, such that
greater intensities of emotion are harder to regulate. Rothbart and Bates (1998) include attention,
activity, variability in arousability, and distress to overstimulation in theories of temperament,
and temperamental characteristics are theorized to be consistent and stable across situations, and
result due to biological dispositions. Compatible with the notion of multifinality, temperamental
characteristics are modifiable with maturation and experience (Marshall et  al., 2000; Rothbart
& Bates, 1998). Thus, different developmental trajectories are possible given similar biological
diatheses. Of importance, different temperamental dispositions may underlie heterogeneity in
childhood disorders.
Optimal development of emotion regulation depends on flexible control of attention and the
ability to shift attention (Rothbart, Posner, & Hershey, 1995). A  high degree of self-╉regulation
is essential when emotions, particularly negative emotions, are frequently experienced intensely
(Eisenberg et al., 1997). That is, negative emotions that are very intensely experienced are more
difficult to manage. Although emotionality and regulation may have additive effects on social
competence (Blair et  al., 2015; Dollar & Stifter, 2012; Rothbart & Bates, 1998), negative emo-
tionality, alone, seems to distinguish those children with behavior problems from those without
(Eisenberg et al., 1997; Nozadi, Spinrad, Eisenberg, & Eggum-╉Wilkens, 2015).
4
3
1

134 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Conduct disorder with callous unemotional traits


Children in the CP + CU group have been characterized as temperamentally fearless with
diminished emotionality, which is suggested to explain their propensity toward lifelong antiso-
cial behavior (Fanti, Panayiotou, Lazarou, et al., 2015; Viding et al., 2012). The lack of emotional
response is thought to explain why children with CU traits do not learn from others’ distress
or, indeed, by cues for punishment (Frick & Viding, 2009; Pardini & Frick, 2013). Physiological
studies have found that children with CU traits have reduced sympathetic reactivity (Muñoz,
Frick, Kimonis, & Aucoin, 2008) and are less responsive to others’ emotional and physical distress
(Fanti, Panayiotou, Kyranides, et al., 2015; Wolf & Centifanti, 2014). Specifically, children with
CP + CU show lower startle potentiation to fearful mental imagery compared to children in the
CP-╉only group (Fanti, Panayiotou, Lazarou, et al., 2015). Neuroimaging studies have supported
this lack of emotionality to others’ distress, finding that children with higher levels of CU traits
have less activation in the amygdala in response to fearful faces (Viding et al., 2012). Additionally,
children with CP + CU have been found to have reduced grey matter volume in the left orbital
frontal cortex and the right anterior cingulate cortex, which are key brain regions for decision-╉
making and empathy (Sebastian et al., 2015). Emerging evidence suggests a neurological continu-
ity from childhood CU to adult psychopathy, with a reduction in white matter in regions of the
limbic system notably occurring from childhood. Atypical neurological development may explain
the hypoactivity to others’ distress, which impairs important brain regions for social and affective
functioning (Breeden, Cardinale, Lozier, VanMeter, & Marsh, 2015; Hoppenbrouwers et al., 2013;
Wolf et al., 2015). These biological influences suggest that a lack of emotional and physiological
reactivity to fearful events could explain why children with CU traits are less receptive to learning
as a result of punitive measures, hindering normative social development, which subsequently
predisposes these children as life-╉course-╉persistent offenders.

Conduct disorder with severe anger dysregulation


Recent evidence has emerged which suggests that the combination of CD with severe anger dys-
regulation (e.g., oppositional and defiant behaviors) has meaningful implications for the diagno-
sis of childhood-╉onset of CD, the pathway into adulthood outcomes, and treatment (Pardini &
Frick, 2013). Although ODD is a common precursor to CD (Stringaris, 2011), and both disorders
share theoretical and empirical overlap, the two are symptomatically distinct (Krieger et al., 2013;
Whelan, Stringaris, Maughan, & Barker, 2013). The core feature of CD is the purposeful viola-
tion of the rights of others and/╉or breaking of major social norms, whereas the nucleus of ODD
is negative emotionality, such as irritability and anger (Lindhiem et al., 2015). It is common for
these two disorders and symptoms to co-╉occur (Copeland, Angold, Costello, & Egger, 2013); how-
ever, note that although most children with ODD receive a diagnosis of CD, the reverse is not so
(Lindhiem et al., 2015). Of importance, children with ODD tend to develop anxiety and depres-
sion later in life (Boylan, Vaillancourt, Boyle, & Szatmari, 2007; Pardini & Fite, 2010), whereas
children with CD who had not received an ODD diagnosis tend to face a trajectory toward antiso-
cial behavior, criminality, and the development of psychopathic personality traits (Burke, Loeber,
& Lahey; Byrd, Loeber, & Pardini, 2012). Unlike the CD + CU group, children with CD and severe
anger dysregulation tend to be hypersensitive to threat and hyper-╉reactive to fear (Pardini & Frick,
2013), which may result in reactive and explosive forms of aggression (Okado & Bierman, 2015).
Although research shows that biological factors contribute to conduct problems for children with
severe emotional dysregulation, evidence suggests that these biological factors are influenced by
exposure to a negative childhood environment (Pardini & Frick, 2013).
5
3
1

Childhood aggression 135

An example of the biological and environmental interplay can be seen in longitudinal studies.
The parasympathetic nervous system facilitates a reduction in heart rate and increases respira-
tory sinus arrhythmia (RSA). When operating effectively, this helps facilitate emotion regula-
tion (Beauchaine, 2015; Hinnant, Erath, & El-╉Sheikh, 2015). Lower resting RSA and less RSA
withdrawal (during a threatening or challenging event) has been associated with poor emotion
regulation, executive control, adjustment problems, and greater levels of parent-╉child aggression
(Beauchaine, 2015; Whitson & El-╉Sheikh, 2003). In a recent study, children whose parents used
harsher parenting strategies had reductions over time in their RSA withdrawal to stress, suggest-
ing that children in a hostile home environment are more likely to suffer from long-╉term physi-
ological alterations which affect their ability to regulate their emotions (Hinnant et al., 2015). In
support of the emotion dysregulation subgroup, CD-╉only children have demonstrated a hyper-
sensitivity to fear, and poor behavioral inhibition (Fanti, Panayiotou, Kyranides, & Avraamides,
2015). Children with poor emotion regulation tend to have elevated levels of hyper-╉vigilance to
threat cues and attributions of hostile intent. Children with higher levels of hostile attributional
bias misinterpret ambiguous social cues as hostile intent which results in the child responding
reactively (Dodge et  al., 2015). To further exacerbate matters, engaging in aggressive behavior
inevitably places the child in hostile social situations, which will likely increase the child’s ten-
dency to attribute hostile intent from peers (Dodge et al., 2015). Children with CD-╉only indeed
show increased heightened emotional reactivity. Because of the emotional instability, a tendency
to misinterpret benign intents as hostile intents, and hypersensitivity to fear, this subgroup of chil-
dren uses reactive (i.e., in response to provocation) aggression as a result of poor emotion regula-
tion (de Wied, van Boxtel, Matthys, & Meeus, 2012; Frick, Cornell, Barry, Bodin, & Dane, 2003).
There are some factors that might distinguish CD-╉only groups from CD + CU groups, making
the former more amenable to intervention. Children with CD-╉only are more receptive to pun-
ishment (Fanti, Panayiotou, Lazarou, et al., 2015) and affectively empathetic and sympathetic to
others (de Wied et al., 2012; Frick et al., 2003; Frick & Morris, 2004). In contrast, CD + CU chil-
dren are punishment insensitive, lack emotionality, and consider deviant strategies (e.g., revenge,
blaming others, aggression [Pardini, 2011; Stickle, Kirkpatrick, & Brush,  2009]) as acceptable
methods to achieve a goal (Frick et al., 2014). The juxtaposition of CD + CU and CD-╉only in
childhood-╉onset illustrates distinguishing features that could affect treatment outcome, hence the
importance of distinguishing childhood-╉onset subgroups.
Recent research supports the principle of equifinality, whereby different developmental mecha-
nisms (e.g., hyper and hyposensitivity to fear) may lead to the same outcome of antisocial behav-
ior (Fanti, Panayiotou, Kyranides, et  al., 2015). However, the way in which a child perpetrates
antisocial behavior may be indicative of the developmental pathway he\she has taken. As with
the two subtypes of childhood-╉onset of CD, emotionality and emotion dysregulation play integral
roles in how aggressive behavior in children is understood.

Childhood aggression
CU traits have been suggested to moderate antisocial behavior for youth with conduct problems
(Helseth, Waschbusch, King, & Willoughby, 2015), including aggression subtypes that differ in
emotionality and emotion regulation. Proactive aggression occurs without provocation and is
typically motivated by intentional purpose (e.g., social dominance, physical goal). By comparison,
reactive aggression occurs in response to a perceived provocation or threat (Dodge & Coie, 1987).
Proactive aggression is characterized as cold-╉blooded, whereas reactive aggression is fueled by
anger or frustration (Dodge, 1991; Teten Tharp et al., 2011). Theoretically, children with CU traits
are more likely to be proactively aggressive, yet, empirically, children with CU traits tend to show
6
3
1

136 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

high levels of both reactive and proactive aggression (Centifanti, Fanti, Thomson, Demetriou, &
Anastassiou-╉Hadjicharalambous, 2015; Muñoz et  al., 2008). For instance, detained adolescents
(13–╉18  years) who reported being high on both reactive and proactive aggression (forming a
“mixed” aggressor group) had higher levels of CU traits, and lower levels of physiological reactivity
when provoked (Muñoz et al., 2008). The mixed group were also more aggressive in a behavioral
task when they experienced no provocation from an opponent (Muñoz et al., 2008). Therefore,
youth with CU traits may respond aggressively in all provocation situations (without provocation,
or in response to low or high provocation), but they seem to have less of an emotional response.

Development of reactive aggression


Reactive aggression is characterized by “hot blooded” anger, driven by hostile responses to minor
or perceived provocation, and intense physiological reactivity (Dodge & Coie, 1987; Dodge,
Lochman, Harnish, Bates, & Pettit, 1997; Hubbard et al., 2002; Zhang & Gao, 2015). One of the
hallmark differences between reactive and proactive aggression (Pardini, Raine, Erickson, &
Loeber, 2014) is a differential association with a deficiency in processing information related to
social threat (Crick & Dodge, 1996). In fact, reactive aggression, as opposed to proactive aggres-
sion, has been related to social information-╉processing deficits and biases at many, if not all,
levels of decision-╉making, but especially during early stage processing (encoding of social cues
and interpretation of social cues [Arsenio, Adams, & Gold, 2009; Dodge et al., 1997; Dodge &
Pettit, 2003; Lemerise & Arsenio, 2000]). Children who engage in mainly reactive aggression have
been shown to attribute greater hostile intent to peer behavior (Dodge & Coie, 1987), which may
cause them to respond aggressively. Indeed, studies have supported this assertion: higher levels
of reactive aggression were associated with greater hostile attributional biases (Hubbard, Dodge,
Cillessen, Coie, & Schwartz, 2001) and with being aggressive in response to low levels of provoca-
tion (Muñoz et al., 2008).
Although Dodge’s theory focuses mainly on social information-╉processing, emotion and
emotion regulation processes also are important to both the development and the expression
of reactive aggression (Lemerise & Arsenio, 2000). For example, de Castro et  al. (2005) found
that reactive aggressive children often reported that they did not know a strategy to regulate
their strong emotions or that another person(s) would have to regulate the emotions for them.
In addition, there is evidence that strong emotions overwhelm reactively aggressive children’s
social information-╉processing, leading to impulsive reactions to threat (Crick & Dodge, 1994; de
Castro, Verhulp, & Runions, 2012; Lemerise & Arsenio, 2000). Vitaro, Brendgen, and Tremblay
(2002) found that only reactive aggression was related to high reactivity, inattention, anxiety, and
depression. A child who is high in negative emotionality and lacks the ability to regulate his or her
emotions would be highly susceptible to evoking negative responses from his or her environment
(Schwartz et al., 1998). Deficits in emotion regulation may stem from an inability to focus or shift
attention (Rothbart et al., 1995) or equally, may stem from a lack of inhibition and careful plan-
ning due to executive functioning deficits (Bridgett, Oddi, Laake, Murdock, & Bachmann, 2013).
Indeed, reactive aggression has been repeatedly related to peer social rejection (Dodge et al., 1997;
Evans, Fite, Hendrickson, Rubens, & Mages, 2015; Schwartz et al., 1998). Moreover, having high
negative emotionality or being easily angered might predispose a child to “cue up” (internally)
past negative situations that resulted in hostility, but which may not necessarily be related to the
child’s current situation (Lemerise & Arsenio, 2000). An aggressive response would thus become
more likely.
Children who show high levels of reactive aggression also show selective attention to social
threat words or cues (e.g., “teased,” “rejected,” “failure,” and “unpopular” [Schippell, Vasey,
7
3
1

Development of proactive aggression 137

Cravens-╉Brown, & Bretveld, 2003]). Aspinwall (1998) reviewed evidence that people in a nega-
tive mood orient more quickly to negative information, but their resources are tied up in regu-
lating their negative emotions; thus, they are unable to fully process the negative information.
Alternatively, attentional systems may become dysregulated by intense emotional arousal, pre-
cluding an adequate processing of relevant cues (Thompson & Calkins, 1996). Aspinwall’s (1998)
conclusion may explain why Schippell et al. (2003) found that reactive aggressors showed selective
attention and suppression of social threat words. A  large stress reaction results in an internal-╉
focus (Thompson & Calkins, 1996), diverting attention away from the stimulus and precluding
adequate processing. In the reactive-╉aggressive child, this diversion of attention may serve to pre-
vent further emotional negativity that typically threatens their self-╉image (Schippell et al., 2003).
In fact, aggressive children have been shown to overestimate their likeability with their peers
(Rudolph & Clark, 2001). Hypervigilance to threat cues may serve to aid in their suppression and
in the protection of esteem or may simply reflect the selection of mood-╉congruent information
(Lemerise & Arsenio, 2000; Schippell et al., 2003).
Cognitive competence is essential for successful development of behavioral regulation (Olson,
Bates, Sandy, & Schilling, 2002) or inhibitory control. A lower verbal intelligence quotient may
lead to a generation of fewer alternatives when generating responses to a situation and a rapid
accessing of aggressive responses, which could lead to aggressive behavior (Dodge & Pettit, 2003;
Lemerise & Arsenio, 2000). Children who are more impulsively aggressive evidence intellec-
tual deficits, particularly verbal deficits (Arsenio et  al., 2009; Babcock, Tharp, Sharp, Heppner,
& Stanford, 2014; Loney, Frick, Ellis, & McCoy, 1998). One study found that children classified
as rejected-╉reactive aggressive were better able to choose a constructive response, when given
response options rather than free choice (Wood & Gross, 2002). The generation and subsequent
selection of an aggressive response may result from deficits in inhibitory control and/╉or planning.
Emotional competence involves being able to control one’s expressivity as well as to express
emotions flexibly and appropriately within the situation (see Lemerise & Arsenio, 2000). In addi-
tion, competence is shown by being able to sensitively respond to others’ emotional cues and
behavioral cues, which are important in providing feedback for the child’s behavior. Of impor-
tance, emotional cues are displayed by the child and by others as the encounter proceeds, which
allow the child to adjust his or her response in-╉line with the current environmental demands
(Lemerise & Arsenio, 2000). During situations of high arousal, this delicate interchange may dis-
integrate. Emotional cues may be missed or misinterpreted, whereby a peer’s positive affective
desire to share a toy may be misconstrued as an angry demand. The result is possibly an angry
reaction or resistance on the part of the reactive-╉aggressive child.
The outcome of these perceptual and interpretive processes is an emotional and behavioral
response. Differences in the expression of emotion have been found to distinguish the two sub-
types of aggression (Hubbard et al., 2002). During a competitive game, a dysregulation of angry
or hostile emotions seemed to characterize reactive but not proactive aggression (Hubbard et al.,
2002). Strong emotional reactions to stressful situations can impede attempts to regulate behav-
ior as well as cognitive attempts to regulate emotion (Lemerise & Arsenio, 2000; Thompson &
Calkins, 1996). All of the deficits discussed thus far increase the likelihood of aggressive behavior
(Crick & Dodge, 1994).

Development of proactive aggression


Proactive aggression, unlike reactive aggression, is not typically associated with verbal defi-
cits or early stage social information-╉processing deficits or biases (cue encoding or attribu-
tions). Proactive aggression, instead, is motivated by rewards or perceived gains, thus the term
8
3
1

138 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

“instrumental” is used in describing this type of aggression (Dodge et  al., 1997; Raine, Fung,
Portnoy, Choy, & Spring, 2014). Proactive aggressors have been shown to prefer instrumental and
dominance goals over relational goals (de Castro et al., 2005; e.g., Salmivalli, Ojanen, Haanpää,
& Peets, 2005), and to evaluate aggressive responses more positively in terms of their effective-
ness and one’s emotional reactions and self-╉efficacy (Arsenio et  al.; Dodge et  al., 1997). Their
preference for instrumental goals over relational goals biases response selection toward aggressive
responses (see Harper, Lemerise, & Caverly, 2010).
Those who use proactive aggression tend to show higher levels of aggression and blunted emo-
tion or emotion that is inconsistent with their behavioral displays (Bobadilla, Wampler, & Taylor,
2012; Hubbard et al., 2002). This is similar to the low emotionality related to CU traits. As sug-
gested above, those who experience distress, such as anxiety and fear, may be more easily social-
ized (Eisenberg et al., 1997; Izard, 1991). Children who were emotionally reactive and more prone
to negative emotion were also high in conscience development (Kochanska, 1991). High reactivity
to transgressions may facilitate the affective and affiliative component of conscience development,
but it also may hinder the enactment of guilt-╉related behavior (such as reparation attempts and
confession [Kochanska et  al.,  1994]). Nonetheless, this pattern of responding was found to be
more characteristic of girls than of boys (Kochanska et al., 1994). Eisenberg et al. (1997) found
that, although both boys’ and girls’ negative emotionality positively relates to behavior problems,
only boys’ anxiety (expressed in the laboratory) was negatively related to behavior problems.
Thus, boys’ low anxiety might be more germane to the development of a cold, unfeeling type of
aggression.
In sum, reactive aggressive children appear to suffer more from information-╉processing errors,
which result in hostile-╉attributional bias, making their behavior amenable to interventions that
focus on regulating their emotional arousal. For example, the Coping Power Program (Lochman,
1992; Lochman & Wells, 2004) specifically focuses on helping the aggressive child to deal with
his/╉her intense anger arising from provocation. It also targets biases that often result in reactive
aggression, such as viewing others’ actions as originating from hostile intentions. Using cognitive-╉
behavioral techniques, these programs target faulty information processing deficits that can lead
to reactive aggression (Boxer & Frick, 2008). The verbal deficits, emotional reactivity, and impul-
sivity that accompany reactive aggression may underlie information processing errors, as well
as also underlying their unsuccessful attempts to regulate their behavior. Admittedly, proactive
aggressive children also have social-╉cognitive biases in prejudicially perceiving positive outcomes
for behaving badly. However, their emotional deficits predispose them to fail to develop complex
cognitive and emotional processes, such as understanding and identifying emotions in others
(specifically, negative emotions) and responding sensitively to them. Complex emotions such as
guilt may fail to develop as a result. Thus, they require different types of treatments, such as inter-
ventions that target the multiple contexts in which children function.

Treatment
Youths who engage in serious antisocial behavior are more likely to experience psychosocial, aca-
demic, and occupational challenges (Frick & Dickens, 2006). The negative influence is not exclu-
sive to the youths’ wellbeing but extends to the community, with extensive economic and social
burden as well as the physical and emotional cost to victims. Therefore, the impact that interven-
tions have is not isolated to the individual or the family involved, but they positively affect the
greater community and society. However, high-╉risk individuals and families (with severe exter-
nalizing behaviors) are considered extremely difficult to treat due to the strength and stability of
personality characteristics (Moffitt, 1993) and exacerbation by the (cross)generational reach of
9
3
1

MULTISYSTEMIC THERAPY AND CONDUCT DISORDER 139

antisocial behavior (Hawkins, Catalano, Kosterman, Abbott, & Hill, 1999). Although there are
a variety of interventions that have demonstrated effectiveness in children and adolescents with
severe antisocial behavior, the remainder of this chapter will focus on three empirically efficacious
interventions designed for youth with conduct problems:  Multisystemic Therapy, Functional
Family Therapy, and The Incredible Years.

Multisystemic therapy
Multisystemic Therapy (MST; Henggeler, Schoenwald, Borduin, Rowland, & Cunningham,
2009)  is a family-╉focused home-╉and agency-╉based intervention designed to treat adolescents
with severe antisocial behavioral problems. As shown in Figure 7.1, MST considers that antisocial
behavior is attributed from multiple social domains including, peers, family, school, and the com-
munity. As the name implies, MST is truly integrative of therapeutic practices, drawing on the use
of cognitive–╉behavioral approaches, behavior therapies, parent training, and family therapies. In
order to address the multiple social domains that affect the child, MST functions as an intensive
therapy tailored to the unique needs of the family and child, and builds on the assistance and
involvement from multiple sources (e.g., teachers, parents, extended family). Based on the func-
tion of social ecology, the MST practitioner delivers the intervention and assessments within the
child’s day-╉to-╉day environment where the maladaptive behavior occurs naturally (e.g., at home,
school), which adds to the ecological validity of MST.

Multisystemic therapy and conduct disorder


The strength of MST comes from the wealth of empirical support for it. After 18 months of MST,
youth with serious antisocial behavior had decreased antisocial symptoms, improved social func-
tioning, and had less out-╉of-╉home placements (Faw, Stambaugh et al., 2007). Compared to com-
munity services, for juvenile offenders who were at imminent risk of placement, MST significantly
decreased substance use and rearrests, and improved school functioning (Timmons-╉Mitchell,
Bender, Kishna, & Mitchell, 2006). Recipients of MST have been shown to be less likely to recidi-
vate (Henggeler et al., 2009) and have improved family relations (Borduin et al., 1995). Further,

Peers

Reduced antisocial
MST Improved family
School behavior and improved
functioning
functioning

Community

Figure 7.1╇ Multisystemic (MST) Therapy Theory of Change


Reproduced from Scott W. Henggeler, Sonja K. Schoenwald, Charles M. Borduin, Melisa D. Rowland, and Phillippe
B. Cunningham, Multisystemic Therapy for Antisocial Behavior in Children and Adolescents, 2e, © Guilford Press, 2009,
with permission.
0
4
1

140 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

four years post-╉MST, adolescents had a reduction in violent crime and drug use (Henggeler,
Clingempeel, Brondino, & Pickrel, 2002). Positive longitudinal results have been found with ado-
lescent sexual offenders who have received MST. When compared to youth receiving commu-
nity services, adolescents who received MST had decreased behavioral problems and symptoms,
decreased sexual and other criminal offending, and improved social relations and academic per-
formance nine years after treatment (Borduin, Schaeffer, & Heiblum, 2009). Based on the high
intensity design of MST, there is strong empirical support for its use in adolescents with conduct
disorder and for treating severe behavioral problems (Brestan & Eyberg, 2010; Curtis, Ronan, &
Borduin, 2004; Henggeler & Sheidow, 2012).

Functional family therapy


Functional family therapy (FFT; Alexander & Parsons, 1973; Sexton & Alexander, 2004) is one
of the oldest and widely applied evidence-╉based therapies for youth displaying severe antisocial
behavior (Henggeler & Sheidow, 2012). FFT is a family-╉based intervention that targets the ado-
lescent’s and family’s maladaptive behaviors. The relational dynamics of the family are central to
FFT, however, FFT employs both cognitive and behavioral strategies. FFT has a focus on how
the behavior of each family member contributes to the environment, which supports the youth’s
behavior. Because all family members are important to the social-╉behavioral model of FFT, all
family members are involved in the intervention. FFT is designed to move through three phases
of change, each of which is designed to support the next phase and increase reciprocity and posi-
tive reinforcement among family members: 1) engagement and motivation, 2) behavioral change,
and 3) generalization. The goals of engagement and motivation are to develop an alliance, improve
communication, engagement, and optimism within the family with the therapist. The aim of the
behavior change is to implement individualized change plans, improve negative behaviors, and
develop relational skills. The final stage, generalization, focuses on relapse prevention and making
use of community support to ensure maintenance of positive gains.

Functional family therapy and conduct disorder


FFT has been recognized as an effective treatment for adolescents with disruptive behavior by
the Centers for Disease Control (CDC) and the Surgeon General. Since the early development of
FFT, it has yielded encouraging results. Compared to three treatment programs, FFT was more
effective at decreasing recidivism and improving family interactions (Alexander & Parsons, 1973).
FFT has been shown to reduce criminal activity for serious juvenile offenders (Barton, Alexander,
Waldron, Turner, & Warburton, 1985). For adolescents who received treatment during their
teens, a reduction in criminality carried over into early adulthood (Gordon, Graves, & Arbuthnot,
1995). A large scale study, including 400 families, showed that compared to a treatment-╉as-╉usual
group, FFT clients had a 38% lower recidivism rate 18-╉months post-╉treatment (Barnowski, 2004).
Further, FFT is considered an effective intervention for the most difficult to treat youths. Although
adolescents with CU traits have been shown to be the most challenging subgroup of adolescents
to treat for behavioral problems (Spain, Douglas, Poythress, & Epstein, 2004), encouraging evi-
dence suggests that FFT is an effective form of treatment (White, Frick, Lawing, & Bauer, 2013).
While youth with CU traits showed higher risk of violence when entering into treatment, FFT
still resulted in improvements over the 20 month treatment period, with decreased risk for vio-
lent offending at six and 12 month post treatment follow-╉ups (White et al., 2013). Because of the
inclusive dynamic of FFT, positive influences on siblings of the target child have been found. FFT
siblings were less likely to have court involvement (20%), compared to the no treatment group
1
4

INCREDIBLE YEARS AND CONDUCT DISORDER 141

(40%), client-╉centered treatment group (59%), and the eclectic-╉dynamic family program group
(63% [Klein, Alexander, & Parsons, 1977]). Overall, there is a wealth of empirical support for the
use of FFT in youth with severe behavioral and offending problems.

Incredible Years
Although research has shown the positive effect that interventions have for teens with CD, com-
pared to early interventions (during childhood) the effectiveness is less reliable (Pardini & Frick,
2013). Therefore, early intervention and prevention during childhood is considered the opti-
mum period for preventing the trajectory to severe antisocial behavior. The Incredible Years (IY;
Webster-╉Stratton, 1984, 2011)  is a well-╉validated set of three programs designed for children,
parents, and teachers (Webster-╉Stratton, Reid, & Hammond, 2004; Webster-╉Stratton, 2016). The
aim of the interlocking series of programs is to prevent, reduce, and treat behavioral problems
and promote social and emotional stability through instruction. The parent training is designed
to target high-╉risk families as well as those families with children with behavior problems. The
parent training programs are age adjusted (toddlers [one to three years], preschoolers [three to
five years], and school-╉age [six to 12  years]) to deliver developmentally appropriate strategies
for increasing child prosocial attitudes and emotional wellbeing, while reducing and preventing
behavior problems (Webster-╉Stratton, 2016). The teacher training program is a six-╉day workshop
designed for educators and school counselors of pupils ages three to ten years. Teachers are taught
classroom management strategies, and how to encourage children’s prosocial behavior and reduce
problematic classroom behavior (e.g., aggression, hostile interpersonal relations). The child pro-
gram promotes friendship, emotion regulation and literacy, and perspective taking for children
ages three to eight years. Children are assessed on one of three “levels” for the most developmen-
tally appropriate class. IY is based on well-╉established behavioral principles, which in applica-
tion are simple and comprehensive to the user, making it a reliable and replicable intervention
(Webster-╉Stratton, Jamila Reid, & Stoolmiller, 2008). IY applies these teaching principles concur-
rently across a variety of environments, supporting prosocial behaviors in “real-╉world” settings,
which essentially covers all areas in which the child socializes (Boxer & Frick, 2008). Programs
such as IY, that apply positive modifications to the child’s environment (e.g., parenting behaviors),
are reliably shown to be effective methods of improving childhood behavioral outcomes (Gridley,
Hutchings, & Baker-╉Henningham, 2015).

Incredible Years and conduct disorder


Thirty years of development and validation with multiple randomized controlled trials supports
IY as one of the most widely used, cost-╉effective, and validated programs for early intervention
and prevention programs for children with conduct problems (McIntyre, 2008; Menting, Orobio
de Castro, & Matthys, 2013). For teachers participating in the IY program, benefits have included
more proactive teaching strategies, more emotional support for pupils, and the use of fewer
critical and coercive tactics (Webster-╉Stratton, Gaspar, & Seabra-╉Santos, 2012). Parents respond
positively to the IY program, which has led to a high attendance rate (Homem, Gaspar, Santos,
Azevedo, & Canavarro, 2015). Parents of children with high levels of externalizing and conduct
problems have sustained improvement in mother-╉child interactions and decreases in children’s
oppositional behaviors 12 months post intervention (Homem et al., 2015). Two years post treat-
ment, parents of children (age four) who were at risk of developing chronic conduct problems
were assessed by observational and self-╉report measures, and compared to a control group who
received “care as usual” (Posthumus, Raaijmakers, Maassen, van Engeland, & Matthys, 2012).
2
4
1

142 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

The parents who participated in the program showed significant improvements compared to the
control group. Parents reported using less harsh and inconsistent discipline and more positive
parenting strategies (e.g., praise, appropriate discipline), and were observed using fewer critical
statements to their child (Posthumus et al., 2012). IY has been shown to aid mothers who were
recently released from prison, with parent-╉and teacher-╉report of improvements in child behavior
(Menting, de Castro, Wijngaards-╉de Meij, & Matthys, 2014). A meta-╉analysis including 50 studies
(see Menting et al., 2013) found that children who entered the program with more severe behav-
ior problems reported the greatest improvements such that prosocial behavior had increased and
oppositional behavior decreased immediately after the intervention, for observations, teacher and
parent ratings.
Children who come from homes with a family history of externalizing behaviors have been
suggested to have a genetic risk as well as an environmental risk for showing antisocial behav-
ior cross-╉generationally (Silberg, Maes, & Eaves, 2012). These children are more likely to have
chronic behavior problems in childhood and develop antisocial personality disorder in adulthood
(Lahey et al., 1988). A recent study (see Presnall, Webster-╉Stratton, & Constantino, 2014) assessed
the effectiveness of the IY program for children (three to eight years) with CD, with and without
a family history of externalizing behaviors (e.g., Antisocial Personality Disorder). Although chil-
dren from families with histories of externalizing behaviors had more severe conduct disorder
symptoms upon entry into the program, both groups benefitted from the intervention showing
a reduction in externalizing behavior (Presnall et al., 2014). Overall, IY has been shown to have
excellent utility in diverse samples, including those who come from homes which pose the great-
est risk for developing severe behavioral problems, which is why IY is considered by the National
Institute of Justice as an effective form of prevention and treatment for children with conduct
problems.

Conclusion
The cost of conduct problems is extensive, causing emotional and physical damage to victims, and
causing financial and community resource burdens. To add to the challenge, these children are
viewed by society as “bad” children rather than children suffering from a mental illness. Further,
treatment is notoriously difficult in this population, especially due to the etiological and devel-
opmental heterogeneity within the disorder. Despite all these setbacks, effective intervention
programs have begun to accumulate evidenced in support of improving behavior for children
with conduct disorder, even in the most challenging subgroups. These interventions offer sig-
nificant life improvement for children and families, and substantial cost-╉savings to society when
compared to children not receiving adequate interventions (Bonin, Stevens, Beecham, Byford,
& Parsonage, 2011). Although intervention programs tend to focus on outcome measures (e.g.,
aggression), further improvement in treatment outcomes can be accomplished by understanding
that conduct disorder is multifaceted and subgroups can be differentiated based on etiological and
developmental differences. Thus, we should aim to tailor treatment and family interventions to
specific subgroups of children with conduct disorder.

References
Alexander, J. F., & Parsons, B. V. (1973). Short-╉term behavioral intervention with delinquent
families: Impact on family process and recidivism. Journal of Abnormal Psychology, 81(3), 219–╉225.
doi:10.1037/╉h0034537
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Washington, DC.
3
4
1

Conclusion 143

Arsenio, W. F., Adams, E., & Gold, J. Social information processing, moral reasoning, and emotion
attributions: relations with adolescents’ reactive and proactive aggression. Child Development, 80(6),
1739–​55. doi:10.1111/​j.1467-​8624.2009.01365.x
Aspinwall, L. G. (1998). Rethinking the Role of Positive Affect in Self-​Regulation. Motivation and Emotion,
22(1), 1–​32. doi:10.1023/​A:1023080224401
Babcock, J. C., Tharp, A. L. T., Sharp, C., Heppner, W., & Stanford, M. S. (2014). Similarities and
differences in impulsive/​premeditated and reactive/​proactive bimodal classifications of aggression.
Aggression and Violent Behavior, 19(3), 251–​262. doi:10.1016/​j.avb.2014.04.002
Bardone, A. M., Moffitt, T. E., Caspi, A., Dickson, N., Stanton, W. R., & Silva, P. A. (1998). Adult
physical health outcomes of adolescent girls with conduct disorder, depression, and anxiety.
Journal of the American Academy of Child and Adolescent Psychiatry, 37(6), 594–​601. doi:10.1097/​
00004583-​199806000-​00009
Barnowski, R. (2004). Outcome evaluation of Washington State’s research-​based programs for juvenile
offenders. Olympia, WA: Washington State Institute for Public Policy.
Barton, C., Alexander, J. F., Waldron, H., Turner, C. W., & Warburton, J. (1985). Generalizing Treatment
Effects on Functional Family Therapy: Three Replications. American Journal of Family Therapy, 13(3),
16–​26. Retrieved from http://​www.safetylit.org/​citations/​index.php?fuseaction=citations.viewdetails&ci
tationIds%5B%5D=citjournalarticle_​415921_​38
Beauchaine, T. P. (2015). Respiratory Sinus Arrhythmia: A Transdiagnostic Biomarker of Emotion
Dysregulation and Psychopathology. Current Opinion in Psychology, 3, 43–​47. doi:10.1016/​
j.copsyc.2015.01.017
Blair, B. L., Perry, N. B., O’Brien, M., Calkins, S. D., Keane, S. P., & Shanahan, L. (2015). Identifying
developmental cascades among differentiated dimensions of social competence and emotion regulation.
Developmental Psychology, 51(8), 1062–​1073. doi:10.1037/​a0039472
Blair, C., & Raver, C. C. (2015). School readiness and self-​regulation: a developmental psychobiological
approach. Annual Review of Psychology, 66, 711–​731. doi:10.1146/​annurev-​psych-​010814-​015221
Blair, R. J. R., Leibenluft, E., & Pine, D. S. (2014). Conduct disorder and callous-​unemotional traits in
youth. New England Journal of Medicine, 371(23), 2207–​2216. doi:10.1056/​NEJMra1315612
Bobadilla, L., Wampler, M., & Taylor, J. (2012). Proactive and Reactive Aggression are Associated with
Different Physiological and Personality Profiles. Journal of Social and Clinical Psychology, 31(5), 458–​
487. doi:10.1521/​jscp.2012.31.5.458
Bohnert, A. M., Crnic, K. A., & Lim, K. G. (2003). Emotional Competence and Aggressive Behavior
in School-​Age Children. Journal of Abnormal Child Psychology, 31(1), 79–​91. doi:10.1023/​
A:1021725400321
Bonin, E.-​M., Stevens, M., Beecham, J., Byford, S., & Parsonage, M. (2011). Costs and longer-​term savings
of parenting programmes for the prevention of persistent conduct disorder: a modelling study. BMC
Public Health, 11, 803. doi:10.1186/​1471-​2458-​11-​803
Borduin, C. M., Mann, B. J., Cone, L. T., Henggeler, S. W., Fucci, B. R., Blaske, D. M., & Williams, R. A.
(1995). Multisystemic treatment of serious juvenile offenders: Long-​term prevention of criminality and
violence. Journal of Consulting and Clinical Psychology, 63(4), 569–​578. doi:10.1037/​0022-​006X.63.4.569
Borduin, C. M., Schaeffer, C. M., & Heiblum, N. (2009). A randomized clinical trial of multisystemic
therapy with juvenile sexual offenders: Effects on youth social ecology and criminal activity. Journal of
Consulting and Clinical Psychology, 77(1), 26–​37.
Boxer, P., & Frick, P. J. (2008). Treating Conduct Problems, Aggression, and Antisocial Behavior in
Children and Adolescents: An Integrated View. In R. G. Steele, D. T. Elkin, & M. C. Roberts (Eds.),
Handbook of Evidence-​Based Therapies for Children and Adolescents: Bridging Science and Practice (pp.
241–​260). New York: Springer Science + Business Media, LLC.
Boylan, K., Vaillancourt, T., Boyle, M., & Szatmari, P. (2007). Comorbidity of internalizing disorders in
children with oppositional defiant disorder. European Child & Adolescent Psychiatry, 16(8), 484–​494.
doi:10.1007/​s00787-​007-​0624-​1
41

144 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Breeden, A. L., Cardinale, E. M., Lozier, L. M., VanMeter, J. W., & Marsh, A. A. (2015). Callous-​
unemotional traits drive reduced white-​matter integrity in youths with conduct problems. Psychological
Medicine, 1–​14. doi:10.1017/​S0033291715000987
Brestan, E. V., & Eyberg, S. M. (2010). Effective psychosocial treatments of conduct-​disordered children
and adolescents: 29 years, 82 studies, and 5,272 kids. Journal of Clinical Child Psychology, 27(2), 180189.
Retrieved from http://​www.tandfonline.com/​doi/​abs/​10.1207/​s15374424jccp2702_​5
Bridgett, D. J., Oddi, K. B., Laake, L. M., Murdock, K. W., & Bachmann, M. N. (2013). Integrating and
differentiating aspects of self-​regulation: effortful control, executive functioning, and links to negative
affectivity. Emotion (Washington, D.C.), 13(1), 47–​63. doi:10.1037/​a0029536
Burke, J. D., Loeber, R., & Lahey, B. B. (2007). Adolescent conduct disorder and interpersonal callousness
as predictors of psychopathy in young adults. Journal of Clinical Child and Adolescent Psychology : The
Official Journal for the Society of Clinical Child and Adolescent Psychology, American Psychological
Association, Division 53, 36(3), 334–​346. doi:10.1080/​15374410701444223
Byrd, A. L., Loeber, R., & Pardini, D. A. (2012). Understanding desisting and persisting forms of
delinquency: the unique contributions of disruptive behavior disorders and interpersonal callousness.
Journal of Child Psychology and Psychiatry, and Allied Disciplines, 53(4), 371–​380. doi:10.1111/​
j.1469-​7610.2011.02504.x
Calkins, S. D. (1994). Origins and outcomes of individual differences in emotion regulation. Monographs of
the Society for Research in Child Development, 59(2-​3), 53–​72. Retrieved from http://​www.ncbi.nlm.nih.
gov/​pubmed/​7984167
Centifanti, L. C. M., Fanti, K. A., Thomson, N. D., Demetriou, V., & Anastassiou-​Hadjicharalambous,
X. (2015). Types of Relational Aggression in Girls Are Differentiated by Callous-​Unemotional Traits,
Peers and Parental Overcontrol. Behavioral Sciences (Basel, Switzerland), 5(4), 518–​536. doi:10.3390/​
bs5040518
Centifanti, L. C. M., Meins, E., & Fernyhough, C. (2015). Callous-​unemotional traits and
impulsivity: distinct longitudinal relations with mind-​mindedness and understanding of others. Journal
of Child Psychology and Psychiatry, and Allied Disciplines, 57(1), 84-​92. doi:10.1111/​jcpp.12445
Cohen, P., Cohen, J., & Brook, J. (1993). An Epidemiological Study of Disorders in Late Childhood and
Adolescence?II. Persistence of Disorders. Journal of Child Psychology and Psychiatry, 34(6), 869–​877.
doi:10.1111/​j.1469-​7610.1993.tb01095.x
Colins, O. F., & Andershed, H. (2015). The DSM-​5 with limited prosocial emotions specifier for conduct
disorder among detained girls. Law and Human Behavior, 39(2), 198–​207. doi:10.1037/​lhb0000108
Copeland, W. E., Angold, A., Costello, E. J., & Egger, H. (2013). Prevalence, comorbidity, and correlates of
DSM-​5 proposed disruptive mood dysregulation disorder. The American Journal of Psychiatry, 170(2),
173–​179. doi:10.1176/​appi.ajp.2012.12010132
Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social information-​processing
mechanisms in children’s social adjustment. Psychological Bulletin, 115(1), 74–​101. doi:10.1037/​
0033-​2909.115.1.74
Crick, N. R., & Dodge, K. A. (1996). Social Information-​Processing Mechanisms in Reactive and Proactive
Aggression. Child Development, 67(3), 993–​1002. doi:10.1111/​j.1467-​8624.1996.tb01778.x
Curtis, N. M., Ronan, K. R., & Borduin, C. M. (2004). Multisystemic treatment: a meta-​analysis of outcome
studies. Journal of Family Psychology : JFP : Journal of the Division of Family Psychology of the American
Psychological Association (Division 43), 18(3), 411–​419. doi:10.1037/​0893-​3200.18.3.411
Dadds, M. R., & Fraser, J. A. (2003). Prevention Programs. In C. A. Essau (Ed.), Conduct and Oppositional
Defiant Disorders Epidemiology, Risk Factors, and Treatment (pp. 193–​222). New Jersey: Lawrence
Erlbaum Associates, Inc., Publishers.
Dalwani, M. S., McMahon, M. A., Mikulich-​Gilbertson, S. K., Young, S. E., Regner, M. F., Raymond,
K. M., … Sakai, J. T. (2015). Female adolescents with severe substance and conduct problems
have substantially less brain gray matter volume. PloS One, 10(5), e0126368. doi:10.1371/​journal.
pone.0126368
5
4
1

Conclusion 145

de Castro, B. O., Merk, W., Koops, W., Veerman, J. W., & Bosch, J. D. (2005). Emotions in social
information processing and their relations with reactive and proactive aggression in referred aggressive
boys. Journal of Clinical Child and Adolescent Psychology : The Official Journal for the Society of Clinical
Child and Adolescent Psychology, American Psychological Association, Division 53, 34(1), 105–​116.
doi:10.1207/​s15374424jccp3401_​10
de Castro, B. O., Verhulp, E. E., & Runions, K. (2012). Rage and revenge: Highly aggressive boys’
explanations for their responses to ambiguous provocation. European Journal of Developmental
Psychology, 9(3), 331–​350. doi:10.1080/​17405629.2012.680304
de Wied, M., van Boxtel, A., Matthys, W., & Meeus, W. (2012). Verbal, facial and autonomic responses to
empathy-​eliciting film clips by disruptive male adolescents with high versus low callous-​unemotional
traits. Journal of Abnormal Child Psychology, 40(2), 211–​223. doi:10.1007/​s10802-​011-​9557-​8
Dodge, K. A. (1991). The structure and function of reactive and proactive aggression. In D. Peppler
& K. Rubin (Eds.), The development and treatment of childhood aggression (pp. 201–​218). Hillside,
NJ: Erlbaum.
Dodge, K. A., & Coie, J. D. (1987). Social-​information-​processing factors in reactive and proactive
aggression in children’s peer groups. Journal of Personality and Social Psychology, 53(6), 1146–​1158.
Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​3694454/​
Dodge, K. A., Lochman, J. E., Harnish, J. D., Bates, J. E., & Pettit, G. S. (1997). Reactive and proactive
aggression in school children and psychiatrically impaired chronically assaultive youth. Journal of
Abnormal Psychology, 106(1), 37–​51. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​9103716/​
Dodge, K. A., Malone, P. S., Lansford, J. E., Sorbring, E., Skinner, A. T., Tapanya, S., … Pastorelli, C.
(2015). Hostile attributional bias and aggressive behavior in global context. Proceedings of the National
Academy of Sciences of the United States of America, 112(30), 9310–​9315. doi:10.1073/​pnas.1418572112
Dodge, K. A., & Pettit, G. S. (2003). A biopsychosocial model of the development of chronic conduct
problems in adolescence. Developmental Psychology, 39(2), 349–​371. Retrieved from http://​www.
pubmedcentral.nih.gov/​articlerender.fcgi?artid=2755613&tool=pmcentrez&rendertype=abstract
Dollar, J. M., & Stifter, C. A. (2012). Temperamental surgency and emotion regulation as predictors of
childhood social competence. Journal of Experimental Child Psychology, 112(2), 178–​194. doi:10.1016/​
j.jecp.2012.02.004
Egger, H. L., & Angold, A. Common emotional and behavioral disorders in preschool
children: presentation, nosology, and epidemiology. Journal of Child Psychology and Psychiatry, and
Allied Disciplines, 47(3-​4), 313–​337. doi:10.1111/​j.1469-​7610.2006.01618.x
Eisenberg, N., Fabes, R. A., Shepard, S. A., Murphy, B. C., Guthrie, I. K., Jones, S., … Maszk, P. (1997).
Contemporaneous and Longitudinal Prediction of Children’s Social Functioning from Regulation and
Emotionality. Child Development, 68(4), 642–​664. doi:10.1111/​j.1467-​8624.1997.tb04227.x
Essau, C. A., Sasagawa, S., & Frick, P. J. (2006). Callous-​unemotional traits in a community sample of
adolescents. Assessment, 13(4), 454–​469. doi:10.1177/​1073191106287354
Esser, G., Schmidt, M. H., & Woerner, W. (1990). Epidemiology and course of psychiatric disorders in
school-​age children-​-​results of a longitudinal study. Journal of Child Psychology and Psychiatry, and
Allied Disciplines, 31(2), 243–​263. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​2312652/​
Evans, S. C., Fite, P. J., Hendrickson, M. L., Rubens, S. L., & Mages, A. K. (2015). The Role of Reactive
Aggression in the Link Between Hyperactive-​Impulsive Behaviors and Peer Rejection in Adolescents.
Child Psychiatry and Human Development. doi:10.1007/​s10578-​014-​0530-​y
Fairchild, G., Hagan, C. C., Passamonti, L., Walsh, N. D., Goodyer, I. M., & Calder, A. J. (2014).
Atypical neural responses during face processing in female adolescents with conduct disorder.
Journal of the American Academy of Child and Adolescent Psychiatry, 53(6), 677–​687.e5. doi:10.1016/​
j.jaac.2014.02.009
Fanti, K. A., Panayiotou, G., Kyranides, M. N., & Avraamides, M. N. (2015). Startle modulation during
violent films: Association with callous–​unemotional traits and aggressive behavior. Motivation and
Emotion. doi:10.1007/​s11031-​015-​9517-​7
6
4
1

146 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Fanti, K. A., Panayiotou, G., Lazarou, C., Michael, R., & Georgiou, G. (2015). The better of two evils?
Evidence that children exhibiting continuous conduct problems high or low on callous-​unemotional
traits score on opposite directions on physiological and behavioral measures of fear. Development and
Psychopathology, 1–​14. doi:10.1017/​S0954579415000371
Fanti, K. A., Panayiotou, G., Lombardo, M. V., & Kyranides, M. N. (2015). Unemotional on all
counts: Evidence of reduced affective responses in individuals with high callous-​unemotional traits
across emotion systems and valences. Social Neuroscience, 1–​16. doi:10.1080/​17470919.2015.1034378
Faraone, S. V., Biederman, J., Keenan, K., & Tsuang, M. T. (1991). Separation of DSM-​III attention deficit
disorder and conduct disorder: evidence from a family-​genetic study of American child psychiatric
patients. Psychological Medicine, 21(01), 109. doi:10.1017/​S0033291700014707
Faw Stambaugh, L., Mustillo, S. A., Burns, B. J., Stephens, R. L., Baxter, B., Edwards, D., & Dekraai,
M. (2007). Outcomes From Wraparound and Multisystemic Therapy in a Center for Mental Health
Services System-​of-​Care Demonstration Site. Journal of Emotional and Behavioral Disorders, 15(3),
143–​155. doi:10.1177/​10634266070150030201
Frick, P. J., Cornell, A. H., Barry, C. T., Bodin, S. D., & Dane, H. E. (2003). Callous-​unemotional traits
and conduct problems in the prediction of conduct problem severity, aggression, and self-​report of
delinquency. Journal of Abnormal Child Psychology, 31(4), 457–​470. Retrieved from http://​www.ncbi.
nlm.nih.gov/​pubmed/​12831233
Frick, P. J., & Dickens, C. (2006). Current perspectives on conduct disorder. Current Psychiatry Reports,
8(1), 59–​72. doi:10.1007/​s11920-​006-​0082-​3
Frick, P. J., & Morris, A. S. (2004). Temperament and developmental pathways to conduct problems.
Journal of Clinical Child and Adolescent Psychology : The Official Journal for the Society of Clinical Child
and Adolescent Psychology, American Psychological Association, Division 53, 33(1), 54–​68. doi:10.1207/​
S15374424JCCP3301_​6
Frick, P. J., Ray, J. V, Thornton, L. C., & Kahn, R. E. (2014). Annual research review: A developmental
psychopathology approach to understanding callous-​unemotional traits in children and adolescents
with serious conduct problems. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 55(6),
532–​548. doi:10.1111/​jcpp.12152
Frick, P. J., Stickle, T. R., Dandreaux, D. M., Farrell, J. M., & Kimonis, E. R. (2005). Callous–​Unemotional
Traits in Predicting the Severity and Stability of Conduct Problems and Delinquency. Journal of
Abnormal Child Psychology, 33(4), 471–​487. doi:10.1007/​s10648-​005-​5728-​9
Frick, P. J., & Viding, E. (2009). Antisocial behavior from a developmental psychopathology perspective.
Development and Psychopathology, 21(4), 1111–​1131. doi:10.1017/​S0954579409990071
Gordon, D. A., Graves, K., & Arbuthnot, J. (1995). The Effect of Functional Family Therapy for
Delinquents on Adult Criminal Behavior. Criminal Justice and Behavior, 22(1), 60–​73. doi:10.1177/​
0093854895022001005
Gottman, J. M., Katz, L. F., & Hooven, C. (1997). Meta-​emotion. Hillside, NJ: Erlbaum.
Gridley, N., Hutchings, J., & Baker-​Henningham, H. (2015). The Incredible Years Parent-​Toddler
Programme and parental language: a randomised controlled trial. Child: Care, Health and Development,
41(1), 103–​111. doi:10.1111/​cch.12153
Harper, B. D., Lemerise, E. A., & Caverly, S. L. (2010). The effect of induced mood on children’s social
information processing: goal clarification and response decision. Journal of Abnormal Child Psychology,
38(5), 575–​586. doi:10.1007/​s10802-​009-​9356-​7
Hawkins, J. D., Catalano, R. F., Kosterman, R., Abbott, R., & Hill, K. G. (1999). Preventing Adolescent
Health-​Risk Behaviors by Strengthening Protection During Childhood. Archives of Pediatrics &
Adolescent Medicine, 153(3), 226–​234. doi:10.1001/​archpedi.153.3.226
Helseth, S. A., Waschbusch, D. A., King, S., & Willoughby, M. T. (2015). Aggression in Children with
Conduct Problems and Callous-​Unemotional Traits: Social Information Processing and Response
to Peer Provocation. Journal of Abnormal Child Psychology, 43(8), 1503–​1514. doi:10.1007/​
s10802-​015-​0027-​6
7
4
1

Conclusion 147

Henggeler, S. W., Clingempeel, W. G., Brondino, M. J., & Pickrel, S. G. (2002). Four-​year follow-​up
of multisystemic therapy with substance-​abusing and substance-​dependent juvenile offenders.
Journal of the American Academy of Child and Adolescent Psychiatry, 41(7), 868–​874. doi:10.1097/​
00004583-​200207000-​00021
Henggeler, S. W., Schoenwald, S. K., Borduin, C. M., Rowland, M. D., & Cunningham, P. B. (2009).
Multisystemic Therapy for Antisocial Behavior in Children and Adolescents. New York: Guilford Press.
Henggeler, S. W., & Sheidow, A. J. (2012). Empirically supported family-​based treatments for conduct
disorder and delinquency in adolescents. Journal of Marital and Family Therapy, 38(1), 30–​58.
doi:10.1111/​j.1752-​0606.2011.00244.x
Hinnant, J. B., Erath, S. A., & El-​Sheikh, M. (2015). Harsh parenting, parasympathetic activity, and
development of delinquency and substance use. Journal of Abnormal Psychology, 124(1), 137–​151.
doi:10.1037/​abn0000026
Hinshaw, S. P., & Lee, S. S. (2003). Oppositional defiant and conduct disorder. In E. J. Mash & R. A. Barkley
(Eds.), Child psychopathology (2nd ed., pp. 144–​198). New York: Guilford Press.
Homem, T. C., Gaspar, M. F., Santos, M. J. S., Azevedo, A. F., & Canavarro, M. C. (2015). Incredible Years
Parent Training: Does it Improve Positive Relationships in Portuguese Families of Preschoolers with
Oppositional/​Defiant Symptoms? Journal of Child and Family Studies, 24(7), 1861–​1875. doi:10.1007/​
s10826-​014-​9988-​2
Hoppenbrouwers, S. S., Nazeri, A., de Jesus, D. R., Stirpe, T., Felsky, D., Schutter, D. J. L. G., … Voineskos,
A. N. (2013). White Matter Deficits in Psychopathic Offenders and Correlation with Factor Structure.
PLoS ONE, 8(8), e72375. doi:10.1371/​journal.pone.0072375
Hubbard, J. A., Dodge, K. A., Cillessen, A. H., Coie, J. D., & Schwartz, D. (2001). The dyadic nature of
social information processing in boys’ reactive and proactive aggression. Journal of Personality and
Social Psychology, 80(2), 268–​280. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​11220445
Hubbard, J. A., Smithmyer, C. M., Ramsden, S. R., Parker, E. H., Flanagan, K. D., Dearing, K. F.,
… Simons, R. F. (2002). Observational, Physiological, and Self-​Report Measures of Children’s
Anger: Relations to Reactive versus Proactive Aggression. Child Development, 73(4), 1101–​1118.
doi:10.1111/​1467-​8624.00460
Izard, C. E. (1991). The Psychology of Emotions. New York: Plenum Press.
Johnson, V. A., Kemp, A. H., Heard, R., Lennings, C. J., & Hickie, I. B. (2015). Childhood-​versus
adolescent-​onset antisocial youth with conduct disorder: psychiatric illness, neuropsychological and
psychosocial function. PloS One, 10(4), e0121627. doi:10.1371/​journal.pone.0121627
Jones, P. B. (2013). Adult mental health disorders and their age at onset. The British Journal of Psychiatry.
Supplement, 54(s54), s5–​s10. doi:10.1192/​bjp.bp.112.119164
Kessler, R. C., Avenevoli, S., Costello, E. J., Georgiades, K., Green, J. G., Gruber, M. J., … Merikangas,
K. R. (2012). Prevalence, persistence, and sociodemographic correlates of DSM-​IV disorders in the
National Comorbidity Survey Replication Adolescent Supplement. Archives of General Psychiatry, 69(4),
372–​380. doi:10.1001/​archgenpsychiatry.2011.160
Kim-​Cohen, J., Arseneault, L., Caspi, A., Tomás, M. P., Taylor, A., & Moffitt, T. E. (2005). Validity of DSM-​
IV conduct disorder in 41/​2-​5-​year-​old children: a longitudinal epidemiological study. The American
Journal of Psychiatry, 162(6), 1108–​1117. doi:10.1176/​appi.ajp.162.6.1108
Kimonis, E. R., Centifanti, L. C. M., Allen, J. L., & Frick, P. J. (2014). Reciprocal influences between
negative life events and callous-​unemotional traits. Journal of Abnormal Child Psychology, 42(8), 1287–​
1298. doi:10.1007/​s10802-​014-​9882-​9
Kimonis, E. R., Frick, P. J., & McMahon, R. J. (2014). Conduct and oppositional defiant disorders. In
E. J. Mash & R. A. Barkley (Eds.), Child psychopathology (3rd ed., pp. 145–​179). New York: The
Guilford Press.
Klein, N. C., Alexander, J. F., & Parsons, B. V. (1977). Impact of family systems intervention on recidivism
and sibling delinquency: A model of primary prevention and program evaluation. Journal of Consulting
and Clinical Psychology, 45(3), 469–​474. doi:10.1037/​0022-​006X.45.3.469
8
4
1

148 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Kochanska, G. (1991). Socialization and Temperament in the Development of Guilt and Conscience. Child
Development, 62(6), 1379–​1392. doi:10.1111/​j.1467-​8624.1991.tb01612.x
Kochanska, G., DeVet, K., Goldman, M., Murray, K., & Putnam, S. P. (1994). Maternal reports of
conscience development and temperament in young children. Child Development, 65(3), 852–​868.
Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​8045172
Krieger, F. V., Polanczyk, V. G., Goodman, R., Rohde, L. A., Graeff-​Martins, A. S., Salum, G., …
Stringaris, A. (2013). Dimensions of oppositionality in a Brazilian community sample: testing the
DSM-​5 proposal and etiological links. Journal of the American Academy of Child and Adolescent
Psychiatry, 52(4), 389–​400.e1. doi:10.1016/​j.jaac.2013.01.004
Lahey, B. B., Piacentini, J. C., McBurnett, K., Stone, P., Hartdaghn, S., & Hynd, G. (1988). Psychopathology
in the parents of children with conduct disorder and hyperactivity. Journal of the American Academy of
Child & Adolescent Psychiatry, 27(2), 163–​170. doi:10.1097/​00004583-​198803000-​00005
Laible, D., & Panfile, T. (2009). Mother-​child reminiscing in the context of secure attachment
relationships: Lessons in understanding and coping with negative emotions. In J. Quas & R. Fivush
(Eds.), Stress and memory development: Biological, social and emotional considerations (pp. 166–​195).
New York: Oxford University Press.
Lemerise, E. A., & Arsenio, W. F. (2000). An Integrated Model of Emotion Processes and Cognition in
Social Information Processing. Child Development, 71(1), 107–​118. doi:10.1111/​1467-​8624.00124
Lemerise, E. A., & Dodge, K. A. (1993). The Development of Anger and Hostile Interactions. In M. Lewis &
J. M. Haviland (Eds.), Handbook of Emotions (pp. 537–​546). New York: Guilford Press.
Lemerise, E. A., & Dodge, K. A. (2008). The Development of Anger and Hostile Interactions. In M. Lewis,
J. M. Haviland-​Jones, & L. Feldman Barrett (Eds.), Handbook of Emotions (3rd ed., pp. 730–​741).
New York: The Guilford Press.
Lemerise, E. A., & Harper, B. D. (2014). Emotional Competence and Social Relations. In K. H. Lagattuta
(Ed.), Children and Emotion (Vol. 26, pp. 57–​66). Basel: S. KARGER AG. doi:10.1159/​000354353
Lindhiem, O., Bennett, C. B., Hipwell, A. E., & Pardini, D. A. (2015). Beyond Symptom Counts for
Diagnosing Oppositional Defiant Disorder and Conduct Disorder? Journal of Abnormal Child
Psychology, 43(7), 1379–​1387. doi:10.1007/​s10802-​015-​0007-​x
Loeber, R., Burke, J. D., Lahey, B. B., Winters, A., & Zera, M. (2000). Oppositional defiant and conduct
disorder: a review of the past 10 years, part I. Journal of the American Academy of Child and Adolescent
Psychiatry, 39(12), 1468–​1484. doi:10.1097/​00004583-​200012000-​00007
Loeber, R., Green, S. M., Keenan, K., & Lahey, B. B. (1995). Which Boys Will Fare Worse? Early Predictors
of the Onset of Conduct Disorder in a Six-​Year Longitudinal Study. Journal of the American Academy of
Child & Adolescent Psychiatry, 34(4), 499–​509. doi:10.1097/​00004583-​199504000-​00017
Loney, B. R., Frick, P. J., Ellis, M., & McCoy, M. G. (1998). Intelligence, Callous-​Unemotional Traits,
and Antisocial Behavior. Journal of Psychopathology and Behavioral Assessment, 20(3), 231–​247.
doi:10.1023/​A:1023015318156
Loney, B. R., & Lima, E. N. (2003). Classification and Assessment. In E. C. A (Ed.), Conduct
and oppositional defiant disorders: Epidemiology, risk factors, and treatment (pp. 3–​32). New
Jersey: Lawrence Erlbaum Associates.
Malatesta, C. Z., Culver, C., Tesman, J. R., & Shepard, B. (1989). The development of emotion expression
during the first two years of life. Monographs of the Society for Research in Child Development, 54(1-​2),
1–​104; discussion 105–​136. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​2770755
Marshall, P. J., Fox, N. A., & Henderson, H. A. (2000). Temperament as an Organizer of Development.
Infancy, 1(2), 239–​244. doi:10.1207/​S15327078IN0102_​5
McCabe, K. M., Hough, R., Wood, P. A., & Yeh, M. (2001). Childhood and Adolescent Onset Conduct
Disorder: A Test of the Developmental Taxonomy. Journal of Abnormal Child Psychology, 29(4), 305–​
316. doi:10.1023/​A:1010357812278
McGee, R., Feehan, M., Williams, S., Partridge, F., Silva, P. A., & Kelly, J. (1990). DSM-​III disorders in a
large sample of adolescents. Journal of the American Academy of Child and Adolescent Psychiatry, 29(4),
611–​619. doi:10.1097/​00004583-​199007000-​00016
9
4
1

Conclusion 149

McIntyre, L. L. (2008). Adapting Webster-​Stratton’s incredible years parent training for children with
developmental delay: findings from a treatment group only study. Journal of Intellectual Disability
Research : JIDR, 52(12), 1176–​1192. doi:10.1111/​j.1365-​2788.2008.01108.x
McMahon, R. J., Witkiewitz, K., & Kotler, J. S. T. C. P. P. R. G. (2010). Predictive validity of callous–​
unemotional traits measured in early adolescence with respect to multiple antisocial outcomes. Journal
of Abnormal Psychology, 119(4), 752–​763. doi:10.1037/​a0020796
Meltzoff, A. N. (2002). Elements of a Developmental Theory of Imitation. In A. N. Meltzoff & P. Wolfgang
(Eds.), The Imitative Mind: Development, Evolution and Brain Bases (pp. 19–​41). New York: Cambridge
University Press.
Menting, A. T. A., de Castro, B. O., Wijngaards-​de Meij, L. D. N. V, & Matthys, W. (2014). A trial of parent
training for mothers being released from incarceration and their children. Journal of Clinical Child and
Adolescent Psychology : The Official Journal for the Society of Clinical Child and Adolescent Psychology,
American Psychological Association, Division 53, 43(3), 381–​396. doi:10.1080/​15374416.2013.817310
Menting, A. T. A., Orobio de Castro, B., & Matthys, W. (2013). Effectiveness of the Incredible Years parent
training to modify disruptive and prosocial child behavior: a meta-​analytic review. Clinical Psychology
Review, 33(8), 901–​913. doi:10.1016/​j.cpr.2013.07.006
Moffitt, T. E. (1993). Adolescence-​limited and life-​course-​persistent antisocial behavior: A developmental
taxonomy. Psychological Review, 100(4), 674–​701. doi:10.1037/​0033-​295X.100.4.674
Moffitt, T. E. (2006). Life-​course persistent versus adolescence-​limited antisocial behavior. In D. Cicchetti &
J. Cohen (Eds.), Developmental psychopathology: Risk, disorder, and adaptation (2nd ed., pp. 570–​598).
New York: Wiley.
Moffitt, T. E., & Caspi, A. (2001). Childhood predictors differentiate life-​course persistent and adolescence-​
limited antisocial pathways among males and females. Development and Psychopathology, 13(2), 355–​
375. doi:10.1017/​S0954579401002097
Moffitt, T. E., Caspi, A., Dickson, N., Silva, P., & Stanton, W. (1996). Childhood-​onset versus adolescent-​
onset antisocial conduct problems in males: Natural history from ages 3 to 18 years. Development and
Psychopathology, 8(02), 399. doi:10.1017/​S0954579400007161
Munoz, L. C., & Frick, P. J. (2012). Callous-​Unemotional Traits and Their Implication for Understanding
and Treating Aggressive and Violent Youths. Criminal Justice and Behavior, 39(6), 794–​813.
doi:10.1177/​0093854812437019
Muñoz, L. C., Frick, P. J., Kimonis, E. R., & Aucoin, K. J. (2008). Types of aggression, responsiveness
to provocation, and callous-​unemotional traits in detained adolescents. Journal of Abnormal Child
Psychology, 36(1), 15–​28. doi:10.1007/​s10802-​007-​9137-​0
Nock, M. K., Kazdin, A. E., Hiripi, E., & Kessler, R. C. (2006). Prevalence, subtypes, and correlates of
DSM-​IV conduct disorder in the National Comorbidity Survey Replication. Psychological Medicine,
36(5), 699–​710. doi:10.1017/​S0033291706007082
Nozadi, S. S., Spinrad, T. L., Eisenberg, N., & Eggum-​Wilkens, N. D. (2015). Associations of Anger and
Fear to Later Self-​Regulation and Problem Behavior Symptoms. Journal of Applied Developmental
Psychology, 38, 60–​69. doi:10.1016/​j.appdev.2015.04.005
Odgers, C. L., Moffitt, T. E., Broadbent, J. M., Dickson, N., Hancox, R. J., Harrington, H., … Caspi, A.
(2008). Female and male antisocial trajectories: from childhood origins to adult outcomes. Development
and Psychopathology, 20(2), 673–​716. doi:10.1017/​S0954579408000333
Okado, Y., & Bierman, K. L. (2015). Differential risk for late adolescent conduct problems and mood
dysregulation among children with early externalizing behavior problems. Journal of Abnormal Child
Psychology, 43(4), 735–​747. doi:10.1007/​s10802-​014-​9931-​4
Olson, S. L., Bates, J. E., Sandy, J. M., & Schilling, E. M. (2002). Early developmental precursors of impulsive
and inattentive behavior: from infancy to middle childhood. Journal of Child Psychology and Psychiatry,
and Allied Disciplines, 43(4), 435–​447. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​12030590/​
Pajer, K. A. (1998). What happens to “bad” girls? A review of the adult outcomes of antisocial adolescent
girls. The American Journal of Psychiatry, 155(7), 862–​870. Retrieved from http://​www.ncbi.nlm.nih.
gov/​pubmed/​9659848/​
0
5
1

150 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Pajer, K. A., Stein, S., Tritt, K., Chang, C.-​N., Wang, W., & Gardner, W. (2008). Conduct disorder in
girls: neighborhoods, family characteristics, and parenting behaviors. Child and Adolescent Psychiatry
and Mental Health, 2(1), 28. doi:10.1186/​1753-​2000-​2-​28
Pardini, D. A. (2011). Perceptions of social conflicts among incarcerated adolescents with callous-​
unemotional traits: “you”re going to pay. It’s going to hurt, but I don’t care.’ Journal of Child Psychology
and Psychiatry, and Allied Disciplines, 52(3), 248–​255. doi:10.1111/​j.1469-​7610.2010.02336.x
Pardini, D. A., & Byrd, A. L. (2012). Perceptions of aggressive conflicts and others’ distress in
children with callous-​unemotional traits: “I”ll show you who’s boss, even if you suffer and
I get in trouble’. Journal of Child Psychology and Psychiatry, 53(3), 283–​291. doi:10.1111/​
j.1469-​7610.2011.02487.x
Pardini, D. A., & Fite, P. J. (2010). Symptoms of conduct disorder, oppositional defiant disorder, attention-​
deficit/​hyperactivity disorder, and callous-​unemotional traits as unique predictors of psychosocial
maladjustment in boys: advancing an evidence base for DSM-​V. Journal of the American Academy of
Child and Adolescent Psychiatry, 49(11), 1134–​1144. doi:10.1016/​j.jaac.2010.07.010
Pardini, D. A., & Frick, P. J. (2013). Multiple developmental pathways to conduct disorder: current
conceptualizations and clinical implications. Journal of the Canadian Academy of Child and
Adolescent Psychiatry = Journal de l’Académie Canadienne de Psychiatrie de L’enfant et de L’adolescent,
22(1), 20–​25. Retrieved from http://​www.pubmedcentral.nih.gov/​articlerender.fcgi?artid=
3565711&tool=pmcentrez&rendertype=abstract
Pardini, D. A., Raine, A., Erickson, K., & Loeber, R. (2014). Lower amygdala volume in men is associated
with childhood aggression, early psychopathic traits, and future violence. Biological Psychiatry, 75(1),
73–​80. doi:10.1016/​j.biopsych.2013.04.003
Pechorro, P., Jiménez, L., Hidalgo, V., & Nunes, C. (2015). The DSM-​5 Limited Prosocial Emotions subtype
of Conduct Disorder in incarcerated male and female juvenile delinquents. International Journal of Law
and Psychiatry, 39, 77–​82. doi:10.1016/​j.ijlp.2015.01.024
Perou, R., Bitsko, R. H., Blumberg, S. J., Pastor, P., Ghandour, R. M., Gfroerer, J. C., … Crosby, A. E.
(2013). Mental health surveillance among children—​United States, 2005–​2011. Morbidity and Mortality
Weekly Report, 62(02), 1–​35.
Posthumus, J. A., Raaijmakers, M. A. J., Maassen, G. H., van Engeland, H., & Matthys, W. (2012).
Sustained effects of incredible years as a preventive intervention in preschool children with conduct
problems. Journal of Abnormal Child Psychology, 40(4), 487–​500. doi:10.1007/​s10802-​011-​9580-​9
Presnall, N., Webster-​Stratton, C., & Constantino, J. N. (2014). Parent training: equivalent
improvement in externalizing behavior for children with and without familial risk. Journal of the
American Academy of Child and Adolescent Psychiatry, 53(8), 879–​887, 887.e1–​e2. doi:10.1016/​
j.jaac.2014.04.024
Raine, A., Fung, A. L. C., Portnoy, J., Choy, O., & Spring, V. L. (2014). Low heart rate as a risk factor for
child and adolescent proactive aggressive and impulsive psychopathic behavior. Aggressive Behavior, 49,
343–​369. doi:10.1002/​ab.21523
Rothbart, M. K., & Bates, J. E. (1998). Temperament. In W. Damon, R. Lerner, & N. Eisenberg (Eds.),
Handbook of Child Psychology (5th ed., pp. 37–​86). New York: Wiley.
Rothbart, M. K., Posner, M. I., & Hershey, K. (1995). emperament, attention and developmental
psychopathology. In D. Cicchetti & J. D. Cohen (Eds.), Manual of developmental psychopathology (pp.
315–​340). New York: Wiley.
Rowe, R., Maughan, B., Moran, P., Ford, T., Briskman, J., & Goodman, R. (2010). The role of callous and
unemotional traits in the diagnosis of conduct disorder. Journal of Child Psychology and Psychiatry, and
Allied Disciplines, 51(6), 688–​695. doi:10.1111/​j.1469-​7610.2009.02199.x
Rudolph, K. D., & Clark, A. G. (2001). Conceptions of relationships in children with depressive and
aggressive symptoms: social-​cognitive distortion or reality? Journal of Abnormal Child Psychology,
29(1), 41–​56. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​11316334
1
5

Conclusion 151

Salmivalli, C., Ojanen, T., Haanpää, J., & Peets, K. (2005). “I’m OK but you’re not” and other peer-​
relational schemas: explaining individual differences in children’s social goals. Developmental
Psychology, 41(2), 363–​375. doi:10.1037/​0012-​1649.41.2.363
Schippell, P. L., Vasey, M. W., Cravens-​Brown, L. M., & Bretveld, R. A. (2003). Suppressed attention to
rejection, ridicule, and failure cues: a unique correlate of reactive but not proactive aggression in youth.
Journal of Clinical Child and Adolescent Psychology : The Official Journal for the Society of Clinical Child
and Adolescent Psychology, American Psychological Association, Division 53, 32(1), 40–​55. doi:10.1207/​
S15374424JCCP3201_​05
Schwartz, D., Dodge, K. A., Coie, J. D., Hubbard, J. A., Cillessen, A. H., Lemerise, E. A., & Bateman, H.
(1998). Social-​cognitive and behavioral correlates of aggression and victimization in boys’ play groups.
Journal of Abnormal Child Psychology, 26(6), 431–​440. Retrieved from http://​www.ncbi.nlm.nih.gov/​
pubmed/​9915650
Sebastian, C. L., De Brito, S. A., McCrory, E. J., Hyde, Z. H., Lockwood, P. L., Cecil, C. A. M., & Viding,
E. (2015). Grey Matter Volumes in Children with Conduct Problems and Varying Levels of Callous-​
Unemotional Traits. Journal of Abnormal Child Psychology. doi:10.1007/​s10802-​015-​0073-​0
Sexton, T., & Alexander, J. (2004). Functional Family Therapy: Principles of clinical intervention, assessment
and implementation. Retrieved from http://​www.functionalfamilytherapy.com/​wp-​content/​uploads/​
FFT-​Clinical-​Manual-​Blue-​Blook-​8.1.08.pdf
Silberg, J. L., Maes, H., & Eaves, L. J. (2012). Unraveling the effect of genes and environment in
the transmission of parental antisocial behavior to children’s conduct disturbance, depression
and hyperactivity. Journal of Child Psychology and Psychiatry, 53(6), 668–​677. doi:10.1111/​
j.1469-​7610.2011.02494.x
Spain, S. E., Douglas, K. S., Poythress, N. G., & Epstein, M. (2004). The relationship between psychopathic
features, violence and treatment outcome: the comparison of three youth measures of psychopathic
features. Behavioral Sciences & the Law, 22(1), 85–​102. doi:10.1002/​bsl.576
Stickle, T. R., Kirkpatrick, N. M., & Brush, L. N. (2009). Callous-​unemotional traits and social information
processing: multiple risk-​factor models for understanding aggressive behavior in antisocial youth. Law
and Human Behavior, 33(6), 515–​529. doi:10.1007/​s10979-​008-​9171-​7
Stringaris, A. (2011). Irritability in children and adolescents: a challenge for DSM-​5. European Child &
Adolescent Psychiatry, 20(2), 61–​66. doi:10.1007/​s00787-​010-​0150-​4
Teten Tharp, A. L., Sharp, C., Stanford, M. S., Lake, S. L., Raine, A., & Kent, T. A. (2011). Correspondence
of aggressive behavior classifications among young adults using the Impulsive Premeditated Aggression
Scale and the Reactive Proactive Questionnaire. Personality and Individual Differences, 50(2), 279–​285.
doi:10.1016/​j.paid.2010.10.003
Thompson, R. A. (1994). Emotion Regulation: A Theme In Search of Definition. Monographs of the Society
for Research in Child Development, 59(2-​3), 25–​52. doi:10.1111/​j.1540-​5834.1994.tb01276.x
Thompson, R. A. (2006). The development of the person: Social understanding, relationships, self,
conscience. In W. Damon & R. M. Lerner (Eds.), Handbook of child psychology (6th ed., pp. 24–​98).
New York: Wiley.
Thompson, R. A., & Calkins, S. D. (1996). The double-​edged sword: Emotional regulation for children at
risk. Development and Psychopathology, 8(01), 163. doi:10.1017/​S0954579400007021
Timmons-​Mitchell, J., Bender, M. B., Kishna, M. A., & Mitchell, C. C. (2006). An independent
effectiveness trial of multisystemic therapy with juvenile justice youth. Journal of Clinical Child
and Adolescent Psychology : The Official Journal for the Society of Clinical Child and Adolescent
Psychology, American Psychological Association, Division 53, 35(2), 227–​236. doi:10.1207/​
s15374424jccp3502_​6
Viding, E., Blair, R. J. R., Moffitt, T. E., & Plomin, R. (2005). Evidence for substantial genetic risk for
psychopathy in 7-​year-​olds. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 46(6),
592–​597. doi:10.1111/​j.1469-​7610.2004.00393.x
2
5
1

152 Emotion Regulation and Conduct Disorder: The Role of Callous-Unemotional Traits

Viding, E., Jones, A. P., Frick, P. J., Moffitt, T. E., & Plomin, R. (2008). Heritability of antisocial behaviour
at 9: do callous-​unemotional traits matter? Developmental Science, 11(1), 17–​22. doi:10.1111/​
j.1467-​7687.2007.00648.x
Viding, E., Sebastian, C. L., Dadds, M. R., Lockwood, P. L., Cecil, C. A. M., De Brito, S. A., & McCrory, E.
J. (2012). Amygdala response to preattentive masked fear in children with conduct problems: the role of
callous-​unemotional traits. The American Journal of Psychiatry, 169(10), 1109–​1116. doi:10.1176/​appi.
ajp.2012.12020191
Vitaro, F., Brendgen, M., & Tremblay, R. E. (2002). Reactively and proactively aggressive
children: antecedent and subsequent characteristics. Journal of Child Psychology and Psychiatry, and
Allied Disciplines, 43(4), 495–​505. Retrieved from http://​www.ncbi.nlm.nih.gov/​pubmed/​12030595
Washburn, J. J., Romero, E. G., Welty, L. J., Abram, K. M., Teplin, L. A., McClelland, G. M., & Paskar,
L. D. (2007). Development of antisocial personality disorder in detained youths: the predictive value
of mental disorders. Journal of Consulting and Clinical Psychology, 75(2), 221–​231. doi:10.1037/​
0022-​006X.75.2.221
Webster-​Stratton, C. (1984). Randomized trial of two parent-​training programs for families with conduct-​
disordered children. Journal of Consulting and Clinical Psychology, 52(4), 666–​678. Retrieved from
https://​pdfs.semanticscholar.org/​9b55/​1ca57240cfe773214eadf62da7da1b1cd96e.pdf
Webster-​Stratton, C. (2011). The Incredible Years Parents, Teachers, and Children’s Training series. Seattle,
WA: Incredible Years, Inc.
Webster-​Stratton, C. (2016). The Incredible Years Parent Programs. In J. J. Ponzetti (Ed.), Evidence-​based
Parenting Education: A Global Perspective (pp. 143–​160). New York: Taylor & Francis Group.
Webster-​Stratton, C., Gaspar, M. F., & Seabra-​Santos, M. J. (2012). Incredible Years® Parent, Teachers and
Children’s Series: Transportability to Portugal of Early Intervention Programs for Preventing Conduct
Problems and Promoting Social and Emotional Competence. Psychosocial Intervention, 21(2), 157–​169.
doi:10.5093/​in2012a15
Webster-​Stratton, C., Jamila Reid, M., & Stoolmiller, M. (2008). Preventing conduct problems and
improving school readiness: evaluation of the Incredible Years Teacher and Child Training Programs
in high-​risk schools. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 49(5), 471–​488.
doi:10.1111/​j.1469-​7610.2007.01861.x
Webster-​Stratton, C., Reid, M. J., & Hammond, M. (2004). Treating children with early-​onset conduct
problems: intervention outcomes for parent, child, and teacher training. Journal of Clinical Child and
Adolescent Psychology : The Official Journal for the Society of Clinical Child and Adolescent Psychology,
American Psychological Association, Division 53, 33(1), 105–​124. doi:10.1207/​S15374424JCCP3301_​11
Welsh, B. C., Loeber, R., Stevens, B. R., Stouthamer-​Loeber, M., Cohen, M. A., & Farrington, D. P. (2008).
Costs of Juvenile Crime in Urban Areas: A Longitudinal Perspective. Youth Violence and Juvenile Justice,
6(1), 3–​27. doi:10.1177/​1541204007308427
Whelan, Y. M., Stringaris, A., Maughan, B., & Barker, E. D. (2013). Developmental continuity of
oppositional defiant disorder subdimensions at ages 8, 10, and 13 years and their distinct psychiatric
outcomes at age 16 years. Journal of the American Academy of Child and Adolescent Psychiatry, 52(9),
961–​969. doi:10.1016/​j.jaac.2013.06.013
White, S. F., Frick, P. J., Lawing, K., & Bauer, D. (2013). Callous-​unemotional traits and response to
functional family therapy in adolescent offenders. Behavioral Sciences & the Law, 31(2), 271–​285.
doi:10.1002/​bsl.2041
Whitson, S. M., & El-​Sheikh, M. (2003). Moderators of Family Conflict and Children’s Adjustment and
Health. Journal of Emotional Abuse, 3(1-​2), 47–​73. doi:10.1300/​J135v03n01_​03
Wolf, R. C., Pujara, M. S., Motzkin, J. C., Newman, J. P., Kiehl, K. A., Decety, J., … Koenigs, M. (2015).
Interpersonal traits of psychopathy linked to reduced integrity of the uncinate fasciculus. Human Brain
Mapping. doi:10.1002/​hbm.22911
3
5
1

Conclusion 153

Wolf, S., & Centifanti, L. C. M. (2014). Recognition of Pain as Another Deficit in Young Males with High
Callous-​Unemotional Traits. Child Psychiatry & Human Development, 45(4), 422–​432. doi:10.1007/​
s10578-​013-​0412-​8
Wood, C. N., & Gross, A. M. (2002). Behavioral Response Generation and Selection of Rejected-​Reactive
Aggressive, Rejected-​Nonaggressive, and Average Status Children. Child & Family Behavior Therapy,
24(3), 1–​19. Retrieved from http://​eric.ed.gov/​?id=EJ657151
Zhang, W., & Gao, Y. (2015). Interactive effects of social adversity and respiratory sinus arrhythmia activity
on reactive and proactive aggression. Psychophysiology, 52(10), 1343–​1350. doi:10.1111/​psyp.12473
4
5
1

Chapter 8

Emotion Regulation
and Anxiety: Developmental
Psychopathology and Treatment
Dagmar Kr. Hannesdóttir & Thomas H. Ollendick

Anxiety
Some years ago, Barlow (1991) put forth the notion that anxiety was a disorder of emotion and
was characterized by problems in regulating those emotions. In doing so, he noted that emotions
themselves were not maladaptive or problematic in and of themselves but rather that the tim-
ing and intensity of these emotions could be problematic. Furthermore, it has long been known
that individuals differ in the ways in which they appraise their emotions (Gross & John, 1995;
1998). Subsequently, these appraisals contribute to whether emotions are perceived as aversive or
nonaversive and whether the person attempts to avoid, escape or embrace them. If the emotion is
perceived as aversive and undesirable, the individual is more likely to attempt to regulate the emo-
tion than if the emotion is viewed as pleasant and desirable. Unfortunately, all attempts to regulate
aversive emotions are not effective—╉some attempts, in fact, lead to undesirable effects and their
exacerbation. Given these relations, it is no surprise that emotion regulation and its associated
deficiencies are intimately associated with the anxiety disorders and have been examined exten-
sively over the years (e.g., Amstadter, 2008; Davidson, 1998; Kring & Werner, 2004).
In this chapter, we will first briefly review the anxiety disorders of childhood and adolescence,
examining the role of emotion regulation in the onset, maintenance and expression of these dis-
orders, and then highlighting evidence-╉based interventions for these disorders that incorporate
emotion and its regulation. In doing so, we hope to illustrate the complexity of these disorders and
to illustrate the promise of emotion-╉based interventions.

Anxiety disorders of childhood and adolescence


Anxiety Disorders are among the most commonly experienced and diagnosed conditions of
childhood (see below, and Grills-╉Taquechel & Ollendick, 2012). Since the most recent Diagnostic
and Statistical Manual of Mental Disorders (DSM-╉5) has recently been published (APA, 2013),
much of the work examining the anxiety disorders has used categorizations from DSM-╉IV-╉TR
(APA, 2000). Importantly, by and large, the core anxiety disorder descriptors have remained
relatively unchanged across these editions. Anxiety Disorders, however, now include two disor-
ders that were previously listed under “Disorders usually first diagnosed in infancy, childhood,
or adolescence” (i.e., Separation Anxiety Disorder, Selective Mutism), separates Panic Disorder
and Agoraphobia into separate diagnoses, and maintains inclusion of Specific Phobia, Social
Anxiety Disorder, and Generalized Anxiety Disorder. In addition, criteria for separate Substance/╉
Medication Induced Disorders and diagnoses “Due to Another Medical Condition” (e.g., Anxiety
51

Anxiety disorders of childhood and adolescence 155

Disorder Due to Another Medical Condition) are included. Obsessive-​Compulsive Disorder and
Posttraumatic Stress Disorder, previously included in the Anxiety Disorders, are now placed in
separate diagnostic categories, Obsessive-​Compulsive and Related Disorders and Trauma-​and
Stressor-​Related Disorders, respectively.
Based on the DSM-​IV-​TR (2000) criteria (as well as those put forth in DSM-​5) the anxiety
disorders diagnosable in childhood and adolescence have several features in common, includ-
ing: 1) Persistent and excessive anxious arousal, and 2) symptoms that cause clinically significant
distress or impairment in social, academic, and other important areas of functioning. The dif-
ferent disorders vary primarily according to the stimuli eliciting anxiety in these disorders. In
addition, although anxiety disorders are among the most commonly diagnosed disorders of child-
hood, prevalence rates vary by disorder. Each of the disorders is briefly described.
Significant anxiety regarding separation from home or individuals to whom the child is attached
is the hallmark of Separation Anxiety Disorder (SAD). Children must show at least three of eight
symptoms, with onset of these symptoms usually before age 18  years, and they must have the
symptoms for at least four weeks to receive a diagnosis of SAD. Associated features include: per-
sistent reluctance to attend school, remain alone, or go to sleep without a major attachment figure
nearby, as well as nightmares involving the theme of separation and the presence of a number of
physical complaints when separation occurs or is anticipated. The prevalence of SAD is reported
to be between 1–​4%, with rates decreasing as children get older (Brückl et al., 2007; Canino et al.,
2004; Egger & Angold, 2006; Merikangas, He, Burstein, et al., 2010).
Social Anxiety Disorder (SOC), previously referred to as Social Phobia, is characterized by
excessive and persistent (typically lasting for six months or more) fear and avoidance of social
situations or situations where scrutiny could lead to embarrassment. Children with SOC may not
be aware that their fears are unreasonable and/​or excessive and may express their distress through
crying, tantrums, freezing, clinging, or shrinking from social situations with unfamiliar people.
The feared stimuli (i.e., social situations) with SOC are typically avoided or endured with intense
distress that may take the form of a panic attack in some cases. To be diagnosed with SOC, a child
needs to demonstrate age-​appropriate social relationships with familiar people and to display
avoidance in interactions involving peers as well as adults. Prevalence rates for SOC are typically
reported between 1–​3% but also vary by age with increasing rates seen in adolescents (Canino
et al., 2004; Egger & Angold, 2006; Essau, Conradt, & Petermann, 2000; Roberts, Roberts, & Xing,
2007; Wittchen, Nelson, & Lachner, 1998).
Often considered similar to SOC in its focus on socialization and interpersonal relationships
(Muris & Ollendick, 2015), Selective Mutism (SM) is diagnosed when a child refuses to speak in
specific social situations (e.g., school, community, clinic) despite the ability to do so. Such refusal
to speak must occur for at least one month with interference occurring in educational/​occupa-
tional, achievement or social communication domains. However, SM is not diagnosed when
symptoms occur only within the first month of school or because of language/​communication
issues. Associated features include social concerns, shyness, or other anxiety symptoms and the
prevalence of SM is thought to be quite small (i.e., <1%; Egger & Angold, 2006).
Once referred to as Overanxious Disorder, Generalized Anxiety Disorder (GAD) is character-
ized by excessive anxiety and worry about several different domains in the child’s life. The worry
experienced by a child with GAD is often reported to be uncontrollable and occurs more days
than not, typically for the past six months. Children must also exhibit at least one of the following
six physical/​somatic symptoms: 1) Restlessness or feeling keyed up or on edge; 2) easily fatigued;
3)  difficulty concentrating or mind going blank; 4)  irritability; 5)  muscle tension; and 6)  sleep
disturbance. The worries reported by children with GAD are similar to the worries of children
without GAD and vary primarily in terms of their frequency, intensity and duration. Some of the
more common worries reported by children concern evaluation by others, perfectionism, health
6
5
1

156 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

of significant others, and catastrophic events. Prevalence rates have been reported to vary widely
1–​4% (Canino et al., 2004; Egger & Angold, 2006; Lavigne, LeBailly, Hopkins, Gouze, & Binns,
2009; Merikangas, He, Burstein, et al., 2010; Wittchen, Zhao, Kessler, & Eaton, 1994).
Specific Phobias (SP) are excessive and persistent (typically lasting 6 months or longer) fears
of explicit objects or situations, which are typically avoided or endured with intense anxiety or
distress. Exposure or anticipation of exposure to the feared stimulus results in extreme anxiety
including panic attacks in some cases. Children may not be cognizant of the unreasonable or
excessive nature of their fears and may express their fear by crying, throwing a tantrum, freez-
ing, or clinging. At a clinical level, phobias tend to be involuntary, inappropriate, and limiting
to a child’s quality of life (Anderson, 1994; Essau et  al., 2000). Specific phobias can be speci-
fied as falling into one of the following subtypes:  Animal (e.g., snakes, spiders, dogs), Natural
Environment (e.g., storms, heights, water), Blood-​Injection-​Injury (e.g., seeing blood, getting an
injection, receiving an injury), Situational (e.g., tunnels, flying, enclosed places), or Other (e.g.,
choking, loud sounds, costumed characters). Prevalence rates vary by child age and gender but are
typically reported to range from 2–​6% (Burstein et al., 2012; Egger & Angold, 2006; Essau et al.,
2000; Wittchen et al., 1998).
Panic Disorder (PD) and Agoraphobia (AG) are each diagnosed separately in DSM-​5.
The hallmark symptom of panic disorder is the recurrence of panic attacks, which are acute
and extreme feelings of anxiety that occur unexpectedly and are followed by one month or
more of persistent concern about having another attack, worry about the consequences of
the attack, or a change in behavior related to the attack. Agoraphobia is characterized by
excessive anxiety resulting from situations in which escape or avoidance may be inhibited
or in which help may not be available if panic symptoms were to occur. Panic Disorder and
Agoraphobia are among the less commonly diagnosed anxiety disorders during childhood,
with prevalence rates of about 1–​3% increasing into adolescence and adulthood (Canino
et al., 2004; Doerfler, Connor, Volungis, & Toscano, 2007; Essau et al., 2000; Merikangas, He,
Brody et al., 2010; Ollendick, Birmaher, & Mattis, 2004; Roberts et al., 2007; Wells et al., 2006;
Wittchen et al., 1998).
Although the stimuli that elicit anxiety and fear in these disorders differ, the disorders all share
commonalities in the expression of anxiety. Fundamentally, anxiety is an emotional state asso-
ciated with heightened physiological arousal and behavioral avoidance that is triggered by the
perception of real or imagined threat. At times, these perceptions can be distorted and the threat
is exaggerated well beyond the real threat imposed by the stimulus. As Barlow (1991) noted, how-
ever, this emotion can be adaptive and can lead to constructive attempts to handle the threat and
to prepare the organism for adaptive change and growth; however, at other times, the emotion is
intense, frequent, and durable as in an anxiety or phobic disorder and it becomes less adaptive
and leads to a host of problems, including academic, behavioral, and social difficulties (Grills-​
Taquechel & Ollendick, 2012).
The etiology of the Anxiety Disorders is complex and not straightforward. Equifinality (i.e.,
multiple pathways to any one outcome; Cicchetti & Toth, 1991)  is the most succinct way to
describe the etiology of anxiety and its disorders in children. Indeed, fears and anxieties in chil-
dren have been described as multiply determined if not over-​determined (Marks, 1987; Ollendick,
1979; Weems & Stickle, 2005). Contemporary etiological models reflect this heterogeneity with
consideration of various influences that cut across personal-​social-​ecological systems and typi-
cally include biological, developmental, psychological, social, and environmental components
(e.g., Grills-​Taquechel & Ollendick, 2012; Hirshfeld-​Becker, Micco, Simoes, & Henin, 2008;
Vasey & Dadds, 2001). Biological contributions have been well documented in familial, twin,
and genetic research studies, as have various connections with early developmental behaviors
7
5
1

Avoidance, Cognitions and Physical Symptoms 157

(e.g., temperament, attachment) and child dispositional characteristics (e.g., anxiety sensitiv-
ity, cognitive biases, and emotion regulation difficulties). Family influences, beyond genetics,
have also received a great deal of research attention in the past several decades. Research in this
domain has generally concentrated on anxious parenting behaviors and child-╉rearing practices.
For example, several investigators have drawn upon Rachman’s (1977) influential work denoting
three common pathways for fear acquisition and suggested parenting practices that may result in
heightened child anxiety. Indeed, using a variety of paradigms (e.g., cross-╉sectional, longitudinal,
observational, experimental), researchers have demonstrated the influences of parental model-
ing of anxious behaviors, conveying anxiety-╉provoking information, and reinforcing anxious
behaviors displayed by their children (cf, Beidel & Turner, 1998; Fisak & Grills-╉Taquechel, 2007;
Grills-╉Taquechel & Ollendick, 2012; Muris, van Zwol, Huijding, & Mayer, 2010), as well as for
parent-╉rearing behaviors characterized by rejection, control, and overprotection (cf., DiBartolo &
Helt, 2007; Ollendick & Benoit, 2012; Rapee, 1997).
Importantly, however, there is no single or direct cause of anxiety disorders in children and ado-
lescents. As noted by Thompson (2001, p. 160), “the action is in the interaction” among multiple
internal and external influences that contribute to their onset and expression. For example, tem-
peramental vulnerability as shown in behavioral inhibition may not determine alone the develop-
ment of an anxiety disorder but in combination with other influences it may play an important
role (Ollendick & Benoit, 2012). Temperamental inhibition may serve to sensitize young children
to anxiety-╉producing stimuli in a manner not seen in children who are not behaviorally inhibited.
So, too, anxiety sensitivity, emotion regulation difficulties, and parenting practices may function
in similar ways—╉they can heighten the aversive response to the feared stimuli and contribute to
an accumulation of risk over time (Thompson, 2001).
Such a conceptualization of how children develop anxiety disorders sheds light on how children
with a biological disposition to experience anxious feelings easily and who avoid situations that
elicit intense emotions might develop anxiety disorders through a dynamic interplay between
parental reactions, feedback from the environment and their own biased cognitions and expecta-
tions of how the world works for them. However, as noted, the process is not a straightforward
one; difficulties with emotion regulation, like other risk factors, can set the stage for the onset of
anxiety disorders but do not directly cause them.
In the next sections, research is reviewed on emotion regulation and the role it plays in the
development and maintenance of anxiety disorders in children and adolescents and how emerg-
ing knowledge on emotion regulation might improve and increase the effectiveness of cognitive-╉
behavioral treatment (CBT) for anxiety. Recent findings on emotion knowledge and emotion
regulation strategies among children with anxiety are reviewed and discussed in light of new
trends in treatment development, such as emotion-╉focused CBT (e.g., Suveg, Kendall, Comer, &
Robin, 2006), so that current treatment programs can become more effective for a larger group of
children. In addition, the role of parents in modeling emotion regulation skills for anxious youths
are discussed in terms of developmental psychopathology and the implications for treatment pro-
grams currently in use are evaluated.

Avoidance, Cognitions and Physical Symptoms


As noted, in the past ten to 15  years, an increased focus has been placed on examining the
role of emotion knowledge, emotion understanding and emotion regulation skills in developing
and maintaining anxiety disorders among children (Hannesdottir & Ollendick, 2007; Southam-╉
Gerow & Kendall, 2002; Suveg & Zeman, 2004; Thompson, 2001). Research has been aimed
at examining negative emotions among children with internalizing problems (anxiety and/╉or
8
5
1

158 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

depression) in general and how they are perceived and handled, such as anxious feelings, sad-
ness, and anger alike. Collective evidence suggests that children with internalizing disorders
seem to have problems managing negative emotions in general (Trosper, Buzzella, Bennett, &
Ehrenreich, 2009), not just emotions relating to their specific problems and the nature of their
disorder.

Cognitive aspects of emotion regulation and anxiety


Initial studies on emotion understanding and regulation in developmental and developmental
psychopathology research revealed that children’s use of planned and goal directed strategies to
control their emotions increased as their ability for effortful control and executive function capac-
ity was realized (Eisenberg & Morris, 2002). The normal course of emotion development suggests
that children’s emotion regulation skills are dependent on knowledge of how emotions work and
what is appropriate to express (Kail, 2007), what emotion regulation strategies are most effective
in certain situations (Southam-╉Gerow & Kendall, 2002), and the modeling of emotion regulation
competence by parents (Eisenberg, 1998).
With regard to children with anxiety, various studies have shown that they tend to interpret
ambiguous situations as more threatening than they actually are (Chorpita, Albano & Barlow,
1996)  and that they are hypervigilant with regards to information relating to potential threats
in their environment (Vasey & MacLeod, 2001). In addition, children with anxiety seem to have
more limited knowledge compared to non-╉anxious children in terms of when it is appropriate
to “hide” their emotions and that it is possible to actually change and control their emotions
(Southam-╉Gerow & Kendall, 2000).
Children with anxiety also seem to experience negative emotions more intensely than non-╉
anxious children. For example, in one study children were asked to engage in reappraisal to
decrease negative emotions after seeing mildly disturbing pictures. Children diagnosed with
anxiety reported experiencing negative emotions in response to such scenarios more intensely
than the non-╉anxious group and had a more difficult time reappraising the situation to improve
their mood (Carthy, Horesh, Apter, Edge, & Gross, 2010). In another study conducted by the
same research group (Carthy, Horesh, Apter, & Gross, 2010), it was shown that anxious youths
not only had difficulties with reappraisal strategies but also engaged in more avoidant (e.g.,
I will ask to go to the bathroom in school as so to miss my turn to talk about my project; I will
pretend my leg hurts so I  cannot go outside where there might be bees or wasps) and inap-
propriate support-╉seeking strategies (e.g., I will ask my dad to do it for me; I will ask my mom
to come and lie down with me to fall asleep). These authors also showed low levels of self-╉
efficacy relating to regulating emotions in these anxious youths. Similar findings have emerged
in other studies where children with anxiety have been shown to have difficulty managing nega-
tive emotions in general (worry, sadness and anger), experience such negative emotions more
strongly, and have reduced self-╉efficacy in their ability to change their emotional state (Suveg
& Zeman, 2004).
Thus, with regard to emotion regulation, children with anxiety tend to engage in select strate-
gies to calm themselves down when faced with emotionally difficult situations. As noted, they
tend to use ineffective emotion regulation strategies for the given situation. They tend to rely too
much on others to help them reduce negative feelings and are less likely to use problem-╉solving
methods in anxiety provoking situations. In many respects, they select emotion regulations strate-
gies that work to reduce negative feelings in the short run but are detrimental in the long run (e.g.,
avoidance or escape; Carthy et al. 2010; Trosper et al., 2009). Thompson and Calkins (1996) refer
to these strategies as “two-╉edged;” they reduce anxiety in the short term but lead to more anxiety
and impairment in the long term.
9
5
1

Avoidance, Cognitions and Physical Symptoms 159

Behavioral factors and emotion regulation skills


Not only do anxious children tend to view negative emotional experiences differently than non-╉
anxious children and have difficulty reappraising emotionally provoking situations, they also
engage in various strategies that help them avoid experiencing these negative situations. As noted,
such strategies can be efficient in terms of regulating negative emotions in the short run but serve
to maintain the anxiety disorder, making it necessary for the child to escape the anxiety provok-
ing situation each time it is encountered. Behavioral emotion regulation strategies (i.e., strategies
intended to prevent or modify negative emotional experiences) are described by Gross (1998) as
situation selection and situation modification strategies. An example of a situation selection emo-
tion regulation strategy for a child with a specific phobia of heights would be that the child might
feign illness on the day of a school trip to an amusement park to avoid riding on a roller coaster
with her friends. Similarly, a child who is socially anxious might avoid social interactions with a
group of peers by insisting that she wants to stay home and finish a book or work on her stamp
collection. Anxious children also try to modify situations in an effort to regulate their emotions.
Examples of such efforts might be the use of safety signals and maintaining routines and rituals.
For example, a child with fears of being alone and separation anxiety might agree to stay in the car
while her father runs into the house to get something only if her one-╉year old brother is also in the
car or if she can stay on the phone with her father constantly while he goes inside. Children with
anxiety usually have a wide range of avoidance and escape strategies that serve to shelter them
from experiencing negative emotions (fear, worry, embarrassment, etc.). While such strategies are
useful for short-╉term goals such as making the child feel better in the situation, these strategies
prevent the child from learning that they can handle this particular situation and that the feared
consequences usually do not occur. Thus, providing anxious children with different behavioral
strategies in treatment and having them practice them in session (e.g., through exposure) and at
home can prove essential for changing the course of their anxiety and the impairment their disor-
der causes on a daily basis due to avoidance.

Physiological factors of emotion regulation and anxiety


In various studies, children with anxiety have reported that they experience negative emotions,
such as anxiety, more intensely than other children (e.g., Carthy et  al., 2010; Suveg & Zeman,
2004). To complicate the picture of whether children with anxiety actually have a lower threshold
in their sympathetic nervous system for experiencing anxiety and negative emotions, anxious
children also report that they are hypersensitive to bodily cues that they believe signal negative
emotions (Thompson, 2001). They also tend to have catastrophic beliefs and lower tolerance
regarding experiencing physiological symptoms of anxiety, that is, high anxiety sensitivity (Reiss,
Silverman, & Weems, 2001). The question that needs to be answered is therefore, whether it has
been shown that children with anxious dispositions or temperament are likely to be more physi-
ologically reactive in stressful situations and whether they have a harder time regulating their
emotions and calming down.
Indeed, various studies have shown that this pattern emerges both for very young and somewhat
older children. For instance, infants and toddlers who show withdrawal behaviors, fearfulness
and distress in new and unfamiliar situations and are considered to have a behaviorally inhib-
ited temperament have been shown to demonstrate greater right frontal EEG symmetry patterns
opposed to more outgoing children (e.g., Fox, Bell & Jones, 1992; Fox, Henderson, Rubin, Calkins,
& Schmidt, 2001). Behaviorally inhibited temperament, especially when stable throughout early
childhood, has been shown to increase the chances of children developing anxiety disorders later
on in childhood (Hirshfeld et al., 1992; Ollendick & Hirshfeld-╉Becker, 2002).
0
6
1

160 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

A few studies have also shown that this pattern of increased reactivity in infancy and early
childhood is predictive of physiological reactivity and emotion regulation difficulties in later
childhood. For instance, McManis and colleagues demonstrated a link between high reactiv-
ity to stress in infancy and right frontal asymmetry activation patterns among these children at
the age of ten to 12 years while performing a stressful speech task (McManis, Kagan, Snidman,
& Woodward, 2002). Similar findings were obtained in another study where right frontal EEG
activation patterns at approximately four years of age predicted increased heart rate and slower
cardiovascular recovery after a stressful speech task at age nine in a small sample of non-╉anxious
children (Hannesdottir, Doxie, Bell, Ollendick, & Wolfe, 2010). Concurrent studies on anxious
youths have also shown this pattern of cortical activation patterns emerging when anxious chil-
dren are faced with stressful situations. Hum and colleagues found that children with anxiety
showed heightened attention and arousal in response to a task with various emotion faces (both
positive and negative) while the non-╉anxious control group showed differential patterns of activa-
tion dependent on whether the faces were angry, sad, or happy (Hum, Manassis, & Lewis, 2013).
These results indicated that children with anxiety experienced heightened reactivity and were
engaged in a process of emotion regulation simply by being shown faces of people; irrespective of
what emotions those faces were showing.
Overall, the studies that have examined physiological reactivity, cerebral activation patterns and
emotion regulation among children with anxiety support the notion that these children experi-
ence heightened arousal easily and have a lower sympathetic activation threshold compared to
less anxious children (cf., Beauchaine, 2015). In addition, these children may need to put more
effort into regulating such intense emotions and into recovering from them, while not being able
to allocate their cognitive resources elsewhere at the same time. This may in turn limit their abil-
ity to perform well in a particular situation (e.g., a socially anxious child speaking in front of
the class) or to think rationally to reduce catastrophic beliefs in the situation (e.g., a separation
anxious child convincing herself that her mother is simply running late and has not been in a car
accident). Thus, strong physiological reactivity to anxiety provoking situations and considerable
emotion regulation efforts in such situations may play a large role in maintaining anxiety disor-
ders among children and reinforce avoidance and escape behaviors.

Parental reactions and child anxiety


Emotion regulation skills in children do not develop in a vacuum. From the day they are born
infants seek out comfort from their caregiver when distressed and through the years learn how
to become more and more independent in terms of understanding and managing positive and
negative emotions in various situations through socialization by their caregivers, teachers and
peers (Dunsmore & Halberstadt, 1997; Eisenberg, 1998; Halberstadt, Denham, Dunsmore, 2001;
Thompson, 2001).
It is also important to keep in mind that from early on caregivers often choose which situa-
tions their children experience (e.g., situation selection, see Gross, 1998). Therefore, when parents
know that their child fears certain situations and believe that their child would not be able to cope
with experiencing fear, worry, sadness or some other negative emotions in particular situations,
parents are likely to help their child avoid or escape such situations. Such situation selection or
assisted avoidance will in turn increase the child’s anxiety regarding the feared object or situation
and deprives the child of the opportunity to obtain new information on the feared object and to
correct catastrophic beliefs. Another reason for encouraged avoidance might be that the parent,
who might be anxious himself, believes he will not be able to cope with seeing his child experience
distress and therefore helps the child avoid these situations all together. This works well for the
1
6

Parental reactions and child anxiety 161

parent while the child is still relatively young (about birth to five years of age) and the parent is
able to control and select situations for the child. However, when children become more indepen-
dent and enter the school system the parent has less and less control over which situations their
children will find themselves in and then anxiety may become more debilitating for them and
more distress will be evident in their daily lives.
Numerous studies have shown that parents of anxious children tend to do exactly this. For
example, parents of anxious children have been found to be more overinvolved and intrusive,
allowing avoidance of uncomfortable situations and less encouraging of autonomy than parents
of non-​anxious children (e.g., Barrett, Rapee, Dadds, & Ryan, 1996; Hudson, Comer, & Kendall,
2008; Hudson & Rapee, 2001; Hurrell, Hudson, & Schniering, 2015). Studies on emotion social-
ization have also shown that parents of anxious children, especially mothers, tend to show greater
intrusiveness even when simply discussing previously experienced emotionally arousing negative
situations (anxiety or anger) in a structured interaction task as opposed to parents of non-​anxious
children or when discussing happy events and positive emotions (Hudson et al., 2008). In addi-
tion, the way children with anxiety interpret ambiguous situations as threatening may impact
subsequent conversations and discussions with parents (Chorpita, Albano, & Barlow, 1996).
Therefore, children who develop anxiety disorders may have had fewer opportunities to prac-
tice discussing and experiencing negative emotions and figuring out ways to manage their emo-
tions since their parents have often been too quick to assist them in avoiding or escaping negative
emotion situations or taking over the situation themselves. When the parent takes over the situ-
ation or removes the child from a mildly threatening or embarrassing situation (e.g., the child is
asked their name at a family party but cannot utter a single word due to shyness), the child learns
indirectly that 1) “this was in fact a dangerous situation since my parent felt the need to rescue
me,” and 2) “I cannot take care of it myself and I need to be rescued from such situations.”
In one recent study, Suveg, Morelen, Brewer, and Thomassin (2010) explored behavioral inhi-
bition (a temperament characteristic associated with anxiety, as seen earlier in the chapter) and
family emotional environment (restricted expressiveness, as seen earlier in the chapter) and their
associations with anxiety in a sample of late adolescents. They also examined whether emotion
regulation mediated these relationships. They argued, as above, that emotional reactivity as seen
in behaviorally inhibited youths would lead these youths to be “keyed up” or “wired,” making
emotional regulation more difficult and anxiety more probable. Similarly, they argued that the
family emotional environment would influence emotion dysregulation through a failure to appro-
priately socialize the emotion understanding and regulation skills necessary for adaptive func-
tioning, also resulting in increased anxiety. Basically, consistent with their hypotheses, they found
that a measure of emotion regulation fully mediated the relationship between behavioral inhibi-
tion and anxiety and partially mediated the relationship between family emotional environment
and anxiety. Thus, these risk factors, in tandem, were related to anxiety but mostly through their
effects on emotion regulation.
In sum, children with anxiety have less knowledge of emotion and emotion regulation strate-
gies, have fewer ways of solving problems in emotionally arousing situations, show higher emo-
tional reactivity, and have a more difficult time calming down. In addition, because the child’s
anxious behavior often elicits overprotective behavior from the parent (Rapee, Lau, & Kennedy,
2010), parents of anxious children are more likely to accept their avoidance behavior, be overly
intrusive, and encourage less autonomy and expression of emotions compared to parents of chil-
dren without anxiety. Therefore, it is important when reviewing the role of emotion regulation in
child anxiety and its implications for treatment that treatment components focusing on emotion
and emotion regulation are included in programs for child anxiety reduction and that their par-
ents are included to some extent in the treatment of their children.
2
6
1

162 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

Current CBT programs for child anxiety


Various CBT programs have been developed to treat children with anxiety and to help them
overcome their fears and modify their maladaptive behaviors. For instance, the Coping Cat pro-
gram (Kendall, 1994; Kendall & Hedtke, 2006) and the Cool Kids program (Lyneham, Abbott,
Wignall, & Rapee, 2003) were developed for treatment of school-╉aged children with various anx-
iety disorders. One-╉session treatment (OST) was developed for treatment of children with a vari-
ety of specific phobias (Ollendick et al., 2009, 2015; Öst, Svensson, Hellström, & Lindwall, 2001).
Recently, less intensive programs have been developed for parents of children with anxiety, such
as Cool Little Kids for preschool children (Rapee, Lau, & Kennedy, 2010) and From Timid to Tiger
(Cartwright-╉Hatton, 2010) for somewhat older children. The focus of these various programs
has been mostly on modifying maladaptive and catastrophic thoughts associated with anxiety
and changing maladaptive, avoidance responses. Little or no attention has been paid to emotions
or emotion regulation skills in these programs. For the most part, the children are not taught
how to manage their emotions more skillfully or how their emotions are entwined with cogni-
tions and behaviors. In order for children to learn how to overcome, manage, or at least endure
anxiety and other negative emotions, some teaching and skill building in the emotion area seems
necessary. As has been pointed out by several researchers (e.g., Hannesdottir & Ollendick, 2007;
Trosper et al., 2009; Suveg et al., 2006), despite a fairly high success rate of these CBT programs in
reducing clinical problems and anxiety (e.g., Kendall, Hudson, Choudhury, Webb, & Pimentel,
2005; Ollendick et  al., 2009, 2015; Rapee, Kennedy, Ingram, Edwards, & Sweeney, 2005), it is
possible that the children who remain symptomatic (up to one-╉third to a half) and are still meet-
ing clinical diagnosis after CBT treatment, might benefit more if an emotion component were
included in their treatment. Treatment gains might more likely be increased if emotions and
emotion regulation skills were targeted more specifically in these well-╉established and evidence-╉
based programs.
Recently, treatment programs have been developed to focus more specifically on emotions and
building emotion regulation skills for children and adolescents with anxiety. In one of the first
efforts, Suveg and colleagues (Suveg et  al., 2006)  developed a modified version of the Coping
Cat program called Emotion-╉Focused Cognitive-╉Behavioral Therapy (ECBT) where an emotion
component was added to each of the 16 sessions of the CBT program. Components such as iden-
tifying and discussing emotions associated with anxiety-╉producing situations, learning how to
identify different kinds of emotions and not just anxiety (e.g., sadness, anger, guilt, jealousy, pride,
and happiness), solving ambiguous emotion vignettes, and learning how to regulate both anxious
feelings and other emotional feelings (e.g., guilt, sadness, etc.), completing exposure tasks that
elicit not only anxiety but these other emotions as well (e.g., sadness, anger), and learning how to
regulate these emotions have all been included (Suveg et al., 2006). Results from a multiple base-
line study with six children on ECBT indicated that the children not only reduced their anxiety
level but they also improved their emotion understanding of various emotions and emotional
states and increased their knowledge of emotion regulation strategies. However, only four of the
six children were diagnosis free—╉67%, a rate almost identical to that observed in the standard
CBT program (e.g., Kendall, 1994). Subsequently, these authors examined changes in emotion-╉
related functioning following the standard Coping Cat program for anxiety (Suveg, Sood, Comer
& Kendall, 2009) and found that standard CBT did not improve children’s broad emotion regula-
tion skills. Thus, although the ECBT program appears promising, it needs to be further evaluated
with a larger sample and to be compared directly to standard CBT for children to evaluate what
changes in emotion knowledge and emotion regulation occur and how critical these changes are
for enhanced outcomes (Suveg et al., 2006).
3
6
1

Current CBT programs for child anxiety 163

Another program focusing specifically on improving emotion understanding and emotion


regulation skills has been developed for adolescents with emotional problems, anxiety and/​or
depression. A  program called the Unified Protocol for the Treatment of Emotional Disorders in
Youth (UP-​Y) was developed by Ehrenreich and colleagues to address the comorbid and not nec-
essarily well defined emotional problems that adolescents often face and have overall difficulties
in managing (Trosper et al., 2009). Emotion components in their program include, among several
others, increased awareness of emotional states, learning how to manage crisis situations, emotion
exposure, and motivational enhancement. Results of an open trial investigation of UP-​Y revealed
a reduction in clinical severity for both adolescents with anxiety or anxiety and depression, a
reduction in overall emotion dysregulation, and improved coping with worry and anger (Trosper
et al., 2009). Again, however, this program too resulted in about 67% of the youth being diagnosis
free following the intervention. This program is now being modified and examined for seven to
12 year old children who have anxiety disorders and depressive symptoms. For this age group the
program is called Emotion Detectives (Ehrenreich-​May & Bilek, 2012).
Two new developments have targeted parents and their emotion regulation difficulties to
address anxiety in children. In the Cool Little Kids program (Rapee, Lau & Kennedy, 2010), a
new anxiety prevention program intended for parents of preschool children who are showing the
first signs of anxiety, parents learn various skills intended to change maladaptive responses they
show with their children. Among various skills the parents learn is how to inhibit themselves
from jumping in too quickly when the child is faced with a mild anxiety-​provoking situations and
therefore, allowing the child to develop tolerance for the negative emotion and practice their own
emotion regulation skills. This particular program has shown beneficial effects in reducing early
anxiety symptoms and the effects are maintained for at least one year following treatment (Rapee,
Kennedy, Ingram, Edwards, & Sweeney, 2005). One of the goals of the Cool Little Kids program
is therefore, to reverse a maladaptive response cycle between parent and child in which the parent
allows the child to avoid or escape anxiety provoking situations and instead the child learns to face
such situations through modeling, practice and simple graduated exposure exercises. Teaching
parents of young anxious children these skills and modifying overprotective parenting tenden-
cies early on seems important to inhibit the development of anxiety disorders since studies have
shown that these parents are more likely to accept their children’s avoidance behavior and dis-
courage autonomy.
The SPACE program addresses similar parenting issues in older children and adolescents diag-
nosed with anxiety disorders (Lebowitz & Omer, 2013). It is based on a family systemic model of
anxiety and addresses the delicate interplay between child and parent interactions that serves to
maintain anxiety in the child. In contrast to standard CBT, it aims to promote coping, minimize
avoidance, facilitate exposure, and address catastrophic thinking in the child, SPACE addresses
parental overprotective and accommodating behaviors. As noted by Lebowitz and Omer, parental
overprotectiveness and accommodation of the anxious child may encourage the ongoing reli-
ance of the child on the parent for regulating and helping them manage their emotional state. In
effect, these parental behaviors serve to maintain the very behaviors they are intended to change.
In this program, the intervention targets the problematic parental behaviors and the need for
parents to model emotion regulation skills themselves. As such, the program focuses entirely on
the parents’ behaviors and emotions. More specifically, parents are encouraged to model use of
self-​regulation skills to better cope themselves with their child’s distress in the anxiety-​producing
situation. Parents are informed that by applying these self-​regulation skills to themselves when
they are feeling overwhelmed by the child’s distress (and likely to overprotect or accommodate),
the cycle linking their behavior and the child’s anxiety can be broken. Initial findings in an open
trial with ten anxious children and their families appear promising as significant improvements
4
6
1

164 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

in child anxiety were evidenced as were improvements in parents’ own self-╉regulation skills and
reductions in parental accommodation of the child’s anxiety (Lebowitz, Omer, Hermes, & Scahill,
2014). Again, however, only six of the ten children were designated as treatment responders—╉a
rate remarkably similar to the other emotion-╉based interventions and the standard evidence-╉
based treatments.

Future direction and conclusion


Although significant progress has been made in the understanding of emotions and the role of
emotion regulation in the onset, expression, and course of anxiety disorders in children and ado-
lescents as well as the treatment of these disorders, much work remains to be accomplished. For
example, how we define and measure emotion regulation and its dysregulation remains contro-
versial (see chapters in this volume; also White, Mazefsky, Dichter, Chiu, Richey, & Ollendick,
2014). As noted by White and colleagues, at the basic science level, there are various social-╉
cognitive (e.g., emotion awareness, recognition, and accurate expression as well as attention,
working memory, and cognitive control processes), physiological (e.g., cardiovascular activity,
skin conductance responses, pupillometric indices), and neural mechanisms (e.g., neural connec-
tivity among socio-╉affective regions in the brain) that underlie emotion regulation deficits in chil-
dren and adolescents. These mechanisms and others are poorly understood but are being actively
explored in greater detail at this time (see Chapters 2–╉5 of this volume). Thus one may conclude
we have much to learn about emotion regulation and how it “works” at the basic science level.
So, too, we need to see advances in how we measure emotion regulation in youths. To date, we
have relied largely upon self-╉report measures to assess emotion regulation difficulties as exem-
plified by the Emotion Regulation Checklist (ERC; Shields & Cicchetti, 1997), the Difficulties
in Emotion Regulation Scale (DERS; Gratz & Roemer, 2004), and the Children’s Emotion
Management Scales (CEMS; Zeman, Shipman, & Penza-╉Clyve, 2001). Although these scales have
been shown to be reliable and related positively to one another, they have not been carefully vali-
dated against behavioral, physiological, or neural measures of dysregulation. As a result, it is diffi-
cult to compare and contrast the various studies. Thus, we need more consistent and standardized
measures that entail multi-╉method and multi-╉informant strategies (see McLeod, Jensen-╉Doss, &
Ollendick, 2013, for more detail).
In the treatment realm, we need randomized control trials comparing any one of the four
innovative and emotion-╉based programs described above to “treatment as usual” and, ideally, to
standard CBT programs. To date, the evidence supporting these “newer” programs is sparse and
consists largely of small pilot studies and single case studies. We really do not know whether they
are more effective than the more standard clinic-╉based or research-╉based interventions. In fact, to
date, these programs do not appear to be more effective than the standard ones. As these studies
are undertaken, it will be important to explore the effective ingredients in these programs through
dismantling strategies. We really do not know what it is about these newer emotion-╉based treat-
ments that contribute to their effectiveness. As but one example, the first eight sessions of the
16-╉session protocol for EBCT are focused primarily on education and the development of skills
to identify and manage broad emotional experiences (Suveg et al., 2006). Are these initial sessions
really necessary? Or, would the program be more effective if greater emphases were placed on
the last half of the 16-╉session treatment which is focused on exposure tasks that provide the child
opportunities to gain mastery over various emotions in addition to anxiety during actual anxiety-╉
producing situations? Such may seem unexpected but this is exactly what has been found with the
standard Coping Cat program in which the first eight sessions are not found to produce reliable
changes in avoidance behaviors (Kendall, 1994).
5
6
1

Future direction and conclusion 165

Furthermore, following these studies, it will be important to compare the relative efficacy of
parent-╉based interventions such as the Supportive Parenting for Anxious Childhood Emotions
(SPACE) program (Lebowitz & Omer, 2013; Lebowitz et al., 2014) to the more child-╉focused
interventions such as Emotion Focused CBT (ECBT; Suveg et al., 2006; Suveg, Davis, & Jones,
2015) and the UP-╉Y; (Ehrenreich-╉May & Bilek, 2012; Trosper et al., 2009). In doing so, it will be
important to examine moderators of change in these distinctly different approaches (see Maric,
Prins, & Ollendick, 2015). For which families does a more parent-╉focused intervention work
better? It certainly seems plausible that the SPACE program might be more appropriate and
thus more effective in families in which the parents accommodate their child’s emotional dis-
plays and show poor emotion regulation strategies themselves. After all, the program is specifi-
cally designed to alter parent behavior, which is then hypothesized to result in changes in their
child’s anxiety. So, too, of course might the UP-╉Y and EBCT programs work best for children
and adolescents who themselves display emotion regulations difficulties. In one of our recent
trials (Ollendick et al., 2015), for example, only about one third of the anxious children exhib-
ited emotion regulation difficulties. For these children adding an emotion regulation compo-
nent makes good sense; for the other two thirds it might not. We might also consider whether
emotion regulation training would benefit some subgroups of children with anxiety disorders
more so than others. For example, since there is considerable overlap in symptomatology for
children with anxiety/╉worry and depression (Cummings, Caporino, & Kendall, 2014), emo-
tion understanding and regulation training might benefit this group of children with comor-
bid symptoms more in terms of learning how to manage and withstand negative emotions in
daily life. In the final analysis and consistent with a developmental psychopathology framework
(Cicchetti & Rogosch, 1996; Lease & Ollendick, 2000), multiple pathways to anxiety disorders
exist and it will be important to tailor our interventions to the specific pathways involved for
any one child with an anxiety disorder.
Thus, although much has been accomplished, continued progress is needed before we fully
understand the parameters and underlying causes of emotion dysregulation in anxious youths
and to develop, evaluate, and eventually disseminate evidence-╉based treatments for them.
Thanks to the pioneering work of Barlow (1991) some 25  years ago, we are moving in the
right direction and the next generation of research holds considerable promise for reaching
our goal.

References
Amstadter, A. (2008). Emotion regulation and anxiety disorders. Journal of Anxiety Disorders, 22, 211–╉221.
American Psychiatric Association (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Washington, DC: American Psychiatric Press.
American Psychiatric Association (2000). Diagnostic and statistical manual of mental disorders: Text
revision (4th ed.). Washington, DC: American Psychiatric Press.
Anderson, J. C. (1994). Epidemiological issues. In T. H. Ollendick, N. J. King, & W. Yule (Eds.),
International handbook of phobic and anxiety disorders in children and adolescents (pp. 43–╉65).
New York, NY: Plenum Press.
Barlow, D. H. (1991). Disorders of emotion. Psychological Inquiry, 2, 58–╉71.
Barrett, P. M., Rapee, R. M., Dadds, M. M., & Ryan, S. M. (1996). Family enhancement of cognitive style in
anxious and aggressive children. Journal of Abnormal Child Psychology, 24, 187–╉203.
Beauchaine, T. P. (2015). Future directions in emotion dysregulation and youth psychopathology. Journal of
Clinical Child and Adolescent Psychology, 44, 875–╉896.
Beidel, D. C., & Turner, S. M. (1998). Shy children, phobic adults: Nature and treatment of social phobia.
Washington, DC: American Psychological Association.
61

166 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

Brückl, T. M., Wittchen, H.-​U., Höfler, M., Pfister, H., Schneider, S., & Lieb, R. (2007). Childhood
separation anxiety and the risk of subsequent psychopathology: Results from a community study.
Psychotherapy and Psychosomatics, 76, 47–​56.
Burstein, M, Georgiades, K., He, J-​P., H., Schmitz, A., Feig, E., Kharazanov, G. K., & Merikangas, K.
(2012). Specific phobia among U.S. adolescents: Phenomenology and typology. Depression and Anxiety,
29, 1072–​1082.
Canino, G., Shrout, P. E., Rubio-​Stipec, M., Bird, H. R., Bravo, M., Ramirez, R., … Martinez-​Taboas,
A. (2004). The DSM-​IV rates of child and adolescent disorders in Puerto Rico: Prevalence, correlates,
service use, and the effects of impairment. Archives of General Psychiatry, 61, 85–​93.
Carthy, T., Horesh, N., Apter, A., Edge, M. D., & Gross, J. J. (2010). Emotional reactivity and cognitive
regulation in anxious children. Behaviour Research and Therapy, 48, 384–​393.
Carthy, T., Horesh, N., Apter, A., & Gross, J. J. (2010). Patterns of emotional reactivity and regulation in
children with anxiety disorders. Journal of Psychopathology and Behavioral Assessment, 32, 23–​36.
Cartwright-​Hatton, S. (2010). From Timid to Tiger: A Treatment Manual for Parenting the Anxious Child.
Chichester: Wiley-​Blackwell.
Chorpita, B. F., Albano, A. M., & Barlow, D.H. (1996). Cognitive processing in children: Relation to
anxiety and family influences. Journal of Clinical Child Psychology, 25, 170–​176.
Cicchetti, D., & Rogosch, F. A. (1996). Equifinality and multifinality in developmental psychopathology.
Development and Psychopathology, 8, 597–​600.
Cicchetti, D., & Toth, S. L. (1991). A developmental perspective on internalizing and externalizing
disorders. Internalizing and Externalizing Expressions of Dysfunction, 2, 1–​19.
Cummings, C. M., Ca/​porino, N. E., & Kendall, P. C. (2014). Comorbidity of anxiety disorders and
depression in children and adolescents: 20 years after. Psychological Bulletin, 140, 816–​845.
Davidson, R. J. (1998). Affective style and affective disorders: Perspectives from affective neuroscience.
Cognition and Emotion, 12(3) 307–​330).
DiBartolo, P., & Helt, M. (2007). Theoretical models of affectionate versus affectionless control in anxious
families: A critical examination based on observations of parent–​child interactions. Clinical Child and
Family Psychology Review, 10, 253–​274.
Doerfler, L. A., Connor, D. F., Volungis, A. M., & Toscano, P. (2007). Panic disorder in clinically referred
children and adolescents. Child Psychiatry and Human Development, 38, 57–​71.
Dunsmore, J. C., & Halberstadt, A. G. (1997). How does family emotional expressiveness affect children’s
schemas? New Directions for Child Development, 77, 45–​68.
Egger, H., & Angold, A. (2006). Common emotional and behavioral disorders in preschool
children: Presentation, nosology, and epidemiology. Journal of Child Psychology and Psychiatry, 47,
313–​337.
Eisenberg, N. (1998). The socialization of socioemotional competence. In D. Pushkar, A. E. Schwartzman,
& W. M. Bukowski (Eds.), Improving competence across the lifespan: Building interventions based on
theory and research (pp. 59–​78). New York, NY: Plenum Press.
Eisenberg, N., & Morris, A. S. (2002). Children´s emotion-​related regulation. Advances in Child
Development and Behavior, 30, 189–​229.
Ehrenreich-​May, J., & Bilek, E. L. (2012). The development of a transdiagnostic, cognitive behavioral group
intervention for childhood anxiety disorders and co-​occurring depression symptoms. Cognitive and
Behavioral Practice, 19, 41–​55.
Essau, C. A., Conradt, J., & Petermann, F. (2000). Frequency, comorbidity, and psychosocial impairment of
anxiety disorders in German adolescents. Journal of Anxiety Disorders, 14, 263–​279.
Fisak, B. R., & Grills-​Taquechel, A. E. (2007). Parental modeling, reinforcement, and information
transfer: Risk factors in the development of child anxiety? Clinical Child and Family Psychology Review,
10, 213–​231.
7
6
1

Future direction and conclusion 167

Fox, N. A., Bell, M. A., & Jones, N. A. (1992). Individual differences in response to stress and cerebral
asymmetry. Developmental Neuropsychology, 8, 161–​184.
Fox, N. A., Henderson, H. A., Rubin, K. H., Calkins, S. D., & Schmidt, L. A. (2001). Continuity and
discontinuity of behavioral inhibition and exuberance: Psychophysiological and behavioral influences
across the first four years of life. Child Development, 72, 1–​21.
Gratz, K. L., & Roemer, L. (2004). Multidimensional assessment of emotion regulation and
dysregulation: Development, factor structure, and initial validation of the Difficulties in Emotion
Regulation Scale. Journal of Psychopathology and Behavioral Assessment, 26, 41–​54.
Grills-​Taquechel, A. E. & Ollendick, T. H. (2012). Phobic and Anxiety Disorders in Children and
Adolescents. Cambridge, MA: Hogrefe Publishers.
Gross, J. J. (1998). The emerging field of emotion regulation: an integrative review. Review of General
Psychology, 2(3), 271.
Gross, J. J., & John, O. P. (1995). Facets of emotional expressivity: Three self-​report factors and their
correlates. Personality and Individual Differences, 29, 555–​568.
Gross, J. J., & John, O. P. (1998). Mapping the domain of expressivity: Multimethod evidence for a
hierarchical model. Journal of Personality and Social Psychology, 74, 170–​191.
Halberstadt, A. G., Denham, S. A., & Dunsmore, J. C. (2001). Affective social competence. Social
Development, 10, 79–​119.
Hannesdottir, D. K., Doxie, J., Bell, M. A., Ollendick, T. H., & Wolfe, C. D. (2010). A longitudinal study
of emotion regulation and anxiety in middle childhood: Associations with frontal EEG asymmetry in
early childhood. Developmental Psychobiology, 52, 197–​204.
Hannesdottir, D. K., & Ollendick, T. H. (2007). The role of emotion regulation in the treatment of child
anxiety disorders. Clinical Child and Family Psychology Review, 10, 275–​293.
Hirshfeld, D. R., Rosenbaum, J. F., Biederman, J., Bolduc, E. A., Faraone, S. V., Snidman, N., et al. (1992).
Stable behavioral inhibition and its association with anxiety disorder. Journal of the American Academy
of Child and Adolescent Psychiatry, 31, 103–​111.
Hirshfeld-​Becker, D., Micco, J., Simoes, N., & Henin, A. (2008). High risk studies and developmental
antecedents of anxiety disorders. American Journal of Medical Genetics, 148, 99–​117.
Hudson, J. L., Comer, J. S., & Kendall, P. C. (2008). Parental responses to positive and negative emotions in
anxious and nonanxious children. Journal of Clinical Child and Adolescent Psychology, 37(2), 303–​313.
Hudson, J. L., & Rapee, R. M. (2001). Parent-​child interactions and anxiety disorders: An observational
study. Behaviour Research and Therapy, 39, 1411–​1427.
Hum, K. M., Manassis, K., & Lewis, M. D. (2013). Neural mechanisms of emotion regulation in childhood
anxiety. Journal of Child Psychology and Psychiatry, 54, 552–​564.
Hurrell, K. E., Hudson, J. L., & Schniering, C. A. (2015). Parental reactions to children’s negative
emotions: Relationships with emotion regulation in children with an anxiety disorder. Journal of
Anxiety Disorders, 29, 72–​82.
Kail, R. V. (2007). Children and their Development (4th ed.). Upper Saddle River: Pearson.
Kendall, P. C. (1994). Treating anxiety disorders in youth: Results of a randomized clinical trial. Journal of
Consulting and Clinical Psychology, 62, 100–​110.
Kendall, P. C., & Hedtke, K. A. (2006). Coping Cat Workbook (2nd edition). Ardmore: Workbook
Publishing.
Kendall, P. C., Hudson, J. L., Choudhury, M., Webb, A., & Pimentel, S. (2005). Cognitive-​behavioral
treatment for childhood anxiety disorders. In E. D. Hibbs & P. S. Jensen (Eds.), Psychosocial treatments
for child and adolescent disorders: Empirically based strategies for clinical practice (2nd ed.) (pp. 47–​73).
Washington, DC: APA.
Kring, A. M., & Werner, K. H. (2004). Emotin regulation and psychopathology. In P. Philippot & R. S.
Feldman (Eds.) The regulation of emotion (pp. 359–​385). Mahwah, NJ: Lawrence Erlbaum Associates.
8
6
1

168 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

Lavigne, J. J., LeBailly, S.A., Hopkins, J., Gouze, K. R., & Binns, H. J. (2009). The prevalence of ADHD,
ODD, depression, and anxiety in a community sample of 4-​year-​olds. Journal of Clinical Child and
Adolescent Psychology, 38, 315–​328.
Lease, C. A., & Ollendick, T. H. (2000). Development and psychopathology. In A. S. Bellack & M. Hersen
(Eds.), Psychopathology in adulthood: An advanced text (2nd ed., pp. 131–​149). Boston: Allyn & Bacon.
Lebowitz, E. R., & Omer, H. (2013). Treating childhood and adolescent anxiety: A guide for caregivers.
New York: John Wiley & Sons.
Lebowitz, E. R., Omer, H., Hermes, H., & Scahill, L. (2014). Parent training for childhood anxiety
disorders: The SPACE program. Cognitive and Behavioral Practice, 21, 456–​469.
Lyneham, H. J., Abbott, M. J., Wignall, A., & Rapee, R. M. (2003). The Cool Kids Program—​Children’s
Workbook. Sydney: Macquarie University.
Maric, M, Prins, P. J. M., & Ollendick, T. H. (Eds.) (2015). Moderators and mediators of youth treatment
outcomes. New York: Oxford University Press.
Marks, I. M. (1987). The development of normal fear: A review. Journal of Child Psychology and Psychiatry,
28, 667–​697.
McLeod, B. D., Jensen-​Doss, A., & Ollendick, T. H. (Eds.) (2013). Diagnostic and behavioral
assessment: A clinical guide. New York: The Guilford Press.
McManis, M. H., Kagan, J., Snidman, N. C., Woodward, S. A. (2002). EEG asymmetry, power, and
temperament in children. Developmental Psychobiology, 41, 169–​177.
Merikangas, K., He, J., Brody, D., Fisher, P. W., Bourdon, K., & Koretz, D. S. (2010). Prevalence and
treatment of mental disorders among US children in the 2001-​2004 NHANES. Pediatrics, 125, 75–​81.
Merikangas, K., He, J., Burstein, M., Swanson, S. A., Avenevoli, S., Cui, L., & … Swendsen, J. (2010).
Lifetime prevalence of mental disorders in U.S. adolescents: Results from the National Comorbidity
Survey Replication-​Adolescent Supplement (NCS-​A). Journal of the American Academy of Child and
Adolescent Psychiatry, 49, 980–​989.
Muris, P., van Zwol, L., Huijding, J., & Mayer, B. (2010). Mom told me scary things about this
animal: Parents installing fear beliefs in their children via the verbal information pathway. Behaviour
Research and Therapy, 48, 341–​346.
Muris, P., & Ollendick, T. H. (2015). Children who are anxious in silence: A review on selective mutism,
the new anxiety disorder in DSM-​5. Clinical Child and Family Psychology Review, 18, 151–​169.
Ollendick, T. H. (1979). Fear reduction techniques with children. In M. Hersen, R. M. Eisler, & P. M. Miller
(Eds.), Progress in behavior modification, 8 (pp. 127–​168). New York: Academic Press.
Ollendick, T. H., & Benoit, K. (2012). A parent-​child interactional model of social anxiety disorder in
youth. Clinical Child and Family Psychology Review, 15, 81–​91.
Ollendick, T. H., Birmaher, B., & Mattis, S. G. (2004). Panic Disorder. In T. L. Morris & J. S. March (Eds.),
Anxiety disorders in children and adolescents (2nd ed.) (pp. 189–​211). New York, NY: Guilford Press.
Ollendick, T. H., & Hirshfeld-​Becker, D. R. (2002). The developmental psychopathology of social anxiety
disorder. Biological Psychiatry, 51, 44–​58.
Ollendick, T. H., Halldorsdottir, T., Fraire, M. G., Austin, K. E., Noguchi, R. J. P., Lewis, K. M., Jarrett,
M. A., Cunningham, N. R., Canavera, K., Allen, K. B., & Whitmore, M.J. (2015). Specific phobias in
youth: A randomized controlled trail comparing one-​session treatment to a parent-​augmented one-​
session treatment. Behavior Therapy, 46, 141–​155.
Ollendick, T. H., Öst, L. -​G., Reuterskiöld, L., Costa, N., Cederlund, R., Sirbu, C., Davis, T. E., & Jarrett,
M. A. (2009). One-​session treatment of specific phobias in youth: A randomized clinical trial in the
United States and Sweden. Journal of Consulting and Clinical Psychology, 77, 504–​516.
Öst, L. G., Svensson, L., Hellström, K., & Lindwall, R. (2001). One-​session treatment of specific phobias in
youths: A randomized clinical trial. Journal of Consulting and Clinical Psychology, 69, 814–​824.
Rachman, S. (1977). The conditioning theory of fear-​acquisition: A critical examination. Behaviour
Research and Therapy, 15, 375–​387.
9
6
1

Future direction and conclusion 169

Rapee, R. M. (1997). Potential role of childrearing practices in the development of anxiety and depression.
Clinical Psychology Review, 17, 47–​67.
Rapee, R. M., Kennedy, S., Ingram, M., Edwards, S., & Sweeney, L. (2005). Prevention and early
intervention of anxiety disorders in inhibited preschool children. Journal of Consulting and Clinical
Psychology, 73, 488–​497.
Rapee, R. M., Lau, E. X., & Kennedy, S. J. (2010). The Cool Little Kids Anxiety Prevention Program—​
Therapist Manual. Centre for Emotional Health, Macquarie University: Sydney.
Reiss, S., Silverman, W. K., & Weems, C. F. (2001). Anxiety sensitivity. In M. W. Vasey & M. R. Dadds
(Eds.), The Developmental Psychopathology of Anxiety (pp. 92–​111). Oxford: Oxford University Press.
Roberts, R. E., Roberts, C., & Xing, Y. (2007). Rates of DSM-​IV psychiatric disorders among adolescents in
a large metropolitan area. Journal of Psychiatric Research, 41, 959–​967.
Shields, A., & Cicchetti, D. (1997). Emotion regulation among school-​aged children: The development and
validation of a new criterion Q-​sort scale. Developmental Psychology, 33, 906–​917.
Southam-​Gerow, M. A., & Kendall, P. C. (2000). A preliminary study of the emotion understanding
of youths referred for treatment of anxiety disorders. Journal of Clinical Child Psychology, 29,
319–​327.
Southam-​Gerow, M. A., & Kendall, P. C. (2002). Emotion regulation and understanding. Implications for
child psychopathology and therapy. Clinical Psychology Review, 22, 189–​222.
Suveg, C., Davis, M., & Jones, A. (2015). Emotion regulation interventions for youth with anxiety
disorders. In T. J. Cleary (Ed.), Self-​regulated learning intervention with at-​risk populations: Academic,
mental health, and contextual considerations (pp. 137–​156). Washington DC: APA Press.
Suveg, C., Kendall, P. C., Comer, J. S., & Robin, J. (2006). Emotion-​focused cognitive-​behavioral therapy
for anxious youth: A multiple-​baseline evaluation. Journal of Contemporary Psychotherapy, 36, 77–​85.
Suveg, C., Morelen, D., Brewer, G. A., & Thomassin, K. (2010). The emotion dysregulation model of
anxiety: A preliminary path analytic examination. Journal of Anxiety Disorders, 24, 924–​930.
Suveg, C., Sood, E., Comer, J. S., & Kendall, P. C. (2009). Changes in emotion regulation following
cognitive-​behavioral therapy for anxious youth. Journal of Clinical Child and Adolescent Psychology, 38,
390–​401.
Suveg, C. & Zeman, J. (2004). Emotion regulation in children with anxiety disorders. Journal of Clinical
Child and Adolescent Psychology, 33, 750–​759.
Thompson, R. A. (2001). Childhood anxiety disorders from the perspective of emotion regulation and
attachment. In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety (pp.
160–​182). Oxford: Oxford University Press.
Thompson, R. A., & Calkins, S. D. (1996). The double-​edged sword: Emotional regulation for children at
risk. Development and Psychopathology, 8, 163–​182.
Trosper, S. E., Buzzella, B. A., Bennett, S. M., & Ehrenreich, J. T. (2009). Emotion regulation in youth
with emotional disorders: Implications for a unified treatment approach. Clinical Child and Family
Psychology Review, 12, 234–​254.
Vasey, M. W. & Dadds, M. R. (2001). An introduction to the developmental psychopathology of anxiety.
In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety (pp. 3–​26).
New York: Oxford University Press.
Vasey, M. W. & MacLeod, C. (2001). Information-​processing factors in childhood anxiety: A review and
developmental perspective. In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of
anxiety (pp. 253–​277). Oxford: Oxford University Press.
Weems, C. F., & Stickle, T. R. (2005). Anxiety disorders in childhood: Casting a nomological net. Clinical
Child and Family Psychology Review, 8, 107–​134.
Wells, J., Browne, M., Scott, K. M., McGee, M. A., Baxter, J., & Kokaua, J. (2006). Prevalence, interference
with life and severity of 12 month DSM-​IV disorders in Te Rau Hinengaro: The New Zealand Mental
Health Survey. Australian and New Zealand Journal of Psychiatry, 40, 845–​854.
0
7
1

170 Emotion Regulation and Anxiety: Developmental Psychopathology and Treatment

White, S. W., Mazefsky, C. A., Dichter, G. S., Chiu, P. H., Richey, J. A., & Ollendick, T. H. (2014).
Social-​cognitive, physiological, and neural mechanisms underlying emotion regulation
impairments: understanding anxiety in autism spectrum disorder. International Journal of
Developmental Neuroscience, 39, 212–​236.
Wittchen, H., Zhao, S., Kessler, R. C., & Eaton, W. W. (1994). DSM-​III—​R generalized anxiety disorder in
the National Comorbidity Survey. Archives of General Psychiatry, 51, 355–​364.
Wittchen, H., Nelson, C. B., & Lachner, G. G. (1998). Prevalence of mental disorders and psychosocial
impairments in adolescents and young adults. Psychological Medicine: A Journal of Research in
Psychiatry and the Allied Sciences, 28, 109–​126.
Zeman, J., Shipman, K., & Penza-​Clyve, S. (2001). Development and validation of the Children’s Emotion
Management Scales. Journal of Nonverbal Behavior, 25, 187–​205.
1
7

Chapter 9

Emotion Regulation and


Depression: Maintaining Equilibrium
between Positive and Negative Affect
Frances Rice, Shiri Davidovich, & Sandra Dunsmuir

Depression
In this chapter we describe the main features of depression in children and adolescents in terms
of symptomatology, epidemiology and risk factors. One way of viewing depression is as an imbal-
ance of positive and negative affect systems. We take a broad view of emotional regulation as
comprising responses to, interpretations of and control of emotional material affecting the bal-
ance between positive and negative affect and can, therefore, influence a person’s thoughts and
beliefs about themselves and the world around them. We describe cognitive behavioral therapy
and behavioral activation as psychological interventions primarily aimed to alter negative and
positive affect systems respectively.

Depression in children and adolescents


What is depression?
Depression is defined by a constellation of co-╉occurring symptoms that cause functional impair-
ment in areas such as family and peer relationships, as well as school work and participation in
activities. Depression can be defined as a disorder that is either present or absent or as a dimen-
sion of symptoms ranging from low to high. We first describe the criteria used to define depressive
disorder in children and adolescents and then discuss the evidence for viewing depression as a
continuum of symptoms. Symptoms of depression include disruptions in mood, cognition and
vegetative functions, which refer to disturbances in sleep and appetite. Both of the major clas-
sification systems (DSM-╉5 and ICD-╉10; American Psychiatric Association, 2013; World Health
Organization, 2015) use the same criteria to diagnose major depressive disorder (MDD) in chil-
dren, adolescents and adults, with the exception that DSM-╉5 allows irritable mood instead of
depressed mood as a core diagnostic symptom in children and adolescents.
The ICD-╉10 diagnostic system specifies the core symptoms of depression as the following:
depressed mood, loss of interest or pleasure in activities and decreased energy. Associated symp-
toms include loss of confidence or self-╉esteem, unreasonable feelings of self-╉reproach or excessive
inappropriate guilt, recurrent thoughts of death or suicide or any suicidal behavior, diminished
ability to think or concentrate, change in psychomotor activity (either agitation or retardation),
sleep disturbance and change in appetite with corresponding change in weight. According to
ICD-╉10 criteria, at least four of these symptoms must be present for at least two weeks to diagnose
a mild depressive episode, six to diagnose a moderate depressive episode, or eight for a severe
depressive episode (these symptoms must include at least two of the core symptoms described
2
7
1

172 Emotion Regulation and Depression

above). Depressive disorder is usually defined by clinical interview and there are fairly rigorous
quantitative and qualitative criteria for operationalizing or defining each symptom. For instance,
in order to meet the criteria for depressed mood, the low mood must be present most of the time
and nearly every day for at least two weeks (i.e., the “quantity” is high) and the low mood should
be qualitatively different from the ordinary ups and downs of mood; for instance, it should be
present to a degree that is definitely abnormal for the particular individual and should be rela-
tively unaffected by external factors. The symptoms must also cause significant distress or impair-
ment in social, educational or other important areas of functioning. The diagnostic criteria for
depression include symptoms of both increased negative affect as well as decreased positive affect.
Symptoms like low mood and irritability index, high levels of negative affect and symptoms like
loss of interest or pleasure index low levels of positive affect and both are indicative of a depres-
sive mood state. One way of viewing depression is as an imbalance between negative and positive
affect systems (Insel et al., 2010).
There is good evidence to support the validity of viewing depression as a continuous dimension
of symptoms. For instance, when depressive symptoms are present but fall short of meeting the
diagnostic threshold, they are still impairing (Angold, Costello, Farmer, Burns, & Erkanli, 1999;
Pickles et al., 2001), increase the risk of later depressive disorder (Pine, Cohen, Cohen, & Brook,
1999) and are associated with a range of poor outcomes similar in severity to that of the diagnosis
of MDD (Wesselhoeft, Sorensen, Heiervang, & Bilenberg, 2013). Depressive disorder and depres-
sive symptoms are associated with a range of adverse short-╉term and long-╉term outcomes includ-
ing deliberate self-╉harm, educational failure and future depressive episodes (Angold et al., 1999;
Pickles et al., 2001; Riglin, Petrides, Frederickson, & Rice, 2014; Skegg, 2005). Thus, high sub-╉
threshold symptoms indicate an increased probability of developing MDD and are also associated
with functional impairment. The usual approach to defining severity of depression is to sum the
number of symptoms endorsed; however, it is also accepted that particular symptoms may only
be present at higher levels of severity and that certain symptoms, such as suicidality, may be of
particular concern (National Institute for Health and Clinical Excellence, 2005). It is worth noting
that there are different methods used to define depressive symptoms including semi-╉structured
interviews, structured interviews and self or parent reported questionnaires. The threshold for
endorsing symptoms differs according to the method used, with the most conservative estimates
of the prevalence (i.e., lower estimates) provided by semi-╉structured interview.

How common is depression in children and adolescents?


The method used to assess depression estimates the rate of depression in a population during a
specified time frame (i.e., prevalence estimates). MDD during childhood is relatively uncommon
and prevalence estimates in a 12-╉month period range from 0.5–╉3% (Birmaher, Ryan, Williamson,
Brent, & Kaufman, 1996; Harrington et al., 1993). During adolescence, the prevalence of MDD and
depressive symptoms falling below the diagnostic threshold increase dramatically (Lewinsohn,
Rohde, & Seeley, 1998; Rushton, Forcier, & Schectman, 2002), particularly in girls. Estimates of
the 12-╉month prevalence of depressive disorder in adolescence range from 2–╉8%. During the
whole period of adolescence, around 20% of individuals will experience an episode of MDD
(Birmaher et al., 1996; Costello, Foley, & Angold, 2006). In childhood, either an equal proportion
of boys and girls are affected, or a slight excess of boys is affected. However, in adolescence, the
ratio of affected females to males is two to one; this reflects the pattern seen in adult life where
females are twice as likely to be depressed as males (Costello et al., 2006). Symptoms of depression
are much more common than MDD and also increase around adolescence (Angold et al., 2002;
Rushton et al., 2002).
3
7
1

Depression in children and adolescents 173

Depression in young people is under-╉recognized; studies across the world show that only a
small minority of children and adolescents meeting diagnostic criteria for depressive disorder
receive any kind of intervention from a health professional (Paula et al., 2014; Sayal, Yates, Spears,
& Stallard, 2014; Zhong et al., 2013). This low rate of access to services is also true of individuals
with known risk factors for depression such as the offspring of depressed parents (Potter et al.,
2012) and young people with previous depressive episodes (Brenner et al., 2015). Various features
may contribute to difficulties in identifying depression in this age group, including the presen-
tation of depression (where mood may be irritable rather than depressed or low mood may be
fluctuating), as well as the presence of other difficulties such as academic impairment, which may
mask the underlying problem.

Risk factors
It is generally accepted that depression has a complex, multi-╉factorial aetiology. This means there
are multiple causal risk factors involved acting in concert with protective factors in complex ways.
These risk and protective factors are varied (e.g., cognitive, biological, contextual) and both genes
and environment are likely to contribute. In this section, we discuss factors that may be useful for
identifying groups at high risk of developing depression. We also include a brief discussion about
what is known about genetic and environmental risk factors. As mentioned earlier, high levels
of sub-╉threshold depressive symptoms increase risk for episodes of MDD. Other important risk
factors for depression in young people are depressive disorder in a parent and exposure to stress-
ful life events. Children of depressed parents are around three times more likely to be diagnosed
with MDD compared to the children of healthy control groups (Rice, Harold, & Thapar, 2002).
Nevertheless, familial risk is not depression-╉specific and there is familial clustering of other types
of psychopathology, such as antisocial behavior (Harrington et  al., 1997)  and anxiety (Rende,
Warner, Wickramarante, & Weissman, 1999; Warner, Weissman, Mufson, & Wickramaratne,
1999; Weissman, Warner, Wickramaratne, Moreau, & Olfson, 1997).
Stressful life events are important in the onset of depressive disorder and 60% of adolescents
with depression experience an acutely disappointing life event in the month prior to the onset of
depression (National Institute for Health and Clinical Excellence, 2005). Stressful events involving
the loss of a significant relationship due to death or separation and acutely disappointing events,
such as failing an exam, appear to be especially important in precipitating depression. Stressful life
events may be particularly important in the first onset of depression, as opposed to recurrences
(Kendler, Thornton, & Gardner, 2000; Monroe, Rohde, Seeley, & Lewinsohn, 1999). Depressed
young people may also “evoke” or “create” stress in their lives because of the way they behave, for
instance, losing a friend following an argument (Hammen, 1991). There is evidence that exposure
to social stress increases around adolescence (Rice, Harold, & Thapar, 2003). Girls appear particu-
larly sensitive to depressive symptoms following social stress (Rudolph, 2002). Consistent with
their important role in the development of depression, sub-╉threshold symptoms, family history of
depression in a parent and exposure to stressful events have all been used as criteria for the selec-
tion of individuals to receive therapeutic intervention, with the aim of reducing symptomatology
(i.e., selective or indicated prevention programs; Horowitz & Garber, 2006; Stice, Shaw, Bohon,
Marti, & Rohde, 2009).
Twin studies show that depressive symptoms and disorder in adolescence are influenced by
genetic factors to a moderate degree (Rice, 2010). It is unclear exactly what is inherited, but it
seems likely that part of the heritable effect is indirect. For instance, genes may operate indirectly
via influences on behavior or personality traits that affect exposure to stress. Interestingly, a num-
ber of twin studies show that depressive symptoms in childhood are not influenced by genetic
4
7
1

174 Emotion Regulation and Depression

factors but environmental influences predominate (Eaves et al., 1997; Rice et al., 2002; Scourfield
et  al., 2003; Thapar & McGuffin, 1994). Longitudinal studies also suggest that childhood and
adolescent depression may differ in rates of continuity into adult life, with childhood depression
showing low rates of continuity with depression in adulthood and adolescent depression showing
higher rates of continuity with depression in adulthood (Harrington et al., 1991). There are dif-
ferences in the prevalence, sex ratio of cases, rates of continuity and aetiology of childhood and
adolescent onset depressive symptoms and disorder. Evidence, to date, suggests that adolescent
depression can be viewed as an early onset form of the adult disorder (Thapar, Collishaw, Pine, &
Thapar, 2012); whether this is the case for childhood onset depression is less clear and it may be
that childhood onset depression is different.
In summary, depression can be viewed both categorically and continuously. There are differ-
ences between depression in childhood and adolescence. Depression in a parent, stressful life
events and sub-╉threshold symptoms are important risk factors for MDD. Depression involves
symptoms that reflect an excess of negative affect, such as low mood and irritability and a dearth
of positive affect, such as a loss of interest.

Emotion regulation difficulties


Psychological theories of depression
Cognitive theories of depression suggest that negative biases, such as selectively attending to and
elaborating on depression-╉related stimuli, are involved in the onset and maintenance of depres-
sion (Beck, 1976; Beck, Rush, Shaw, & Emery, 1979). Other cognitive theories highlight that indi-
viduals who become depressed may have difficulty employing adaptive strategies to regulate mood
and “recovering” from low mood; they therefore, respond to low mood in ways that encourage
its persistence (Nolen-╉Hoeksema, Wisco, & Lyubomirsky, 2008; Teasdale, 1988). Recent cognitive
theories of depression have also highlighted the importance of lower level cognitive processes in
forming “higher level” negative beliefs about the self, the world and the future (Roiser, Elliott, &
Sahakian, 2012). Such lower level cognitive processes underlie the bleak and pessimistic views
that depressed individuals exhibit and include cognitive control processes. Indeed, it is increas-
ingly recognized that the control of emotional material is compromized in depression (Gotlib
& Joormann, 2010). To some extent, interventions for depression such as Cognitive Behavioral
Therapy (CBT), encourage participants to employ cognitive or executive control over thoughts,
feelings and behavior (Siegle, Ghinassi, & Thase, 2007). Cognitive theories of depression and its
treatment have predominated research and practice in recent years. However, it is also important
to note that behavioral theories of depression take a different stance and emphasize low levels
of engagement in enjoyable activities as important in the onset and maintenance of depression
(Lewinsohn, 1975; Lewinsohn & Graf, 1973). There has been a resurgence of interest in the impor-
tance of positive information and rewarding activities in depression, given strong neuro-╉scientific
evidence that this is compromized in depression (Pizzagalli, 2014).

Emotion regulation in childhood and adolescent depression


Depressed individuals exhibit a number of difficulties in effectively processing and regulating
emotional responses. In this section, we describe the role of key emotional regulation difficul-
ties present in children and adolescents that are depressed. Emotional regulation difficulties
can be organized as relating to a preference or bias for negative information, or an absence of
a preference or bias for positive information. Such emotion regulation difficulties can apply to
5
7
1

Emotion regulation difficulties 175

attention, to memory, to decision making, to attitudes and to the cognitive/╉executive control


of emotional information. We discuss each of these in turn. We also discuss the role of emo-
tion regulation problems in the onset and maintenance of depression in children and young
people.
Various methodologies and designs have been used to assess the role of emotion regulation dif-
ficulties in the aetiology of depression. Tests of whether emotion regulation difficulties are present
only during the actively depressed state or whether they persist following remission have involved
comparing currently depressed and remitted groups. Studies investigating whether emotion regu-
lation problems increase risk for depression have involved comparing high risk groups, such as
the offspring of depressed parents, to low risk groups. Nevertheless, in the absence of longitudinal
follow-╉up, such group based designs lack the important test of whether observed differences are
related to the development of later psychopathology. Whilst there is a general lack of longitudinal
studies, a handful of such studies have examined the role of emotion regulation difficulties in
the onset and persistence of symptoms over time. In our review of emotion regulation, where
possible, we focus on research that uses behavioral tasks to assess aspects of cognition involved
in emotion regulation. We do this because performance-╉based measures allow for the measure-
ment of cognitive biases that may not be open to introspection, whereas, self-╉report measures
do not (Harmer, O’Sullivan, et al., 2009; Rawal, Collishaw, Thapar, & Rice, 2013b). A focus on
performance-╉based indicators of cognitive bias and emotion regulation also lessens the likelihood
that associations with depression are due to shared method variance, which is possible when the
same informant rates a risk factor (e.g., cognitive bias) and an outcome (e.g., depressive symp-
toms) (Rutter, Pickles, Murray, & Eaves, 2001).

Attention and emotion regulation


Cognitive theories of depression suggest that affective biases, whereby individuals pay attention
to depression-╉related stimuli and have difficulty in disengaging from them, are involved in the
onset and maintenance of depression (Beck, 1976). The deployment of attention is also posited
as an important early mechanism of emotional regulation (John & Gross, 2007). For instance,
shifting attention away from an emotionally distressing stimulus may help regulate emotion. An
affective bias towards negative stimuli would involve being more accurate, quicker to respond to
and slower to disengage from negative stimuli compared to positive or neutral stimuli. Such a bias
is likely to affect an individual’s view of themself, the future and the world around them. There
is evidence that such a bias exists in depressed adults. (Gotlib & Joormann, 2010; Mathews &
MacLeod, 2005). However, the evidence for a negative attentional bias in depressed children and
adolescents is inconsistent. Several studies report a negative affective bias (Gibb, Benas, Grassia,
& McGeary, 2009; Joormann, Talbot, & Gotlib, 2007b; Kyte, Goodyer, & Sahakian, 2005) while
others report a preference for positive material or reduced accuracy for negative material, which
is inconsistent with a negative affective bias (Harrison & Gibb, 2014; Kilford et  al., 2015). For
instance, one study found that depressed young people avoided sad stimuli or preferred happy
stimuli (Harrison & Gibb, 2014) and another found that depressed adolescents made more errors
for sad compared to happy stimuli (Kilford et al., 2015). Findings are also fairly inconsistent for
emotional recognition tasks (Joormann, Gilbert, & Gotlib, 2010; Lewinsohn, Zeiss, & Duncan,
1989; Lopez-╉Duran, Kuhlman, George, & Kovacs, 2013). Thus, one study found evidence that
depressed young people may be more sensitive to sad stimuli, in that they detected sad faces at
lower levels of intensity exposure compared to healthy individuals (Schepman, Taylor, Collishaw,
& Fombonne, 2012). Another study also reported a similar finding in high-╉risk, unaffected boys
6
7
1

176 Emotion Regulation and Depression

(Lopez-╉Duran et al., 2013) but a different pattern of results where shown in a study of high-╉risk
girls, where they appeared less sensitive to sad stimuli (Joormann et al., 2010). Studies of remit-
ted, depressed adolescents (Maalouf et al., 2011) and a longitudinal study examining the affective
processing of depressed adolescents, prior to the onset of depression (Kilford et al., 2015), showed
that any observed affective bias appeared to be state dependent and was not observed in remitted
adolescents. Collectively, these results do not suggest a clear pattern of attentional bias for emo-
tional information in depressed children and adolescents.
Studies that have reported a preference for happy stimuli in depressed cases have suggested that
some depressed young people may use this as a rudimentary form of emotion regulation (Cohn
& Tronick, 1983). Thus, allocation of attention away from a negative stimulus (such as a sad face)
and towards a happy one may serve to reduce emotional distress. There is evidence that this may
occur in infants of depressed mothers and the use of this strategy is related to the infant’s own
affect (Cohn & Tronick, 1983; Termine & Izard, 1988).
In summary, there is inconsistent evidence for negative affective bias as measured by atten-
tional tasks in depressed children and adolescents. This contrasts with the pattern of results for
depressed adults and illustrates the importance of considering developmental differences. Results
of one longitudinal study suggested that sad and happy material may differentially interrupt task
performance in depressed and healthy individuals and that it may be important to consider the
role of cognitive control of emotional material (Kilford et al., 2015). We deal with the control of
emotional material later in this chapter.

Memory and emotion regulation


Classic studies of depressed adults show that they remember the past differently to non-╉
depressed individuals. Depressed adults show a skewed recollection of negative experiences,
whereby autobiographical memories that are negative are recalled more often and more quickly
(Gotlib & Joormann, 2010; Lloyd & Lishman, 1975). Studies have used the Self Referent
Encoding Task (SRET) in children and adolescents. This task involves asking individuals to
judge whether a set of positive and negative adjectives are self-╉descriptive, followed by a sur-
prise free recall task where participants are asked to recall as many of the words as possible.
Depressed young people have been reported as recalling fewer positive and more negative self-╉
referent words (Prieto, Cole, & Tageson, 1992; Timbremont & Braet, 2004a; Zupan, Hammen,
& Jaenicke, 1987), or fewer positive self-╉referent words only (Gençöz, Voelz, Gençöz, Pettit, &
Joiner, 2001; Hammen & Zupan, 1984). Interestingly, two independent studies report that fewer
recalled positive words predict later increases in depressive symptoms over time (Connolly,
Abramson, & Alloy, 2015; Goldstein et al., 2014), suggesting that a relative lack of memory for
positive information may be involved in the onset of depression. Thus, memory bias for recall-
ing less positive information is present in the depressed state and may also predict the develop-
ment of depression over time.
Another memory bias present in depressed children and adolescents is over-╉general autobio-
graphical memory (AM). AM refers to the recollection of experiences and is important because
it relates to the sense of self and influences social problem solving and functioning (Goddard,
Dritschel, & Burton, 1996; Raes et al., 2005; Williams et al., 2007). One quality of AM that has
been closely linked with depression is the specificity of the recalled material, that is, recalling
events that occurred at a (specific) particular time and place. There is good evidence that depres-
sion in adults is associated with a tendency to recall autobiographical memories that are over-╉
general and refer to extended periods of time or repeated events (Williams et al., 2007). Thus,
autobiographical memory in depressed adults lacks contextual detail. Similar evidence exists for
71

Emotion regulation difficulties 177

depressed children and young people (Hipwell, Sapotichne, Klostermann, Battista, & Keenan,
2011; Rawal & Rice, 2012a). This pattern of response distinguishes children with depression, not
only from healthy controls, but also from children with other psychiatric disorders (Rawal & Rice,
2012a). Interestingly, there is evidence that over-╉general memory for negative cue words specifi-
cally predicts later depression (Rawal & Rice, 2012a) and distinguishes high-╉risk from low-╉risk
children (Woody, Burkhouse, & Gibb, 2015). It is important to note that the valence (i.e,. happy
or sad) effect is observed for the type of word used as a cue to prompt the recall of a memory
rather that the valence of the autobiographical memory itself. One longitudinal study reported
that currently depressed adolescents were more over-╉general than healthy and psychiatric control
groups for both positive and negative cue words, but that the predictive influence of OGM on
later depressive symptoms and disorder was only observed for negative cue words (Rawal & Rice,
2012a). This provides indirect evidence that, OGM may be a process that begins initially with
negative material (i.e., prior to a depressive episode) but becomes generalized to other material
with time (i.e., once a depressive episode has begun) (Williams, 1996). Thus, OGM may indi-
rectly measure a type of emotion regulation strategy where specific details of the personal past are
avoided in order to reduce distress (Crane et al., 2014; Williams, 1996). Such a strategy is unlikely
to be adaptive in the long-╉term. Over-╉general autobiographical memory may also interfere with
using memory for other, more effective, forms of emotion regulation.
Recalling positive autobiographical memories has been shown to be an effective strategy to
repair low mood. This phenomenon has also been shown in experimental research, where acti-
vating the neural substrate of a positive memory reduces depressive symptoms (Ramirez et al.,
2015). It seems likely that the inability to recall specific details of the personal past will affect a
person’s ability to use positive memories as a form of emotion regulation (Dalgleish & Werner-╉
Seidler, 2014). There is an indication that depressed individuals are less effective in using the recall
of positive events as a strategy to repair sad mood (Joormann, Siemer, & Gotlib, 2007). It is not
exactly clear why this might be; however, it is possible that recalling memories that lack detail and
vividness may fail to have the desired mood-╉repair effect. Indeed, one study showed that adults
with remitted MDD recalled less vivid positive memories than healthy controls (Werner-╉Seidler
& Moulds, 2011). An intervention study found that training in positive imagery improved symp-
toms of anhedonia in currently depressed adults (Blackwell et al., 2015). A preventive, universal
intervention that reduced depressive symptoms in healthy adolescents also included a session
based on recalling positive autobiographical memories (Rice et al., 2015).
In summary, evidence suggests that recalling low levels of positive information is associ-
ated with the onset of depressive symptoms over time. There is reason to believe that recalling
positive autobiographical memories is an effective emotion regulation strategy. Studies of adults
suggest this effective mood repair strategy goes awry in depression. Further research is required
to test if this finding also applies to children and adolescents with depression. Another feature
of autobiographical memory that is important in depression occurring in young people is the
level of specificity that is recalled. Difficulties in recalling specific details of the personal past is
likely to interfere with the ability to use happy memories as an effective mood repair strategy.
Collectively, results illustrate the importance of positive memories in regulating and repairing
low mood.

Decision making and emotion regulation


A body of research suggests that depressed young people have difficulty in making decisions
involving the opportunity for reward. In particular, depressed young people fail to show a norma-
tive bias towards stimuli associated with a more likely chance of obtaining reward. This type of
8
7
1

178 Emotion Regulation and Depression

research most often uses behavioral tasks involving rewards, such as winning points or money;
although, tasks involving other sorts of reward, including social reward are beginning to be devel-
oped and used. Lowered reward responsiveness may be involved in the maintenance of depressive
symptoms, as it may lead to diminished engagement in pleasurable activities and reduced motiva-
tion to pursue outcomes that are usually enjoyable, such as social events, interpersonal relation-
ships and activities like exercise (Depue & Iacono, 1989; Forbes & Dahl, 2005; Lewinsohn, 1975).
Consistent with this idea, lowered reward-​responsiveness, as measured by fMRI, and behavioral
tasks are correlated with affect in every-​day life (Forbes et al., 2009) and engagement in positive
daily activities, such as exercise and extra-​curricular activities (Rawal et al., 2013b). Several stud-
ies have shown that depressed young people “play it safe” and bet less of their points compared to
healthy individuals and psychiatric controls when the chances of winning are very high (Forbes,
Shaw, & Dahl, 2007; Guyer et al., 2006; Rawal et al., 2013b). This pattern of reward decision mak-
ing has also been shown to predict depressive symptoms and new onset depressive disorder over
time when controlling for prior symptom severity (Rawal et al., 2013b). Similarly, lowered activa-
tion in the ventral striatum, a brain area involved in reward processing, has been found to predict
the symptom of anhedonia/​loss of interest over time (Stringaris et al., 2015).
Thus, MDD in children and adolescents appears to be characterized by a reduced expectation
of future reward, a diminished ability to change behavior according to the likelihood of a reward
and alterations in the functioning of the brain’s reward circuit (Pizzagalli, 2014; Stringaris et al.,
2015). Reduced sensitivity to reward is thought to be a factor underlying the symptoms of depres-
sion that index low positive affect, such as anhedonia/​loss of interest and social withdrawal. These
sorts of symptoms may be particularly important markers of severity and prognosis. For instance,
anhedonia has been reported to predict severity and relapse in treatment resistant adolescent
depression (McMakin et al., 2012). Family studies have suggested that reductions in the capacity
for positive affect may distinguish children at low and high familial risk for depression (Olino
et al., 2011). Kovacs & Lopez-​Duran (2010) posit that an attenuated capacity for positive affect
(which may stem in part from blunted reward sensitivity) is likely to interfere with the ability of
young people to engage in effective mood repair strategies such as doing something fun, doing an
enjoyable, distracting activity or focusing on happy memories. An intervention involving training
in reward decision making and strategies to enhance positive affect reduced adolescent depressive
symptoms; the change in reward decision making appeared to explain the reduction in symptoms
(Rice et al., 2015). Thus, the lack of a normative positive bias, as indicated by reduced sensitivity
to reward, may also impair effective emotion regulation strategies.
In line with the idea that depressed young people fail to show a normative positive bias are
studies of probability judgements. It is widely established that healthy individuals show an opti-
mism bias where they judge negative events as more likely to happen to others than to themselves
(Sharot, 2011). Depressed children and adolescents made more balanced judgements and (cor-
rectly) judged these events as equally likely to occur to themselves as to another person (Dalgleish
et al., 1997). This effect appeared to be mood-​dependent as it was not observed in a recovered
depressed group (Dalgleish et al., 1998).
In summary, there is good evidence that decision making, when there is the potential for
reward, is impaired in individuals with depression and those at increased risk of developing
depression. There is also good evidence that the neural correlates of reward processing are affected
in depression and recent evidence also shows that this predicts depression over time (Stringaris
et al., 2015) consistent with previous longitudinal behavioral studies (Forbes et al., 2007; Rawal
et al., 2013b). Depressed and vulnerable young people appear not to expect future reward and are
inflexible in modifying their behavior according to the likelihood of obtaining a reward (Rawal
et  al., 2014). It is possible that low expectations of future reward, as well as lower capacity for
9
7
1

Emotion regulation difficulties 179

positive affect, may interfere with young people’s ability to engage in effective emotion regulation
strategies, particularly those that involve up-╉regulating positive affect.

Attitudes and emotion regulation


Depressed young people show a range of explicitly negative patterns of thinking, including beliefs
about themselves and interpretations about the world around them (Hankin & Abramson, 2001).
Dysfunctional attitudes about the self are common in depression as are negative interpretations
about the causes, consequences or self-╉implications of stressful events (Beck et al., 1979; Teasdale
et al., 2002). Thus, depressed young people tend to interpret the meaning of a negative event in a
negatively biased way as indexed by their attributional style (Schepman, Fombonne, Collishaw, &
Taylor, 2014; Seligman et al., 1984) and under-╉estimate their own abilities relative to their peers
and teachers (Cole, Martin, Peeke, Seroczynski, & Fier, 1999). Nonetheless, while negative styles
of thinking are elevated during an episode of depression, they are not usually elevated in those
whose depression has remitted (Teasdale, 1988), despite the fact that individuals with previous
depression are at heightened risk of experiencing future episodes (Solomon et al., 2000). It has
been suggested that negative thinking remains latent in remitted individuals until activated by
depressive mood or stress and is not accessible in non-╉depressed mood (Miranda & Persons,
1988; Wenzlaff & Bates, 1998). Indeed, vulnerability-╉provoking procedures, such as sad mood
inductions, appear to “activate” or elicit negative thinking in young people at risk of depression
(the offspring of depressed parents or those with a previous episode of depression) (Dearing &
Gotlib, 2009; Ingram & Ritter, 2000; Kelvin, Goodyer, Teasdale, & Brechin, 1999; Taylor & Ingram,
1999; Timbremont & Braet, 2004b). For instance, one study showed that children of depressed
mothers exhibited negative interpretations of their performance on a task only during a mildly
stressful situation and not in the absence of stress (Murray, Woolgar, Cooper, & Hipwell, 2001).
Using a different approach, one study used reaction time as a more implicit measure of negative
thinking, where speed to endorse a dysfunctional attitude was taken as a measure of the “strength”
of the attitude (Rawal, Collishaw, Thapar, & Rice, 2013a). In that study, being relatively quicker
to endorse dysfunctional attitudes differentiated remitted and control individuals and predicted
depressive symptoms over time.
Individuals who are depressed may also respond and relate to low mood differently than non-╉
depressed individuals. Depressed children and adolescents can become captured by and elabo-
rate on negative information, consistently responding to low mood in ways that encourage its
persistence. Rumination, a response to sad mood which involves analysing and questioning the
reasons for low mood (Nolen-╉Hoeksema, 1991; Watkins, 2008), has been associated with depres-
sive symptoms both cross-╉sectionally and longitudinally, suggesting that it is involved in the onset
of depression as well as the current depressive state (Rood, Roelofs, Bogels, Nolen-╉Hoeksema, &
Schouten, 2009). Nonetheless, it should be noted that there are relatively few longitudinal studies
that control for prior depression and those that do, report a substantial reduction compared to the
uncorrected correlation between depression and rumination (Rood et al., 2009). Thus, depressed
individuals certainly ruminate, but whether this predicts the onset of depression over time, inde-
pendently of depressive symptoms, is not completely clear due to a lack of studies using designs
that allow this this question to be addressed.
In summary, depressed individuals show pessimistic interpretations about themselves and the
world. This negative style of thinking may remain latent when mood is normal but may increase
vulnerability to depressed mood following exposure to a stressor. Rumination as a response to
low mood is also increased in depressed young people and may increase vulnerability for future
episodes.
0
8
1

180 Emotion Regulation and Depression

Executive functioning and emotion regulation


Thus far, we have presented evidence showing that individuals with depression show patterns in
cognitive operations such as attention, memory, decision making and attitudes that can affect
the balance of the accessibility of negative and positive emotional information. In this section,
we discuss how executive functioning processes are involved in processing and exerting control
over emotional material. Executive functions (EF) are a set of mental processes that enable goal
directed behavior, planning and adaptive response to novel or challenging situations (Diamond,
2013; Hughes, Graham, & Grayson, 2004). Core executive functions include inhibitory control,
working memory and mental flexibility. In this section, we discuss how executive functions that
involve control and flexibility over emotional material may be important in emotion regulation
(Schmeichel & Tang, 2014) and may be compromised in depression. The development of execu-
tive functions starts in the first years of life and continues throughout childhood and adolescence,
reaching maturity only in early adulthood (Best & Miller, 2010).
A large body of research has shown that depression in adults is associated with impairments in
execution functions, especially in tasks or situations that involve processing emotional informa-
tion (Gotlib & Joormann, 2010; Wagner, Doering, Helmreich, Lieb, & Tadić, 2012). Although
there are fewer studies examining executive functioning in depression in children and adoles-
cents, a recent meta-╉analysis (Wagner, Müller, Helmreich, Huss, & Tadić, 2014)  showed that
children and adolescents with MDD performed significantly worse on measures of a range of
executive functioning tasks. Initial findings also suggest that executive functioning difficulties,
in the context of emotional material, may be a cognitive risk factor for developing depression
(Davidovich et al., 2015; Joormann, Talbot, & Gotlib, 2007a; Kilford et al., 2015). Kilford and col-
leagues (2015) report findings suggesting that sad stimuli interfere with cognitive or behavioral
control in currently depressed young people and in those who later go on to develop depres-
sion. Davidovich and colleagues (2015) report that better executive functioning in the offspring of
depressed parents protects against depressive symptoms in the children themselves.
There are several pathways through which executive functioning may be associated with emo-
tion regulation and risk for depression. One possibility is that the negative affective biases and
mood regulation difficulties that characterize depressed individuals may be underpinned by
impairments in executive functions (Harmer, Goodwin, & Cowen, 2009; Roiser et  al., 2012).
Depressed individuals, as well as those at risk for depression, show impaired inhibition of emo-
tional information (Gotlib & Joormann, 2010; Kilford et  al., 2015; Kujawa et  al., 2011). Such
impairments may contribute to a tendency to get “captured” by negative thoughts and mood,
impede mood regulation and lead to sustained negative affect. In line with this notion, difficul-
ties in inhibiting responses and shifting cognitive sets have been found to be associated with
higher levels of rumination, a cognitive style which perpetuates low mood (De Lissnyder, Koster,
Derakshan, & De Raedt, 2010; Joormann & Quinn, 2014).
Effective mood regulation strategies, such as reappraisal, and examining an issue from different
perspectives, may rely on elements of executive functioning. Employing such strategies requires
inhibition in order to disengage from a particular interpretation or perspective, mental flexibility,
in order to generate and shift to another interpretation or change perspectives and working mem-
ory, in order keep the different options accessible in memory and manipulate them (Kross, Ayduk,
& Mischel, 2005; McRae, Jacobs, Ray, John, & Gross, 2012; Ochsner & Gross, 2008). Preliminary
findings indicated that the ability to use cognitive reappraisal as an emotion regulation strategy
is associated with aspects of executive functioning such as working memory and set shifting in
adults (McRae et al., 2012).
1
8

Evidence-based intervention approaches for depression 181

Reduced executive functioning capacities might interfere with retrieving specific autobiograph-
ical memories, which may serve as an important resource when coping with negative or stressful
events. Difficulties in retrieving specific autobiographical memories (i.e., overgeneral autobio-
graphical memory; [OGM]) have been shown to characterize currently depressed children and
adults (Kuyken, Howell, & Dalgleish, 2006; Park, Goodyer, & Teasdale, 2002; Vrielynck, Deplus,
& Philippot, 2007; Williams et al., 2007). Longitudinal studies also demonstrate that overgeneral
autobiographical memory increases the risk for developing depression (Hipwell et al., 2011; Rawal
& Rice, 2012a). Several theorists suggest that impairment in retrieving specific autobiographical
memories may be at least partially due to reduced executive functions (Conway & Pleydell-╉Pearce,
2000; Hertel & Hardin, 1990; Williams et  al., 2007; Zacks & Hasher, 1994). Studies conducted
with adults and children give support to the role of executive functioning in the retrieval of spe-
cific autobiographical memories (Picard, Reffuveille, Eustache, & Piolino, 2009; Raes, Verstraeten,
Bijttebier, Vasey, & Dalgleish, 2010; Rawal & Rice, 2012b; Williams et al., 2007). This may inter-
fere with the ability to use autobiographical memory as a form of emotion regulation strategy as
described above.
Finally, it is possible that executive functioning could affect how individuals respond to preven-
tive or therapeutic interventions. It has been suggested that psychological interventions used for
depression, including Cognitive Behavioral Therapy(CBT), encourage the patient to exercise cog-
nitive control over thoughts and emotional responses. These strategies attempt to change how the
person relates and responds to their thoughts by generating alternative interpretations, switch-
ing between thoughts and interpretations or examining thoughts from a distanced perspective
(Brewin, 2006; Hayes, Luoma, Bond, Masuda, & Lillis, 2006; Siegle et al., 2007). While such inter-
ventions may improve executive functioning by practicing these skills, it is also possible that those
with executive functioning difficulties might find these tasks challenging.
In summary, depressed children and adolescents show a range of difficulties in processing emo-
tional material that may compromise their ability to use effective emotion regulation techniques.
These difficulties include biases for negative information, an absence of bias for positive informa-
tion and difficulties in exerting control over emotional material. The best evidence of emotion
regulation difficulties predicting the onset of depressive disorder and symptoms comes from low
levels of bias for positive information. Later we review the interventions that are used to amelio-
rate depression in young people and relate these to the emotion regulation difficulties described
previously.

Evidence-based intervention approaches for depression


Psychological therapies are the intervention of choice for depression in children and young people
(National Institute for Health and Clinical Excellence, 2005). Until recently, CBT was considered the
“gold-╉standard” in psychological therapy for depression. However, it is now believed that there is
little good quality evidence to suggest that any one psychological treatment is better than another
(National Institute for Health and Care Excellence, 2015). The results of a large trial comparing differ-
ent types of psychological treatment support the view that no one psychological treatment is superior
to another (Goodyer et al., 2017; National Institute for Health and Care Excellence, 2015). We focus
on describing CBT as an evidence-╉based intervention as there have been more studies of CBT than
other sorts of psychological intervention, such as interpersonal therapy. We also describe modifica-
tions of CBT that focus on the behavioral components of CBT, with the aim of targeting the low levels
of positive affect seen in depression in young people. We also briefly consider practical issues and
adaptations that may need to be made when working with depressed children and young people.
2
8
1

182 Emotion Regulation and Depression

Finally, we discuss psychological approaches to reducing and preventing symptoms in young people
without depressive disorder.
CBT focuses on altering dysfunctional styles of thinking and behavior through challenging
negative thoughts and beliefs. CBT programs encourage and support young people in identifying
and evaluating negative thoughts and cognitive distortions with the aim of encouraging more
balanced, less negative, more reflective beliefs about the self, the world and the future. CBT is,
therefore, aiming to target the negative feelings, thoughts and beliefs that are common in depres-
sion. CBT also encourages individuals to exert greater executive control over automatic emotional
reactions (Siegle et al., 2007). It is possible that young children, or those whose meta-​cognitive
skills (i.e., the ability to think about thoughts) are not well developed and may struggle with this
aspect of CBT.
A number of CBT-​based programs have been used in preventive interventions. Preventive inter-
ventions can be delivered to high-​risk groups, including those with known risk factors for a disor-
der or those with high symptoms that fall below the traditional diagnostic (selective or indicated
prevention). Preventive interventions can also be delivered to all members of a group, regardless
of symptoms (universal prevention). CBT programs appear to be effective in preventing depres-
sive symptoms and disorder in those at high-​risk (Garber et al., 2009; Horowitz & Garber, 2006;
Merry, McDowell, Hetrick, Bir, & Muller, 2004; Stice et al., 2009). In particular, the Coping with
Adolescent Stress program has been shown to be effective (Clarke et al., 1995; Garber et al., 2009).
In contrast, CBT-​based programs do not appear to be effective in universal prevention programs
with a very large randomized controlled trial finding no benefit (Stallard et al., 2012).
In our review of the emotion regulation difficulties seen in depression in children and ado-
lescents, we have identified low levels of positive affect as important in predicting the onset of
depression and in interfering with adaptive styles of emotion regulation. To that end, it is also
worth considering behavioral activation and the more behavioral elements of CBT that are rec-
ommended for interventions with depressed young people. Behavioral activation involves activ-
ity monitoring and scheduling and aims to encourage individuals to engage in interesting and
pleasurable activities (Dichter et  al., 2009). Meta-​analysis has indicated that behavioral activa-
tion is effective in reducing symptoms in depressed adults and is as effective as CBT or antide-
pressant treatment (Cuijpers, van Straten, & Warmerdam, 2007). Activity scheduling may be the
“active” element of behavioral activation packages as this alone is effective in reducing depression
in adults (Cuijpers et al., 2007; Jacobson et al., 1996). An appealing aspect of behavioral activation
is that it may be simpler for individuals to understand and for therapists to deliver then cognitive
behavioral therapy. A preliminary trial suggested that behavioral activation may be efficacious in
depressed adolescents (McCauley et al., 2015). Behavioral activation has been shown to alter the
responding of brain areas involved in reward processing (Dichter et al., 2009).
We developed TRY (see later), which incorporated CBT and behavioral activation and focused
on enhancing reward-​processing and tested it as a classroom-​based universal prevention program
(Rice et al., 2015). TRY aimed to enhance reward-​processing through activities such as illustrating
the use of rewarding experiences to lift mood and evaluating potential risk and rewards involved
in day-​to-​day decision making. We also measured reward-​decision making with a behavioral
task pre and post intervention and compared TRY to two other psychological therapies (cogni-
tive behavioral therapy and mindfulness based cognitive therapy) as well as a comparison group.
TRY was the only intervention associated with a reduction in depressive symptoms at follow-​
up. Reward-​seeking increased following TRY. In the TRY program, which focused on increas-
ing sensitivity to rewarding activities, reward seeking increased; this increase was associated
with decreased depressive symptoms. There is, therefore, preliminary evidence that behavioral
3
8
1

Interventions for depression: CBT with children and adults 183

activation may be worth considering as an adjunct to therapeutic and preventive interventions in


young people (McCauley et al., 2015; Rice et al., 2015).

Interventions for depression: CBT with children and adults


The National Institute of Clinical Excellence (NICE) has evaluated CBT as an effective interven-
tion for depression in both children and young people (National Institute for Health and Clinical
Excellence, 2005) and adults (National Institute for Health and Clinical Excellence, 2009). Beck
(1976) proposed that cognition has a key role in the maintenance of depression, and distinguished
between three levels of cognition: core beliefs (or schema), dysfunctional assumptions (or rules
for living) and negative automatic thoughts (NATs; Beck et al., 1979). Specific psychotherapeutic
techniques have been developed in practice with adults to facilitate cognitive change (Greenberger
& Padesky, 1995).
It is less certain whether the three levels of cognition, proposed by Beck, are distinguishable in
children and young people, due to the paucity of studies into the developmental origins of core
beliefs and a theoretical vacuum with regard to the relationship between core beliefs, schemas
and early memories (James, Southam, & Blackburn, 2004). Beck et al. (1979) propose that cog-
nitions are internalized from previous experiences and that an individual’s response to a current
situation is embedded in behavioral and cognitive responses from the past. Therapy with adults
involves seeking to modify these patterns of response, through changing conceptualisations;
however, children are still in the midst of childhood, and therefore, lack responsibility and
the autonomy to instigate changes within their relationships or environments. Mental health
difficulties in children are more likely to be related to the current social context, not to past
relationships where the dysfunctional cognitions developed. Furthermore, cognitive change is
difficult to justify to a child living in a context where their negative beliefs fulfil an adaptive
function. For example, a child that is constantly criticized by adults may develop beliefs related
to feelings of anger, expressed as “it’s not me; it’s not my fault; you are wrong.” To some extent,
such thoughts have a protective and adaptive function in relation to the child’s self-╉concept,
although it is likely to be maladaptive with regard to the development and maintenance of
harmonious relationships. Social cognition, therefore, develops throughout childhood, with
the child gradually internalising thoughts and beliefs during interaction with others (Sharp,
Fonagy, & Goodyer, 2008).
CBT interventions for depression have a number of common features. The aims are to enable
participants to develop an understanding of the associations between cognitions, emotions,
physiology and behavior, to identify “dysfunctional thoughts” or “thinking errors” and become
motivated to develop more realistic, balanced cognitions (Padesky & Mooney, 1990). However,
many of the strategies used in adult therapy require adaptation for children and young people.
Interventions need to take into account a child’s age, cognitive capabilities and educational experi-
ence (Doherr, Reynolds, Wetherly, & Evans, 2005) as all these factors influence their capacity to
engage with cognitive techniques. Younger children tend to use cognitive strategies (e.g., mne-
monic ones) with less frequency and effectiveness than older children (Bjorklund & Douglas,
1997). Similarly, there is a developmental progression in a child’s ability to generalize learned cog-
nitive strategies to other settings or adapt them to different experiences (Crowley & Siegler, 1999;
Siegler, 1996). It is also important to ensure that CBT related activities are scaffolded to support
new understandings (Vygotsky, 1980; Wood, Bruner, & Ross 1976). The concept has been devel-
oped to capture how adults, or more capable peers, provide temporary and adjustable support
during adjustments to materials, presentation and linguistic support. These aspects all influence
4
8
1

184 Emotion Regulation and Depression

the nature of the scaffold, which in CBT with children may be varied according to the child’s
interests and needs.
A central tenet of CBT for depression is psychoeducation and, therefore, there is a need for
therapy to support children’s learning about internal processes linked with cognitive and emo-
tional aspects of functioning. There has been substantial research that has looked at the teaching
approaches most associated with effective learning (see Hattie, 2009). There are clear indications
that, in order for children to grasp new ideas, information needs to link with pre-╉existing under-
standings and be presented in a variety of ways, using multi-╉media to reinforce concepts where
possible (Mayer, 2005). Spaced or distributed practice (i.e., short, multiple practice sessions inter-
spersed with other activities) is important for efficient learning and is much more effective than
practicing a new skill less frequently, but for longer periods (Walker, Greenwood, Hart, & Carta,
1994). The implication of this is that, for example, shorter daily practice is preferable to more
extended weekly practice. Some children may have difficulty with the abstract nature of CBT and
discussion of concepts that are not in the “here and now.” Therefore, another important adaptation
of CBT for children and young people is to present concrete examples in parallel with abstract
concepts, to ensure transfer (Gentner, Rattermann, & Forbus, 1993). This may involve video, role-╉
play, use of stories, model characters, pictures and so on. CBT for children and young people
needs to provide engaging, stimulating activities that are developmentally appropriate and align
with current interests (Fuggle, Dunsmuir, & Curry, 2012).
CBT for children generally combines behavioral and cognitive components (Weisz & Kazdin,
2010), although the degree to which each component is effective, is contestable (Weisz, McCarty,
& Valeri, 2006). Similarly, both cognitive and behavioral methods are central to CBT with adults
for depression, although the extent to which outcomes are influenced by the balance of these
two components is uncertain (Gortner, Gollan, Dobson, & Jacobson, 1998). CBT for children;
therefore, needs to be adapted to ensure that session content is accessible and addresses the level
of metacognitive and executive functioning of the individual. This will require careful planning,
adaptation and simplification of methods, and for many children, an emphasis on more concrete,
behavioral components. As mentioned earlier, one common technique that has been success-
fully integrated within small scale adolescent interventions for depression with positive out-
comes is behavioral activation (BA) (Ritschel, Ramirez, Jones, & Craighead, 2011). BA involves
the young person monitoring and recording daily activity and then exploring the relationship
between mood and activity levels. Social factors are also taken into account due to their impact
on emotional state.

The thinking about reward in young people program (TRY)


We will now describe a recently developed intervention that offers a promising approach to read-
dress the disequilibrium occurring between positive and negative affect in adolescent depression.
The (TRY; Cobbald & Dunsmuir, 2013; Rice et al., 2015) program combines BA within a modi-
fied CBT model that seeks to build awareness of the relationship between physiological signs,
thoughts, feelings and decisions. TRY encourages young people to consider alternatives to the
negative cycles that can become established in affective disorders through focusing on positive
affect and rewarding activities. There is an emphasis on the associations between cognitions and
decision-╉making. The intervention was developed following research suggesting that reward pro-
cessing was altered in, and predicted risk for, adolescent depression, as previously described.
The TRY intervention aimed to incorporate elements of rational reward-╉seeking behavior,
such as encouraging young people to consider the likelihood of good or bad outcomes in their
reward-╉seeking behavior, based on evidence that low reward-╉seeking may be a causal risk factor
for adolescent depression (Forbes et al., 2007; Rawal et al., 2013b). The intervention also aimed
5
8
1

Interventions for depression: CBT with children and adults 185

to encourage the use of happy autobiographical memories and engagement in enjoyable activi-
ties, including social activities as emotion regulation strategies. TRY was developed for delivery
to universal populations of young people aged 14–​15 years, attending co-​educational mainstream
schools. It consists of eight 60-​minute sessions, designed to be delivered on a weekly basis by a
facilitator during Personal Social and Health Education (PSHE) lessons. The session-​by-​session
outline of the TRY program is detailed below:
1 . Introduction; psychoeducation about stress and depression; rationale for the TRY program
2. Goal setting; introduction to the modified CBT model
3. Identifying rewarding experiences and happy memories
4.
Identifying and evaluating thoughts
5.
Decision making with regard to rewarding experiences and the impact on mood
6.
Evaluation of positive experience and how this informs future decision making and practicing
the TRY decision making process
7 . The role of social support and managing conflict
8. Review of TRY program and personal goals
Additional information and multi-​media resources used in the TRY intervention can be accessed
at http://​www.ucl.ac.uk/​educational-​psychology/​try.html.
It is important to recognize that in order to deliver universal therapeutic interventions in schools,
facilitators require a range of competencies. These include interpersonal skills, such as sensitivity
and the ability to consult and negotiate with school staff (Kratochwill, Elliott, & Callan-​Stoiber,
2002; Zins & Erchul, 2002), as well as the ability to develop and maintain good relationships with
young people (Shirk, Karver, & Brown, 2011). Facilitators also need to be appropriately trained,
experienced and knowledgeable about the intervention and have good group management skills
(Lendrum, Humphrey, Kalambouka, & Wigelsworth, 2009), factors all positively related to better
pupil outcomes (Humphrey et al., 2008).
To ensure appropriate standards of TRY program delivery, facilitators were all qualified educa-
tional psychologists, regularly providing services to schools, who had received additional train-
ing in CBT who had attended a one-​day training session in TRY implementation. In addition,
and consistent with recommendations from the British Association of Behavioural and Cognitive
Psychotherapies (BABCP), the Health and Care Professions Council (HCPC) standards of con-
duct, performance and ethics (2012), the British Psychological Society (BPS) Code of Ethics and
Conduct (2009), facilitators of the TRY intervention attended scheduled, two hour group supervi-
sion meetings with an experienced CBT trainer three times during the eight week intervention.
Supervision frameworks were based on several models of CBT supervision (Liese & Alford, 1998;
Liese & Beck, 1997; Pretorius, 2006), within a structured, practical format and incorporating a
didactic function to ensure consistency in the delivery of the intervention, and to share experi-
ences of managing group dynamics and school personnel.
As reported earlier, in a preliminary study, TRY was associated with a reduction in symptoms
of depression and an increase in reward-​seeking and was more effective than standard CBT and
mindfulness based CBT (Rice et  al., 2015). There is, therefore, preliminary evidence that this
intervention, incorporating CBT and behavioral activation and focusing on enhancing reward-​
processing, may be an effective, accessible intervention that can be delivered in universal settings.
Although psychological therapies are the intervention of choice for treating and preventing
depression in children and young people, there is no single “gold-​standard” intervention package
or approach. Basic research on emotion regulation has informed the development of interven-
tion programs. It may be useful to consider the balance of both positive and negative affect when
seeking to reduce depressive symptomatology and the impact of symptoms in children and young
people.
6
8
1

186 Emotion Regulation and Depression

References
American Psychiatric Association. (2013). Diagnostic and Statistical Manual of Mental Disorders (5th ed.).
Washington, DC: APA.
Angold, A., Costello, E. J., Farmer, E. M., Burns, B. J., & Erkanli, A. (1999). Impaired but undiagnosed. J
Am Acad Child Adolesc Psychiatry, 38(2), 129–╉137. doi:10.1097/╉00004583-╉199902000-╉00011
Angold, A., Erkanli, A., Silberg, J., Eaves, L., Costello, E.J. (2002). Depression scale scores in 8-╉17-╉year-╉
olds: effects of age and gender. J Child Psychol Psychiatry, 43(8), 1052–╉1063.
Beck, A. T. (1976). Cognitive therapy and the emotional disorders. Oxford, England: International
Universities Press.
Beck, A. T., Rush, A. J., Shaw, B. F., & Emery, G. (1979). Cognitive therapy of depression. New York: Guilford Press.
Best, J. R., & Miller, P. H. (2010). A developmental perspective on executive function. Child development,
81(6), 1641–╉1660.
Birmaher, B., Ryan, N. D., Williamson, D. E., Brent, D. A., & Kaufman, J. (1996). Childhood and
adolescent depression: a review of the past 10 years. Part II. J Am Acad Child Adolesc Psychiatry, 35(12),
1575–╉1583. doi:10.1097/╉00004583-╉199612000-╉00008
Bjorklund, D. F., & Douglas, R. N. (1997). The development of memory strategies. In N Cowan (Ed.) The
development of memory in childhood (pp. 201–╉246). Hove, England: Psychology Press/╉Erlbaum (UK)
Taylor & Francis; England.
Blackwell, S. E., Browning, M., Mathews, A., Pictet, A., Welch, J., Davies, J., … Holmes, E. A. (2015).
Positive Imagery-╉Based Cognitive Bias Modification as a Web-╉Based Treatment Tool for Depressed
Adults: A Randomized Controlled Trial. Clin Psychol Sci, 3(1), 91–╉111. doi:10.1177/╉2167702614560746
Brenner, S. L., Burns, B. J., Curry, J. F., Silva, S. G., Kratochvil, C. J., & Domino, M. E. (2015). Mental
health service use among adolescents following participation in a randomized clinical trial for
depression. J Clin Child Adolesc Psychol, 44(4), 551–╉558. doi:10.1080/╉15374416.2014.881291
Brewin, C. R. (2006). Understanding cognitive behaviour therapy: A retrieval competition account.
Behaviour Research and Therapy, 44(6), 765–╉784.
British Psychological Society (2009). Code of Ethics and Conduct. Leicester: BPS.
Clarke, G. N., Hawkins, W., Murphy, M., Sheeber, L. B., Lewinsohn, P. M., & Seeley, J. R. (1995).
Targeted prevention of unipolar depressive disorder in an at-╉risk sample of high school
adolescents: A randomized trial of a group cognitive intervention. Journal of the American Academy of
Child & Adolescent Psychiatry, 34(3), 312–╉321.
Cobbald, A., & Dunsmuir, S. (2013). The Thinking about Reward in Young People Facilitator Manual.
UCL: Educational Psychology Group. London.
Cohn, J. F., & Tronick, E. Z. (1983). Three-╉month-╉old infants’ reaction to simulated maternal depression.
Child Dev, 54(1), 185–╉193.
Cole, D. A., Martin, J. M., Peeke, L. A., Seroczynski, A. D., & Fier, J. (1999). Children’s over-╉and
underestimation of academic competence: a longitudinal study of gender differences, depression, and
anxiety. Child Dev, 70(2), 459–╉473.
Connolly, S., Abramson, L., & Alloy, L. (2015). Information processing biases concurrently and
prospectively predict depressive symptoms in adolescents: Evidence from a self-╉referent encoding task.
Cognition & Emotion, 1–╉11.
Conway, M. A., & Pleydell-╉Pearce, C. W. (2000). The construction of autobiographical memories in the
self-╉memory system. Psychological Review, 107(2), 261.
Costello, E. J., Foley, D. L., & Angold, A. (2006). 10-╉year research update review: the epidemiology of
child and adolescent psychiatric disorders: II. Developmental epidemiology. J Am Acad Child Adolesc
Psychiatry, 45(1), 8–╉25. doi:10.1097/╉01.chi.0000184929.41423.c0
Crane, C., Heron, J., Gunnell, D., Lewis, G., Evans, J., & Williams, J. (2014). Childhood traumatic events
and adolescent overgeneral autobiographical memory: Findings in a UK cohort. Journal of Behavior
Therapy and Experimental Psychiatry, 45(3), 330–╉338.
7
8
1

Interventions for depression: CBT with children and adults 187

Crowley, K., & Siegler, R. S. (1999). Explanation and generalization in young children’s strategy learning.
Child development, 70(2), 304–​316.
Cuijpers, P., van Straten, A., & Warmerdam, L. (2007). Behavioral activation treatments of depression: a
meta-​analysis. Clin Psychol Rev, 27(3), 318–​326. doi:10.1016/​j.cpr.2006.11.001
Dalgleish, T., Neshat-​Doost, H., Taghavi, R., Moradi, A., Yule, W., Canterbury, R., & Vostanis, P. (1998).
Information processing in recovered depressed children and adolescents. J Child Psychol Psychiatry,
39(7), 1031–​1035.
Dalgleish, T., Taghavi, R., Neshat-​Doost, H., Moradi, A., Yule, W., & Canterbury, R. (1997). Information
processing in clinically depressed and anxious children and adolescents. J Child Psychol Psychiatry,
38(5), 535–​541.
Dalgleish, T., & Werner-​Seidler, A. (2014). Disruptions in autobiographical memory processing in
depression and the emergence of memory therapeutics. Trends Cogn Sci, 18(11), 596–​604. doi:10.1016/​
j.tics.2014.06.010
Davidovich, S., Collishaw, S., Thapar, A. K., Harold, G., Thapar, A., & Rice, F. (2015). Do better executive
functions buffer the effect of current parental depression on adolescent depressive symptoms? Manuscript
submitted for publication.
De Lissnyder, E., Koster, E. H., Derakshan, N., & De Raedt, R. (2010). The association between
depressive symptoms and executive control impairments in response to emotional and non-​emotional
information. Cognition and Emotion, 24(2), 264–​280.
Dearing, K. F., & Gotlib, I. H. (2009). Interpretation of ambiguous information in girls at risk for
depression. J Abnorm Child Psychol, 37(1), 79–​91. doi:10.1007/​s10802-​008-​9259-​z
Depue, R. A., & Iacono, W. G. (1989). Neurobehavioral aspects of affective disorders. Annu Rev Psychol, 40,
457–​492. doi:10.1146/​annurev.ps.40.020189.002325
Diamond, A. (2013). Executive functions. Annual review of psychology, 64, 135.
Dichter, G. S., Felder, J. N., Petty, C., Bizzell, J., Ernst, M., & Smoski, M. J. (2009). The effects of
psychotherapy on neural responses to rewards in major depression. Biol Psychiatry, 66(9), 886–​897.
doi:10.1016/​j.biopsych.2009.06.021
Doherr, L., Reynolds, S., Wetherly, J., & Evans, E. H. (2005). Young children’s ability to engage in
cognitive therapy tasks: associations with age and educational experience. Behavioural and Cognitive
Psychotherapy, 33(02), 201–​215.
Eaves, L. J., Silberg, J. L., Meyer, J. M., Maes, H. H., Simonoff, E., Pickles, A., Rutter, M., Neale, M.
C., Reynolds, C. A., Erikson, M. T., Heath, A. C., Loeber, R., Truett, K. R., & Hewitt, J. K. (1997).
Genetics and developmental psychopathology: 2. The main effects of genes and environment on
behavioral problems in the Virginia Twin Study of Adolescent Behavioral Development. J Child Psychol
Psychiatry, 38, 965–​980.
Forbes, E. E., & Dahl, R. E. (2005). Neural systems of positive affect: relevance to understanding child and
adolescent depression? Dev Psychopathol, 17(3), 827–​850. doi:10.1017/​S095457940505039X
Forbes, E. E., Hariri, A. R., Martin, S. L., Silk, J. S., Moyles, D. L., Fisher, P. M., … Dahl, R. E. (2009).
Altered striatal activation predicting real-​world positive affect in adolescent major depressive disorder.
Am J Psychiatry, 166(1), 64–​73. doi:10.1176/​appi.ajp.2008.07081336
Forbes, E. E., Shaw, D. S., & Dahl, R. E. (2007). Alterations in reward-​related decision making in boys with
recent and future depression. Biol Psychiatry, 61(5), 633–​639. doi:10.1016/​j.biopsych.2006.05.026
Fuggle, P., Dunsmuir, S., & Curry, V. (2012). CBT with children, young people and families. London:
SAGE.
Garber, J., Clarke, G. N., Weersing, V. R., Beardslee, W. R., Brent, D. A., Gladstone, T. R., … Hollon, S. D.
(2009). Prevention of depression in at-​risk adolescents: a randomized controlled trial. Jama, 301(21),
2215–​2224.
Gençöz, T., Voelz, Z. R., Gençöz, F., Pettit, J. W., & Joiner, T. E. (2001). Specificity of information
processing styles to depressive symptoms in youth psychiatric inpatients. Journal of Abnormal Child
Psychology, 29(3), 255–​262.
81

188 Emotion Regulation and Depression

Gentner, D., Rattermann, M. J., & Forbus, K. D. (1993). The roles of similarity in transfer: Separating
retrievability from inferential soundness. Cognitive Psychology, 25(4), 524–​575.
Gibb, B. E., Benas, J. S., Grassia, M., & McGeary, J. (2009). Children’s attentional biases and 5-​HTTLPR
genotype: potential mechanisms linking mother and child depression. J Clin Child Adolesc Psychol,
38(3), 415–​426. doi:10.1080/​15374410902851705
Goddard, L., Dritschel, B., & Burton, A. (1996). Role of autobiographical memory in social problem
solving and depression. Journal of Abnormal Psychology, 105(4), 609.
Goodyer, I.M., Reynolds, S., Barrett, B., Byford, S., Dubicka, B., Hill, J., … Fonagy, P. (2017). Cognitive
behavioural therapy and short-term psychoanalytical psychotherapy versus a brief psychosocial
intervention in adolescents with unipolar major depressive disorder (IMPACT): a multicentre,
pragmatic, observer-blind, randomised controlled superiority trial. Lancet Psychiatry, 4, 109–119.
Goldstein, B. L., Hayden, E. P., Klein, D. N. (2014). Stability of self-​referent encoding task performance and
associations with change in depressive symptoms from early to middle childhood. Cogn Emot, 22, 1–​11.
Gortner, E. T., Gollan, J. K., Dobson, K. S., & Jacobson, N. S. (1998). Cognitive–​behavioral treatment for
depression: Relapse prevention. Journal of Consulting and Clinical Psychology, 66(2), 377.
Gotlib, I. H., & Joormann, J. (2010). Cognition and depression: current status and future directions. Annual
Review of Clinical Psychology, 6, 285.
Greenberger, D., & Padesky, C. A. (1995). Mind over mood: Guilford: New York.
Guyer, A. E., Kaufman, J., Hodgdon, H. B., Masten, C. L., Jazbec, S., Pine, D. S., & Ernst, M.
(2006). Behavioral alterations in reward system function: the role of childhood maltreatment
and psychopathology. J Am Acad Child Adolesc Psychiatry, 45(9), 1059–​1067. doi:10.1097/​
01.chi.0000227882.50404.11
Hammen, C. (1991). Generation of stress in the course of unipolar depression. J Abnorm Psychol, 100(4),
555–​561.
Hammen, C., & Zupan, B. A. (1984). Self-​schemas, depression, and the processing of personal information
in children. Journal of Experimental Child Psychology, 37(3), 598–​608.
Hankin, B. L., & Abramson, L. Y. (2001). Development of gender differences in depression: An elaborated
cognitive vulnerability-​transactional stress theory. Psychological Bulletin, 127(6), 773–​796.
Harmer, C. J., Goodwin, G. M., & Cowen, P. J. (2009). Why do antidepressants take so long to work?
A cognitive neuropsychological model of antidepressant drug action. The British Journal of Psychiatry,
195(2), 102–​108.
Harmer, C. J., O’Sullivan, U., Favaron, E., Massey-​Chase, R., Ayres, R., Reinecke, A., … Cowen, P. J.
(2009). Effect of acute antidepressant administration on negative affective bias in depressed patients.
Am J Psychiatry, 166(10), 1178–​1184. doi:10.1176/​appi.ajp.2009.09020149
Harrington, R., Fudge, H., Rutter, M., Pickles, A., & Hill, J. (1991). Adult outcomes of childhood and
adolescent depression: II. Links with antisocial disorders. J Am Acad Child Adolesc Psychiatry, 30(3),
434–​439. doi: 10.1097/​00004583-​199105000-​00013
Harrington, R., Rutter, M., Weissman, M., Fudge, H., Groothues, C., Bredenkamp, D., … Wickramaratne,
P. (1997). Psychiatric disorders in the relatives of depressed probands I. Comparison of prepubertal,
adolescent and early adult onset cases. Journal of Affective Disorders, 42(1), 9–​22.
Harrington, R. C., Fudge, H., Rutter, M. L., Bredenkamp, D., Groothues, C., & Pridham, J. (1993). Child
and adult depression: a test of continuities with data from a family study. Br J Psychiatry,
162, 627–​633.
Harrison, A. J., & Gibb, B. E. (2014). Attentional Biases in Currently Depressed Children: An Eye-​
Tracking Study of Biases in Sustained Attention to Emotional Stimuli. J Clin Child Adolesc Psychol, 1–​7.
doi:10.1080/​15374416.2014.930688
Hattie, J. (2009). Visible learning: A synthesis of 800 + meta-​analyses on achievement. Abingdon: Routledge.
Hayes, S. C., Luoma, J. B., Bond, F. W., Masuda, A., & Lillis, J. (2006). Acceptance and commitment
therapy: Model, processes and outcomes. Behaviour Research and Therapy, 44(1), 1–​25.
9
8
1

Interventions for depression: CBT with children and adults 189

Health and Care Professions Council (2012). Standards of Conduct, Performance and Ethics.
London: HCPC Retrieved from http://​www.hpc-​uk.org/​aboutregistration/​standards/​
standardsofconductperformanceandethics/​.
Hertel, P. T., & Hardin, T. S. (1990). Remembering with and without awareness in a depressed
mood: evidence of deficits in initiative. Journal of Experimental Psychology: General, 119(1), 45.
Hipwell, A. E., Sapotichne, B., Klostermann, S., Battista, D., & Keenan, K. (2011). Autobiographical
memory as a predictor of depression vulnerability in girls. Journal of Clinical Child & Adolescent
Psychology, 40(2), 254–​265.
Horowitz, J. L., & Garber, J. (2006). The prevention of depressive symptoms in children and adolescents: A
meta-​analytic review. Journal of Consulting and Clinical Psychology, 74(3), 401.
Hughes, C., Graham, A., & Grayson, A. (2004). Executive functions in childhood: Development
and disorder. In Cognitive and language development in children (pp. 207–​228). Oxford: Oxford
University Press.
Humphrey, N. A., Kalambouka, J., Bolton, A., Lendrum, M., Wigelsworth, C. L., & Farrell, P. (2008).
Primary social and emotional aspects of learning: Evaluation of small group work. (Research Report
RR064). Nottingham: DCSF.
Ingram, R. E., & Ritter, J. (2000). Vulnerability to depression: cognitive reactivity and parental bonding in
high-​risk individuals. J Abnorm Psychol, 109(4), 588–​596.
Insel, T., Cuthbert, B., Garvey, M., Heinssen, R., Pine, D. S., Quinn, K., … Wang, P. (2010). Research
domain criteria (RDoC): toward a new classification framework for research on mental disorders. Am J
Psychiatry, 167(7), 748–​751. doi:10.1176/​appi.ajp.2010.09091379
Jacobson, N. S., Dobson, K. S., Truax, P. A., Addis, M. E., Koerner, K., Gollan, J. K., … Prince, S. E.
(1996). A component analysis of cognitive-​behavioral treatment for depression. J Consult Clin Psychol,
64(2), 295–​304.
James, I. A., Southam, L., & Blackburn, I. M. (2004). Schemas revisited. Clinical Psychology &
Psychotherapy, 11(6), 369–​377.
John, O. P., & Gross, J. J. (2007). Individual Differences in Emotion Regulation. In J. J. Gross (Ed.),
Handbook of emotion regulation (pp. 351–​372). New York, NY: Guilford Press; US.
Joormann, J., Gilbert, K., & Gotlib, I. H. (2010). Emotion identification in girls at high risk for depression.
J Child Psychol Psychiatry, 51(5), 575–​582. doi: 10.1111/​j.1469-​7610.2009.02175.x
Joormann, J., & Quinn, M. E. (2014). Cognitive processes and emotion regulation in depression. Depression
and anxiety, 31(4), 308–​315.
Joormann, J., Siemer, M., & Gotlib, I. H. (2007). Mood regulation in depression: Differential effects
of distraction and recall of happy memories on sad mood. J Abnorm Psychol, 116(3), 484–​490.
doi:10.1037/​0021-​843X.116.3.484
Joormann, J., Talbot, L., & Gotlib, I. H. (2007a). Biased processing of emotional information in girls at risk
for depression. Journal of Abnormal Psychology, 116(1), 135.
Joormann, J., Talbot, L., & Gotlib, I. H. (2007b). Biased processing of emotional information in girls at risk
for depression. J Abnorm Psychol, 116(1), 135–​143. doi:10.1037/​0021-​843X.116.1.135
Kelvin, R. G., Goodyer, I. M., Teasdale, J. D., & Brechin, D. (1999). Latent negative self-​schema and high
emotionality in well adolescents at risk for psychopathology. J Child Psychol Psychiatry, 40(6), 959–​968.
Kendler, K. S., Thornton, L. M., & Gardner, C. O. (2000). Stressful life events and previous episodes in the
etiology of major depression in women: an evaluation of the “kindling” hypothesis. Am J Psychiatry,
157(8), 1243–​1251.
Kilford, E. J., Foulkes, L., Potter, R., Collishaw, S., Thapar, A., & Rice, F. (2015). Affective bias and current,
past and future adolescent depression: A familial high risk study. Journal of Affective Disorders, 174,
265–​271.
Kovacs, M., & Lopez-​Duran, N. (2010). Prodromal symptoms and atypical affectivity as predictors of
major depression in juveniles: implications for prevention. J Child Psychol Psychiatry, 51(4), 472–​496.
doi:10.1111/​j.1469-​7610.2010.02230.x
0
9
1

190 Emotion Regulation and Depression

Kratochwill, T. R., Elliott, S. N., & Callan-​Stoiber, K. (2002). Best Practices in School-​Based Problem-​
Solving Consultation. In A.Thomas & J. Grimes (Eds.) Best practices in school psychology IV (Vol 1, Vol
2) (pp. 583–​608). Washington, DC: National Association of School Psychologists; US.
Kross, E., Ayduk, O., & Mischel, W. (2005). When asking “why” does not hurt distinguishing rumination
from reflective processing of negative emotions. Psychological Science, 16(9), 709–​715.
Kujawa, A. J., Torpey, D., Kim, J., Hajcak, G., Rose, S., Gotlib, I. H., & Klein, D. N. (2011). Attentional
biases for emotional faces in young children of mothers with chronic or recurrent depression. Journal of
Abnormal Child Psychology, 39(1), 125–​135.
Kuyken, W., Howell, R., & Dalgleish, T. (2006). Overgeneral autobiographical memory in depressed
adolescents with, versus without, a reported history of trauma. Journal of Abnormal Psychology,
115(3), 387.
Kyte, Z. A., Goodyer, I. M., & Sahakian, B. J. (2005). Selected executive skills in adolescents with
recent first episode major depression. J Child Psychol Psychiatry, 46(9), 995–​1005. doi:10.1111/​
j.1469-​7610.2004.00400.x
Lendrum, A., Humphrey, N., Kalambouka, A., & Wigelsworth, M. (2009). Implementing primary
Social and Emotional Aspects of Learning (SEAL) small group interventions: Recommendations for
practitioners. Emotional and Behavioural Difficulties, 14(3), 229–​238.
Lewinsohn, P. M. (1975). Engagement in pleasant activities and depression level. J Abnorm Psychol, 84(6),
729–​731.
Lewinsohn, P. M., & Graf, M. (1973). Pleasant activities and depression. J Consult Clin Psychol, 41(2),
261–​268.
Lewinsohn, P. M., Rohde, P., & Seeley, J. R. (1998). Treatment of adolescent depression: frequency of
services and impact on functioning in young adulthood. Depress Anxiety, 7(1), 47–​52.
Lewinsohn, P. M., Zeiss, A. M., & Duncan, E. M. (1989). Probability of relapse after recovery from an
episode of depression. J Abnorm Psychol, 98(2), 107–​116.
Liese, B. S., & Alford, B. (1998). Recent advances in cognitive therapy supervision. Journal of Cognitive
Psychotherapy, 12, 91–​94.
Liese, B. S., & Beck, J. S. (1997). Cognitive therapy supervision. In C. E. Watkins (Ed.), Handbook of
psychotherapy supervision (pp. 114–​133). Hoboken, NJ: John Wiley & Sons Inc; US.
Lloyd, G., & Lishman, W. (1975). Effect of depression on the speed of recall of pleasant and unpleasant
experiences. Psychological Medicine, 5(02), 173–​180.
Lopez-​Duran, N. L., Kuhlman, K. R., George, C., & Kovacs, M. (2013). Facial emotion expression
recognition by children at familial risk for depression: high-​risk boys are oversensitive to sadness. J
Child Psychol Psychiatry, 54(5), 565–​574. doi:10.1111/​jcpp.12005
Maalouf, F. T., Brent, D., Clark, L., Tavitian, L., McHugh, R. M., Sahakian, B. J., & Phillips, M. L. (2011).
Neurocognitive impairment in adolescent major depressive disorder: state vs. trait illness markers. J
Affect Disord, 133(3), 625–​632. doi:10.1016/​j.jad.2011.04.041
Mathews, A., & MacLeod, C. (2005). Cognitive vulnerability to emotional disorders. Annu Rev Clin Psychol,
1, 167–​195. doi: 10.1146/​annurev.clinpsy.1.102803.143916
Mayer, R. E. (2005). The Cambridge handbook of multimedia learning: Cambridge University Press.
McCauley, E., Gudmundsen, G., Schloredt, K., Martell, C., Rhew, I., Hubley, S., & Dimidjian, S. (2015).
The Adolescent Behavioral Activation Program: Adapting Behavioral Activation as a Treatment for
Depression in Adolescence. J Clin Child Adolesc Psychol, 1–​14. doi:10.1080/​15374416.2014.979933
McMakin, D. L., Olino, T. M., Porta, G., Dietz, L. J., Emslie, G., Clarke, G., … Birmaher, B. (2012).
Anhedonia predicts poorer recovery among youth with selective serotonin reuptake inhibitor
treatment–​resistant depression. Journal of the American Academy of Child & Adolescent Psychiatry,
51(4), 404–​411.
McRae, K., Jacobs, S. E., Ray, R. D., John, O. P., & Gross, J. J. (2012). Individual differences in reappraisal
ability: Links to reappraisal frequency, well-​being, and cognitive control. Journal of Research in
Personality, 46(1), 2–​7.
1
9

Interventions for depression: CBT with children and adults 191

Merry, S., McDowell, H., Hetrick, S., Bir, J., & Muller, N. (2004). Psychological and/​or educational
interventions for the prevention of depression in children and adolescents (Cochrane review). The
cochrane library, 2 CD003380.
Miranda, J., & Persons, J. B. (1988). Dysfunctional attitudes are mood-​state dependent. J Abnorm Psychol,
97(1), 76–​79.
Monroe, S. M., Rohde, P., Seeley, J. R., & Lewinsohn, P. M. (1999). Life events and depression in
adolescence: relationship loss as a prospective risk factor for first onset of major depressive disorder. J
Abnorm Psychol, 108(4), 606–​614.
Murray, L., Woolgar, M., Cooper, P., & Hipwell, A. (2001). Cognitive vulnerability to depression in
5‐year‐old children of depressed mothers. Journal of Child Psychology and Psychiatry, 42(7), 891–​899.
National Institute for Health and Care Excellence (2015). Depression in children and young
people: identification and management in primary, community and secondary care. NICE guideline
(CG28.1). Retrieved from http://​www.nice.org.uk/​guidance/​cg28/​evidence
National Institute for Health and Clinical Excellence (2005). Depression in children and young
people: identification and management in primary, community and secondary care. NICE guideline
(CG28). Retrieved from http://​www.nice.org.uk/​guidance/​cg28/​chapter/​1-​guidance
National Institute for Health and Clinical Excellence (2009). Depression in adults: the treatment and
management of depression in adults. NICE guideline (CG90). Retrieved from https://​www.nice.org.uk/​
guidance/​cg90
Nolen-​Hoeksema, S. (1991). Responses to depression and their effects on the duration of depressive
episodes. J Abnorm Psychol, 100(4), 569–​582.
Nolen-​Hoeksema, S., Wisco, B. E., & Lyubomirsky, S. (2008). Rethinking Rumination. Perspect Psychol Sci,
3(5), 400–​424. doi:10.1111/​j.1745-​6924.2008.00088.x
Ochsner, K. N., & Gross, J. J. (2008). Cognitive emotion regulation insights from social cognitive and
affective neuroscience. Current Directions in Psychological Science, 17(2), 153–​158.
Olino, T. M., Lopez-​Duran, N. L., Kovacs, M., George, C. J., Gentzler, A. L., & Shaw, D. S. (2011).
Developmental trajectories of positive and negative affect in children at high and low familial risk for
depressive disorder. J Child Psychol Psychiatry, 52(7), 792–​799. doi:10.1111/​j.1469-​7610.2010.02331.x
Padesky, C., & Mooney, K. (1990). Clinical tip: Presenting the cognitive model to clients. International
Cognitive Therapy Newsletter, 6(1), 13–​14.
Park, R. J., Goodyer, I., & Teasdale, J. (2002). Categoric overgeneral autobiographical memory in
adolescents with major depressive disorder. Psychological Medicine, 32(02), 267–​276.
Paula, C. S., Bordin, I. A., Mari, J. J., Velasque, L., Rohde, L. A., & Coutinho, E. S. (2014). The mental
health care gap among children and adolescents: data from an epidemiological survey from four
Brazilian regions. PLoS One, 9(2), e88241. doi:10.1371/​journal.pone.0088241
Picard, L., Reffuveille, I., Eustache, F., & Piolino, P. (2009). Development of autonoetic autobiographical
memory in school-​age children: genuine age effect or development of basic cognitive abilities?
Consciousness and Cognition, 18(4), 864–​876.
Pickles, A., Rowe, R., Simonoff, E., Foley, D., Rutter, M., & Silberg, J. (2001). Child psychiatric
symptoms and psychosocial impairment: relationship and prognostic significance. Br J Psychiatry, 179,
230–​235.
Pine, D. S., Cohen, E., Cohen, P., & Brook, J. (1999). Adolescent depressive symptoms as predictors of adult
depression: moodiness or mood disorder? Am J Psychiatry, 156(1), 133–​135.
Pizzagalli, D. A. (2014). Depression, stress, and anhedonia: toward a synthesis and integrated model. Annu
Rev Clin Psychol, 10, 393–​423. doi:10.1146/​annurev-​clinpsy-​050212-​185606
Potter, R., Mars, B., Eyre, O., Legge, S., Ford, T., Sellers, R., … Thapar, A. K. (2012). Missed
opportunities: mental disorder in children of parents with depression. Br J Gen Pract, 62(600), e487–​
e493. doi:10.3399/​bjgp12X652355
Pretorius, W. M. (2006). Cognitive behavioural therapy supervision: recommended practice. Behavioural
and Cognitive Psychotherapy, 34(04), 413–​420.
2
9
1

192 Emotion Regulation and Depression

Prieto, S. L., Cole, D. A., & Tageson, C. W. (1992). Depressive self-​schemas in clinic and nonclinic children.
Cognitive Therapy and Research, 16(5), 521–​534.
Raes, F., Hermans, D., Williams, J. M. G., Demyttenaere, K., Sabbe, B., Pieters, G., & Eelen, P. (2005).
Reduced specificity of autobiographical memory: A mediator between rumination and ineffective social
problem-​solving in major depression? Journal of Affective Disorders, 87(2), 331–​335.
Raes, F., Verstraeten, K., Bijttebier, P., Vasey, M. W., & Dalgleish, T. (2010). Inhibitory control mediates the
relationship between depressed mood and overgeneral memory recall in children. Journal of Clinical
Child & Adolescent Psychology, 39(2), 276–​281.
Ramirez, S., Liu, X., MacDonald, C. J., Moffa, A., Zhou, J., Redondo, R. L., & Tonegawa, S. (2015).
Activating positive memory engrams suppresses depression-​like behaviour. Nature, 522(7556), 335–​
339. doi: 10.1038/​nature14514
Rawal, A., Collishaw, S., Thapar, A., & Rice, F. (2013a). A direct method of assessing underlying
cognitive risk for adolescent depression. J Abnorm Child Psychol, 41(8), 1279–​1288. doi:10.1007/​
s10802-​013-​9760-​x
Rawal, A., Collishaw, S., Thapar, A., & Rice, F. (2013b). “The risks of playing it safe”: a prospective
longitudinal study of response to reward in the adolescent offspring of depressed parents. Psychol Med,
43(1), 27–​38. doi:10.1017/​S0033291712001158
Rawal, A., & Rice, F. (2012a). Examining overgeneral autobiographical memory as a risk factor for
adolescent depression. Journal of the American Academy of Child & Adolescent Psychiatry, 51(5),
518–​527.
Rawal, A., & Rice, F. (2012b). A longitudinal study of processes predicting the specificity of
autobiographical memory in the adolescent offspring of depressed parents. Memory, 20(5), 518–​526.
Rawal, A., Riglin, L., Ng-​Knight, T., Collishaw, S., Thapar, A., & Rice, F. (2014). A longitudinal high-​risk
study of adolescent anxiety, depression and parent-​severity on the developmental course of risk-​
adjustment. J Child Psychol Psychiatry, 55(11), 1270–​1278. doi:10.1111/​jcpp.12279
Rende, R., Warner, V., Wickramarante, P., & Weissman, M. M. (1999). Sibling aggregation for psychiatric
disorders in offspring at high and low risk for depression: 10-​year follow-​up. Psychol Med, 29(6),
1291–​1298.
Rice, F. (2010). Genetics of childhood and adolescent depression: insights into etiological heterogeneity and
challenges for future genomic research. Genome Med, 2(9), 68.
Rice, F., Harold, G., & Thapar, A. (2002). The genetic aetiology of childhood depression: a review. J Child
Psychol Psychiatry, 43(1), 65–​79.
Rice, F., Harold, G. T., & Thapar, A. (2003). Negative life events as an account of age‐related differences
in the genetic aetiology of depression in childhood and adolescence. Journal of Child Psychology and
Psychiatry, 44(7), 977–​987.
Rice, F., Rawal, A., Riglin, L., Lewis, G., Lewis, G., & Dunsmuir, S. (2015). Examining reward-​seeking,
negative self-​beliefs and over-​general autobiographical memory as mechanisms of change in classroom
prevention programs for adolescent depression. Journal of Affective Disorders, 186, 320–​327.
Riglin, L., Petrides, K. V., Frederickson, N., & Rice, F. (2014). The relationship between emotional
problems and subsequent school attainment: a meta-​analysis. J Adolesc, 37(4), 335–​346. doi:10.1016/​
j.adolescence.2014.02.010
Ritschel, L. A., Ramirez, C. L., Jones, M., & Craighead, W. E. (2011). Behavioral activation for depressed
teens: A pilot study. Cognitive and Behavioral Practice, 18(2), 281–​299.
Roiser, J. P., Elliott, R., & Sahakian, B. J. (2012). Cognitive mechanisms of treatment in depression.
Neuropsychopharmacology, 37(1), 117–​136.
Rood, L., Roelofs, J., Bogels, S. M., Nolen-​Hoeksema, S., & Schouten, E. (2009). The influence of emotion-​
focused rumination and distraction on depressive symptoms in non-​clinical youth: a meta-​analytic
review. Clin Psychol Rev, 29(7), 607–​616. doi 10.1016/​j.cpr.2009.07.001
Rudolph, K. D. (2002). Gender differences in emotional responses to interpersonal stress during
adolescence. Journal of Adolescent Health, 30(4), 3–​13.
3
9
1

Interventions for depression: CBT with children and adults 193

Rushton, J. L., Forcier, M., & Schectman, R. M. (2002). Epidemiology of depressive symptoms in the
National Longitudinal Study of Adolescent Health. J Am Acad Child Adolesc Psychiatry, 41(2), 199–​205.
doi:10.1097/​00004583-​200202000-​00014
Rutter, M., Pickles, A., Murray, R., & Eaves, L. (2001). Testing hypotheses on specific environmental causal
effects on behavior. Psychol Bull, 127(3), 291–​324.
Sayal, K., Yates, N., Spears, M., & Stallard, P. (2014). Service use in adolescents at risk of depression
and self-​harm: prospective longitudinal study. Soc Psychiatry Psychiatr Epidemiol, 49(8), 1231–​1240.
doi:10.1007/​s00127-​014-​0843-​y
Schepman, K., Fombonne, E., Collishaw, S., & Taylor, E. (2014). Cognitive styles in depressed
children with and without comorbid conduct disorder. J Adolesc, 37(5), 622–​631. doi:10.1016/​
j.adolescence.2014.04.004
Schepman, K., Taylor, E., Collishaw, S., & Fombonne, E. (2012). Face emotion processing in depressed
children and adolescents with and without comorbid conduct disorder. J Abnorm Child Psychol, 40(4),
583–​593. doi: 10.1007/​s10802-​011-​9587-​2
Schmeichel, B. J., & Tang, D. (2014). The relationship between individual differences in executive
functioning and emotion regulation: A comprehensive review. In J. P. Forgas & E. Harmon-​Jones (Eds.),
The control within: Motivation and its regulation (pp. 133–​151). New York: Psychology Press.
Scourfield, J., Rice, F., Thapar, A., Harold, G. T., Martin, N., & McGuffin, P. (2003). Depressive symptoms
in children and adolescents: changing aetiological influences with development. Journal of Child
Psychology and Psychiatry, 44(7), 968–​976.
Seligman, M. E., Peterson, C., Kaslow, N. J., Tanenbaum, R. L., Alloy, L. B., & Abramson, L. Y. (1984).
Attributional style and depressive symptoms among children. J Abnorm Psychol, 93(2), 235–​238.
Sharot, T. (2011). The optimism bias. Curr Biol, 21(23), R941–​R945. doi: 10.1016/​j.cub.2011.10.030
Sharp, C., Fonagy, P., & Goodyer, I. (2008). Social cognition and developmental psychopathology
Social cognition and developmental psychopathology (pp. xvi, 480). New York, NY: Oxford University
Press; US.
Shirk, S. R., Karver, M. S., & Brown, R. (2011). The alliance in child and adolescent psychotherapy.
Psychotherapy, 48(1), 17–​24.
Siegle, G. J., Ghinassi, F., & Thase, M. E. (2007). Neurobehavioral therapies in the 21st century: Summary
of an emerging field and an extended example of cognitive control training for depression. Cognitive
Therapy and Research, 31(2), 235–​262.
Siegler, R. S. (1996). Emerging minds: The process of change in children’s thinking. New York: Oxford
University Press.
Skegg, K. (2005). Self-​harm. The Lancet, 366(9495), 1471–​1483.
Solomon, D. A., Keller, M. B., Leon, A. C., Mueller, T. I., Lavori, P. W., Shea, M. T., … Endicott, J. (2000).
Multiple recurrences of major depressive disorder. Am J Psychiatry, 157(2), 229–​233.
Stallard, P., Sayal, K., Phillips, R., Taylor, J. A., Spears, M., Anderson, R., … Montgomery, A. A. (2012).
Classroom based cognitive behavioural therapy in reducing symptoms of depression in high risk
adolescents: pragmatic cluster randomised controlled trial. BMJ, 345, e6058. doi:10.1136/​bmj.e6058
Stice, E., Shaw, H., Bohon, C., Marti, C. N., & Rohde, P. (2009). A meta-​analytic review of depression
prevention programs for children and adolescents: factors that predict magnitude of intervention
effects. Journal of Consulting and Clinical Psychology, 77(3), 486.
Stringaris, A., Vidal-​Ribas Belil, P., Artiges, E., Lemaitre, H., Gollier-​Briant, F., Wolke, S., … Consortium,
I. (2015). The Brain’s Response to Reward Anticipation and Depression in Adolescence: Dimensionality,
Specificity, and Longitudinal Predictions in a Community-​Based Sample. Am J Psychiatry, 172(12),
1215–​1223.
Taylor, L., & Ingram, R. E. (1999). Cognitive reactivity and depressotypic information processing in
children of depressed mothers. J Abnorm Psychol, 108(2), 202–​210.
Teasdale, J. D. (1988). Cognitive vulnerability to persistent depression. Cognition and Emotion, 2,
247–​274.
4
9
1

194 Emotion Regulation and Depression

Teasdale, J. D., Moore, R. G., Hayhurst, H., Pope, M., Williams, S., & Segal, Z. V. (2002). Metacognitive
awareness and prevention of relapse in depression: empirical evidence. J Consult Clin Psychol, 70(2),
275–​287.
Termine, N. T., & Izard, C. E. (1988). Infants Responses to Their Mothers Expressions of Joy and Sadness.
Developmental Psychology, 24(2), 223–​229. doi: 10.1037/​0012-​1649.24.2.223
Thapar, A., Collishaw, S., Pine, D. S., & Thapar, A. K. (2012). Depression in adolescence. Lancet, 379(9820),
1056–​1067. doi:10.1016/​S0140-​6736(11)60871-​4
Thapar, A., & McGuffin, P. (1994). A twin study of depressive symptoms in childhood. The British Journal
of Psychiatry, 165(2), 259–​265.
Timbremont, B., & Braet, C. (2004a). Cognitive vulnerability in remitted depressed children and
adolescents. Behaviour Research and Therapy, 42(4), 423–​437.
Timbremont, B., & Braet, C. (2004b). Cognitive vulnerability in remitted depressed children
and adolescents. Behav Res Ther, 42(4), 423–​437. doi:10.1016/​S0005-​7967(03)00151-​7
S0005796703001517 [pii]
Vrielynck, N., Deplus, S., & Philippot, P. (2007). Overgeneral autobiographical memory and depressive
disorder in children. Journal of Clinical Child and Adolescent Psychology, 36(1), 95–​105.
Vygotsky, L. S. (1980). Mind in society: The development of higher psychological processes. Harvard: Harvard
university press.
Wagner, S., Doering, B., Helmreich, I., Lieb, K., & Tadić, A. (2012). A meta‐analysis of executive
dysfunctions in unipolar major depressive disorder without psychotic symptoms and their changes
during antidepressant treatment. Acta Psychiatrica Scandinavica, 125(4), 281–​292.
Wagner, S., Müller, C., Helmreich, I., Huss, M., & Tadić, A. (2014). A meta-​analysis of cognitive functions
in children and adolescents with major depressive disorder. European Child & Adolescent Psychiatry,
24(1), 5–​19.
Walker, D., Greenwood, C., Hart, B., & Carta, J. (1994). Prediction of school outcomes based on early
language production and socioeconomic factors. Child Development, 65(2), 606–​621.
Warner, V., Weissman, M. M., Mufson, L., & Wickramaratne, P. J. (1999). Grandparents, parents, and
grandchildren at high risk for depression: a three-​generation study. Journal of the American Academy of
Child & Adolescent Psychiatry, 38(3), 289–​296.
Watkins, E. R. (2008). Constructive and unconstructive repetitive thought. Psychol Bull, 134(2), 163–​206.
doi:10.1037/​0033-​2909.134.2.163
Weissman, M. M., Warner, V., Wickramaratne, P., Moreau, D., & Olfson, M. (1997). Offspring of
depressed parents: 10 years later. Archives of General Psychiatry, 54(10), 932–​940.
Weisz, J. R., & Kazdin, A. E. (2010). Evidence-​based psychotherapies for children and adolescents.
New York: Guilford Press.
Weisz, J. R., McCarty, C. A., & Valeri, S. M. (2006). Effects of psychotherapy for depression in children and
adolescents: a meta-​analysis. Psychological Bulletin, 132(1), 132.
Wenzlaff, R. M., & Bates, D. E. (1998). Unmasking a cognitive vulnerability to depression: how lapses in
mental control reveal depressive thinking. J Pers Soc Psychol, 75(6), 1559–​1571.
Werner-​Seidler, A., & Moulds, M. L. (2011). Autobiographical memory characteristics in depression
vulnerability: formerly depressed individuals recall less vivid positive memories. Cogn Emot, 25(6),
1087–​1103. doi:10.1080/​02699931.2010.531007
Wesselhoeft, R., Sorensen, M. J., Heiervang, E. R., & Bilenberg, N. (2013). Subthreshold depression
in children and adolescents—​a systematic review. J Affect Disord, 151(1), 7–​22. doi:10.1016/​
j.jad.2013.06.010
Williams, J. M. G. (1996). Depression and the specificity of autobiographical memory. In D. C. Rubin (Ed.),
Remembering our past: Studies in autobiographical memory (pp. 244–​267). New York, NY: Cambridge
University Press; US.
5
9
1

Interventions for depression: CBT with children and adults 195

Williams, J. M. G., Barnhofer, T., Crane, C., Herman, D., Raes, F., Watkins, E., & Dalgleish, T. (2007).
Autobiographical memory specificity and emotional disorder. Psychological Bulletin, 133(1), 122.
Wood, D., Bruner, J. S., & Ross, G. (1976). The role of tutoring in problem solving. Journal of Child
Psychology and Psychiatry, 17(2), 89–​100.
Woody, M. L., Burkhouse, K. L., & Gibb, B. E. (2015). Overgeneral autobiographical memory in children
of depressed mothers. Cognition and Emotion, 29(1), 130–​137.
World Health Organization (2015). International statistical classification of diseases and related health
problems (10th ed.). Retrieved from http://​apps.who.int/​classifications/​icd10/​browse/​2015/​en
Zacks, R. T., & Hasher, L. (1994). Directed ignoring: Inhibitory regulation of working memory.In D.
Dagenbach & T. Carr (Eds.) Inhibitory processes in attention, memory, and language, (pp. 241–​264).
San Diego, CA, US: Academic Press.
Zhong, B. L., Ding, J., Chen, H. H., Li, Y., Xu, H. M., Tong, J., … Zhu, J. H. (2013). Depressive disorders
among children in the transforming China: an epidemiological survey of prevalence, correlates, and
service use. Depress Anxiety, 30(9), 881–​892. doi:10.1002/​da.22109
Zins, J., & Erchul, W. (2002). Best practices in school consultation. In A. Thomas & J. Grimes (Eds.), Best
practices in school psychology (4th ed., pp. 625–​644). Bethesda, MD: National Association of School
Psychologists.
Zupan, B. A., Hammen, C., & Jaenicke, C. (1987). The effects of current mood and prior depressive history
on self-​schematic processing in children. Journal of Experimental Child Psychology, 43(1), 149–​158.
6
9
1

Chapter 10

Emotion Regulation and Eating


Disorders
Julian Baudinet, Lisa Dawson, Sloane Madden,
& Phillipa Hay

Eating disorders
Eating disorders in children and adolescents differ from those in adults in prevalence of clinical
syndromes and in the psychopathology of disorders. Bulimia nervosa and binge eating disorder
have a later age of onset than anorexia nervosa and Avoidant/╉Restrictive Food Intake Disorder
(ARFID), which are the more common childhood problems. However, very little is known about
ARFID and its treatment. Thus, this chapter will focus on anorexia nervosa and describe a new
approach to emotion regulation in its management in adolescents. The diagnostic clinical features
of common eating disorders (American Psychiatric Association, 2013) are listed in Box 10.1.
Anorexia nervosa occurs in around one in 400 adolescent girls. In pre-╉pubertal children eat-
ing disorders are less common and have a different sex distribution where almost a quarter of
presentations manifest in boys (Madden, Morris, Zurynski, Kohn, & Elliot, 2009). ARFID is also
common in children and boys; however, it is unclear if ARFID, or at least a proportion of ARFID
cases of early onset, are a predecessor for anorexia nervosa or other eating disorders, or whether it
is a distinct eating disorder with longitudinal studies needed to elucidate this. In children it often
presents with significant weight loss in the context of somatic concerns such as nausea and full-
ness, and in adults with a specific anxiety related to eating.
In addition to differences in sex distribution, eating disorders of early onset also appear to
have a greater responsivity to treatment and, in clinical samples, better long term outcomes
than when onset presents in adulthood (Hay, 2015). A a strong body of research supports the
efficacy of treatment in young people with eating disorders (See for example, Forsberg & Lock,
2015). The causes of eating disorders in children and adolescents are similar to those in adults;
these include a family history of eating, mood and substance abuse disorders as well as obesity.
Exposure to “anorexogenic” environments such as classical ballet and high levels of criticism
and parental expectations are often associated with the onset and maintenance of disorders
(Hay et  al., 2014; Hay & Claudino, in press; Zipfel, Giel, Bulik, Hay, & Schmidt, 2015). The
pathway into an eating disorder is a complex interplay of biological/╉genetic, psychological vul-
nerabilities and societal factors (Mitchison & Hay, 2014)  and recent research has pointed to
the importance of gene-╉environment interactions and the role of epigenetics (Campbell, Mill,
Uher, & Schmidt, 2011).

Emotion regulation and eating disorders


Interest in emotion regulation and processing in anorexia nervosa has been long-╉standing.
Difficulty in identifying and describing emotions has been noted in individuals since 1973 (Bruch,
7
9
1

Emotion regulation and eating disorders 197

BOX 10.1╇ Clinical features of eating disorders in children


and adolescents
Anorexia nervosa
Under-╉weight for age and height
Intense fear of fatness/╉gaining weight or behaviours to avoid weight gain
Overvalued ideas of body weight/╉shape on self-╉view or denial of seriousness of low weight

Bulimia nervosa
Recurrent binge eating-╉uncontrolled overeating
Use of extreme measures to control their weight—╉purging/╉non-╉purging
Overvalued ideas of body weight/╉shape on self-╉view
Normal weight or over weight

Binge Eating Disorder


Recurrent and distressing binge eating-╉uncontrolled overeating
No regular use of extreme measures to control their weight—╉purging/╉non-╉purging
Overvalued ideas of body weight/╉shape on self-╉view not required
Normal weight or over weight

Avoidant/╉Restrictive Food Intake Disorder


Extreme disinterest in eating or food and/╉or food is avoided because of its colour, smell, taste
or other sensory quality and/╉or or there is a fear of an eating consequence, e.g. choking
Severe weight loss and/╉or specific nutritional deficiency and/╉or medically supported feeding
and/╉or associated psychosocial functional impairment is present
There is no associated weight/╉shape overvaluation or body image concern
Absence of another medical or psychiatric disorder or cultural circumstance that explains the
food avoidance

1973). This construct, defined as alexithymia, has been consistently reported by clinicians and
demonstrated by researchers with rates as high 77.1% in individuals with anorexia nervosa com-
pared to 6.7% in healthy matched controls. In addition, rates of comorbid depression and anxiety
in anorexia nervosa are high (Hatch, Madden et al. 2010).
Recent models of anorexia nervosa are increasingly emphasising the role of maladaptive
emotion regulation strategies and difficulties with emotion identification as key precursors to
the development and maintenance of anorexia nervosa (Haynos & Fruzzetti, 2011; Oldershaw,
Lavendar, Sallis, Stahl, & Schmidt, 2015; Lavendar, et al., 2015). There is a small but growing body
of evidence indicating that people suffering from anorexia nervosa show a greater use of maladap-
tive emotion regulation strategies, such as avoidance, emotion suppression, inhibition, repression,
rumination and self-╉destructive behaviors (Haynos & Fruzzetti, 2011). Furthermore, people strug-
gling with anorexia nervosa use fewer adaptive strategies compared to healthy controls (Haynos &
8
9
1

198 Emotion Regulation and Eating Disorders

Fruzzetti, 2011; Oldershaw, et al., 2015). This is important because less adaptive emotion regula-
tion strategies are suggested to result in more overall emotional problems whilst contributing to
psychological co-╉morbidity (Haynos & Fruzzetti, 2011)—╉a factor now recognized as a barrier to
family based treatment (FBT) outcomes (Lock, Courtourier, Bryson, & Agras, 2006). Preliminary
findings suggest emotion regulation difficulties may persist following weight restoration (Haynos,
Roberto, Martinez, Attia, & Fruzzetti, 2014). Furthermore, they can moderate against treatment
efficacy, contributing to the maintenance of anorexia nervosa (Racine & Wildes, 2015), with poor
emotion regulation techniques related to relapse (Federici & Kaplan, 2008).

Evidence-╉based treatments
The first treatments developed for eating disorders were for anorexia nervosa. Early trials
included adults as well as adolescents. The seminal study of psychological therapies were those
of Russell, Szmukler, Dare, and Eisler (1987) and Eisler, Dare, Russell, Szmukler, le Grange, and
Dodge (1997). This was a post weight-╉restoration outpatient psychotherapy trial where indi-
vidual therapy was compared to family therapy. While there were no differences in outcomes
between the two treatment arms, secondary analysis revealed better outcomes with family ther-
apy in participants who had an eating disorder for less than three years and were under the age
of 18. Although predominantly a trial of anorexia nervosa, this study also included participants
with bulimia nervosa. Two further trials have looked at treatment interventions in study samples
of adults and adolescents (Crisp, Norton, et al. 1991; Ball and Mitchell 2004), while there have
been nine randomized controlled trials that have specifically studied the efficacy of psychologi-
cal therapies for children and adolescents with anorexia nervosa which have included weight
restoration in aims and outcomes.
There have been five randomized control trials of manualized family based therapies with a
predominant behavioral focus, as in FBT. The first of these was by Robin et al., (1999). This was
a small non-╉blind trial with unclear allocation concealment and thus, had a high risk of bias.
Thirty-╉seven participants were randomized to either a families systems therapies or to an ego-╉
orientated individual therapy. Those in the family therapy arm had significantly greater weight
gain at the end of treatment and at a one-╉year follow up. Similarly, the second (Eisler et al. 2000;
Eisler, Simic, Russell, & Dare, 2007) also had risk of bias in that there was unclear allocation,
concealment and blinding. In this study, forty participants were randomized to either family
based treatment, conjointly or individualized therapy where the parents were seen separately
from the child with anorexia nervosa. Similarly in this study, there were no differences between
the groups in outcomes at any point up to a five-╉year follow up with the exception that where
there were maternal criticism participants showed significantly higher levels of improvement
when therapy was separated.
Three trials that controlled for bias with adequate allocation concealment that involved inde-
pendent or blind outcome assessments were conducted by Lock, Agras, Bryson and Kraemer
(2005), Lock et al. (2010) and Agras et al. (2014). In Lock et al. (2005), 86 participants were ran-
domized to either ten sessions over six months or 20 sessions over 12 months of FBT. Although
there were no between group differences, the longer treatment led to greater improvements in
people with higher levels of obsessive compulsive symptoms and those with non-╉intact families.
Agras et al. (2014) compared FBT with systemic family therapy in 164 participants. There were
no differences in weight or other primary outcomes. However, there was earlier weight regain
and fewer hospitalizations in participants who were treated with family based treatment. On the
other hand, systemic family therapy led to better outcomes with those who had higher levels of
obsessive compulsive symptoms. Finally, Lock et al. (2010) randomized 121 participants to FBT
91

Evidence-based treatments 199

or individual, adolescent focused controlled psychotherapy. This singular study reported higher
remission rates and greater weight gain at both end of treatment and a one-​year follow up in those
randomized to family treatment.
Other studies done in the treatment of children and adolescents include research by Geist,
Heinmaa, Stephens, Davis, and Katzman (2000) who randomized participants to family therapy
where the families were seen for eight sessions with the patient, the patients parents and siblings
or to a family group psycho-​education arm where groups of families were seen in a workshop
design for eight sessions. In this study of 25 participants, there were no significant differences
between groups. This study also had risk of bias, as there was no blinding. Another study by
Gowers et al. (2007) compared a specialist outpatient, manualized cognitive behavioral thearapy
treatment intervention with separate parental counselling and non-​manualized supportive and
family care. This study found no differences between groups. Godart et al. (2006), in a high quality
randomized control trial, compared a non-​manualized psychodynamic systemic family therapy
to usual specialist care in 60 adolescent participants. This trial reported significantly improved
weight-​gain and other outcomes in those who received the additional family therapy.
These trials of FBT have formed the basis for the leading evidence based therapy in chil-
dren and adolescents with anorexia nervosa. It is notable that the majority was conducted with
female participants and that only two found significant improvements in primary outcomes.
It is also important to note that although there were minimal differences in symptomatic out-
comes, family based treatment in the trial by Agras et  al. (2014) was associated with lower
financial costs and hospitalization rates. When compared with other non-​family based treat-
ments, FBT also demonstrated improved remission rates at follow-​up (Courtourier, et al., 2010
Forsberg & Lock, 2015). It has to be acknowledged however, that although FBT and other fam-
ily therapies have strong evidence for treatment of children and adolescents with anorexia ner-
vosa (Zipfel et al., 2015) their efficacy for adolescents with bulimia nervosa is less established
with mixed or inconsistent findings. In addition, there have been no trials in ARFID or binge
eating disorder (Hay et al., 2014).
As previously highlighted, FBT has become established as the leading treatment for adolescents
with anorexia nervosa. The treatment has been manualized (Lock & Le Grange, 2015) and dis-
seminated internationally. The treatment includes three phases. Phase I focuses on empowering
parents to manage all anorexia nervosa related behavior until the adolescent is weight restored.
Following a period of weight maintenance, Phase II focuses on working with the adolescent to
return to an appropriate level of control over food and eating. Phase III then focuses on life cycle
events that may have been interrupted by the eating disorder.
Research indicates that FBT is effective for anywhere from approximately 30–​60% of young
people struggling with anorexia nervosa at the end of treatment, with these findings improved
upon at follow-​up (Forsberg & Lock, 2015). While this data is encouraging, particularly when
compared to poor response rates to adult treatments (Bulik, Berkman, Brownley, Sedway, & Lohr,
2007), it is now clear that FBT is not effective for a substantial minority of young people who con-
tinue to struggle or do not complete treatment. Furthermore, remission is often defined within
the literature as reaching a specified weight range, which does not always correspond with full
psychological recovery. Given a lack of effective alternative treatments, this leaves a substantial
proportion of adolescents with anorexia nervosa at high risk of becoming chronically unwell.
Research is beginning to investigate possible factors associated with poor treatment responses
or drop out in FBT. Initial findings have identified a range of family factors related to emotional
expression and management that are associated with poorer outcomes or dropout. These include
high expressed emotion, family conflict and criticism (Eisler, Simic, Russell & Dare, 2007; Lock,
Coutourier, Bryson, & Agras, 2006; Russell, Szmukler, Dare, & Eisler; Le Grange, Eisler, Dare,
02

200 Emotion Regulation and Eating Disorders

& Russell, 1992), with parental warmth being related to good outcomes (Le Grange, Hoste,
Lock, & Bryson, 2011). These findings are important, as Phase I of FBT can be very stressful and
emotionally challenging for all family members as parents actively and consistently challenge
the symptoms of anorexia nervosa. Accordingly, this often results in young people and families
being faced with extremely distressing events on a regular basis with, potentially, reduced emo-
tion regulation capacities. The treatment often requires this to be repeated consistently for many
weeks to months.
Several individual factors have also been identified that are associated with poorer outcomes in
FBT. Recent findings suggest that adolescents with more severe eating disorder psychopathology
and those struggling with co-╉morbid Axis I  and/╉or emerging Axis II psychological difficulties
have a greater likelihood of dropout and may require a longer duration of treatment (Forsberg
& Lock, 2015). This is important, as co-╉morbidity rates in anorexia nervosa remain high, with
more than 50% experiencing a co-╉morbid anxiety disorder (Kaye, Bulik, Thornton, Barbarich,
& Masters 2004)  and between 50 and 70% and experiencing a major mood disorder (Godart,
et al., 2006).

Emotion regulation and eating disorder treatments


The development of emotion regulation difficulties is hypothesized to be the result of numerous
factors. These include biological factors, attachment and attunement difficulties within family sys-
tems (Zeman, Cassano, Perry-╉Parish, & Stegall, 2006), as well as traumatic childhood events (Dvir,
Ford, Hill, & Frazier, 2014). Preliminary investigations are now being conducted into how best
to include emotion regulation interventions into anorexia nervosa treatment. This has included
the development of new treatments, such as the Maudsley Model of Anorexia Nervosa Treatment
for Adults (MANTRA; Schmidt et al., 2012) and Emotion Acceptance Behavior Therapy (EABT;
Wildes, Marcus, Cheng, McCabe, & Gaskill, 2014), as well as the modification or adaption of
evidence based treatments from other areas of clinical psychology, such as Dialectical Behavior
Therapy (DBT; Linehan, 1993). Research into the use of DBT for eating disorders in adults is
promising, although yet to be rigorously researched and tested in adolescents (Bankoff, Karpel,
Forbes, & Pantalone, 2012). Radically-╉Open DBT (RO-╉DBT) is a modified DBT treatment spe-
cifically designed to target emotion expression and the maladaptive strategy of emotional over
control in anorexia nervosa (Lynch et al., 2013). The efficacy of RO-╉DBT has also yet to be tested
in anorexia nervosa.
These emotion-╉focused interventions are being trialed in various modes of delivery ranging
from individual and group, to family and multi-╉family interventions. Some examples include
adding specific DBT informed emotion focused modules as an adjunct to FBT (Robertson,
Alford, Wallis, & Miskovic-╉Wheatley, 2015), integrating emotion regulation ideas and techniques
throughout FBT (Federici & Wisniewski, 2012; Robinson, Dolhanty, & Greenberg, 2015) or offer-
ing skills training group based interventions as one of a range of interventions provided in the
context of a day program (Girz, Robinson, Foroughe, Jasper, & Boachie, 2013).
Emotion regulation difficulties have been identified as maintenance factors in anorexia nervosa
and as a barrier to effective FBT. Given emotion regulation is influenced and impacted by envi-
ronments and relationships (Oldershaw et al., 2015), it has been proposed that simultaneously
addressing established interpersonal patterns may be an important part of treatment (Treasure &
Schmidt, 2013; Oldershaw et al., 2015). Thus, taken together the literature suggests that while FBT
is effective for the majority of adolescents, there are compelling arguments that adjunctive treat-
ments targeting emotion regulation in the young person, as well as family members, may help in
improving outcomes for those identified as at risk of poor treatment outcomes.
1
0
2

THE CHILDREN’S HOSPITAL AT WESTMEAD TREATMENT PROGRAM 201

The Children’s Hospital at Westmead Treatment Program


The day program at The Children’s Hospital at Westmead operates five days per week. It is a pro-
gram designed for adolescents and those who are not responding to outpatient FBT. The pro-
gram offers three supported meals per day, adolescent group therapy each afternoon, weekly
family therapy, individual therapy as needed, medical and psychiatric monitoring, a weekly par-
ent group, multi-╉family groups, and educational support. The program is designed to provide
increased intensity for adolescents and their families who are not responding to FBT by providing
additional inputs to care. It is designed as an adjunct to FBT to facilitate its effectiveness, rather
than a new treatment.
Accordingly, interventions for emotional regulation are not the sole focus of treatment, but are
rather an essential component of a much larger treatment program. The aim of the program is to
broadly address emotional regulation difficulties across a number of domains in the context of
treating anorexia nervosa using FBT-╉informed systemic framework.
The program is designed to specifically address disturbances in the experience of emotions for
children and adolescents with anorexia nervosa. The program is based on increasing emotional
awareness and reducing emotional avoidance by supporting young people to develop adaptive
emotional regulation strategies and the capacity to select how and when to implement these.
Based on the dimensions of Gratz and Roemer’s (2004) multidimensional model of emotion regu-
lation and dysregulation, the program specifically aims to help young people to:
1 . Appropriately and flexibly managing distress
2. Maintain behavioral control in the face of distress
3. Increase emotional awareness, clarity, and acceptance
4. Explore willingness to tolerate difficult emotions in order to pursue a fulfilling life
A multidimensional approach to addressing these goals is used, including:
1 . Specific skills-╉based groups (drawn from DBT, CBT, and ACT)
2. Experiential/╉behavioral opportunities to manage distress
3. Effectively utilising process and the structure of a day program (e.g., creating the right milieu
for the group, engagement, validation, boundary setting, normalising, use of language)
4. Encouraging a life beyond anorexia nervosa /╉adolescent development

Appropriately and flexibly managing distress


Patients are taught a variety of specific techniques to modulate the duration and intensity
of their emotional responses. Mindfulness is a key intervention in emotion regulation, with
young people introduced to mindfulness at the commencement of their treatment and prac-
ticed daily at the beginning of all therapeutic groups. Mindfulness practice includes mindful
breathing, mindfulness of objects, music, and mindful games. How mindfulness can be an
effective tool for managing distress is discussed and the ways it can be employed is problem
solved by staff and patients. It is particularly encouraged after more challenging meals as a way
to be in the present moment, non-╉judgementally, and not to dwell on the meal that has just
past or is upcoming.
Participants are also taught other ways to manage distress including distraction, self-╉soothing,
using intense sensations, and radical acceptance. Young people create their own “self-╉soothing
kits” by collecting objects that can be accessed in the moment of distress or remind them of strat-
egies to use. Using such multiple emotion regulation methods flexibly (not relying on only one
method) is encouraged, with patients assisted to match challenging situations with appropriate
emotion regulation strategies.
2
0

202 Emotion Regulation and Eating Disorders

Central to this, is teaching young people how to identify the early signs of their distress and
the importance of intervening early in distress management. Creating “distress thermometers,”
where participants map the physical, psychological and emotional changes that occur as their
distress increases from zero out of ten (no distress) to ten out of ten (high distress), is part of this.
Following this, young people then match strategies with differing levels of distress intensity. As
distress becomes exponentially harder to manage the more it intensifies, staff focus on and sup-
port young people to intervene early in their distress management. Young people are encouraged
to involve family and other support as needed.
Additionally, patients’ beliefs about their capacity to effectively manage emotions are chal-
lenged. Many young people in the program have strong beliefs that they do not possess adequate
skills to effectively and adaptively modulate emotional experiences—╉a common experience in
anorexia nervosa (Lavender, et al., 2015). Situations where young people effectively manage dis-
tress are identified and amplified by staff with young people supported to identify and label their
own skills to help increase self-╉efficacy.

Maintaining behavioral control


The ability to maintain behavioral control in the context of heightened negative emotional arousal
is an important component of emotional regulation. While within-╉session experiential emotional
arousal has been argued as essential for addressing emotional dysregualtion (Greenberg & Pavio,
1997), creating such experiences in outpatient treatment is generally challenging, not least because
those with anorexia nervosa are often highly emotionally avoidant, scared of emotion, and lack
motivation to change (Hoetzel, von Brachel, Schlossmacher, & Vocks, 2013). In traditional FBT
the meal session is a good opportunity for this, however, such a session generally occurs just
once in treatment. The meal session is also often based around a food challenge and offers a lot of
coaching by the therapist for parents, but less for the young person.
The three meals per day provided in the day program provide patients with extensive practice at
experiential emotional arousal in a therapeutic context. Meals are utilized not only as a means of pro-
viding young people with anorexia nervosa with sufficient energy requirements, but also as an oppor-
tunity to provide young people with repeated exposure to distressing events and a means to practice
and utilize emotion regulation skills learned in the program. This is done explicitly by reminding and
encouraging young people to use specific emotion regulation in response to food and challenging
eating disorder rules or behaviors during meals, and implicitly by establishing an environment that
aims to be less clinical than traditional inpatient settings. While clear boundaries remain, humour
and engagement is utilized to create an adolescent-╉appropriate environment. For example, meal times
always include music, conversation, jokes, or games with all young people being equally engaged.
While all eating disorder inpatient treatments provide meals, there are benefits to providing this
in combination with a complementary psychological treatment (for those who are well enough
to no longer require inpatient care). This is supported by Oldershaw and colleagues (2015), who
suggest that interventions for people with anorexia nervosa should include and seek a balance
between behavioral or experiential components and cognitive components. The meals provide
opportunities to help young people manage their distress in vivo by drawing on psycho-╉education
and distress tolerance skills. During meals, the young people must inhibit dysfunctional behaviors
when distressed, such as refusing to eat, absconding, eating slowly, or hiding food. If someone
does become distressed they are encouraged to draw upon skills they have learnt and other young
people are invited to also encourage the young person struggling and provide advice.
3
0
2

THE CHILDREN’S HOSPITAL AT WESTMEAD TREATMENT PROGRAM 203

Beyond food and eating, behavioral/╉experiential opportunities to manage distress are also
encouraged, planned for and debriefed. Young people are encouraged to practice managing many
distressing situations, particularly those related to adolescent development e.g., turning to par-
ents when distressed (practicing openness) and attending social events that might be anxiety-╉
provoking. Young people are also encouraged to tolerate changes to their body occurring with
weight gain.

Emotional awareness, clarity, and acceptance


Many young people with anorexia nervosa are inattentive to their own emotional experiences and
struggle to understand their emotions and the ability to differentiate between affective states. They
also often have difficulty accepting emotions and often reject emotional experiences. Accordingly,
therapeutic groups that specifically target understanding and accepting your feelings are an essen-
tial component of the program.
Groups to support this focus on identifying and labelling emotions, differentiating between
the intensity of different emotions, exploring primary and secondary emotions, investigating the
function of emotions, the pros and cons of having feelings, and challenging myths about emotions
are utilized. Young people have opportunities to practice recognising their own emotions as well
as those of others.
Outside of specific skills-╉based groups, participants are encouraged regularly to identify, label,
and reflect on their emotional experiences. Each morning young people are asked about the pre-
vious night and asked to describe how they felt, label the feeling, rate the intensity of the feeling,
identify how it affected them, how the feeling was managed and how they would like to manage
it next time. Experiencing emotions, including distress, is normalized. Staff model healthy emo-
tional regulation by helping young people label their emotions as well as normalising and vali-
dating the young person’s emotional responses (e.g., “I can understand why you felt that,” “when
I have that feeling it also feels pretty bad”).
Psycho-╉education about the impact of suppressing and avoiding emotions is provided. This
includes the functional aspects of using starvation to avoid experiencing difficult emotions and
the long-╉term consequences of this and how rejecting emotional experiences can result in second-
ary negative affective states regarding the primary emotional response.

Willingness to experience emotional distress


Throughout the program, normal adolescent development is encouraged, as staff actively sup-
port young people to value and remain on their normal adolescent developmental trajectory.
For example, conversations are had about gaining independence from parents, learning to drive,
attending school dances and what life after high school may look like. Staff emphasize the impor-
tance of being willing and able to tolerate aversive emotional experiences in the context of pur-
suing activities that are meaningful to the individual. Young people are encouraged to explore
what it will mean to them now and in their future lives if they do not pursue recovery by avoiding
emotional distress and contrast this with the impact of pursuing meaningful life activities (e.g.,
school, study, careers, sport, friendships and relationships). Activities are linked to specific eat-
ing disorder symptoms, for example, what will it mean for young people if they are unable to
eat in front of or with others. Activities include exploring the pros and cons of having an eating
disorder, living a valued life and completing pie charts for now and the future without anorexia
nervosa.
4
0
2

204 Emotion Regulation and Eating Disorders

Case example
The case of Emma outlined below is a combination of several patients and their families who
have completed the day program. The case is used for two purposes; firstly, to give an example
of the way emotion regulation interventions can enhance standard family based treatments and
secondly, to illustrate the importance of not only providing young people with a forum to learn
skills, but also highlighting the importance of ensuring there are the appropriate structures and
therapeutic processes working in tandem to facilitate skill implementation.

Emma
Emma first presented with anorexia nervosa when she was 16-╉years-╉old. She weighed 41kg, was
156 cm tall and presented with medical complications of her weight loss including bradycardia
(low heart rate) and hypothermia (low temperature). She resided with her mother, Leanne. Her
father had been living in a separate house since the acrimonious breakdown of her parents’ mar-
riage 18 months prior her 19 year-╉old brother had moved out of the family home at the comple-
tion of high school. Emma had an eight-╉month history of food restriction and compulsive exercise
resulting in a 7kg weight loss. She reported a six-╉month history of amenorrhea.
Emma described experiencing significant mood difficulties for the previous nine to twelve
months, with reduced sleep, increased social isolation and anhedonia. Emma had been engag-
ing in deliberate self-╉harm of superficial cutting on her wrists and hip up to twice a week for the
previous three months. She described passive suicidal ideation, denying any active plans or will
to commit suicide. Her presentation occurred in the context of ongoing, severe bullying at school
and her grandmother passing away nine months prior with bowel cancer. Emma also described
experiencing separation anxiety from her mother, Leanne, up until early primary school.
Emma was admitted to an inpatient paediatric ward for medical stabilisation and psychologi-
cal containment. Following her discharge FBT was provided by a clinical psychologist. After nine
months of treatment Emma’s weight had slowly been reducing and family conflict continued to
escalate. This had resulted in several occurrences of Emma running away from home during
meals and physically intimidating her parents by threatening to hit them and on two occasions
assaulting them.
At this point the day program was offered to Emma and her family to contain Emma’s weight
loss, stop the escalation of dangerous behavior, provide skills training to Emma around emotion
regulation and increase systemic empathy and understanding. The aim of the admission to day
program was to break the vicious cycle of Emma’s experience of parental invalidation resulting in
Emma’s emotional distress and behavioral escalation.
Over the course of her admission in the day program, emotion regulation and distress tolerance
were key treatment interventions for Emma. Emma’s goal was to find ways of not becoming so
angry that she needed to run away during meals or become threatening to her mother; something
that made her feel very guilty. The daily adolescent group provided the most direct method of
equipping Emma with the specific knowledge and skills around how to regulate her emotions.
Box 10.2 outlines the selected skills Emma was taught in the group. Emma initially struggled with
participating in group, often saying very little or saying she had tried everything and it did not
help. While this was challenging to staff initially, through validation, encouragement and genuine
interest in her difficulties staff were able to engage Emma in the process of group discussions, even
though content remained difficult to engage with.
In tandem with skills group, the day program context provided an opportunity for staff to
prompt the early identification of signs of distress, the communication of difficult emotions and
the appropriate use of skills in different contexts throughout the day, such as during difficult meals
5
0
2

Case example 205

BOX 10.2  Skills Group Content Provided to Emma


Skills below are shared with parents during family sessions to ensure families are actively involved
in skill implementation outside of program hours.
Mindfulness ◆ Mindful observation using ◆ Anchoring oneself to the
5 senses present moment (e.g., using
breathing
◆ counting)
Finding “wise mind”

Emotion Psycho-​education around the function of emotion


Identification Guided practice on noticing internal experiences


Emotion labelling, observation, distancing and acceptance


Riding the wave of emotion


Communicating Psycho-​education on communication styles


Emotions Using words vs behaviours to communicate emotions


Practice with role playing helpful communication styles and


behavioural experiments
Distraction ◆ Group discussion around the difference between skilful
distraction and unhelpful avoidance
Generation of a list of distraction techniques and pairing with

appropriate times they can be used (e.g., drawing, board


games/​cards, fidget toys/​kinetic sand, time outs and engaging
in conversation)
Self-​Soothing ◆ Design and creation of toolbox to have at home filled
with items to self-​soothe using all five senses (e.g., music,
motivational statements, meaningful gifts/​items, hand cream,
nail polish, perfume, etc.)
Opposite Action ◆ Identification of usual responses to each particular emotion and
practice doing the opposite
Replacements Holding ice

Techniques for Cold shower


Self-​Harm Rubber band on wrist


or following stressful events. By ensuring group sessions involved practical elements, in vivo tasks
or experiments and homework tasks, the environment ensured Emma experimented with skills,
albeit begrudgingly. This slowly allowed her to experience some mild benefits from skill imple-
mentation, which then allowed her to generalize them from their use on the program to life out-
side of the program. Staff consistency in their relationship with Emma facilitated a safe space for
her to feel accepted, despite frequent emotional outbursts. This allowed her to feel more comfort-
able in trying new things and reduced feeling of shame or embarrassment.
Importantly, other therapeutic aspects of the program, which were not directly related to emo-
tion regulation skill development, were seen as key to helping Emma improve her ability to regu-
late her emotions. Family and multi-​family sessions provided an opportunity for Emma to involve
Leanne in emotion regulation skill development and planning could be done with the therapist
6
0
2

206 Emotion Regulation and Eating Disorders

as to how Leanne could best support Emma with skill use. Additionally, with Leanne spending
less time providing meals for Emma, this enabled her to plan more specifically around the meals
she was supervising. This allowed her to feel more prepared and confident, leaving her more
able to tune into Emma’s needs, reduce criticism and provide much needed validation during the
meals. Similarly, multi-╉family groups and meals were also beneficial for Emma as they provided
repeated opportunity for staff to model and coach Leanne on how to support Emma with consis-
tent warmth and firmness, as well as skill use and implementation.
Effective skill implementation only really began to result in noticeable changes for Emma four
to five weeks after she commenced treatment. Through the process of staff using a firm but kind
approach, with consistent boundaries and communication across all activities, Emma settled
enough to attempt learnt techniques. She described finding it helpful being “checked-╉in” with
frequently and said it provided the opportunity to test out expressing her more difficult emotions.
She also said the experience of interacting with staff in multiple therapeutic context (e.g. meals,
groups, family therapy) was beneficial. She said this exposure to staff across settings, as well as
staff being able to engage in adolescent appropriate conversations, use humour, model appropriate
eating, and tolerate high affect allowed Emma to feel able to accept and engage in the program. It
was then through this connection that Emma described feeling able to experiment with alternate
ways of managing her emotions and tolerating feelings of worthlessness and hopelessness.
Emma was discharged from the day program after completing 11 weeks. She was discharged
within her healthy weight range after having gained four kilograms. While she continued to feel
distressed around meals and many eating disorder behaviors persisted, she and her family said
they felt much better equipped to continue to make gains in outpatient treatment. Both Emma
and Leanne said that it was the combination of Emma learning new ways to regulate her emotions
with Leanne being able to validate, understand and provide support around skill use that made
them feel less stuck and able to move forward in treatment.

Summary
A growing body of literature indicates FBT is an effective treatment for adolescents with anorexia
nervosa. Nevertheless, FBT is not effective for everyone, with a significant minority continuing to
respond poorly to even the best available treatments. Given the role emotion regulation difficul-
ties are hypothesized to play in the development and maintenance of anorexia nervosa, modifica-
tions to FBT that target emotion dysregulation are emerging. The case of Emma highlights a few
key factors to consider when designing and implementing emotion regulation focused adjuncts
or modifications to treatment. It highlights the importance of matching skills training with a
consistent program structure and a positive group milieu. It is through the combination of these
three factors that progress in treatment is hypothesized to occur. In the case of Emma, without
the structure or milieu, skills training was unlikely to have been meaningfully attempted poten-
tially adding to her feelings of hopelessness and helplessness. It was through the combination of
all three elements that psychoeducation was delivered in a format and environment that allowed
Emma to make meaningful treatment gains.
This approach to improving emotional regulation in adolescents with anorexia nervosa is in
the early stages of assessment and further investigation and controlled trials are needed. Further
research is also indicated to investigate the best approach to young people with other eating
disorders including bulimia nervosa and binge eating disorder where individuals may have con-
comitant problems with impulsivity and emotion regulation. There is a small body of research
supporting the efficacy of a modified individual outpatient form of dialectical behavior ther-
apy in adults with bulimia nervosa or binge eating disorder (Safer, Telch & Agras 2001; Safer
7
0
2

Summary 207

Robinson, & Jo, 2010) and trials are now being run in adolescents. Although research is promis-
ing, it is in the early stages and further investigation is required involving large-╉scale unbiased
studies. However, it is important to note, treatment outcomes have high success rates (20–╉60%)
when eating disorders are treated in childhood and adolescence; which is imperative, as adult
anorexia nervosa is one of the most challenging psychiatric illnesses to treat effectively with one
of the highest morbidity rates.

References
American Psychiatric Association (APA). (2013). Diagnostic and statistical manual of mental disorders (5th
Ed.). Arlington, VA: American Psychiatric Publishing.
Agras, W. S., Lock, J., Brandt, H., Bryson, S. W., Dodge, E., Halmi, K. A., … & Woodside, B. (2014).
Comparison of 2 family therapies for adolescent anorexia nervosa: a randomized parallel trial. Journal
of the American Medical Association Psychiatry, 71(11), 1279–╉1286.
Ball, J., & Mitchell, P. (2004). A randomized controlled study of cognitive behavior therapy and behavioral
family therapy for anorexia nervosa patients. Brunner-╉Mazel Eating Disorders Monograph Series, 12(4),
303–╉314.
Bankoff, S. M., Karpel, M. G., Forbes, H. E. & Pantalone, D. W. (2012). A systematic review of dialectical
behaviour therapy for the treatment of eating disorders. Eating Disorders, 20, 196–╉215.
Bruch, H. (1973). Eating disorders. Obesity, anorexia nervosa and the person within. New York: Basic Books.
Bulik, C., Berkman, N. D., Brownley, K. A., Sedway, J. A. & Lohr, K. N. (2007). Anorexia Nervosa
Treatment: A Systematic Review of Randomized Controlled Trials. International Journal of Eating
Disorders, Vol. 40 (pp. 310–╉320).
Campbell, I. C., Mill, J., Uher, R., & Schmidt, U. (2011). Eating disorders, gene–╉environment interactions
and epigenetics. Neuroscience Biobehavior Reviews, 35(3), 784–╉793.
Couturier, J., Isserlin, L. & Lock, J. (2010). Family based treatment for adolescents with anorexia
nervosa: A dissemination study. Eating Disorders, Vol. 18 (pp. 199–╉209).
Crisp, A. H., Norton, K., Gowers, S., Halek, C., Bowyer, C., Yeldham, D., … & Bhat, A. (1991). A
controlled study of the effect of therapies aimed at adolescent and family psychopathology in anorexia
nervosa. British Journal of Psychiatry, 159(3), 325–╉333.
Dvir, Y., Ford, J. D., Hill, M., & Frazier, J. A. (2014). Childhood Maltreatment, Emotional Dysregulation,
and Psychiatric Comorbidities. Harvard Review of Psychiatry, 22(3). 149–╉161.
Eisler, I., Dare, C., Hodes, M., Russell, G., Dodge, E., & Le Grange, D. (2000). Family therapy for
adolescent anorexia nervosa: The results of a controlled comparison of two family interventions.
Journal of Child Psychology and Psychiatry, 41, 727–╉736.
Eisler, I., Dare, C., Russell, G., Szmukler, G., le Grange, D., & Dodge, E. (1997). Family and individual
therapy in anorexia nervosa: A 5–╉year follow-╉up. Archives of General Psychiatry, 54, 1025–╉1030.
Eisler, I., Simic, M., Russell, G. F. M., & Dare C. (2007). A randomised controlled treatment trial of
two forms of family therapy in adolescent anorexia nervosa: a five-╉year follow-╉up. Journal of Child
Psychology and Psychiatry, 48, 552–╉560.
Eisler, I., Simic, M., Russel, G. & Dare, C. (2007). A randomised controlled treatment trial of two forms of
family therapy in adolescent anorexia nervosa: a five-╉year follow-╉up. Journal of Child Psychology and
Psychiatry, 48(6), 552–╉560.
Federici, A. & Kaplan, A. (2008). The Patient’s Account of Relapse and Recovery in Anorexia
Nervosa: A Qualitative Study. European Eating Disorders Review, 16, 1–╉10.
Federici, A. & Wisniewski, L. (2012). Integrating dialectical behavioral therapy and family-╉based treatment
for multidiagnostic adolescent patients. In Alexander, J. & Treasure, J. (Eds.), A collaborative approach
to eating disorders (pp. 177–╉188). New York: Routledge.
Forsberg, S. & Lock, J. (2015). Family-╉based treatment of child and adolescent eating disorders. Child and
Adolescent Clinics of North America, 24, 617–╉629.
8
0
2

208 Emotion Regulation and Eating Disorders

Geist, R., Heinmaa, M., Stephens, D., Davis, R., & Katzman, D. K. (2000). Comparison of family therapy
and family group psychoeducation in adolescents with anorexia nervosa. Canadian journal of
psychiatry. Revue canadienne de psychiatrie, 45(2), 173–​178.
Girz, L., Robinson, A. L., Foroughe, M., Jasper, K. & Boachie, A. (2013). Adapting family-​based therapy to
a day hospital programme for adolescenets with eating disorders: preliminary outcomes and trajectories
of change. Journal of Family Therapy, 35 (Supp 1), 102–​120.
Godart, N. T., Perdereau, F., Curt, F., Rein, Z., Lang, F., Venisse, J. L., … & Flament, M. F. (2006). Is major
depressive episode related to anxiety disorders in anorexics and bulimics? Comprehensive Psychiatry,
47, 91–​98.
Gowers, S. G., Clark, A., Roberts, C., Griffiths, A., Edwards, V., Bryan, C., …& Barrett, B. (2007). Clinical
effectiveness of treatments for anorexia nervosa in adolescents. British Journal of Psychiatry, 191,
427–​435.
Gratz, K. L., & Roemer, L. (2004). Multidimensional assessment of emotion regulation and
dysregulation: Development, factor structure, and initial validation of the difficulties in ER scale.
Journal of Psychopathology and Behavioral Assessment, 26(1), 41–​54.
Greenberg, L. S., & Paivio, S. (1997). Working with emotions: Changing core schemes.
New York: Guildford Press.
Hatch, A., Madden, S., Kohn, M., Clarke, S., Touyz, S., & Williams, L. M. (2010). Anorexia
nervosa: towards an integrative neuroscience model. European Eating Disorders Review, 18(3), 165–​179.
Hay, P. (2015). Course and Outcome of Eating Disorders. In Wade,T. (ed), Encyclopedia of Feeding and
Eating Disorders. ISBN: 978-​981-​287-​087-​2 (Online) http://​link.springer.com/​referenceworkentry/​
10.1007/​978-​981-​287-​087-​2_​123-​1).Accessed 15 December 2015.
Hay, P., & Claudino, A. (in press). Evidence-​Based Treatment for the Eating Disorders. In Agras, S. (Ed),
Oxford Handbook of Eating Disorders, 2nd Edition. Oxford University Press, New York.
Hay, P., Chinn, D., Forbes, D., Madden, S., Newton, R., Sugenor, L., … & Ward, W. (2014). Royal
Australian and New Zealand College of Psychiatrists clinical practice guidelines for the treatment of
eating disorders. Australian and New Zealand Journal of Psychiatry, 48(11), 977–​1008.
Haynos, A. F., Roberto, C. A., Martinez, M. A., Attia, E., Fruzzetti, A. E. (2014). Emotion regulation
difficulties in anorexia nervosa before and after inpatient weight restoration. International Journal of
Eating Disorders, 47, 888–​891.
Haynos, A. F. & Fruzzetti, A. E. (2011). Anorexia nervosa as a disorder of emotion dysregulation: Theory,
evidence, and treatment implications. Clinical Psychology: Science and Practice, 18, 183–​202.
Hoetzel, K., von Brachel, R., Schlossmacher, L., & Vocks, S. (2013). Assessing motivation to change in
eating disorders: A systematic review. Journal of Eating Disorders, 1, 1–​9. http://​dx.doi.org/​10.1186/​
2050-​2974-​1-​38
Kaye, W. H., Bulik, C. M., Thornton, L., Barbarich, N., & Masters, K. (2004). Comorbidity of anxiety
disorders with anorexia and bulimia nervosa. American Journal of Psychiatry, 161, 2215–​2221.
Lavendar, J., Wonderlich, S., Engel, S., Gordon, K., Kaye, W., & Mitchell, J. (2015). Dimensions of emotion
dysregulation in anorexia nervosa and bulimia nervosa: A conceptual review of the empirical literature.
Clinical Psychology Review, 40, 111–​122.
Le Grange, D., Eisler, I., Dare, C., & Russell, G. (1992). Evaluation of family treatments in adolescent
anorexia nervosa: a pilot study. International Journal of Eating Disorders, 12, 347–​357.
Lock, J., & Le Grange, D. (2015). Treatment Manual for Anorexia Nervosa A Family-​Based Approach.
Second Edition. The Guilford Press: New York. ISBN 978146252346
Lock, J., Agras, S., Bryson, S., & Kraemer, H. (2005). A comparison of short and long term family therapy
for adolescent anorexia nervosa. Journal of the American Academy of Child and Adolescent Psychiatry,
47, 632–​638.
Lock, J., Le Grange, D., Agras, W. S., Moye, A., Bryson, S. W., & Jo, B. (2010). Randomized clinical trial
comparing family-​based treatment with adolescent-​focused individual therapy for adolescents with
anorexia nervosa. Archives of general psychiatry, 67(10), 1025–​1032.
9
0
2

Summary 209

Le Grange, D., Reinecke-​Hoste, R., Lock, J., & Bryson, S. W. (2011). Parental expressed emotion of
adolescents with anorexia nervosa: outcome in family-​based treatment. International Journal of Eating
Disorders, 44, 731–​734.
Linehan, M. M. (1993). Skills Training Manual for Treating Borderline Personality Disorder.
New York: Guilford Press.
Lock, J., Coutourier, J. Bryson, S. & Agras, S. (2006). Predictors of Dropout and Remission in Family
Therapy for Adolescent Anorexia Nervosa in a Randomized Clinical Trial. International Journal of
Eating Disorders, 39, 639–​647.
Lynch, T. R., Gray, K. L. H., Hempel R. J., Titley, M., Chen, E. Y. & O’Mahen, H. A. (2013). Radically
open-​dialectical behavior therapy for adult anorexia nervosa: feasibility and outcomes from an
inpatient program. BMC Psychiatry, 13, 293.
Madden, S., Morris, A., Zurynski, Y. A., Kohn, M., & Elliot, E. J. (2009). Burden of eating disorders in 5–​
13-​year-​old children in Australia. Medical Journal of Australia, 190(8), 410–​414.
Mitchison, D., & Hay, P. J. (2014). The epidemiology of eating disorders: genetic, environmental, and
societal factors. Clinical Epidemiology, 6, 89–​97.
Oldershaw, A., Lavender, T., Sallis, H., Stahl, D., Schmidt, U. (2015). Emotion generation and regulation in
anorexia nervosa: A systematic review and meta-​analysis of self-​report data. Clinical Psychology Review,
39, 83–​95.
Racine, S. E. & Wildes, J. E. (2015). Dynamic longitudinal relations between emotion regulation difficulties
and anorexia nervosa symptoms over the year following intensive treatment. Journal of Consulting and
Clinical Psychology, 83(4), 785–​795.
Robertson, A., Alford, C., Wallis, A., & Miskovic-​Wheatley, J. (2015). Using a brief family-​based DBT
adjunct with standard FBT in the treatment of Anorexia Nervosa. Journal of Eating Disorders, 11(3),
(Suppl 1). doi: 10.1186/​2050-​2974-​3-​S1–​O39
Robin, A., Siegel, P., Moye, A., Gilroy, M., Dennis, A., & Sikand, A. (1999). A controlled comparison
of family versus individual therapy for adolescents with anorexia nervosa. Journal of the American
Academy of Child & Adolescent Psychiatry, 38, 1482–​1489.
Robinson, A., Dolhanty, J. & Greenberg, L. (2015). Emotion-​Focused Family Therapy for Eating Disorders
in Children and Adolescents. Clinical Psychology and Psychotherapy, 22, 75–​82.
Russell, G. F. M., Szmukler, G. I., Dare, C., & Eisler, I. (1987). An evaluation of family therapy in anorexia
nervosa and bulimia nervosa. Archives of General Psychiatry, 44, 1047–​1056.
Safer, D. L., Telch, C. F., & Agras, W. S. (2001). Dialectical behavior therapy for bulimia nervosa. American
Journal of Psychiatry, 158, 632–​634.
Safer, D. L., Robinson, A. H., & Jo, B. (2010). Outcome from a randomized controlled trial of group
therapy for binge eating disorder: comparing dialectical behavior therapy adapted for binge eating to an
active comparison group therapy. Behavior Therapy, 41(1), 106–​120.
Schmidt, U., Oldershaw, A., Jichi, F., Sternheim, L., Startup, H., McIntosh, V., … & Treasure, J. (2012).
Out-​patient psychological therapies for adults with anorexia nervosa: Randomised controlled trial.
British Journal of Psychiatry, 201(5), 392–​399.
Treasure, J., & Schmidt, U. (2013). The cognitive-​interpersonal maintenance model of anorexia nervosa
revisited: a summary of the evidence for cognitive, socio-​emotional and interpersonal predisposing and
perpetuating factors. Journal of Eating Disorders, 1(13), 10.
Wildes, J. E., Marcus, M. D., Cheng, Y., McCabe, E. B. & Gaskill, J. A. (2014). International Journal of
Eating Disorders, 47, 870–​873.
Zeman, J., Cassano, M., Perry-​Parrish, C. & Stegall, S. (2006). Emotion regulation in children and
adolescents. Developmental and Behavioural Pediatrics, 27(2), 155–​168.
Zipfel, S., Giel, K., Bulik, C. M., Hay, P., & Schmidt, U. (2015). Anorexia nervosa: aetiology, assessment and
treatment. The Lancet Psychiatry, 2(12), 1099–​1111.
Chapter 11

Emotion Regulation and Substance Use


Disorders in Adolescents
Thomas A. Wills, Jeffrey S. Simons, Olivia Manayan,
& M. Koa Robinson

Substance use disorders


This chapter considers the role of emotion regulation with regards to vulnerability to substance
use disorders in adolescence. While a number of young people only experiment at low levels with
tobacco, alcohol, or other substances in early adolescence (11–╉14 years of age), a proportion of these
escalate their frequency and intensity of use over time (Colder et al., 2002; White et al., 2002; Windle
& Wiesner, 2004). In later adolescence (15–╉18 years of age), frequent substance users are more likely
to transition to substance use disorder, though a sizable proportion do not (Harrison, Fulkerson,
& Beebe, 1998; Simons, Carey, & Wills, 2009; Wills, Sandy, & Yaeger, 2002). Emotion regulation
processes are implicated in the likelihood of transitioning from frequent use to the development
of a disorder. Yet, while theoretical papers have outlined conceptual approaches to understanding
how emotion regulation relates to risk for substance abuse (Khantzian, 1990; Southam-╉Gerow &
Kendall, 2002), there is still little direct evidence on this question from adolescents.
The purpose of this chapter is to provide a conceptual approach to understanding how emo-
tion regulation is involved in vulnerability vs. protection for adolescent substance abuse. It is
important to emphasize that many questions about the relation of emotion and substance use
are not settled at this time. For example, it remains unclear how emotional distress is causally
related to disorder (Cheetham et al., 2010; Swendsen & Le Moal, 2011), how negative affect and
substance use are related in daily life (see for ex. Kassel & Veilleux, 2010; Simons, Wills, & Neal,
2014), and how the physical effects of substances may contribute to the dysregulation of emotion
(Baker et al., 2004; Koob & Le Moal, 2008). These topics will be illuminated by discussing current
questions about emotion and its regulation and then showing how these are relevant for clinical
research and practice with adolescents. An initial review of the data will highlight the prevalence
of tobacco, alcohol, and marijuana use among adolescents, summarize the prevalence of sub-
stance use disorders in late adolescence, and discuss clinical issues in diagnosing these disorders
in the adolescence populace. Once prevalence rates have been addressed, the emotion regulation
processes relevant for substance use disorders will be discussed and reviewed. Clinical implica-
tions will be outlined, with a focus on a preventive intervention based on emotion regulation
concepts that has demonstrated efficacy in reducing substance abuse issues. Finally, the current
state of the area will be summarized and directions for future research will be discussed.

Prevalence of substance use
Data on the prevalence of substance use among adolescents is available from several US national
studies. The Monitoring the Future (MTF) project has been conducted annually since 1975
(Johnston et al., 2015). This project is a repeated series of cross-╉sectional school-╉based surveys,
with the same set of questions given every year to comparable age groups. Trends for prevalence
of tobacco, alcohol, and marijuana use delineated in the MTF survey have also been observed in
other national surveys with different sampling and data collection methods, such as the Youth
Risk Behavior Surveillance Study (Frieden et al., 2014) and the National Survey on Drug Use and
Health (Center for Behavioral Health Statistics, 2015).
The 2014 MTF survey included about 41,600 students who were in eighth, tenth, or twelfth
grade in public or private secondary schools across different areas of the United States. In this
survey, the lifetime prevalence for any alcohol use was 27%, 49% and 66% for eighth, tenth, and
twelvth graders, respectively. Rates of alcohol use within the past 30 days, an index of regular use,
were 9%, 24% and 37% for these same age groups. Hence, the prevalence of alcohol use was shown
to be substantial, particularly in later adolescence. Consistent with other surveys, the MTF study
has found marijuana to be the most widely used illicit drug over the 40-╉year history of the survey
(Johnston et al., 2015). In the 2014 survey, the prevalence for ever-╉use of marijuana or hashish
ranged from 12% to 35% among eighth to twelfth graders, and the prevalence of use during the
past 30 days was 6%, 17%, and 21% for these same age groups. Thus, it is observed that usage rates
increase steadily by age for all substances. The study also provided information regarding differ-
ing patterns of use with regards to gender, ethnicity, and socioeconomic status; however, detailed
discussion of this data is beyond the scope of the present chapter.
With regard to secular trends, the absolute level of cigarette use has declined steadily over the
last decade. The most recent MTF survey found that cigarette use among adolescents is now at
the lowest level recorded in the history of the survey (30-╉day prevalence of 4%, 7%, and 14% for
eighth, tenth, and twlveth graders, respectively). Although MTF data on adolescents shows rates
of marijuana use that have remained stable in recent years, Hasin et al. (2015) noted that the rate
of marijuana use in the U.S. adult population has doubled in the past ten years. MTF data have
shown a steady decline in perceived risk of marijuana use among adolescents, and because of
policy changes in some U.S. states, rates of teenage use are being watched with concern.

Prevalence of substance use disorder


Although many adolescents experiment with alcohol or marijuana use, only a proportion prog-
ress to developing a substance use disorder (SUD; Wills et al., 2002). The differentiating charac-
teristics of substance use disorders involve physiological, behavioral, and cognitive changes in the
user and continued use despite negative, substance-╉related consequences (American Psychiatric
Association, 2013). Studies conducted during the past two decades to determine the prevalence
of disorders in the U.S. adolescent population are summarized in Table 11.1. The researchers used
DSM-╉III or DSM-╉IV criteria for diagnoses and reported separate prevalences for substance abuse
and substance dependence. (The term Substance Use Disorder has been recently adopted as the
sole diagnostic term in the DSM-╉5.)
Kilpatrick et al. (2000) were the first to assess the prevalence of substance use and abuse, as
defined by DSM-╉IV guidelines, in a U.S. national sample. Data were collected through telephone
interviews with a sample of adolescents ages 12–╉17 years, recruited through random-╉digit dial-
ing. Hard drugs were defined as cocaine, heroin, inhalants, LSD, or prescription drugs. This study
found a 12-╉month prevalence of 8% for alcohol abuse/╉dependence, 7% for marijuana abuse/╉
dependence, and 2% for hard drug abuse/╉dependence.
In a study by Merikangas et al. (2010), face-╉to-╉face interviews were conducted in households
by trained research staff with a sample of 10,123 adolescents ages 13–╉18 years. This study used a
modified version of the World Health Organization (WHO) Composite International Diagnostic
Interview (CIDI) designed to be better suited to the lifestyles and experiences of adolescents. This
2
1

212 Emotion Regulation and Substance Use Disorders in Adolescents

Table 11.1╇ Studies on Prevalence of Substance Use Disorder among Adolescents

Citation Sample N Alcohol Marijuana Illicit Drug Any


Abuse/ Abuse/ Abuse/ Substance
Dependence Dependence Dependence Disorder
Kilpatrick et al., 2000 National 4,023 8%A 7% 2% —╉
Merikangas et al., 2010 National 10,123 6%B n.a. 9% 11%
SAMHSA, 2014 National 17,046 2.7A 2.7 3.5 5.0
A 12-╉month prevalence.
B Lifetime prevalence.

study revealed the most-╉reported illegal drugs used were marijuana, cocaine, and illicit prescrip-
tion drugs (Swendsen et al., 2012). These investigators reported a lifetime prevalence of 6% for
alcohol abuse/╉dependence, 7% for illicit drug abuse/╉dependence, and 11% for any substance use
disorder. Thus, consistent with Kilpatrick et  al. (2000), there was an appreciable prevalence of
substance use disorder observed in the adolescent population.
The National Survey on Drug Use and Health (NSDUH) is an annual survey of the US
population and most recently has used DSM-╉IV criteria with a 12-╉month time frame. Data
for the 2014 survey for adolescents aged 12–╉17 years (Center for Behavioral Health Statistics,
2015) showed that 2.7% of the sample had an alcohol abuse diagnosis, 2.7% had a marijuana
abuse diagnosis, and 3.5% had an illicit drug abuse diagnosis (including marijuana but also
cocaine, heroin, inhalants, and non-╉prescribed prescription drugs). Overall, 5.0% of the ado-
lescent population was indicated as having any kind of substance use disorder. NSDUH data
have shown declines in rates of disorder from 2002 to 2012, though most rates have been stable
for 2013 and 2014.
Across studies, a higher prevalence of SUD was found for adolescent boys when compared to
adolescent girls of the same age demographic. All studies saw an increase in SUD prevalence with
age, particularly between the ages of 14–╉17. Overall, lower rates of substance abuse and depen-
dence were found in minority ethnic groups compared to Caucasians (Kilpatrick et al., 2000,
Merikangas et al., 2010). Although rates of substance use disorders among adolescents are not
as high as for depression and anxiety disorders, studies have shown a significant comorbidity of
SUDs and other psychopathological conditions (Cheetham et al., 2010).

Diagnosis of substance use disorder: Adolescents


The latest diagnostic criteria in DSM-╉5 (American Psychiatric Association, 2013) define SUD as
a pattern of substance use that substantially interferes with social, occupational, or interpersonal
functioning. The occurrence of only two symptoms from a heterogeneous set of 11 symptoms is
required. Symptoms may be broadly clustered into domains of 1) impaired control over use (e.g.,
using larger amounts or over a longer period of time than intended; craving), 2) social impairment
(e.g., failure to fulfill role obligations in work, school, or home; interpersonal conflict), 3) hazard-
ous use (e.g., drinking and driving), and 4) pharmacological indicators of tolerance or withdrawal.
The pharmacological criteria notwithstanding, the syndrome may be broadly conceptualized as
an inability to effectively regulate substance use. In this regard, SUD is characterized by failed
efforts to control consumption, resulting in interference with social functioning, accumulating
negative consequences, and repeated, substantial risks to the self.
3
1
2

Emotion regulation strategies 213

There has been debate in the clinical literature about diagnosing SUD in adolescence. There
are two primary concerns regarding the appropriateness of SUD criteria to younger adolescents.
Firstly, younger adolescents generally have less opportunity to use substances due to parental
monitoring and restricted access; hence, use may be less frequent and more sporadic, which
substantially reduces the likelihood of symptoms such as withdrawal and of use interfering with
social role obligations (Kaminer & Winters, 2015). Moreover, the hazardous-╉use criteria often
reflect drinking and driving, which younger adolescents typically are precluded from (Winters,
2013). Finally, several criteria require fairly sophisticated executive functions and self-╉awareness,
such that individuals need to identify and set use limits; and impaired control is then inferred by
failure to meet these limits. However, the cognitive development of children and younger adoles-
cents reflects a period of heightened reward-╉seeking and socio-╉emotional functioning prior to the
maturation of executive control functions (Steinberg, 2008). Hence among adolescents, substance
use may be severely dysregulated in the sense that they do not have high control over it, yet they
are not using more than “intended” nor displaying failed efforts to “cut down or control” use
(Chung & Martin, 2005). Thus the standard diagnostic criteria could miss patterns of substance
use in children and adolescents that are of clinical relevance.
Second, even modest substance use by children and adolescents is highly likely to cause con-
flict with parents, educators, and law enforcement. This may be particularly pronounced with
female adolescents, who are more likely to report drinking despite interpersonal problems than
their older peers (Harford, Grant, Yi, & Chen, 2005). Furthermore, children and adolescents are
in the midst of a period of shifting peer groups, struggles with emotional and behavioral regula-
tion, and transitions in their involvement in school and recreational activities; and some have
yet to develop long-╉term commitments to educational and occupational pursuits. Thus, there is
the risk that experimental, essentially normative, substance use during this dynamic period may
be mislabeled as an SUD (Winters, 2013). Indeed, many individuals mature out of risky sub-
stance use patterns in young adulthood (Jochman, Fromme, & Scheier, 2010; Reich, Cummings,
Greenbaum, Moltisanti, & Goldman, 2015); for example, the frequency of binge drinking tends
to decline in the mid-╉20s (Reich et al., 2015). Although substance use frequency may decline,
some research suggests that observed declines in SUD with age reflect decreases in new cases and
lower risk of relapse rather than developmental changes in the persistence of SUD once estab-
lished (Verges et al., 2013). In other words, transition in and out of problematic substance use is
common (Compton, Dawson, Conway, Brodsky, & Grant, 2013), but the risk of developing new
problematic patterns of use tends to decline with age. In summary, diagnosis and recognition of
substance-╉related problems in youth requires a balanced consideration of the potential increased
vulnerability of the developing brain to substances (Spear, 2010), the potential for diagnostic cri-
teria to underestimate the severity of the problem (Kaminer & Winters, 2015), and conversely a
need to recognize that there is a certain amount of age-╉appropriate drug experimentation that
may not warrant costs associated with diagnosis and intervention (Winters, 2013).

Emotion regulation strategies


Research on specific emotion regulation strategies has focused on the regulation of negative emo-
tions (anxiety and depression). There is of course a reason for this because there is a substantial
comorbidity of affective disorders with substance use disorder (Kober, 2014). However, it should
be noted that there is still debate about the causal interpretation of the comorbidity (Cheetham
et al., 2010) and evidence is mixed on the relationship between negative affect and substance use
in laboratory studies and in daily life (see Kassel, Hussong, et al., 2010; Mohr et al., 2010; Shrier,
Ross, & Blood, 2014; Simons, Dvorak, Batien, & Wray, 2010; Sher & Grekin, 2007). It is possible
4
1
2

214 Emotion Regulation and Substance Use Disorders in Adolescents

that substance disorder develops because of elevated negative affect. However, it is also possible
that both substance use and affective disorder are attributable to an underlying, transdiagnostic
vulnerability factor (dysregulation of behavior and emotion being a plausible candidate) and/╉
or that the biological and social disruptions occasioned by substance abuse themselves produce
negative affect (Koob & Le Moal, 2008; Swendsen & Le Moal, 2011). A credible body of theory
also points to deficiencies in positive affect as an important but understudied influence on the
development of disorder (Gilbert, 2012). Low positive affect may occur because of a dispositional
deficiency in the ability to experience positive mood (reward deficiency syndrome or hedonic
capacity: Audrain-╉McGovern et al., 2012; Yacubian & Buchel, 2009) or because of lack of access to
alternative reinforcers (Audrain-╉McGovern et al., 2010).

Specific strategies for emotion regulation


Emotion regulation is a multifaceted domain ranging from relatively automatic biological pro-
cesses (e.g., vagal tone, amygdala) to effortful coping processes (e.g., reappraisal, breathing exer-
cises etc.) and meta-╉emotional constructs such as mindfulness (Brewer, Elwafi, & Davis, 2013;
McRae et al., 2012). There have been a considerable number of studies conducted in which a
specific emotion regulation strategy is related in a laboratory setting to a measure of a particular
emotion (Webb et al., 2012). This research has produced a number of theory-╉testing findings,
for example, the finding that distracting oneself from the stressor can have beneficial effects; but
generalization of the laboratory paradigms to risk for drug use or abuse remains unknown. Aldao,
Noelen-╉Hoeksema, and Schweizer (2010) have considered studies conducted with distressed
clinical samples of adults, some of which included substance use as an outcome (e.g., Noelen-╉
Hoeksema et al., 2007). In short, the strategies they examined include: Reappraisal, problem solv-
ing, acceptance, avoidance, and suppression. Findings on these will be summarized below.
Reappraisal involves changing one’s perception of a problem so that it is perceived as less serious
or less threatening. The meta-╉analysis by Aldao et al. (2010) concluded that on average this was
a moderately effective strategy. Problem solving involves getting information about the problem,
considering alternative solutions to a problem, and making an active effort to change the situation.
Though it is behavioral, not emotional in nature, meta-╉analysis has supported problem solving as
an effective emotion regulation strategy (Aldao et al., 2010). Acceptance involves non-╉judgmental
acceptance of the distressed emotion and has been posited to prevent negative emotions from tak-
ing over and overwhelming other coping efforts; it involves accepting sensations, such as craving,
for what they are. Meta-╉analyses so far have not shown strong evidence across psychopathology
groups for acceptance as a single emotional regulation strategy (Aldao et  al., 2010). However,
studies focused on drug abuse have shown mindfulness to be a significant protective factor against
substance use and relapse (Brewer et al., 2013; Elwafi et al., 2013).
Avoidance of negative emotions has been studied in clinical settings and across pyschopathol-
ogy groups; it has been shown to have a strong adverse effect, creating emotional distress rather
than reducing it (Aldao et al., 2010). Indeed, reducing avoidance is a central component of the
unified protocol for treating emotion disorders (Moses & Barlow, 2006). However, although per-
sistent avoidance appears detrimental, distraction appears to be an effective strategy for the initial
response to high-╉intensity emotional stimuli (Sheppes et al., 2014). Suppression is a response-╉
focused emotion regulation strategy that has been linked to negative health outcomes. Trying
to suppress unwanted thoughts and emotions has been examined in a number of laboratory
and clinical studies and is consistently found to be an ineffective strategy (Aldao et al., 2010).
Paradoxically, attempts to suppress emotion usually increase the occurrence of the unwanted
thoughts and feelings (Webb et al., 2012).
5
1
2

General attributes of emotion regulation 215

General attributes of emotion regulation


There are more studies that have related general attributes of emotional regulation to risk for
substance use or disorder. In this section five key attributes will be considered, namely: reactivity,
soothability, distress tolerance, affective variability, and emotional inertia accompanied by rumi-
nation. These are summarized with sample measurement items in Table 11.2. It should be noted
that these attributes can be delineated separately, but while recognizing that they are conceptually
distinct, we note that they may be empirically intercorrelated.
In relation to reactivity, the origins of vulnerability to substance use disorder are believed to
be rooted in early temperament characteristics (Tarter et al., 1999; Wills, Sandy, & Yaeger, 2000).
Indeed, temperament characteristics measured at three to five years predict risk for substance
abuse and other psychopathology dimensions at age 21  years (Caspi et  al., 1996; Kirisci et  al.,
2015). Temperament research with young children has distinguished irritability and reactivity to
aversive stimulation as a basic dimension of temperament, which is predictive of substance use
but may, however, be modulated by cognitive control (Rothbart, Ahadi, & Evans, 2000; Wills et al.,
2000). Initial temperamental reactivity can be increased by early adversity (e.g., poverty, child
abuse), which works to change hypothalamic–╉pituitary–╉adrenal (HPA) functioning so as to make

Table 11.2╇ Sample Items for General Attributes of Emotion Regulation or Dysregulation

Reactivity (Angerability)
When I have a problem at school or at home:
I get mad at people.
I yell and scream at someone.
Anger Control
When I am angry or upset:
I stay calm and “keep my cool” when I’m feeling mad.
I try to calmly deal with what is making me mad.
Soothability
I can easily calm down when I am excited or “wound up.”
If I get upset or distressed, I can recover quickly.
Sadness Control
When I am feeling sad or down:
I can control my sadness and carry on with things.
I stay calm and don’t let sad things get to me.
Distress Tolerance
I can’t handle feeling distressed or upset. (Disagree)
When I feel distressed or upset, I must do something about it immediately. (Disagree)
Affective Lability
My moods change a lot from day to day.
I shift back and forth from feeling calm to feeling tense and “jittery.”
Rumination
I often find myself thinking about things that have made me angry.
I get angry thinking about things that have happened in the past.
Sources: Oliver & Simons, 2004; Simons & Gaher, 2005; Wills et al., 2006, 2011, 2013.
6
1
2

216 Emotion Regulation and Substance Use Disorders in Adolescents

some individuals even more reactive to stress (Andersen & Teicher, 2009). Other things equal,
persons who are more reactive to stress are at an increased risk for substance abuse and other
disorders (Siegel, 2010, 2015; Sinha, 2008).
Soothability, an additional key element related to substance abuse, is defined as the ability to
reduce aversive arousal states through one’s own efforts. Khantzian (1990) originally noted that
clients with drug use disorder experienced difficulty soothing or calming themselves when they
were in stress-​provoking situations. Khantzian (1990) suggested, in his self-​medication model,
that this was a basic process in perpetuating drug consumption because persons with low
soothability turned to drugs for more immediate relief. Effective implementation of self-​soothing
produces lowered arousal, which makes it easier to pursue effortful, active coping efforts and has
the additional benefit of not alienating supporters through lashing out in anger. Soothability, as
a general attribute, may involve several of the specific strategies outlined above (e.g., attentional
focusing, distraction, reappraisal). Research has demonstrated that measures of soothability are
positively correlated with other indices of emotional control ability and are inversely related to
substance use in early adolescence (Wills et al., 2006).
Distress tolerance is another key element that is intricately tied to risk for substance use disor-
der. The ability to reflect on feeling states and engage in adaptive coping responses may depend,
in part, on an individual’s ability to tolerate distress. In this regard, distress tolerance may be
considered a meta-​emotion construct that incorporates the perceived ability to withstand dis-
tress; appraisal of distress; efforts to stop distress; and the tendency to become absorbed by dis-
tress (Simons & Gaher, 2005). Low tolerance for distress has been linked to substance use (Leyro,
Bernstein, Vujanovic, McLeish, & Zvolensky, 2011; Wray, Simons, Dvorak, & Gaher, 2012), to
other indicators of dysregulated affect, such as deliberate self harm (Arens, Gaher, Simons, &
Dvorak, 2014), and to psychopathology syndromes linked to dysregulated affect such as posttrau-
matic stress disorder and borderline personality disorder (Gaher, Hofman, Simons, & Hunsaker,
2013; Vujanovic, Marshall-​Berenz, & Zvolensky, 2011). In addition, reduced ability to differen-
tiate, label, and understand the source of emotion states (i.e., alexithymia) has been inversely
associated with distress tolerance (Gaher et al., 2013). Both low distress tolerance and deficits in
emotional awareness have been associated with poor behavioral control, especially when nega-
tively aroused (Emery, Simons, Clarke, & Gaher, 2014; Gaher et  al., 2013; Shishido, Gaher, &
Simons, 2013). Thus, the lack of understanding of emotional states and the inability to tolerate or
accept aversive feeling states may increase the likelihood of incurring substance-​related problems
(Buckner, Keough, & Schmidt, 2007; Emery et al., 2014; Shishido et al., 2013), whilst also inter-
fering with engagagement in substance use treatment (Daughters et al., 2005). Taken together,
the findings suggest that the clarity of emotional experience and the ability to mindfully accept
emotions are indicative of adaptive emotion regulation. In contrast, poor tolerance for distress
and limited awareness of emotional experience (e.g., poor differentiation, poor labeling or under-
standing) promote impulsive responding, efforts to suppress emotion, and maladaptive substance
use outcomes.
An additional factor that impacts substance use disorder is affective variability. In recent years,
there has been increased understanding of the importance of the dynamic time course of emo-
tion in studies of emotion regulation (Ebner-​Priemer, Eid, Kleindienst, Stabenow, & Trull, 2009;
Fairbairn & Sayette, 2013; Simons, Wills, & Neal, 2014). Affective lability refers to the speed, fre-
quency, and range of changes in affective states (Oliver & Simons, 2004). Studies on borderline
personality disorder and depression have highlighted the importance of instability of affect, over
and above mean levels, in contributing to pathology (Jahng et al., 2011). Similarly, research on
substance use problems has indicated significant effects of affective variability, often over and
above mean affect level, such that individuals who are more variable in mood show more alcohol
7
1
2

Emotion regulation and substance use/abuse in adolescence 217

and marijuana related problems and tobacco use (Dvorak & Simons, 2008; Mohr, Arpin, &
McCabe, 2015; Simons & Carey, 2006; Simons et al., 2014; Weinstein & Mermelstein, 2013a).
In respect to emotion regulation strategies for reducing affective variability, research indicates
that the source (e.g., bottom-╉up vs. top-╉down) as well as the intensity of emotions influences the
selection and effectiveness of emotion regulation strategies (McRae, Misra, Prasad, Pereira, &
Gross, 2012; Sheppes et al., 2014). In this regard, cognitive strategies, such as reappraisal, appear
to be more effective for emotions that are, in part, the result of cognitive evaluations (McRae et
al., 2012). Cognitive reappraisal is more often utilized when individuals are experiencing rela-
tively low-╉intensity emotions. In contrast, when emotions are of high intensity, strategies such as
distraction, which minimizes emotional processing, are efficacious (Sheppes et al., 2014). Taken
together, these results suggest that individuals who experience intense, unpredictable shifts in
emotion may be predisposed towards relying on regulatory strategies that minimize awareness
and processing of the emotional stimuli, making substance use an attractive option. This view
is consistent with Khantzian’s (1990) hypothesis that the negative consequences experienced by
individuals with a drug use disorder are less salient than the fact that he/╉she can control emotional
states through drug use.
A final factor influencing substance abuse is emotional inertia and rumination. Emotional iner-
tia (i.e., the autocorrelation of emotion across time) has also been identified as a central construct
in research on affect dysregulation (Kuppens et al., 2012) and substance use (Fairbairn & Sayette,
2013). In contrast, affective lability indicates an inability to maintain homeostasis and continu-
ity in emotional responding. Emotional inertia is indicative of dysregulation in emotion, such
that emotion is likely fixed by an inward, ruminative focus rendering the person disconnected
from important contextual stimuli that the emotion is expected to vary in response to (Fairbairn
& Sayette, 2013; Koval, Kuppens, Allen, & Sheeber, 2012). Research suggests that some of the
reinforcing properties of alcohol may stem from alcohol’s ability to disrupt emotional inertia
(Fairbairn & Sayette, 2013). Alcohol myopia theory (Steele & Josephs, 1988) predicts that tension-╉
reduction properties of alcohol are due, in part, to the effects of alcohol on limiting focus to
immediately salient stimuli.

Emotion regulation and substance use/╉abuse in adolescence


In this section representative studies relating emotion regulation attributes to risk for substance
use or use disorder in young persons are highlighted (Table 11.3). There have been few studies
conducted in early or middle adolescence (ages 12–╉16 years); based on this fact, this chapter will
include studies conducted in late adolescence or young adulthood (ages 17–╉25 years). Initially,
studies that have related a single emotional measure to substance use will be discussed. Following
this, studies that used composite scores based on a dual-╉process approach (Gibbons et al., 2009;
Wills et al., 2013) will be delineated. This approach posits distinct systems of regulation and dys-
regulation, which have different antecedents and different consequences.

Emotion dysregulation and motives for use


Motives for substance use are a significant predictor of disorder (Wills & Ainette, 2010) and may
derive from deficits in emotion regulation. Several recent studies have tested how the Difficulties
in Emotion Regulation Scales (DERS; Gratz & Roemer, 2004; Weinberg & Klonsky, 2009) is
related to high-╉risk motives for use, particularly using drugs to deal with personal distress (i.e.,
coping motives). The DERS assesses difficulties in controlling emotion as well as dimensions of
suppression and non-╉acceptance of emotion. Dvorak et al. (2014) found that difficulties with
controlling negative emotion were related to higher frequency of drinking and to more adverse
8
1
2

Table 11.3  Studies of Emotion Regulation and Substance Use, For Single Strategies and Dual-​Process
Constructs

Citation Mean age (yrs) Findings


General emotion regulation difficulties
Dvorak et al. (2014) 20.5 Difficulties with ER related to adverse alcohol
consequences.
Veilleux et al. (2014) 19.7 Limited ER strategies related to drinking to cope.
Distress tolerance
Buckner et al. (2007) 18.7 Tolerance related to fewer alcohol and cannabis
problems.
Winward et al. (2014) 17.7 Heavy-​drinking adolescents had lower distress
tolerance.
Emotion dysregulation and PTSD
Weiss et al. (2013) 35.5 Emotion dysregulation related to PTSD in SUD
sample.
Gaher et al. (2014) 28.9 Emotional intelligence inversely related to
impulsivity, PTSD.
Affective lability
Simons et al. (2009) 19.6 Affective lability predicted alcohol dependence,
but not abuse.
Weinstein & 15.7 Affective lability predicted smoking escalation
Mermelstein (2013b) among girls.
Hedonic capacity
Audrain-​McGovern 15.7 Low hedonic capacity was related to escalation of
et al. (2012) smoking.
Goelz et al. (2014) 49.7 Successful smoking quitters had more alternative
reinforcers.
Trait mindfulness
Tarantino et al. (2015) 19.9 Mindfulness tendency inversely related to drug
use/​problems.
Black et al. (2012) 16.2 Mindfulness had indirect effect to smoking
through less depressive affect, perceived stress.
Dual-​process studies
Dvorak, Simons, & Wray (2011) 20.2 Dysregulation had less impact on alcohol problems
among persons scoring higher on self-​control.
Wills et al. (2011) 16.0 Emotional dysregulation had direct + indirect
effects to dependence problems.
Wills et al. (2016) 12.5 Emotional dysregulation had direct + indirect
effects to both externalizing and internalizing
symptomatology. Emotional self-​control had direct
effect to positive well-​being.
9
1
2

Emotion regulation and substance use/abuse in adolescence 219

consequences (i.e., alcohol abuse) among college students. In addition, non-╉acceptance of emo-
tion was related to more adverse consequences among the most problematic drinkers. Veilleux et
al. (2014) found that lack of emotional control strategies and lack of clarity about emotion were
both predictive of high-╉risk alcohol use (drinking to cope). Furthermore, Simons et al. (2005a)
have shown that expectancies about one’s competencies in the regulation of negative mood were
inversely related to coping motives for marijuana use. Additional studies have related emotion
regulation difficulties to substance-╉related problems in samples of college students (Chandley et
al., 2014; Messman-╉Moore & Ward, 2014) and clinical samples of individuals with substance use
disorder (Buckholdt et al., 2014; Fox et al., 2008). Wong et al. (2013) found emotional suppression
was positively related to prescription drug misuse and illicit drug use. In contrast, persons who
actively coped with negative emotions had less illicit drug use and were more likely to also use
other adaptive coping strategies, such as support-╉seeking and positive reappraisal.

Distress tolerance
Overall emotional experience may depend on the ability to tolerate negative emotions. This con-
struct is often measured using the Distress Tolerance Scale (DTS, Simons & Gaher, 2005), which
includes subscales termed Tolerance, Absorption, Appraisal, and Regulation (needing to act
immediately when distressed). Buckner et al. (2007) found that a better ability to tolerate distress
was related to lower frequency of alcohol use and was inversely related to alcohol and cannabis
problems. They suggested that individuals with low distress tolerance are more likely to use sub-
stances for regulating negative emotions. Winward et al. (2014) studied adolescents with heavy
episodic drinking (HED) and matched controls. The HED adolescents initially had low distress
tolerance but showed decreased emotional reactivity when they became abstinent.
Other studies have shown distress tolerance inversely related to coping motives in a sample of
current marijuana smokers (Zvolensky et al., 2009) and a sample of Posttraumatic Stress Disorder
(PTSD)-╉affected young adults (Marshall-╉Berenz et  al., 2011). One study found a laboratory
measure of distress tolerance inversely related to alcohol use in a sample of younger adolescents
(Daughters et al., 2009). These studies complement findings showing that low distress tolerance
predicts lapse and relapse among smoking cessation clients (e.g., Brown, Lejuez, & Kahler, 2002).

Emotion dysregulation, PTSD, and substance abuse


Several studies have implicated poor emotion regulation in stress-╉ related symptomatology
and substance abuse in college students (Gaher et al., 2013; Goldsmith et al., 2013; Weiss et al.,
2012) and victimized populations (Hellmuth et al. 2013; Sullivan et al., 2012). Of particular note
from a clinical standpoint are studies showing how maltreatment in early childhood contributes
to an increased risk for substance use disorder in adolescence. Studies have now implicated under-
mining of self-╉regulation as a pathway for effects of maltreatment, alternately showing mediation
of these effects through decreased problem solving (Shin, Hong, & Wills, 2012), through more
affect dysregulation and PTSD (Oshri et al., 2015; Rosenkranz, Muller, & Henderson, 2014), or
through alexithymia and impulsivity (Hahn, Simons, & Simons, 2015).
Studies of trauma-╉exposed adult samples have consistently shown PTSD to co-╉occur with ele-
vated rates of substance abuse (e.g., Hellmuth et al., 2012, 2013). Recent studies have clarified
how difficulties in emotion regulation may contribute to this comorbidity. Weiss et al. (2013)
studied SUD inpatients using the DERS. Patients with PTSD scored higher on emotion dysregu-
lation, with the strongest differences on the subscales for nonacceptance of emotion and limited
emotion regulation strategies. Gaher et al. (2013) included an emotional intelligence measure in
an experience sampling study with heavy-╉drinking military veterans. Between-╉person analyses
0
2

220 Emotion Regulation and Substance Use Disorders in Adolescents

showed alcohol problems were related to lower emotional intelligence and positively correlated
with PTSD symptoms. On a repeated-╉measures basis, occurrence of PTSD symptoms during the
day was related to increased alcohol use and associated problems the same night.

Affective lability
Some investigations of affective lability have used a dispositional measure, the Affective Lability
Scales (ALS, Oliver & Simons, 2004); others have constructed statistical indices of variability in
mood from repeated-╉measures data. The ALS was used in a longitudinal study with heavy-╉drinking
college students by Simons, Carey, and Wills (2009). A structural modeling test of the influence
of affective lability, controlling for behavioral dysregulation, found lability was not related to level
of alcohol consumption but it did predict change in alcohol dependence symptoms. Behavioral
dysregulation, in contrast, predicted abuse symptoms but not dependence symptoms. This find-
ing was replicated in an experience sampling study in which daily reports were obtained over a
two-╉year period and variability was indexed by a statistical algorithm (Simons, Wills, & Neal,
2014). Here, lability in negative mood showed a direct relation to the likelihood of alcohol depen-
dence symptoms. This research shows the value of distinguishing variability in mood over time
from mean level of mood (see also Simons & Carey, 2002, 2006; Simons et al., 2005b). Notably,
trait positive mood was related to a lower proportion of drinking days over the study period (cf.
Gilbert, 2012); whereas, trait negative mood was related to a higher proportion of drinking days.
Weinstein and Mermelstein (2013b) studied a sample of high school students through obtain-
ing experience sampling reports on palmtop computers. Reports of smoking were obtained over
one-╉week periods on two occasions and variability in negative mood was determined by a statisti-
cal algorithm. Greater mood variability predicted escalation of smoking in the subsample of girls,
while mean level of negative mood (but not its variability) predicted escalated smoking among
boys who scored higher on coping motives for smoking. This shows the value of including both
gender and motives for use in research designs.

Hedonic capacity and alternative reinforcers


Emotion regulation processes may be particularly relevant for persons with lower capacity to
experience pleasure from natural reinforcers. Low positive mood could be the stimulus for sub-
stance use (Mohr et  al., 2008; Wills et  al., 1999)  but may be countered through active coping
mechanisms such as a search for alternative reinforcers (Audrain-╉McGovern et al., 2010). A dis-
positional measure of hedonic capacity, the Snaith-╉Hamilton Pleasure Scale (SHAPS, Franken
et al., 2007) has been used in some studies.
Audrain-╉McGovern et al. (2012) followed a sample of adolescents over four waves of observa-
tion, measuring hedonic capacity with the SHAPS and indexing cigarette smoking at each assess-
ment. Results showed lower hedonic capacity was related to greater likelihood of smoking in the
past month and a greater rate of increase in smoking over time. Among adult smokers attempting
to quit, Goelz et al (2014) found that those who successfully quit smoking showed higher levels
of alternative reinforcers compared with persons who relapsed. These results complement find-
ings showing declining levels of alternative reinforcers related to increases in smoking over time
(Audrain-╉McGovern et al., 2010) and anhedonia related to a lower likelihood of smoking cessa-
tion (Leventhal et al., 2009).

Mindfulness and substance abuse


Mindfulness training as an approach to emotion regulation has been tested in several intervention
studies for smoking and alcohol use (Brewer et al., 2013). Dispositional measures are also avail-
able including the Mindfulness Attention Scale (MAAS, Black et al., 2012).
1
2

Composite constructs and dual-process theory 221

Tarantino et  al. (2015) surveyed a sample of college students with the MAAS as a predictor
and with measures of drug use and drug problems as outcome variables. Higher mindfulness
tendency was correlated with behavioral self-╉control and self-╉reinforcement, and was related to
lower levels of drug use and problems independently of other coping strategies. A clinical study
with a mindfulness-╉based smoking cessation program (Brewer et al., 2011) found that those who
practiced meditation more frequently showed a lower relation between craving and cigarette
use (Elwafi et al., 2013). Also noteworthy is evidence showing neurological changes suggesting
enhanced self-╉regulation as a result of mindfulness training (Tang et al., 2012).

Composite constructs and dual-╉process theory


Dual-╉process theory posits that there are two distinct systems for dealing with the environment,
which have different antecedents and different consequences. The system for self-╉control (also
termed controlled processing, reflection, or reasoned processing) is more deliberate, conscious,
fact-╉based, and slower. The system of reactive processing (also termed impulsiveness, disinhi-
bition, or automatic processing) is faster but is based more on images, heuristics, and reward
reactions rather than deductive reasoning. Dual-╉process theories have been developed for sub-
stance abuse (Gerrard et al., 2008; Gibbons et al., 2009; Volkow & Baler, 2012; Wiers et al., 2007;
Wills & Dishion, 2004) as well as for other health behaviors (Hoffman, Friese, & Strack, 2009;
Rothman et al., 2009) and for depression (Beevers, 2005). The main propositions relevant to the
present discussion focused on emotion regulation are that 1)  emotional self-╉control and emo-
tional dysregulation will make independent contributions to outcomes (in different directions);
and 2) emotional self-╉control and emotional dysregulation will have different types of pathways
to outcomes.
Studies of younger adolescents have confirmed these propositions for behavioral measures,
showing that self-╉control and dysregulation constructs make independent contributions to entry-╉
level substance use and have different types of pathways to substance use outcomes (e.g., Wills
et al., 2001). Pearson et al. (2013) showed that individuals scoring high on behavioral self-╉control
had fewer alcohol problems because they applied more protective behavioral strategies when
drinking. In contrast, behavioral dysregulation was directly related to more alcohol-╉related prob-
lems, independent of level of use.
A study by Wills et al. (2006) added constructs of emotional self-╉control and emotional dysreg-
ulation to behavioral constructs and found that these constructs showed similar types of relations
to substance use among both younger adolescents and older adolescents. A subsequent study with
high school students (Wills et  al., 2011)  used composite constructs for emotional self-╉control
(soothability, sadness management, and anger management) and emotional dysregulation (affec-
tive lability, angerability, and rumination). Behavioral and emotional constructs made indepen-
dent contributions to outcomes and had different types of pathways. Emotional dysregulation
had indirect effects to more substance problems through lower academic competence and more
negative life events; it also had direct effects to more dependence and abuse problems.
Is emotion regulation related to psychopathology before the onset of substance use problems?
This question was examined by Wills et al. (2016) with a large sample of 12 to 13-╉year old ado-
lescents. Composite constructs for emotional self-╉control and emotional dysregulation were
obtained, similar to those in previous studies, together with measures of externalizing and inter-
nalizing symptomatology, both of which predict substance use disorder at later ages (Colder et al.,
2010; Wills et al., 2005). Analyses including behavioral self-╉control as a covariate indicated there
were both direct and indirect effects, outlined schematically in Figure 11.1.
Emotional dysregulation was related to psychopathology partly via indirect effects (through
lower academic competence and more negative life events) and partly through direct effects to
2

222 Emotion Regulation and Substance Use Disorders in Adolescents

Emotional Deviance-prone Externalizing


dysregulation attitudes symptomatology

Negative Internalizing
life events symptomatology

Emotional Developed Positive


self-control competencies well-being

Figure 11.1╇ A conceptual model of direct and indirect effects of emotion dysregulation and
emotional self-control. Model emphasizes effects of emotional regulation via both socio-
environmental and cognitive/attitudinal constructs. See Wills et al., 2016.

more externalizing and internalizing symptomatology. Independently, emotional self-╉control had


several significant indirect effects (through less life stress and deviance-╉prone attitudes) and it
also had a large direct effect to positive well-╉being. This research indicates that these types of
psychopathology are in some sense disorders of self-╉regulation but it is also true that emotional
regulation shapes level of exposure to environmental risk and protective factors.

Example of emotion regulation intervention


There have been few interventions conducted for substance use based directly on self-╉regulation
constructs and most of these were done with adults (see Wills, Simons, & Gibbons, 2015). There
will be discussion on a school-╉based intervention for adolescents that was centered on emotion
regulation strategies. School-╉based interventions are an attractive venue for prevention research
because they can access a general population and provide a unique environment for students to
interact with facilitators (i.e., teachers) in both structured and unstructured settings (Diamond &
Lee, 2011; Skara & Sussman, 2003). Conrod et al. have focused on a range of psychopathological
symptoms, utilizing a preventive approach that aimed to prevent substance abuse by lessening its
precursors. The risk factors considered in these studies were more individual in nature as opposed
to contextual (O’Leary-╉Barrett et al., 2013).
All studies took place at high schools in the United Kingdom and Canada. A trained, school-╉
based facilitator and co-╉facilitator administered the intervention, which consisted of two manual-╉
guided 90-╉minute sessions. The manuals consisted of three components: 1) A psychoeducational
component, 2)  a motivational component, and 3)  a cognitive behavioral therapy component.
Later editions of the manual also included real-╉life scenarios shared by high-╉risk British youth in
focus group sessions (Conrod, Castellanos-╉Ryan, & Strang, 2010). Data were collected through
self-╉report questionnaires distributed at baseline, at the completion of the intervention, and at
six-╉month intervals for two years post-╉intervention.
Four personality profiles were targeted in these studies:  Anxiety Sensitivity, Hopelessness,
Impulsivity, and Sensation Seeking (O’Leary-╉Barrett et  al., 2013). Students who scored at least
3
2

Directions for further work 223

one standard deviation above the school average in one of these four subscales were categorized as
high-╉risk. In earlier studies, all high-╉risk students were invited to participate in the intervention.
In later studies, all consenting high-╉risk students were randomly assigned to either the interven-
tion group or the control group. Two studies also included low-╉risk students in both the interven-
tion and control groups (Conrod, Castellanos-╉Ryan, & Strang, 2010; Conrod, Castellanos-╉Ryan,
& Mackie, 2011; Conrod et al., 2013). Internalizing symptom severity was measured using the
Depression and Anxiety subscales from the Brief Symptom Inventory (BSI). Externalizing
symptom severity was measured using the Conduct subscale from the Strengths and Difficulties
Questionnaire (SDQ; O’Leary-╉Barrett et al., 2013).
Conrod et al. (2010) aimed to improve various cognitive and behavioral problems associated
with specific high-╉risk personality factors. For instance, the cognitive distortion termed over-
generalization (i.e., when one makes universal assumptions based on experiences through one
specific situation) is often found in individuals prone to depression (O’Leary-╉Barrett et al., 2013).
By targeting cognitive distortions such as these, it was proposed that a variety of internalizing
and externalizing symptoms that may underlie substance use could be effectively addressed.
Interventions were focused on assisting students in each personality profile to adopt more adap-
tive coping mechanisms, so as to help students manage their personality risk in a way that does
not promote problematic substance use (Conrod, Castellanos-╉Ryan, & Strang, 2010).
The first part of the intervention involved a goal-╉setting exercise that encouraged behavior
change and development of new coping methods. Students were taught about their target person-
ality variable and associated non-╉adaptive coping responses (e.g., avoidance, aggression). They
then learned about the cognitive-╉behavioral therapy model and practiced applying it by analyzing
emotional responses in sample scenarios, as well as their personal experiences. Students were
encouraged to recognize and confront personality-╉specific cognitive distortions that could induce
risk behaviors. Post-╉exercise discussions were held to focus on personality-╉specific thoughts,
emotions, and behaviors (Conrod et al., 2010, 2013).
Across all studies, significantly lower frequency and quantity of overall alcohol consumption,
as well as binge drinking and rate of growth of binge drinking, were found at the completion
of intervention for high-╉risk adolescents assigned to the intervention group. In the 2013 study,
benefits of the intervention were also apparent at the 24-╉month follow-╉up, as displayed through
lower growth in drinking quantity and binge drinking frequency, compared to non-╉intervention
adolescents (Conrod et al., 2013). The intervention was also associated with a reduced likelihood
of marijuana and cocaine use (Conrod, Castellanos-╉Ryan, & Strang, 2010).
Conrod’s approach did not directly target substance misuse. Rather, the intervention targeted
individuals who displayed personality risk factors previously shown to correlate with substance
use disorders. By using specific personality profiles, the individuals learned which strategies of
emotion regulation were most beneficial, as well as which strategies were more detrimental in
relation to their personalities. By doing so, Conrod et al. addressed, at a more selective level, the
emotion dysregulation that is often proximal to substance use, or is involved with motivational
processes that inspire substance use.

Directions for further work
In this chapter, both specific strategies for controlling emotion and general attributes of emotion
regulation have been discussed and the available evidence about emotion regulation and sub-
stance use disorder has been reviewed. Although there is a sizable body of evidence on emotion
regulation from laboratory studies, unsettled questions remain about its applicability to clinical
and community samples. Progress has been made in research on substance use among adolescents
4
2

224 Emotion Regulation and Substance Use Disorders in Adolescents

but a strong understanding of exactly how emotion regulation contributes to the development of
substance use disorder has not yet been fully delineated. Further, while emotion regulation has
often been considered as a single dimension, a body of evidence indicates that emotional self-╉
control and emotional dysregulation are distinct constructs, not opposite ends of a single dimen-
sion. The following section discusses major themes in the chapter and their clinical implications.

Emotion regulation and its relation to risk


This chapter has highlighted the considerable evidence that demonstrates a connection between
emotion dysregulation and substance use disorder in adolescence and young adulthood. Findings
on prospective associations between early temperament characteristics and later risk for substance
use also suggest that emotional dysregulation may be present well before the onset of disorder,
thereby making it a true predisposing factor (Wills et al., 2000, 2016). The new research discussed
here has been consistent with the proposition that substance use disorder is one manifestation of
behavioral dysregulation and that dysregulation in emotion, behavior, and cognition are interre-
lated processes. Furthermore, this chapter has focused on constructs such as negative life events,
social functioning, and academic engagement as intermediate factors, such that these environ-
mental consequences of dysregulation partly mediate the association between emotion regulation
and disorder (Wills et al., 2011; 2016).
The latest research has revealed complexities that were not evident in earlier studies. Several
aspects of emotion regulation and dysregulation have been delineated, including emotional
reactivity, affective variability, and low distress tolerance as key risk factors, while emotional
soothability, alternative reinforcers and positive affect may serve as protective factors (Leventhal
& Zvolensky, 2015; Simons et al., 2014). Emotional dysregulation may create dependent stress
(Liu & Alloy, 2010), creating reciprocal associations between emotional dysregulation and envi-
ronmental context. Aside from individual differences, the environmental context appears to
affect drug involvement, both as a source of drug-╉related opportunities and as a source of drug-╉
free reinforcement (Audrain-╉McGovern et al., 2010; Yurasek et al., 2015). In addition, cognitive
research has shown how excessive drug use may bias responses toward drug-╉related reinforcers,
interfering with the normative development of emotional regulation (Lisdahl, Gilbart, Wright, &
Shollenbarger, 2013; Wills et al., 2015).

Early experience is important


Evidence regarding the impact of early experiences continues to accumulate. It is clear from sev-
eral types of research that early adversity (e.g., family poverty or child maltreatment) is related
to increased risk for substance abuse in adolescence and adulthood (Andersen & Teicher, 2009,
Sinha, 2008; Swendsen & Le Moal, 2011). The impact of early adversity extends to psychological
disorders including PTSD, which itself has a strong co-╉occurrence with substance abuse (e.g.,
Gaher et al., 2013; Goldsmith et al., 2013). Recent research is beginning to clarify how the disrup-
tion of behavioral and emotional regulation may contribute to this effect (Hahn et al., 2015; Oshri
et al., 2015; Shin et al., 2012).
What is less clear is what aspects of emotion dysregulation are modifiable factors and how they
should be targeted to reduce substance use in at-risk adolescents. There is reason to believe that
family dynamics shape the development of self-╉regulation (Farley & Kim-╉Spoon, 2014; Siegel,
2013; Wills, Forbes, & Gibbons, 2014). However, there is also evidence that individual differ-
ences in self-╉control and dysregulation stem, in part, from genetic predisposition (Kreek et al.,
2005; Yacubian & Buechel, 2009; Yamagata et al., 2005). Yet, how genetic vulnerabilities interact
with early environmental context and parenting characteristics to shape self-╉regulation is not well
5
2

Summary 225

studied, and understanding biological variables as well as social learning factors will be impor-
tant for informing the design of future prevention and treatment efforts (Beauchaine, 2015). The
extent to which substance use disorder in adolescents can be reduced via training individuals to
have better emotion regulation skills is an open question, though there are promising results in
this regard (see for example: Conrod et al., 2013; Southam-╉Gerow, 2013; Wills et al., 2015).

Behavioral and emotional regulation


A variety of excellent measures have been developed for studying specific aspects of behavioral and
emotional regulation or dysregulation. However, empirical studies that include both types of mea-
sures have consistently shown a high correlation between the behavioral and emotional domains
(Wills et al., 2011; 2013, 2016). From a methodological standpoint, this is an important issue because
studies of emotion regulation typically have not controlled for behavioral regulation, hence they may
overestimate the effect sizes for emotional regulation variables (Wills et al., 2016). This still leaves
unanswered the question: Why are behavioral and emotional regulation so strongly correlated in the
first place? This issue may in part be a matter of definition (e.g., Aldao et al. (2010) recognized that
problem solving, a cognitive-╉behavioral process, has a significant role in emotion regulation) but it
also raises the question of why individuals who are behaviorally dysregulated (e.g., impatient, dis-
tractible, having short time horizons) also show substantial dysregulation of emotion. Such research
is needed to help inform the design and implementation of prevention programs.
From a conceptual standpoint, the behavioral-╉emotional correlation raises important questions
for the design of prevention and treatment programs, particularly in view of the fact that there
are several aspects of emotional self-╉control or dysregulation. For example, should a prevention
program focus more on improving emotional self-╉control, or on implementing cognitive and
behavioral strategies to mitigate the impact of a tendency for dysregulation. If aiming to enhance
overall emotion regulation, should one focus more on characteristics related to mean level of
affect (e.g., distress tolerance) or would it be better to focus on reducing variability in affect?
A dialogue on these kinds of questions would be most valuable for helping to link prevention
researchers with adolescent treatment specialists. Previous prevention programs have focused
largely on behavioral regulation and there is evidence of their effectiveness (Bukoski, 2015; Griffin
& Botvin, 2010). Given the close association between behavioral and emotional regulation and the
importance of related contextual factors (e.g., peer groups, dependent stress, academic engage-
ment) in promoting or reducing substance use, it is likely that prevention programs may be most
effective when they take a holistic approach that promotes self-╉regulation (e.g., incorporating
both behavioral and emotional regulation aspects) and provide opportunities for, and the skills to
access, the development of social environments conducive to positive outcomes (e.g., high degrees
of drug-╉free reinforcers, healthy peer norms).

Summary
Adaptive emotion regulation is essential for well-╉being and successful adaption in all aspects of
life. Hence, programs that successfully enhance emotional regulation in youth have the potential
to have far-╉reaching benefits in addition to reducing substance use behavior. For this reason, it
is an important target for intervention programs. The multifaceted nature of emotion regulation
provides both challenges and opportunities for the development of prevention programs. Thus,
future research is needed to identify specific components of emotion regulation that are the most
amenable to intervention and have the broadest impact on non-╉targeted emotion regulatory pro-
cesses. Phrased differently, the key question remains: How can research find the optimal interven-
tion target that has the largest cascading effects throughout the regulatory system?
6
2

226 Emotion Regulation and Substance Use Disorders in Adolescents

References
Aldao, A., Nolen-╉Hoeksema, S., & Schweizer, S. (2010). Emotion-╉regulation strategies across
pychopathology: A meta-╉analytic review. Clinical Psychology Review, 30(2), 217–╉237.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Washington, DC: American Psychiatric Association.
Andersen, S. L., & Teicher, M. H. (2009). Developmental stress exposure and subsequent risk for substance
use. Neuroscience & Biobehavioral Reviews, 33(4), 515–╉524.
Arens, A. M., Gaher, R. M., Simons, J. S., & Dvorak, R. D. (2014). Child maltreatment and deliberate
self-╉harm: A negative binomial hurdle model for explanatory constructs. Child Maltreatment, 19(3–╉4),
168–╉177.
Audrain-╉McGovern, J., Rodriguez, D., Leventhal, A. M., Cuevas, J., Rodgers, K., & Sass, J. (2012). Where
is the pleasure in that? Low hedonic capacity predicts smoking onset and escalation. Nicotine & Tobacco
Research, 14(10), 1187–╉1196.
Audrain-╉McGovern, J., Rodriguez, D., Rodgers, K., & Cuevas, J. (2010). Declining alternative reinforcers
link depression to young adult smoking. Addiction, 106(1), 178–╉187.
Audrain-╉McGovern, J., Rodriguez, D., Leventhal, A. M., Cuevas, J., Rodgers, K., & Sass, J. (2012).
Low hedonic capacity predicts smoking onset and escalation. Nicotine & Tobacco Research, 14(1),
1187–╉1196.
Baker, T. B., Piper, M. E., McCarthy, D. E., Majeskie, M. R., & Fiore, M. C. (2004). Addiction motivation
reformulated: an affective processing model of negative reinforcement. Psychological Review,
111(1), 33–╉51.
Beauchaine, T. P. (2015). Future directions in emotion dysregulation and youth psychopathology. Journal of
Clinical Child & Adolescent Psychology, 44(5), 875–╉896.
Beevers, C. G. (2005). Cognitive vulnerability to depression: A dual-╉process model. Clinical Psychology
Review, 25(7), 975–╉1002.
Black, D. S., Sussman, S., Johnson, C. A., & Milam, J. (2012). Testing the indirect effect of trait mindfulness
on adolescent cigarette smoking through negative affect and perceived stress mediators. Journal of
Substance Use, 17(5–╉6), 417–╉429.
Brewer, J. A., Elwafi, H. M., & Davis, J. H. (2013). Psychological models and neurobiological mechanisms
of mindfulness training as treatment for addictions. Psychology of Addictive Behaviors, 27(2), 366–╉379.
Brewer, J. A., Mallik, S., Babuscio, T. A., Nich, C., Johnson, H. E., Deleone, C. M., & Rounsaville, B. J.
(2011). Mindfulness training for smoking cessation: Results from a randomized controlled trial. Drug
and Alcohol Dependence, 119(1), 72–╉80.
Brown, R. A., Lejuez, C. W., Kahler, C. W., & Strong, D. R. (2002). Distress tolerance and duration of past
smoking cessation attempts. Journal of Abnormal Psychology, 111(1), 180–╉185.
Buckner, J. D., Keough, M. E., & Schmidt, N. B. (2007). Problematic alcohol and cannabis use among
young adults: The roles of depression and discomfort and distress tolerance. Addictive Behaviors, 32(9),
1957–╉1963.
Buckholdt, K. E., Parra, G. R., Anestis, M. D., Lavender, J. M., Jobe-╉Shields, L. E., Tull, M. T., & Gratz, K.
L. (2014). Emotion regulation difficulties and maladaptive behaviors: Examination of deliberate self-╉
harm, disordered eating, and substance misuse in two samples. Cognitive Therapy and Research, 39(2),
140–╉152.
Bukoski, W. (2015). A history of drug abuse prevention science. In L. A. Scheier (Ed.), Handbook of
adolescent drug use prevention (pp. 31–╉48). Washington, DC: American Psychological Association.
Caspi, A., Moffitt, T., Newman, D., & Silva, P. (1996). Behavioral observations at age 3 years predict adult
psychiatric disorders. Archives of General Psychiatry, 53(11), 1033–╉1039.
Center for Behavioral Health Statistics and Quality (2015). Behavioral health trends in the United
States: Results from the 2014 National Survey on Drug Use and Health. Rockville, MD: Substance Abuse
and Mental Health Services Administration.
7
2

Summary 227

Chandley, R. B., Luebbe, A. M., Messman-​Moore, T. L., & Ward, R. M. (2014). Anxiety sensitivity, coping
motives, emotion dysregulation, and alcohol-​related outcomes in college women: A moderated-​
mediation model. Journal of Studies on Alcohol and Drugs, 75(1), 83–​92.
Cheetham, A., Allen, N.B., Yücel, M., & Lubman, D.I. (2010). The role of affective dysregulation in drug
addiction. Clinical Psychology Review, 30(6), 621–​634.
Chung, T., & Martin, C. S. (2005). Adolescents’ interpretations of DSM-​IV alcohol dependence symptom
queries and implications for diagnostic validity. Drug and Alcohol Dependence, 80(2), 191–​200.
Colder, C. R., Campbell, R. T., Ruel, E., Richardson, J. L., & Flay, B. R. (2002). A finite mixture model of
growth trajectories of adolescent alcohol use: Predictors and consequences. Journal of Consulting and
Clinical Psychology, 70(4), 976–​985.
Colder, C. R., Chassin, L., Lee, M. R., & Villalta, I. K. (2010). Developmental perspectives on affect and
adolescent substance use. In J. D. Kassel (Ed.), Substance abuse and emotion (pp. 109–​135). Washington,
DC: American Psychological Association.
Compton, W. M., Dawson, D. A., Conway, K. P., Brodsky, M., & Grant, B. F. (2013). Transitions in illicit
drug use status over 3 years: a prospective analysis of a general population sample. American Journal of
Psychiatry, 170(8), 660–​670.
Conrod, P. J., Castellanos-​Ryan, N., & Mackie, C. (2011). Long-​term effects of a personality-​targeted
intervention to reduce alcohol use in adolescents. Journal of Consulting and Clinical Psychology, 79(3),
296–​306.
Conrod, P. J., Castellanos-​Ryan, N., & Strang, J. (2010). Brief, personality-​targeted coping skills
interventions and survival as a non-​drug user over a 2-​year period during adolescence. Archives of
General Psychiatry, 67(1), 85–​93.
Conrod, P. J., O’Leary-​Barrett, M., Newton, N., Topper, L., Castellanos-​Ryan, N., Mackie, C., & Girard, A.
(2013). Effectiveness of a selective, personality-​targeted prevention program for adolescent alcohol use
and misuse. JAMA Psychiatry, 70(3), 334–​342.
Daughters, S. B., Lejuez, C., Bornovalova, M. A., Kahler, C. W., Strong, D. R., & Brown, R. A. (2005).
Distress tolerance as a predictor of early treatment dropout in a residential substance abuse treatment
facility. Journal of Abnormal Psychology, 114(4), 729–​734.
Daughters, S. B., Reynolds, E. K., MacPherson, L., Kahler, C. W., Danielson, C. K., Zvolensky, M., &
Lejuez, C. W. (2009). Distress tolerance and early adolescent externalizing and internalizing symptoms.
Behavior Research and Therapy, 47(3), 198–​205.
Diamond, L., & Lee, K. (2011). Interventions shown to aid executive function development in children 4 to
12 years old. Science, 333(6045), 959–​964.
Dvorak, R. D., & Simons, J. S. (2008). Affective differences among daily tobacco users, occasional users,
and non-​users. Addictive Behaviors, 33(1), 211–​216.
Dvorak, R. D., Simons, J. S., & Wray, T. B. (2011). Alcohol problem severity: Associations with dual
systems of self-​control. Journal of Studies on Alcohol and Drugs, 72(4), 678–​684.
Dvorak, R. D., Sargent, E. M., Kilwein, T. M., Stevenson, B. L., Kuvaas, N. J., & Williams, T. J. (2014).
Alcohol use and alcohol-​related consequences: Associations with emotion regulation difficulties. The
American Journal of Drug and Alcohol Abuse, 40(2), 125–​130.
Ebner-​Priemer, U. W., Eid, M., Kleindienst, N., Stabenow, S., & Trull, T. J. (2009). Analytic strategies
for understanding affective (in)stability and other dynamic processes in psychopathology. Journal of
Abnormal Psychology, 118(1), 195–​202.
Elwafi, H. M., Witkiwitz, K., Mallik, S., Thornhill, T. A., & Brewer, J. A. (2013). Mindfulness training for
smoking cessation: Moderation of the relationship between craving and cigarette use. Drug and Alcohol
Dependence, 130(1), 222–​229.
Emery, N. N., Simons, J. S., Clarke, C. J., & Gaher, R. M. (2014). Emotion differentiation and alcohol-​
related problems: the mediating role of urgency. Addictive Behaviors, 39(10), 1459–​1463.
Fairbairn, C. E., & Sayette, M. A. (2013). The effect of alcohol on emotional inertia: A test of alcohol
myopia. Journal of Abnormal Psychology, 122(3), 770–​781.
8
2

228 Emotion Regulation and Substance Use Disorders in Adolescents

Farley, J. P., & Kim-​Spoon, J. (2014). The development of adolescent self-​regulation: Reviewing the role of
parent and peer relationships. Journal of Adolescence, 37(4), 433–​440.
Fox, H. C., Hong, K. A., & Sinha, R. (2008). Difficulties in emotion regulation and impulse control in
recently abstinent alcoholics compared with social drinkers. Addictive Behaviors, 33(2), 388–​394.
Franken, I. H., Rassin, E., & Muris, P. (2007). The assessment of anhedonia in clinical and non-​clinical
populations: Further validation of the Snaith-​Hamilton Pleasure Scale. Journal of Affective Disorders,
99(1), 83–​89.
Frieden, T. R., Jaffe, H. W., Cono, J, et al. (2014). Youth Risk Behavior Surveillance System: United States,
2013. Morbidity and Mortality Weekly Reports, 63(2). Centers for Disease Control and Prevention,
Department of Health and Human Services.
Gaher, R. M., Hofman, N. L., Simons, J. S., & Hunsaker, R. (2013). Emotion regulation deficits as mediators
between trauma exposure and borderline symptoms. Cognitive Therapy and Research, 37(3), 466–​475.
Gaher, R. M., Simons, J. S., Hahn, A. M., Hofman, N. L., Hansen, J., & Buchkoski, J. (2014). An experience
sampling study of PTSD and alcohol-​related problems. Psychology of Addictive Behaviors, 28(4),
1013–​1025.
Gerrard, M., Gibbons, F. X., Houlihan, A.E., Stock, M. L., & Pomery, E.A. (2008). A dual-​process
approach to health decision making. Developmental Review, 28(1), 29–​61.
Gibbons, F. X., Houlihan, A. E., & Gerrard, M. (2009). Reason and reaction: The utility of a dual-​focus
perspective on prevention of adolescent health risk behavior. British Journal of Health Psychology, 14(2),
231–​248.
Gilbert, K. E. (2012). The neglected role of positive emotion in adolescent psychopathology. Clinical
Psychology Review, 32(6), 467–​481.
Goelz, P. M., Audrain-​McGovern, J. E., Hitsman, B., Leone, F. T., Veluz-​Wilkins, A., Jepson, C., &
Schnoll, R. A. (2014). The association between changes in alternative reinforcers and short-​term
smoking cessation. Drug and Alcohol Dependence, 138(1), 67–​74.
Goldsmith, R. E., Chesney, S. A., Heath, N. M., & Barlow, M. R. (2013). Emotion regulation difficulties
mediate associations between betrayal trauma and symptoms of posttraumatic stress, depression, and
anxiety. Journal of Traumatic Stress, 26(3), 376–​384.
Gratz, K. L., & Roemer, L. (2004). Multidimensional assessment of emotional regulation and
dysregulation: Development and validation of the Difficulties in Emotion Regulation Scale. Journal of
Psychopathology and Behavioral Assessment, 26(1), 41–​54.
Griffin, K. W., & Botvin, G. J. (2010). Evidence-​based interventions for preventing substance use disorders
in adolescents. Child and Adolescent Psychiatric Clinics of North America, 19(3) 505–​526.
Hahn, A. M., Simons, R. M., & Simons, J. S. (2016). Childhood maltreatment and sexual risk taking: The
mediating role of alexithymia. Archives of Sexual Behavior, 45(1), 53–​62.
Harford, T. C., Grant, B. F., Yi, H. Y., & Chen, C. M. (2005). Patterns of DSM-​IV alcohol abuse and
dependence criteria among adolescents and adults in the 2001 National Household Survey on Drug
Abuse. Alcoholism, Clinical and Experimental Research, 29(5), 810–​828.
Harrison, P. A., Fulkerson, J. A., & Beebe, T. J. (1998). DSM-​IV substance use disorder criteria for
adolescents: A critical examination based on a statewide school survey. American Journal of Psychiatry,
155(4), 486–​492.
Hasin, D. S., Saha, T. D., Kerridge, B. T., …. & Grant, B. T. (2015). Prevalence of marijuana use disorders
in the United States, between 2001-​2002 and 2012-​2013. JAMA Psychiatry, 72(12), 1232–​1245.
Hellmuth J. C., Stappenbeck C. A., Hoerster K. D., & Jakupcak, M. (2012). Modeling PTSD symptom
clusters, alcohol misuse, anger, and depression as they relate to aggression and suicidality in returning
U.S. Veterans. Journal of Traumatic Stress, 25(5), 527–​534.
Hellmuth, J. C., Jaquier, V., Young-​Wolff, K., & Sullivan, T. P. (2013). Posttraumatic stress disorder
symptom clusters, alcohol misuse, and women’s use of intimate partner violence. Journal of Traumatic
Stress, 26(4), 451–​458.
9
2

Summary 229

Hoffman, W., Friese, M., & Strack, F. (2009). Impulse and self-​control from a dual-​systems perspective.
Perspectives on Psychological Science, 4(2), 162–​175.
Jahng, S., Solhan, M. B., Tomko, R. L., Wood, P. K., Piasecki, T. M., & Trull, T. J. (2011). Affect and alcohol
use: an ecological momentary assessment study of outpatients with borderline personality disorder.
Journal of Abnormal Psychology, 120(3), 572–​584.
Jochman, K. A., Fromme, K., & Scheier, P. L. (2010). Maturing out of substance use: The other side
of etiology. In L. A. Scheier (Ed.), Handbook of drug use etiology (pp. 565–​578). Washington, DC
US: American Psychological Association.
Johnston, L. D., O’Malley, P. M., Miech, R. A., Bachman, J.G., Schulenberg, J. E. (2015). Monitoring the
Future national survey results on drug use: 1975-​2014: Overview, key findings on adolescent drug use.
Ann Arbor: Institute for Social Research, The University of Michigan.
Kaminer, Y., & Winters, K. C. (2015). DSM-​5 criteria for youth substance use disorders: Lost in translation?
Journal of the American Academy of Child and Adolescent Psychiatry, 54(5), 350–​351.
Kassel, J. D., Hussong, A. M., Wardle, M. C., Veilleux, J. C., Heinz, A., Greenstein, J. E., & Evatt, D. P.
(2010). Affective influences in drug use etiology. In L. Scheier (Ed.), Handbook of drug use etiology (pp.
183–​205). Washington, DC US: American Psychological Association.
Kassel, J. D., & Veilleux, J. C. (2010). Introduction: The complex interplay between substance abuse and
emotion. In J. D. Kassel (Ed.), Substance abuse and emotion (pp. 3–​12). Washington, DC: American
Psychological Association.
Khantzian, E. J. (1990). Self-​regulation and self-​medication factors in alcoholism and the addictions. In M.
Galanter (Ed.), Recent developments in alcoholism (Vol. 8, pp. 255–​271). New York: Plenum Press.
Kilpatrick, D. G., Acierno, R., Saunders, B., Resnick, H. S., Best, C. L., & Schnurr, P. P. (2000). Risk factors
for adolescent substance abuse and dependence: Data from a national sample. Journal of Consulting and
Clinical Psychology, 68(1), 19–​30.
Kirisci, L., Tarter, R., Ridenour, T., Reynolds, M., Horner, M., & Vanyukov, M. (2015). Externalizing
behavior and emotion dysregulation are indicators of transmissible risk for substance use disorder.
Addictive Behaviors, 42, 57–​62.
Kober, H. (2014). Emotion regulation in substance use disorders. In: J. J. Gross (Ed.), Handbook of emotion
regulation (2nd ed., pp. 428–​446). New York: Guilford.
Koob, G. F., & Le Moal, M. (2008). Addiction and the brain anti-​reward system. Annual Review of
Psychology, 59, 29–​53.
Koval, P., Kuppens, P., Allen, N. B., & Sheeber, L. (2012). Getting stuck in depression: the roles of
rumination and emotional inertia. Cognition and Emotion, 26(8), 1412–​1427.
Kreek, M. J., Nielsen, D. A., Butelman, E. R., & LaForge, K. S. (2005). Genetic influences on impulsivity,
risk taking, stress responsivity and vulnerability to drug abuse and addiction. Nature Neuroscience,
8(11), 1450–​1457.
Kuppens, P., Sheeber, L. B., Yap, M. B. H., Whittle, S., Simmons, J. G., & Allen, N. B. (2012). Emotional
inertia prospectively predicts the onset of depressive disorder in adolescence. Emotion, 12(2), 283–​289.
Leventhal, A. M., Waters, A. J., Kahler, C. W., Ray, L. A., & Sussman, S. (2009). Relations between
anhedonia and smoking motivation. Nicotine & Tobacco Research, 11(9), 1047–​1054.
Leventhal, A. M., & Zvolensky, M. J. (2015). Anxiety, depression, and cigarette smoking: A transdiagostic
vulnerability framework to understanding emotion-​smoking comorbidity. Psychological Bulletin,
141(1), 176–​212.
Leyro, T. M., Bernstein, A., Vujanovic, A. A., McLeish, A. C., & Zvolensky, M. J. (2011). Distress Tolerance
Scale: A confirmatory factor analysis among daily cigarette smokers. Journal of Psychopathology and
Behavioral Assessment, 33(1), 47–​57.
Lisdahl, K. M., Gilbart, A. R., Wright, N. E., & Shollenbarger, S. (2013). Dare to delay? The impact
of adolescent alcohol and marijuana use on cognition, brain structure, and function. Frontiers in
Psychiatry, 4, 53.
0
3
2

230 Emotion Regulation and Substance Use Disorders in Adolescents

Liu., R. T., & Alloy, L. B. (2010). Stress generation in depression: A review of empirical literature and
recommendations for further study. Clinical Psychology Review, 30(5), 582–​593.
Marshall-​Berenz, E. C., Vujanovic, A. A., & MacPherson, L. (2011). Impulsivity and alcohol use coping
motives in a trauma-​exposed sample: The mediating role of distress tolerance. Personality and
Individual Differences, 50(5), 588–​592.
McRae, K., Misra, S., Prasad, A. K., Pereira, S. C., & Gross, J. J. (2012). Bottom-​up and top-​down emotion
generation: implications for emotion regulation. Social Cognitive and Affective Neuroscience, 7(3), 253–​262.
Merikangas, K. R., He, J., Burstein, M., Swanson, S. A., Avenevoli, S., Cui, L., & Swendsen, J. (2010).
Lifetime prevalence of mental disorders in U.S. adolescents: Results from the National Comorbidity
Survey Replication: Adolescent Supplement (NCS-​A). Journal of the American Academy of Child &
Adolescent Psychiatry, 49(10), 980–​989.
Messman-​Moore, T. L., & Ward, R. M. (2014). Emotion dysregulation and coping drinking motives in
college women. American Journal of Health Behavior, 38(4), 553–​559.
Mohr, C. D., Brannan, D., Mohr, J., Armeli, S., & Tennen, H. (2008). Evidence for positive mood buffering
among college student drinkers. Personality and Social Psychology Bulletin, 34(9), 1249–​1259.
Mohr, C., Armeli, S., Tennen, H., & Todd, M. (2010). The complexities of modeling mood-​drinking
relationships: Lessons learned from daily process research. In J. D. Kassel (Ed.), Substance abuse and
emotion (pp. 189–​216). Washington, DC: APA.
Mohr, C. D., Arpin, S., & McCabe, C. T. (2015). Daily affect variability and context-​specific alcohol
consumption. Drug and Alcohol Review, 34(6), 581–​587.
Moses, E. B., & Barlow, D. H. (2006). A unified treatment approach for emotional disorders based on
emotion science. Current Directions in Psychological Science, 15(3), 146–​150.
Noelen-​Hoeksema, S., Stice, E., Wade, E., & Bohon, C. (2007). Reciprocal relations between rumination
and bulimic, substance abuse, and depressive symptoms in female adolescents. Journal of Abnormal
Psychology, 116(1), 198–​207.
O’Leary-​Barrett, M., Topper, L., Khudhairy, N. Al-​, Pihl, R. O., Castellanos-​Ryan, N., Mackie, C. J., &
Conrod, P. J. (2013). Two-​year impact of personality-​targeted, teacher-​delivered interventions on youth
internalizing and externalizing problems. Journal of the American Academy of Child and Adolescent
Psychiatry, 52(9), 911–​920.
Oliver, M. N. I., & Simons, J. S. (2004). The Affective Lability Scales: Development of a short-​form measure.
Personality and Individual Differences, 37(6), 1279–​1288.
Oshri, A., Sutton, T. E., Clay-​Warner, J., & Miller, J. D. (2015). Child maltreatment types and risk
behaviors: Associations with attachment style and emotion regulation dimensions. Personality and
Individual Differences, 73, 127–​133.
Pearson, M. R., Kite, B. A., & Henson, J. M. (2013). Predictive effects of good self-​control and poor
regulation on alcohol-​related outcomes: Do protective behavioral strategies mediate? Psychology of
Addictive Behaviors, 27(1), 81–​89.
Reich, R. R., Cummings, J. R., Greenbaum, P. E., Moltisanti, A. J., & Goldman, M. S. (2015). The temporal
“pulse” of drinking: Tracking 5 years of binge drinking in emerging adults. Journal of Abnormal
Psychology, 124(3), 635–​647.
Rosenkranz, S. E., Muller, R. T., & Henderson, J. L. (2014). The role of complex PTSD in mediating
childhood maltreatment and substance abuse severity among youth seeking substance abuse treatment.
Psychological Trauma: Theory, Research, Practice, and Policy, 6(1), 25–​33.
Rothbart, M. K., Ahadi, S. A., & Evans, D. E. (2000). Temperament and personality: Origins and outcomes.
Journal of Personality and Social Psychology, 78(1), 122–​135.
Rothman, A., Sheeran, P., & Wood, W. (2009). Reflective and automatic processes in the initiation and
maintenance of diet change. Annals of Behavioral Medicine, 38(Suppl. 1), S4–​S17.
Sheppes, G., Scheibe, S., Suri, G., Radu, P., Blechert, J., & Gross, J. J. (2014). Emotion regulation
choice: A conceptual framework and supporting evidence. Journal of Experimental Psychology: General,
143(1), 163–​181.
1
3
2

Summary 231

Sher, K. J., & Grekin, E. R. (2007). Alcohol and affect regulation. In J. J. Gross (Ed.), Handbook of emotion
regulation (pp. 560–​580). New York: Guilford.
Shin, S., Hong, H. G., & Wills, T. A. (2012). Examination of pathways from childhood maltreatment to
adolescent binge drinking. American Journal on Addictions, 21(3), 202–​209.
Shishido, H., Gaher, R. M., & Simons, J. S. (2013). I don’t know how I feel, therefore I act: alexithymia,
urgency, and alcohol problems. Addictive Behaviors, 38(4), 2014–​2017.
Shrier, L. A., Ross, C. S., & Blood, E. A. (2014). Momentary positive and negative affect preceding
marijuana use in youth. Journal of Studies on Alcohol and Drugs, 75(5), 781–​789.
Siegel, J. P. (2010). Stop overreacting: Effective strategies for calming your emotions. Oakland, CA: New
Harbinger Publications.
Siegel, J. P. (2013). Breaking the links in intergenerational violence: An emotional regulation perspective.
Family Process, 52(2), 163–​178.
Siegel, J. P. (2015). Emotional regulation in adolescent substance use disorders: Rethinking risk. Journal of
Child and Adolescent Substance Abuse, 24(2), 67–​79.
Simons, J. S., & Carey, K. B. (2002). Risk and vulnerability for marijuana use problems: The role of affect
dysregulation. Psychology of Addictive Behaviors, 16(1), 72–​75.
Simons, J. S., & Carey, K. B. (2006). An affective and cognitive model of marijuana and alcohol problems.
Addictive Behaviors, 31(9), 1578–​1592.
Simons, J. S., Carey, K. B., & Wills, T. A. (2009). Alcohol abuse and dependence symptoms: A multidimensional
model of common and specific etiology. Psychology of Addictive Behaviors, 23(3), 415–​427.
Simons, J. S., Dvorak, R. D., Batien, B. D., & Wray, T. B. (2010). Event-​level associations between affect,
alcohol intoxication, and acute dependence symptoms: Effects of urgency, self-​control, and drinking
experience. Addictive Behaviors 35(12), 1045–​1053.
Simons, J. S., & Gaher, R. M. (2005). The Distress Tolerance Scale: Development and validation of a self-​
report measure. Motivation and Emotion, 29(2), 83–​102.
Simons, J. S., Wills, T. A., & Neal, D. J. (2014). The many faces of affect: A multilevel model of drinking
frequency/​quantity and alcohol dependence symptoms among young adults. Journal of Abnormal
Psychology, 123(3), 676–​694.
Simons, J. S., Gaher, R. M., Oliver, M. N. I., Bush, J. A., & Palmer, M. A. (2005a). An experience sampling
study of associations between affect and alcohol use and problems among college students. Journal of
Studies on Alcohol, 66(4), 459–​469.
Simons, J. S., Oliver, M. N. I., Gaher, R. M., Ebel, G., & Brummels, P. (2005b). Methamphetamine and
alcohol abuse and dependence symptoms: Associations with affect lability and impulsivity in a rural
treatment population. Addictive Behaviors, 30(7), 1370–​1381.
Sinha, R. (2008). Chronic stress, drug use, and vulnerability to addiction. Annals of the New York Academy
of Sciences, 1141(1), 105–​130.
Skara, S., & Sussman, S. (2003). A review of 25 long-​term adolescent tobacco and other drug use
prevention program evaluations. Preventive Medicine, 37(5), 454–​474.
Southam-​Gerow, M. A., & Kendall, P. C. (2002). Emotional regulation: Implications for child
psychopathology and therapy. Clinical Psychology Review, 22(2), 189–​222.
Southam-​Gerow, M. A. (2013). Emotion regulation in children and adolescents: A practitioner’s guide.
New York: Guilford.
Spear, L. (2010). The behavioral neuroscience of adolescence. New York: W.W. Norton.
Steele, C. M., & Josephs, R. A. (1988). Drinking your troubles away: An attention-​allocation model of
alcohol’s effect on psychological stress. Journal of Abnormal Psychology, 97(2), 196–​205.
Steinberg, L. (2008). A social neuroscience perspective on adolescent risk taking. Developmental Review,
28(1), 78–​106.
Sullivan, T. P., Ashare, R. L., Jaquier, V., & Tennen, H. (2012). Risk factors for alcohol problems among
victims of partner violence. Substance Use & Misuse, 47(6), 673–​685.
2
3

232 Emotion Regulation and Substance Use Disorders in Adolescents

Swendsen, J., Burstein, M., Case, B., Conway, K. P., Dierker, L., He, J., & Merikangas, K. R. (2012). Use
and abuse of alcohol and illicit drugs in US adolescents: Results of the National Comorbidity Survey.
Journal of the American Medical Association, 69(4), 390–​398.
Swendsen, J., & Le Moal, M. (2011). Individual vulnerability to addiction. Annals of the New York Academy
of Sciences, 1216(1), 73–​85.
Tang, Y.-​Y., Yang, L., Leve, L., & Harold, G. T. (2012). Improving executive function and its neurobiological
mechanisms through a mindfulness-​based intervention: Advances within the field of developmental
neuroscience. Child Development Perspectives, 6(4), 361–​366.
Tarantino, N., Lamis, D. A., Ballard, E. D., Masuda, A., & Dvorak, R. D. (2015). Parent-​child conflict and
drug use in college women: A moderated mediation model of self-​control and mindfulness. Journal of
Counseling Psychology, 62(2), 303–​313.
Tarter, R., Vanyukov, M., Giancola, P., Dawes, M., Blackson, T., Mezzich, A., & Clark, D. (1999).
Etiology of early age onset substance use disorder: A maturational perspective. Development and
Psychopathology, 11(4), 657–​683.
Veilleux, J. C., Skinner, K. D., Reese, E. D., & Shaver, J. A. (2014). Negative affect intensity influences
drinking to cope through facets of emotion dysregulation. Personality and Individual Differences, 59,
96–​101.
Verges, A., Haeny, A. M., Jackson, K. M., Bucholz, K. K., Grant, J. D., Trull, T. J., … Sher, K. J. (2013).
Refining the notion of maturing out: Results from the National Epidemiologic Survey on Alcohol and
Related Conditions. American Journal of Public Health, 103, e67–​e73.
Volkow, N. D., & Baler, R. D. (2012). To stop or not to stop? Science, 335(6068), 546–​548.
Vujanovic, A. A., Marshall-​Berenz, E. C., & Zvolensky, M. J. (2011). Posttraumatic stress and alcohol
use motives: A test of the incremental and mediating role of distress tolerance. Journal of Cognitive
Psychotherapy, 25(2), 130–​141.
Webb, T. L., Miles, E., & Sheeran, P. (2012). A meta-​analysis of strategies derived from the process model of
emotion regulation. Psychological Bulletin, 138(4), 775–​808.
Weinberg, A., & Klonsky, E. D. (2009). Measurement of emotion dysregulation in adolescents. Psychological
Assessment, 21(4), 616–​621.
Weinstein, S. M., & Mermelstein, M. (2013a). Associations of negative mood and smoking in the
development of smoking in adolescence. Journal of Clinical Child and Adolescent Psychology, 42(5),
629–​642.
Weinstein, S. M., & Mermelstein, R. J. (2013b). Influences of mood variability and negative moods on
adolescent smoking. Psychology of Addictive Behaviors, 27(4), 1068–​1078.
Weiss, N. H., Tull, M. T., Anestis, M. D., & Gratz, K. L. (2013). The relative and unique contributions of
emotion dysregulation and impulsivity to posttraumatic stress disorder among substance dependent
inpatients. Drug and Alcohol Dependence, 128(1), 45–​51.
Weiss, N. H., Tull, M. T., Davis, L. T., Dehon, E. E., Fulton, J. J., & Gratz, K. L. (2012). Examining the
association between emotion regulation difficulties and probable posttraumatic stress disorder within a
sample of African Americans. Cognitive Behaviour Therapy, 41(1), 5–​14.
White, H. R., Pandina, R. J., & Chen, P-​H. (2002). Developmental trajectories of cigarette use from
adolescence into adulthood. Drug and Alcohol Dependence, 65(2), 167–​178.
Wiers, R. W., Bartholow, B. D., van den Wildenberg, E., Thush, C., Engels, R. C. M. E., Sher, K. J., …
Stacy, A. W. (2007). Automatic and controlled processes and the development of addictive behaviors in
adolescents. Pharmacology, Biochemistry and Behavior, 86(2), 263–​283.
Wills, T. A., & Ainette, M. G. (2010). Temperament, self-​control, and adolescent substance use: A two-​
factor model of etiological processes. In L. M. Scheier (Ed.), Handbook of drug use etiology (pp. 127–​
146). Washington, DC: American Psychological Association.
32

Summary 233

Wills, T. A., Bantum, E. O., Pokhrel, P., Maddock, J. E., Ainette, M. G., Morehouse, E., & Fenster, B.
(2013). A dual-​process model of early substance use: Tests in two diverse populations of adolescents.
Health Psychology, 32(5), 533–​542.
Wills, T. A., Cleary, S. D., Filer, M., Shinar, O., Mariani, J., & Spera, K. (2001). Temperament and self-​
control related to early-​onset substance use: Test of a developmental model. Prevention Science, 2(3),
145–​163.
Wills, T. A., & Dishion, T. J. (2004). Temperament and adolescent substance use: A transactional analysis of
emerging self-​control. Journal of Clinical Child and Adolescent Psychology, 33(1), 69–​81.
Wills, T. A., Forbes, M., & Gibbons, F. X. (2014). Parental and peer support: An analysis of their relations
to adolescent substance use. In L. M. Scheier & W. B. Hansen (Eds.), Parenting and teen drug use (pp.
148–​165). New York: Oxford.
Wills, T. A., Pokhrel, P., Morehouse, E., & Fenster, B. (2011). Behavioral and emotional regulation and
adolescent substance use problems: A test of moderation effects in a dual-​process model. Psychology of
Addictive Behaviors, 25(2), 279–​292.
Wills, T. A., Sandy, J. M., & Yaeger, A. (2000). Temperament and adolescent substance use: An epigenetic
approach to risk and protection. Journal of Personality, 68(6), 1127–​1152.
Wills, T. A., Sandy, J. M., & Yaeger, A. (2002). Moderators of the relation between substance use level
and problems: Test of a self-​regulation model in middle adolescence. Journal of Abnormal Psychology,
111(1), 3–​21.
Wills, T. A., Sandy, J. M., Shinar, O., & Yaeger, A. (1999). Contributions of positive and negative affect to
adolescent substance use: Test of a bidimensional model in a longitudinal study. Psychology of Addictive
Behaviors, 13(4), 327–​338.
Wills, T. A., Simons, J. S., Sussman, S., & Knight, R. A. (2016). Emotional self-​control and
dysregulation: A dual-​process analysis of pathways to externalizing and internalizing symptomatology
and positive well-​being in younger adolescents. Drug and Alcohol Dependence, 163, S37–​S45.
Wills, T. A., Simons, J. S., & Gibbons, F. X. (2015). Self-​control and substance use prevention. In L.A.
Scheier (Ed.), Handbook of adolescent drug use prevention (pp. 121–​139). Washington, DC: American
Psychological Association.
Wills, T. A., Walker, C., Mendoza, D., & Ainette, M. G. (2006). Behavioral and emotional self-​
control: relations to substance use in samples of middle and high school students. Psychology of
Addictive Behaviors, 20(3), 265–​278.
Wills, T. A., Walker, C., & Resko, J. A. (2005). Longitudinal studies of drug use and abuse. In Z. Sloboda
(Ed.), Epidemiology of drug abuse (pp. 177–​192). New York: Springer.
Windle, M., & Wiesner, M. (2004). Trajectories of marijuana use from adolescence to adulthood: Predictors
and outcomes. Development and Psychopathology, 16(4), 1007–​1027.
Winters, K. C. (2013). Advances in the science of adolescent drug involvement: implications for assessment
and diagnosis. Current Opinion in Psychiatry, 26(4), 318–​324.
Winward, J. L., Bekman, N. M., Hanson, K. L., Lejuez, C. W., & Brown, S. A. (2014). Changes in emotional
reactivity and distress tolerance and heavy-​drinking adolescents during sustained abstinence.
Alcoholism: Clinical and Experimental Research, 38(6), 1761–​1769.
Wong, C. F., Silva, K., Kecojevic, A., Schrager, S. M., Bloom, J. J., Iverson, E., & Lankenau, S. E. (2013).
Emotion regulation profiles predict nonmedical prescription drug and illicit drug use among high-​risk
young adults. Drug and Alcohol Dependence, 132(1), 165–​171.
Wray, T. B., Simons, J. S., Dvorak, R. D., & Gaher, R. M. (2012). Trait-​based affective processes in alcohol-​
involved risk behaviors. Addictive Behaviors, 37(11), 1230–​1239.
Yacubian, J., & Buchel, C. (2009). The genetic basis of individual differences in reward processing and the
link to addictive behavior and social cognition. Neuroscience, 164(1), 55–​71.
4
3
2

234 Emotion Regulation and Substance Use Disorders in Adolescents

Yamagata, S., Takahasi, Y., Kijima, N., Maekawa, H., Ono, Y., & Ando, J. (2005). Genetic/​environmental
etiology of effortful control. Twin Research and Human Genetics, 8(4), 300–​306.
Yurasek, A. M., Dennhardt, A. A., & Murphy, J. G. (2015). A randomized controlled trial of a behavioral
economic intervention for alcohol and marijuana use. Experimental and Clinical Psychopharmacology,
23(5), 332–​338.
Zvolensky, M. J., Marshall, E. C., Johnson, K., Hogan, J., Bernstein, A., & Bonn-​Miller, M. O. (2009).
Relations between anxiety sensitivity, distress tolerance, and fear reactivity to bodily sensations to
coping and conformity marijuana use motives among young adult marijuana users. Experimental and
Clinical Psychopharmacology, 17(1), 31–​42.
5
3
2

Chapter 12

Emotion Regulation in Autism


Spectrum Disorder
Jonathan A. Weiss, Priscilla Burnham Riosa,
Carla A. Mazefsky, & Renae Beaumont

Autism spectrum disorder


Autism spectrum disorder (ASD) is a pervasive neurodevelopmental disorder characterized by
deficits in social communication, along with restricted, repetitive patterns of behavior, interests,
or activities (American Psychiatric Association [APA], 2013). These symptoms must be present in
the early developmental period, though may not become impairing until later in life. The changing
nosology from DSM-╉IV to DSM-╉5 involved a considerable shift in the amalgamation of Pervasive
Developmental Disorder subtypes into one diagnosis of ASD (Lord & Bishop, 2015). The focus on
autism as a “spectrum” largely reflects the lack of reliability found in distinguishing past subtypes
(e.g., Asperger’s disorder vs. PDD-╉NOS vs. Autistic disorder with no intellectual impairment)
(Bennett et al., 2008; Lord et al., 2012; Woodbury-╉Smith et al., 2005). As well, DSM-╉IV subtypes
were largely being used as proxies for patient overall severity of impairment in multisite trials,
rather than consistently taking specific diagnostic subtype criteria into consideration (Lord et al.
2012). Long before the publication of the DSM-╉5, authors were observing a preference in using
the singular term “Autism Spectrum Disorder” in the scientific literature over the reporting of
DSM-╉IV subtypes (Bebko et al., 2008).
The notion of a spectrum is an important consideration in understanding the expression of
ASD and of emotion dysregulation in this population. The primary conceptualization of a spec-
trum relates to the overall severity of ASD symptomatology and the level of impairment in these
domains, as evidenced in the three severity level specifiers proposed in the DSM-╉5 (APA, 2013).
Level Three (“requiring very substantial support”) is reserved for severe deficits in social commu-
nication, along with extreme restricted/╉repetitive behaviors. Level Two is less severe (“requiring
substantial support”), with marked deficits in social communication and frequent difficulties with
restricted, repetitive behaviors. Level One is meant to reflect the least severe degree of impairment
(“requiring support”), where social communication difficulties are still apparent when there is
a lack of support, and when there are difficulties with behaviors that are not considered as fre-
quent or as severe as in Level Two. There are considerable reservations in approaching severity in
such a discrete ordinal manner, given that severity categorizations across adaptive, cognitive, and
autism symptomatology have been found to be discrepant at mild or moderate levels (Weitlauf
et al., 2014), leaving authors to caution for the need for greater research in categorical specifier
use (Lord & Bishop, 2015). It is clear though that the expression of ASD symptom severity ranges
from mild to severe across the population, and to some extent has variability within the individual
over time and across contexts.
The spectrum also applies to considerations of intellectual functioning and related adap-
tive behavior. Intellectual disability (ID) is characterized by significant limitations in cognitive
6
3
2

236 Emotion Regulation and Autism Spectrum Disorder

functioning (approximately two standard deviations below what would be expected based on
someone of the same chronological age), as well as commensurate deficits in adaptive behavior
across social, conceptual, and practical life areas. These impairments must have occurred prior
to age 18 and be observed within the appropriate developmental period. Over time, prevalence
estimates of ID in the ASD population have decreased, the result of greater numbers of youth
being diagnosed with ASD without ID as well as potential diagnostic substitution. In the 2006
surveillance year, the Centers for Disease Control and Prevention (CDC, 2009) reported that 41%
of eight-​year-​olds with ASD had intellectual functioning in the ID range (ranging from 29% to
51% across sites). This rate decreased to 31% in the 2010 surveillance year (ranging from 18% to
37% across sites; CDC, 2014), though this latest surveillance suggests that an additional 24% of
youth with ASD had low levels of cognitive functioning but not in the range of ID (with 23% in the
borderline ID range). The spectrum can also be applied to one’s level of adaptive functioning—​the
ability to function independently in age appropriate domains of every-​day life, including social
(e.g., forming and maintaining interpersonal relationships, play and leisure skills), practical (e.g.,
dressing, grooming, toileting), and conceptual spheres (e.g., reading, writing, understanding time
and money).
There has been an apparent rise in prevalence rates in youth diagnosed with ASD over the last
decade, with ASD affecting between one in 110 in 2006, to one in 68 in 2010 (CDC, 2007; 2014).
Concerns over the methodology and interpretation of the latest rates have been noted, including
the substantial cross-​site variability (Mandell & Lecavalier, 2014). Lord and Bishop (2015) suggest
a number of reasons for the increase in prevalence. First, given that the majority of the increase
comprises youth without ID, higher rates could reflect the broader criteria now employed in the
diagnosis of ASD (King & Bearman, 2009). Second, higher rates could reflect the greater availabil-
ity of training and tools at clinicians’ disposal in the assessment of ASD. Third, many of these tools
have low specificity in distinguishing ASD from other psychiatric or genetic disorders (Charman
et al., 2007, DiGuiseppi et al., 2010, Hus et al., 2013), which could lead to children being misdiag-
nosed with ASD. At the same time, Lord and Bishop (2015) suggest that prevalence rates of ASD
may still continue to rise, as future work addresses the possible under-​representation of females
with ASD, the increasing awareness that can lead to more referrals, and the known disparities in
rates across demographic characteristics (e.g., race, neighborhood).
Emotional and behavioral problems, though not diagnostic, are also often associated charac-
teristics in the clinical presentation of ASD. There is considerable data indicating that individuals
with ASD are more likely to have clinically significant levels of these problems compared to both
typically developing individuals of the same age, and individuals of the same cognitive ability
(Leyfer et al., 2006; Simonoff et al., 2008; Totsika et al., 2011a; Totsika et al., 2011b). Using second-
ary data analysis on a UK national survey (Millennium Cohort Study), Totsika and colleagues
(2011a) examined rates of borderline/​clinically significant mental health problems in 14,807 typi-
cally developing five-​year-​olds, compared to 82 children with only ASD, 432 children with only
ID, and 32 children with both ASD and ID, using a parent report questionnaire. Children with
developmental disabilities, across diagnoses, were more likely to have significant hyperactivity,
conduct problems, and emotional problems compared to the typically developing comparison
group. In the ASD samples, 59% of five-​year-​olds with no ID had hyperactivity problems, 46% had
conduct problems, and 38% had emotional problems. Higher rates were reported with regard to
hyperactivity (88%) and similar rates for conduct problems (57%) and emotional problems (39%)
in the youth with both ASD and ID. In another large scale UK population study, Totsika and col-
leagues (2011b) compared mental health problems among 17,727 typically developing youth, to
those of 47 youth with only ASD, 590 youth with only ID, and 51 youth with both ASD and ID,
again using a parent report questionnaire about symptoms in the last six months. Compared to
the expected 20% of the typically developing comparison group, youth with only ID had a two-​to
7
3
2

Emotion regulation and autism spectrum disorder 237

three-╉fold increase in rates of emotional and behavior problems, and youth with ASD had a three-╉
to four-╉fold increase. Approximately 70% of youth with ASD had clinically significant emotional
problems (compared to 42% of youth with only ID without ASD), and 65% had clinically signifi-
cant conduct problems (compared to 46% of youth with only ID without ASD).
Given the high rates observed in community samples, it is not surprising that many individu-
als with ASD experience multiple co-╉occurring mental health problems. In fact, 70% of youth
with ASD meet criteria for at least one additional psychiatric disorder, and 40% for two or more,
when using an adaptive structured interview to distinguish ASD symptoms from other psychi-
atric symptoms (Leyfer et al., 2006; Simonoff et al., 2008). Given the overlapping manifestation
among behaviorally-╉based disorders, there is some concern that discrete psychiatric diagnoses
could be overestimated, though even after focused examinations are taken into account, ASD
symptom severity and psychiatric rates remain much higher than expected in the general popula-
tion (Mazefsky et al., 2012). In community and clinically referred samples, levels of internalizing
symptoms (i.e., depression, anxiety) are often moderately correlated with the expression of exter-
nalizing symptoms (i.e., noncompliance, aggressive behavior, irritability; Gadow et al., 2006; Kim
et al., 2000; Lecavalier, Gadow, DeVincent, & Edwards, 2009; Weisbrot et al., 2005), leaving many
authors to suggest they may have an underlying basis in emotion dysregulation (Mazefsky et al.,
2013; Weiss, 2014).
Thompson (1994) defines emotion regulation as “the extrinsic and intrinsic processes respon-
sible for monitoring, evaluating, and modifying emotional reactions, especially their intensive
and temporal features, to accomplish one’s goals” (pp. 27–╉28). This conceptualization is useful as
it emphasizes the multiple strategies that may result in emotion regulation, and the notion that
an emotion regulation process can be either adaptive or maladaptive, depending on whether it is
successful in achieving the appropriate affective state and does not have negative long-╉term costs
(Campbell-╉Sills & Barlow, 2007). Gross and Thompson (2007) outline five broad classes of emo-
tion regulation strategies:  Situation Selection, Situation Modification, Attentional Deployment,
Cognitive Control, and Response Modulation. The first four classes are considered “antecedent”
to emotional disruption. If we are able to successfully influence the situations that may cause dis-
tress, focusing on the most useful stimuli and having a good understanding of emotions, or alter-
ing how we think about our experiences, then we are able to maintain emotion regulation without
our affect ever becoming disrupted in the first place. However, when those strategies fail, and we
end up with high levels of negative affect (i.e., dysregulation), then we are required to employ the
last class, modifying our experience to regulate. Adaptive emotion regulation strategies are typi-
cally voluntary in nature, being consciously initiated with effort and control, whereas maladap-
tive strategies are often involuntary, being applied automatically in response to emotion-╉eliciting
stimuli (Aldao & Nolen-╉Hoeksema, 2010; Mazefsky, Borue, Day & Minshew, 2014).

Emotion regulation and autism spectrum disorder


Emotion regulation difficulties are common in youth and adults with ASD, including the funda-
mental task of understanding emotion. Uljarevic and Hamilton (2013) conducted a meta-╉analysis
of 48 studies investigating differences in recognizing others’ basic emotions (fear, surprise, anger,
disgust, happiness, and surprise) between children and adults with ASD and typically develop-
ing peers. The majority of studies have indicated that individuals with ASD perform significantly
worse than their typically developing counterparts on emotion recognition tasks, regardless of age
or IQ, though emotion recognition difficulties may not be as evident as previously thought. For
instance, Wright and colleagues (2008) found differences between youth with and without ASD
on recognizing anger and happiness but not sadness, fear, surprise, or disgust. Interestingly, diffi-
culty recognizing happiness was related to greater global emotion recognition difficulties. Beyond
8
3
2

238 Emotion Regulation and Autism Spectrum Disorder

recognition, children with ASD report more difficulties with insight into their own emotional
functioning and in interpreting their own emotional experiences, which is known as emotion
awareness (Losh & Capps, 2006; Rieffe, Terwogt, & Kotronopoulou, 2007); difficulties that also
extend to adults with ASD. Hill and colleagues (2004) surveyed 27 adults with ASD about three
aspects of cognitive processing of emotions: Identifying emotions (e.g., “When I am upset, I don’t
know if I am sad, frightened or angry.”), describing feelings (e.g., “I find it hard to describe how
I feel about people.”), and externally oriented thinking (e.g., “I find examination of my feelings
useful in solving personal problems.”), and found that compared to typically developing adults,
adults with ASD reported significantly greater difficulties with overall emotion processing.
Several studies have examined the specific strategies children and adults with ASD use to regu-
late their emotions, with considerable evidence demonstrating they are less effective than their
peers. Youth and young adults with ASD use adaptive strategies (e.g., problem solving, social
support, cognitive reappraisal, cognitive distraction, acceptance, exercise, and relaxation) less
frequently and maladaptive strategies (i.e., avoidance, expressive suppression) more frequently
compared to typically developing peers (Konstantareas & Stewart, 2006; Samson, et  al., 2015).
Self-╉reports of using more maladaptive coping strategies have also been found among children
and adolescents with ASD compared to peers, and are associated with higher self-╉reported lev-
els of psychopathology (Mazefsky et al., 2014; Pouw et al., 2013). Jahromi and colleagues (2012)
reported a higher ratio of maladaptive to adaptive strategies among children with ASD compared
to controls, when employing a frustration task and coding ensuing strategies. Emotion regulation
strategies used by children with and without ASD were coded as constructive/╉adaptive (e.g., help-╉
seeking) and maladaptive (e.g., venting, avoidance). Not only did children with ASD use more
maladaptive strategies across the frustration tasks, but they also more quickly resigned the task
compared to their typically developing peers, and even more so when their parents were not avail-
able to assist them, suggesting the potential role of parental co-╉regulation for children with ASD.

The interplay between emotion regulation and ASD


characteristics and impairment
The accumulating evidence indicative of ineffective or maladaptive emotion regulation in ASD
supports the need for a greater focus on emotion regulation as a core feature of the disorder.
Although, like ID, emotion dysregulation is not a universal characteristic in ASD, its presence has
important implications for the expression of ASD and functional outcomes. Adopting an emotion
regulation framework can provide a mechanistic explanation for the many associated emotional
and behavioral concerns in ASD, which may be more informative for developing interventions
than attributing these concerns solely to psychiatric comorbidity (Mazefsky et al., 2013; Weiss,
2014). Further, by focusing on emotion regulation as the core disrupted process, the likely under-
lying biological and cognitive factors contributing to poor emotional and behavioral functioning
become more readily identifiable.

Biological differences related to emotion regulation


Although there are few neurobiological studies of emotion regulation in ASD, neuroimaging
studies suggest atypical functioning and connectivity in regions that are involved in emotion reg-
ulation, such as the amygdala and prefrontal cortex (Mazefsky et al., 2013). In addition, studies
have identified disparate processing of emotional information at the neural level between children
and adults with ASD and their peers (Harms, Martin, & Wallace, 2010; Kana et al., 2015). For
example, Silani et al. (2008) examined fMRI activation in 15 adults with ASD and 15 age-╉matched
peers without ASD who viewed positive, neutral, and negatively valenced stimuli, while rating the
9
3
2

The interplay between emotion regulation and ASD characteristics and impairment 239

valence of each image. Adults with ASD showed decreased brain activation in areas associated
with emotional awareness compared to adults without ASD in their sample. Further, although the
specific findings are inconsistent, studies largely indicate atypical patterns of neural reactivity dur-
ing emotional face processing (Harms et al., 2010) and emotional language processing (Lartseva,
Dijkstra, & Buitelaar, 2014) in ASD.
The only two neuroimaging studies to explicitly investigate emotion regulation in ASD both
utilized cognitive reappraisal tasks (Pitskel, Bolling, Kaiser, Pelphrey, & Crowley, 2014; Richey
et al., 2015). Cognitive reappraisal tasks provide an opportunity to isolate neural reactivity that
is due to regulation of emotion by requiring participants to increase or decrease their emotional
reaction to various stimuli. Importantly, behavioral valence ratings from both youth with ASD
(Pitskel et  al., 2014)  and adults with ASD (Richey et  al., 2015)  revealed significant differences
between conditions which supports the use of cognitive reappraisal tasks as an emotion regula-
tion probe in ASD. A consistent finding from both studies, despite the use of different stimuli, was
the decreased ability to suppress amygdala activation in ASD compared to typically-​developing
controls, which may provide one mechanistic account for impaired voluntary emotion regulation
in ASD.
Another possibility is that problems with emotion regulation stem from underlying dif-
ferences in physiological arousal. There are a number of different measures of physiological
reactivity that have been used to investigate arousal in ASD, including pupillometry, heart
rate, heart rate variability, respiratory sinus arrhythmia, blood pressure, electrodermal activity,
and cortisol. One question that has been explored using a variety of these methods is whether
individuals with ASD have atypical baseline levels of physiological arousal, with some theories
suggesting baseline levels would be increased and others hypothesizing decreased baseline
arousal. A review of studies focused on baseline physiological reactivity found that approxi-
mately half found no difference compared to controls across a variety of physiologic indica-
tors, while the other studies were in both directions (Lydon, Healy, Reed, Mulhern, Hughes,
& Goodwin, 2014).
Currently, we may also infer conclusions about physiological differences associated with emo-
tion regulation based on studies of related constructs such as the processing of emotional faces
and response to stressors. One interesting conclusion from this work is that individuals with ASD
may perceive their physiological activity differently than their typically-​developing peers. For
instance, Shalom et al. (2006) measured the concordance between a physiological measure of
skin conductance in ten children with ASD, and ten children without ASD, who were presented
with pleasant, unpleasant, and neutral pictures, and how pleasant and interesting they rated the
images. While physiological results across the image types were comparable across children with
ASD and those without, self-​reported affect ratings were significantly different across the two
groups, such that children with ASD reported more similar answers when rating the pleasant-
ness and interestingness of unpleasant, pleasant, and neutral pictures when compared to children
without ASD. These findings suggest that children with ASD may perceive and consciously report
different experiences of emotional stimuli even though physiologically, they are comparable to
children without ASD. Recently, a study of high-​functioning adults supported the importance
of the individual’s perception, by finding the perception of stress, but not physiological reactivity
during a social stressor, was related to social outcomes (Bishop-​Fitzpatrick, Minshew, Mazefsky,
& Eack, 2016).
It is difficult to make firm conclusions about specific physiological reactivity differences in ASD
as there have only been a few small studies employing different tasks, mostly with inconsistent
findings. A focus on the higher quality studies does however, indicate discernable differences in
physiological responses during stressor tasks in ASD and, in particular, suggests that further study
0
4
2

240 Emotion Regulation and Autism Spectrum Disorder

of the limbic-╉hypothalamic-╉pituitary-╉adrenal (LHPA) axis system is warranted (Lydon, Healy,


Reed, et al. 2014). This is of great relevance to emotion regulation, as the LHPA system is respon-
sible for regulating the body’s response to stress and restoring homeostasis (Smith & Vale, 2006).
Also intriguing is the notion that there may be high-╉and low-╉arousal subgroups in ASD (Schoen,
Miller, Brett-╉Green, & Hepburn, 2008), which may play a role in the observed within-╉ASD vari-
ability in emotion regulation capacity.

Cognitive and behavioral influences


Although there is limited empirical evidence of an association between the cognitive and infor-
mation processing style of individuals with ASD and emotion regulation impairments, theoreti-
cal models have highlighted some of the characteristics that are likely to play a role (Mazefsky &
White, 2014). Some of the most likely factors in ASD include alexithymia (difficulty identifying
and labeling one’s own emotions), difficulty perceiving social and emotional cues, poor problem-╉
solving and reasoning ability, cognitive rigidity, and sensitivity to change and environmental
stimulation (see Figure 12.1).
Some have argued that emotional problems in ASD may be attributed entirely to the well-╉
documented occurrence of alexithymia (Bird & Cook, 2013). Indeed, the accurate identifica-
tion of one’s own feelings is necessary in order to employ response-╉focused voluntary emotion
regulation strategies, and accurate emotion identification and emotional language is essential for
specific strategies such as talking to others or joint problem solving (Mazefsky & White, 2014).
However, alexithymia may not be as relevant for more automatic and involuntary emotion regula-
tion processes, and emotion regulation deficits have been found in ASD even when controlling
for alexithymia (Samson, Huber, & Gross, 2012). This argues for alexithymia playing an important
role in emotion regulation but not being the sole explanation for aberrant emotional functioning,
as Bird and Cook (2013) proposed.
Difficulty accurately perceiving other types of emotional and social cues in ASD may also
interfere with emotion regulation. One prime example is deficits in theory of mind, or the

Poor
problem-
solving & Difficulty
abstract reading
Lower reasoning social &
inhibition emotional
cues

Sensitivity to
Cognitive
change &
rigidity; Poor
environmental
flexibility
stimulation

Biological
Alexithymia, predisposition
Emotion
Limited (physiological
Dysregulation
emotional arousal, neural
in ASD
language circuitry,
genetics)

Figure 12.1╇ Characteristics of ASD that may contribute to emotion dysregulation


Reprinted from Child and Adolescent Psychiatric Clinics of North America, 23 (1), Carla A. Mazefsky and Susan W. White,
Emotion Regulation Concepts & Practice in Autism Spectrum Disorder, pp. 15–24, http://dx.doi.org/10.1016/j.chc.2013.07.002
Copyright © 2014 Elsevier Inc., with permission from Elsevier.
1
4
2

Interventions to improve emotion regulation 241

ability to take other’s perspectives (Samson et al., 2012). Poor perspective taking could lead
to misreading other’s intentions as negative or hostile, a perception bias that has been linked
to increased negative affect in non-╉ASD studies (Schultz, Izard, & Bear, 2004). Even if an
individual with ASD knows the right emotion regulation strategies to use, they face several
barriers in the effective implementation of these strategies. For example, problems with inhib-
itory control (Geurts, van den Bergh, & Ruzzano, 2014), a tendency to be rigid and have dif-
ficulty shifting (Granader et al., 2014), impaired abstract reasoning and poor problem solving
(Williams, Mazefsky, Walker, Minshew, & Goldstein, 2014), may interfere with the timing
of efforts as well as the generation of flexible solutions during novel situations (Mazefsky &
White, 2014).
A recent study found that repetitive behaviors were strongly correlated with the presence of
emotion dysregulation among children and adolescents with ASD (Samson, Phillips, Parker, Shah,
Gross, & Hardan, 2014). This is consistent with hypotheses that the tendency to be perseverative
and difficulty shifting may lead to sustained emotional reactions and difficulty down-╉regulating
negative affect in ASD (Mazefsky, Pelphrey, & Dahl, 2012). Heightened sensory sensitivity may
also interfere by decreasing the threshold for the experience of negative affect as well as creating
a physiologic state of stress which makes voluntary emotion regulation efforts challenging (Dunn
& City, 1997). Although this implies causality, the relationship is likely reciprocal. For example,
increased repetitive behaviors or unusual sensory reactions may also occur as a consequence of
problems with emotion regulation, and some have even argued that repetitive behaviors may be a
coping mechanism (Turner, 1999), further complicating the dynamic relationship between regu-
lation and expression of ASD symptomatology.
Experiencing highly dysregulated emotion undoubtedly further impairs one’s ability to
attend to and process information from the environment, including social information. Wood
and Gadow (2010) proposed a model describing the reciprocal relation between ASD and anxi-
ety that further illustrates this point. Specifically, they suggest that ASD-╉related stressors (e.g.,
social confusion; peer rejections and victimization; prevention or punishment of preferred
behaviors or interests; and frequent aversive sensory experiences) lead to increased overall
negative affectivity, anxiety disorders (specific type dependent on the experiences), or depres-
sion. In turn, such negative affectivity contributes to more ASD symptoms, conduct problems,
and personal distress, which lead to even further ASD-╉related stressors, completing a negative
cycle. Although Wood and Gadow (2010) focused their model on the development of anxiety
or depression, this dynamic cycle also applies to emotion dysregulation more broadly. As such,
if interventions can effectively improve emotion regulation, there is an opportunity not only to
support emotional well-╉being and decrease distress, but also to improve the course of ASD and
functional outcomes.

Interventions to improve emotion regulation


A variety of treatment strategies have been employed to address problems with emotion regula-
tion in ASD, largely focused on targeting a particular manifestation of poor emotion regulation
(e.g., irritability) or the presence of an anxiety disorder. There is a growing interest in transdiag-
nostic treatment approaches that target the underlying mechanism (in this case, emotion regu-
lation), rather than a specific disorder. Going beyond a disorder-╉specific approach can make it
easier to address the common factors that can lead to multiple emotional problems and emotion-
ally driven behaviors (Barlow et al., 2011), making it more likely to be effective in situations when
patients exhibit combinations of anger, sadness, and anxiety. The primary forms of treatment that
have received research attention thus far include psychopharmacological approaches and psycho-
therapy. The current evidence for each is described below.
2
4

242 Emotion Regulation and Autism Spectrum Disorder

There are only two psychopharmacologic medications with U.S. Federal Drug and Safety
Administration approval for use in ASD. They are both indicated for the treatment of irri-
tability, which is arguably a manifestation of emotion regulation failure. These medications
include risperidone and ariprazole, antipsychotics with well-​established evidence support-
ing their efficacy despite some significant adverse effects including weight gain and metabolic
changes (Marcus et  al., 2011; McCracken et  al., 2002). Off-​label use of other medications is
common in ASD with an estimated two-​thirds of children with ASD on at least one psychotro-
pic medication and over one-​third on multiple medications (Spencer et al., 2013). Although the
pace of clinical trials research has increased recently, a review concluded that there was insuf-
ficient evidence to support the efficacy of mood stabilizers or serotonin reuptake inhibitors in
ASD (Siegel & Beaulieu, 2012). Further, studies of serotonin reuptake inhibitors in ASD have
focused on repetitive behaviors as the target symptom rather than anxiety, and there have actu-
ally been no randomized controlled trials of psychotropic medication use for anxiety in ASD
(Vasa et al., 2014).
In contrast, there is now considerable evidence that cognitive behavior therapy (CBT) is effi-
cacious in reducing symptoms of anxiety in youth with ASD who do not have ID, and there
is emerging evidence of its efficacy in adults with ASD. Ung and colleagues’ (2015) systematic
review and meta-​analysis of 14 CBT trials for children and adults with ASD indicated moderate
treatment effect sizes across studies. Studies compared CBT to a wait-​list (N = 8), treatment as
usual course (N = 3), or alternate treatment (N = 1). However, it is important to note, two of the
14 studies were open trials with no control condition. Seven studies were delivered individually
either with or without parents, six studies were delivered in a group format, and one study incor-
porated a combined individual and group format approach. Parents were involved in 11 of the
studies. No difference in treatment response was found based on informant (parent, child, clini-
cian) and modality (group and individual formats with and without parents). In summary, treat-
ment effects typically translated into clinically significant changes for 50–​70% of participants (Vasa
et al., 2014). An earlier meta-​analysis of eight randomized controlled trials of CBT with children
with ASD and at least average intellectual functioning reported large effect sizes for parent-​and
clinician-​reported child anxiety, and small effect sizes for child-​reported anxiety (Sukhodolsky,
Bloch, Panza, & Reichow, 2013). Danial and Wood (2013) reviewed intervention studies focus-
ing on mental health problems for children five to 18 years of age with high-​functioning autism
spectrum disorder (HFASD), and included randomized trials, group comparison studies, and
multiple baseline designs. Similar to other systematic reviews, they found CBT for anxiety to be
a promising treatment. Reviews also suggest that we know almost nothing about the long-​term
effectiveness of CBT treatment in this population, even with reference to anxiety, where up to
44% of youth who improve post intervention show some reduction in gains at follow-​up ten to
26 months later (Selles et al., 2015).
CBT studies to date have almost exclusively focused on anxiety. Only one randomized trial
exists on the efficacy of CBT to address anger problems in youth with ASD. Sofronoff and col-
leagues (2007) reported on a waitlist controlled trial of a six-​week CBT intervention for anger
management among 45 ten-​to 14-​year-​old children with ASD with at least average IQ, while
parents attended a concurrent parent group to review session material. Parents in the interven-
tion group reported significant decreases in child anger and increased confidence in managing
child anger compared to parents in the waitlist group. Children in the intervention generated
more appropriate coping strategies when given the hypothetical scenario, “Dylan is being Teased”
(Attwood, 2004) compared to those on the waitlist. Until further well-​controlled studies are done
though, the efficacy and effectiveness of CBT for externalizing problems among youth with ASD
remains unclear (Danial & Wood, 2013).
3
4
2

Interventions to improve emotion regulation 243

Research on the psychotherapeutic treatment of emotional concerns in adults with ASD lags
behind research on children and adolescents. A recent systematic review identified six studies that
used CBT to treat mental health conditions in adults with ASD, and included case studies and
case series, quasi-​experimental designs, and randomized controlled trials (Spain, Sin, Chadler,
Murphy, & Happe, 2015). All participants were diagnosed with ASD with at least average intel-
lectual functioning, and treatment was focused on treating anxiety disorders, mood disorders,
or self-​harm. Four studies provided CBT individually and two delivered it in a group format.
No parents or caregivers were reported to be involved. In their narrative synthesis of the find-
ings, Spain et al. (2015) reported decreases in self-​and clinician-​reported mental health symptom
severity. However, because of the lack of methodological rigor of the extant literature, results must
be interpreted with caution.
Although most of the psychotherapeutic studies in ASD have focused on CBT as the treatment
modality, there is emerging evidence in support of mindfulness-​based interventions for anxiety
and depression in ASD. Spek, van Ham, and Nyklicek (2013) conducted a randomized waitlist
controlled trial to examine changes in anxiety, depression, and rumination following a nine-​week
mindfulness-​based therapy for adults ages 18 to 65 years old (M = 44 years) with ASD and average
verbal ability. Two clinicians facilitated the group and session content included body scans, medi-
ations, breathing exercises, and movement exercises. Compared to adults in the waitlist group,
adults in the mindfulness group reported a reduction in anxiety, depression, and rumination and
increases in positive affect.
Research examining the effects of mindfulness-​based strategies with children and youth is also
surfacing. De Bruin, Blom, Smit, van Steensel, and Bogels (2015) evaluated a mindfulness training
intervention for youth 11 to 23 years old (M = 15.8 years) and their parents who participated in a
parallel mindful parenting group, using a pre-​post design without a control condition. Youth and
their parents completed measures of mindfulness, worry, depressed mood, ruminations, quality
of life, and ASD symptoms before the intervention, after the intervention, and at follow-​up. Youth
reported an increase in quality of life and a decrease in rumination following the intervention,
but no change in worrying, ASD symptoms, or mindfulness. Parents reported increases in mind-
ful parenting (i.e., attentive listening, emotional awareness, self-​regulation, and nonjudgmental
acceptance) at post-​intervention and follow-​up. They also reported significant improvements in
their children’s social responsiveness at follow-​up.
In a multiple-​baseline design across participants, Singh et  al. (2011) examined the effects of
Meditation on the Soles of the Feet on aggressive behavior of three youth (all male) ages 14, 16,
and 17 years with ASD without ID. At the start of the training, youth-​mother dyads practiced
the meditations together. Once the youth learned the basics of the meditation, an audiotape was
provided for self-​guided practice. Over the course of the meditation training, results indicated a
decreasing trend in aggressive behavior across all three participants, according to parent and sib-
ling reports of observed aggression. Once participants met the criterion of no aggressive behav-
iors for four weeks, there were reports of only one or two aggressive acts within the three-​year
follow-​up period. Results from these preliminary evaluations provide support for mindfulness-​
based skills training as a component of interventions to address emotion regulation difficulties
among youth with ASD.
In addition to studies examining mindfulness practices with youth with ASD themselves,
indirect effects of parent mindfulness training on child externalizing problems have also been
reported. For example Singh, Lancioni, et al. (2006) investigated the effects of a 12-​week parent
mindfulness training intervention on child behavior problems. They found that following par-
ent training, parents reported decreases in their children’s externalizing problems. These results
highlight that the children’s externalizing symptoms were reported to decrease, without any direct
42

244 Emotion Regulation and Autism Spectrum Disorder

intervention with the children. It is possible that the intervention influenced parent perceptions of
their child including unconditional acceptance and non-​judgment rather than the externalizing
symptomatology itself. Nonetheless, this shift in perspective appears to have positive implications
on the parent-​child dynamic and additional well-​controlled studies are needed to elucidate how
parent mindfulness-​based interventions help individuals with ASD and their parents.
Given high rates of ID among those with ASD, additional studies are needed to examine
treatments to help children and adults with ASD and ID cope with stressors and manage their
emotions. Frequently using single-​subject designs, behavioral interventions using exposure and
desensitization have consistently shown to be effective in treating anxiety in this population (Lang
et al., 2011). Jennett and Hagopian (2008) reviewed treatments for phobic avoidance among chil-
dren and adults with ID; approximately one third of the participants in included studies also had
ASD, and all studies involved exposure to the feared stimulus and reinforcement of appropriate
behaviors, specifically approaching the feared stimulus. Results consistently indicate a high degree
of effectiveness, leaving the authors to conclude that behavioral treatments can be considered well-​
established for phobic avoidance in this population. More recently, Lydon, Healy, O’Callaghan,
et al. (2014) systematically reviewed treatments for fears and phobias, including three studies (one
case study and two single-​subject designs) in which participants had both ASD and ID. Authors
reported positive treatment outcomes with all of the behaviorally-​based interventions: Reinforced
practice (Chok et al., 2010), contingent reinforcement and systematic exposure (Schmidt et al.,
2013), and stimulus fading and differential reinforcement (Shabani & Fisher, 2006). Behavioral
interventions have been used to teach relaxation strategies to address disruptive behaviors as well.
In a multi-​element single subject design, Mullins and Christian (2001) examined the effects of
progressive muscle relaxation strategies to decrease disruptive behaviors of a 12-​year-​old male
with ASD and ID during unstructured leisure activities. There was a decrease in the duration of
disruptive behaviors when relaxation strategies were cued before leisure activities, suggesting that
the training had a positive impact on his problem behaviors.
Given the central role that emotion regulation plays in the onset or maintenance of emotion
disorders, and the high degree of emotion regulation difficulties in individuals with ASD, it is
surprising that few studies exist on evaluating the treatment of emotion regulation deficits. The
vast majority of studies focus exclusively on symptoms of anxiety in ASD, neglecting the need to
also target co-​occurring depression or anger. Scarpa and Reyes (2011) evaluated a modified CBT
program to address emotion dysregulation in five six- to eight-year-old children with ASD and
average intellectual functioning compared to six children in a delayed treatment condition. Using
one-​tailed tests of significance, the authors suggest improvements in child reported coping strate-
gies in response to vignettes, and parent reported child negativity/​lability and emotion regulation.
More recently, Thomson, Burnham Riosa, and Weiss (2015) evaluated the feasibility of the Secret
Agent Society-​Operation Regulation (SAS-​OR; described in detail below) in addressing emotion
regulation in 13 youth with ASD, eight to 12 years of age, and their parents. All children had at
least average intellectual functioning and parent-​reported problems with anxiety, depression or
anger/​aggression. The SAS-​OR program was found to have high participant and therapist satisfac-
tion and therapeutic alliance, excellent treatment adherence across sessions (including homework
completion and child engagement), and face validity. Results revealed parents reported significant
improvements in child emotional lability, internalizing symptoms, behavioral dysregulation, and
adaptive behavior. Independent clinician evaluation indicated significant improvements in child
overall severity and number of psychiatric diagnoses. Children reported significant improvement
in inhibition and dysregulation across anger, anxiety, and sadness, as well as in generating emo-
tion regulation strategies to hypothetical vignettes. An RCT of the SAS-​OR program is currently
underway (ISRCTN67079741), and the details regarding each session and the treatment approach
are articulated below.
5
4
2

PROMOTING SKILL LEARNING AND APPLICATION 245

The Secret Agent Society-╉Operation Regulation program


The SAS-╉OR program is a spy-╉themed emotion regulation intervention delivered individually to
children and their parent(s). The program, comprised of ten weekly one-╉hour sessions, aims to
teach children with ASD a core set of mindfulness-╉based CBT emotion regulation skills that can
be applied to a range of feelings in different contexts. Specifically, it targets the potentially modifi-
able characteristics of ASD that may contribute to emotion regulation difficulties, as shown in
Figure 12.1. Program content aims to teach children how to recognize and understand emotional
experiences in themselves and others by attending to and accurately interpreting social-╉emotional
clues, to think more flexibly about distressing situations, to use mindfulness strategies to tolerate
unpleasant sensory experiences and uncomfortable situations, and to successfully apply a step-╉by-╉
step formula to solve social dilemmas.
Therapists work collaboratively with children and their parents to determine core goals for ther-
apy at the outset. These typically involve symptom reduction and/╉or better coping skills in one
or more emotional domains (i.e., anger, anxiety, or depression). The SAS-╉OR Facilitator Manual
guides the therapist in how to tailor the program to individual client needs and preferences (e.g.,
focusing on anger management strategies for children with behavioral challenges versus anxiety
management for children with comorbid anxiety disorders). The SAS-╉OR program content draws
from and expands on the emotion recognition and emotion regulation strategies and multime-
dia resources featured in the evidence-╉based Secret Agent Society group social skills program for
children with ASD (Beaumont, 2010), which is published and distributed worldwide by Social
Skills Training Pty Ltd. (www.sst-╉institute.net), a subsidiary of the Cooperative Research Centre
for Living with Autism (www.autismcrc.com.au). The program and its variants are designed for
children with ASD who have an IQ at least within the average range.

Promoting skill learning and application


SAS-╉OR features several core elements to promote children’s skill learning and application to
home and school. Visually engaging child resources (e.g., the SAS-OR Cadet Handbook that
features full-╉color character illustrations of skill steps and concepts; pocket sized collectable
Relaxation Gadget Code Cards featuring color images of relaxation and mindfulness techniques;
the SAS compute game), active session activities, skill demonstrations and discussions aim to
optimize children’s interest and engagement in the learning content by catering to all learning
styles (visual, verbal, and kinesthetic). The SAS computer game (a resource from the Secret Agent
Society group social skills training program) also capitalizes on the affinity that individuals on the
spectrum often have for technology (Odom et al., 2014).
As skill generalization is a common challenge for children with ASD (Church et al., 2015), SAS-╉
OR engages parents and school staff as co-╉therapeutic agents to optimize children’s skill applica-
tion to daily life. Where possible, parents are involved in session activities and games to help
them to become fluent in the program “language” and to provide them with a model of how to
talk to their child about emotions. Parents are provided with the SAS-╉OR Parent Workbook that
summarizes core concepts and provides tips and strategies on how to help children to apply skills
from the program outside of formal therapy sessions (e.g., through the use of incidental teaching,
modelling and rewarding child skill usage). Weekly teacher tip sheets are also included in the
program materials to provide a child’s teachers and other school support staff with a brief sum-
mary of the skills that they are learning in the program, and tips on how they can support their
student in applying these skills in class and at break time. An example Teacher Tip Sheet is shown
in Figure 12.2.
A Home-╉School Diary (see Figure 12.3) is also included in the program materials to allow par-
ents and school staff to track children’s progress in using their emotion regulation skills at home
6
4
2

Figure 12.2  An example Teacher Tip Sheet from the SAS-​OR Program


Reprinted with the permission of Social Skills Training Pty Ltd (www.sst-​institute.net)
7
4
2
Figure 12.3  The Home School Diary used in the SAS-​OR Program
Reprinted with the permission of Social Skills Training Pty Ltd (www.sst-​institute.net)
8
4
2

248 Emotion Regulation and Autism Spectrum Disorder

and at school each day. Parents and teachers are encouraged to award a child a diary point when
they use their target skill for the week at home and at school (based on the child’s self-​report). If
the child reaches their daily points “target,” which is negotiated between the parent and therapist
at each session, their points can be exchanged for a home-​based reward. The child is also given
weekly “home missions” to complete with parent support between sessions, which involve prac-
ticing the skills that they are learning at home, at school and when they are out. After completing
each mission, the child is asked to answer questions in the Secret Agent Journal section of their
SAS-​OR Cadet Handbook. These questions help the child to understand the benefits of using their
SAS-​OR skills, building their intrinsic motivation to do so in the future when the Home-​School
Diary reward system is gradually phased out.
Children are also given full-​color pocket-​sized collector “Code Cards” throughout the program,
and a Code Card holder. Each card features an image of a different relaxation and/​or mindfulness
technique, and a description of what level of anxiety or anger it is best used for and where it is best
used (at home, at school and/​or when the child is out). Children are encouraged to secretly refer to
their Code Cards to remind them of the strategies they can use to feel happier, calmer, and braver
and to make smart choices just before entering situations where they are likely to feel uncomfort-
able or distressed (with adult help where available). An example Relaxation Gadget Code Card is
shown in Figure 12.4.
Each SAS-​OR session typically commences with a ten-​minute review of how the child and
parent(s) did with completing the home missions and using the Home-​School Diary during the
week. Any challenges are problem-​solved to set the child and family up for success in completing

Figure 12.4  An example Relaxation Gadget Code Card from the SAS-​OR Program (Front and Back of
Card shown)
Reprinted with the permission of Social Skills Training Pty Ltd (www.sst-​institute.net)
9
4
2

Promoting skill learning and application 249

their home missions and using the Home-╉School Diary the following week. For example, if a child
struggled to differentiate feelings of anger from anxiety and rates the intensity of their emotions
one week, they may just be asked to practice detecting when they feel happy or upset the follow-
ing week.
After reviewing home missions and the Home-╉School Diary with the child and parent(s), the
therapist facilitates a series of espionage-╉themed games and activities (including playing assigned
sections of the SAS computer game) that teach the child how to recognize and manage their emo-
tions (see below for further details). Based on literature showing that children with ASD often
struggle to accurately identify emotions in themselves and others (e.g., Rieffe et al., 2007; Uljarevic
& Hamilton, 2013), initial session content predominantly focuses on teaching children emotion
identification skills, as they are a prerequisite for them to be able to use their emotion regulation
strategies when needed. Sessions end with a review of the home missions to be completed during
the coming week and discussing the target behavior(s) and points target for the Home-╉School
Diary with the child’s parent(s).

Session content
In the first session of the program, the child and parent(s) are introduced to the aims of the inter-
vention (i.e., “to help them to feel happier, calmer, and braver”) and create a Challenge Card fea-
turing a picture of the child’s favorite fictional or real life hero or role model who has to be brave,
face their fears and do things that they do not want to do at times. On the card, the child creates
a hierarchy of boring, scary, or difficult things that they have to do in their own life (ranked from
“least” to “most” challenging). As the program progresses, the child is asked to use the emotion
regulation skills that they are learning to cope with unpleasant feelings and to face these fears or
challenges between sessions. Session rules and an in-╉session reward system for following the ses-
sion rules are negotiated in the first session. Children receive points for following the rules in each
session (e.g., trying their best) and if they reach their points target for the session, their points are
exchanged for an end of session reward. A rewards menu is negotiated with the child in the first
session, listing up to five different types of rewards that they would like to receive if they reach
their session points target. Recommended points targets for each session are provided in the SAS-╉
OR Facilitator Manual, although these can be adjusted as the therapist sees fit.
In Session One, the child is introduced to activities in the SAS computer game that teach them
how to recognize emotions in other people from facial expression and body posture/╉movement
clues (“Spot the Suspect” and “The Line Up” games), fostering emotional awareness based on cues
and context. They also practice the mindfulness technique of being aware of their own breath
(“The Breath Analyzer”), introducing them to attention shifting from negative emotions. At the
end of the first session, the therapist spends time alone with the child’s parent(s) explaining the
Home-╉School Diary monitoring and reward system. Home missions for the coming week include
the child practicing their mindful breathing (the Breath Analyzer Mission) and detecting how
other people feel from face and body clues (the Secret Spy Mission) with parent and school staff
support.
After the Home-╉School Diary and home mission review, Session Two commences with a cha-
rades game (“Detection of the Expression”) where the child, therapist and parent(s) take turns
acting out and guessing how each other feels from facial expression and body posture clues. The
child subsequently plays a SAS computer game activity (“Voice Verification”) that involves detect-
ing how people feel from the pitch, pace, and volume of their voice. They then play a game that
involves saying and listening to different secret messages said in different voice tones with their
parent(s) and the therapist, before learning the mindfulness technique of scanning their bod-
ies for physical sensations in the moment (“Body Scan”). Home missions for the week include
0
5
2

250 Emotion Regulation and Autism Spectrum Disorder

practicing body-​scanning and detecting how other people feel from their tone of voice with the
help of their parent and teacher mentors.
Session Three involves the child playing computer game activities that teach them the body
clues and thoughts that signal emotions of happiness, sadness, anxiety and anger within them-
selves (“Detective Laboratory”) and learning that different body clues signal different emotional
intensities (“Degrees of Delight and Distress”). Children draw their own body clues that signal
target emotions on body outlines, and play a game similar to statues with their therapist and
parent(s) to practice quickly identifying emotion body clues (“The Body Clues Freeze Game”).
They repeat the Body Scan activity introduced in Session Two before planning to practice this
skill for their weekly home mission with parent support, practicing awareness of their own arousal
through a focus on their internal cues.
In Session Four, the child creates anxiety-​, anger-​and/​or sadness “Emotionometers” (pocket-​
sized emotion scales featuring stickers showing the body clues and situations where they feel low,
moderate and high levels of the target emotion). They then learn how to piece together face-​,
voice-​, body-​and situational clues that signal how someone is feeling in the “Secret Agent Viewing
Panel” SAS computer game activity, and plan how they can use their Emotionometers to detect
the type and strength of their emotions during the week. Learning to rate degrees of emotions is
common across all CBT interventions, and here is applied to internalizing and externalizing emo-
tional states to promote understanding and differentiation across emotions.
Session Five involves the child learning about relaxation “gadgets” to help them to feel hap-
pier, calmer and braver and to make smart choices. This concept is initially introduced to the
child through the first virtual reality mission in Level Three of the SAS computer game. Each
of the virtual reality missions in Level Three of the game involve the child choosing how their
avatar can cope with a challenge (e.g. trying something new, performing poorly at a game or
competition), with both appropriate (take some slow breaths, think helpful thoughts) and less
appropriate (scream and shout, run away) choices available (see the screenshot in Figure 12.5).
The child discovers the consequences of different response options for their avatar, allowing
them to learn through a depersonalized self-​discovery process that using skills to stay calm
and cope typically leads to better outcomes than aggressive outbursts, avoidance, or escape
behaviors.
After the child finishes playing the first Level Three virtual reality mission in Session Five, the
therapist helps them to choose Relaxation Gadget Code Cards illustrating the gadgets or strate-
gies that they would like to use to feel happier and calmer in the situations shown on their anxi-
ety, anger and/​or sadness Emotionometers. The child and parent(s) learn and rehearse the “O2
Regulator Gadget” (slow, mindful breathing) in session, before planning how they will use this
and other Relaxation Gadgets during the week in situations featured on the child’s Challenge Card
(created in Session One).
In Session Six, the child plays the second Level Three virtual reality mission in the SAS computer
game, which introduces them to the “Fire Engine” Relaxation Gadget (doing a physical activity to
burn up anxious or angry energy when really upset). They also learn about the “Relaxation Radar”
gadget—​being on high alert for relaxing or friendly things around you (e.g., friends smiling at you
while you are giving a talk in front of the class). The possible role of sensory items (e.g., a piece of
fabric or scratch and sniff stickers) as Relaxation Gadgets is also explained to children in this ses-
sion, transforming sensory-​seeking behaviors or interests that are common in children with ASD
(Little, Ausderau, Sideris & Baranek, 2015) into an adaptive coping skill. The child is introduced
how to use their five “super-​senses” (sight, touch, smell, taste, and sound) to be aware of their sur-
roundings as well as their internal body sensations with the “Enviro-​Body Scan” gadget, and they
attach stickers illustrating their chosen Relaxation Gadgets to the backs of their Emotionometers.
1
5
2

Promoting skill learning and application 251

Figure 12.5  Screenshot from a Level 3 Virtual Reality Mission in the SAS Computer Game
Reprinted with the permission of Social Skills Training Pty Ltd (www.sst-​institute.net)

Their home mission for the coming week involves using their Relaxation Gadgets to continue
climbing the hierarchy of situations featured on their Challenge Card.
Cognitive strategies for coping with emotional discomfort are taught in Session Seven of the
SAS-​OR program, aimed at enhancing cognitive control. These include “shooting down” unhelp-
ful thoughts with more helpful alternatives (the “Helpful Thought Missile” Gadget) and being
aware of unhelpful thoughts and allowing them to pass when ready, as if they were printed on the
wings of an SAS spy plane or blimp (the “Thought Tracker”). Children continue rehearsing the
Enviro-​Body Scan mindfulness activity introduced in Session Six, practicing their awareness of
physiological arousal and body cues as well as cues in the environment. They also plan their home
missions of being on “unhelpful thought alert” and continuing to progress up their Challenge
Card situational hierarchy, using their Relaxation Gadgets to help them cope.
Session Eight involves the child learning how to be a “Losing Champion”—​that is, learning how
to use their Relaxation Gadgets to stay calm when they are losing at a game or competition, just
like their role model or hero does at times. Coping with uncontrollable situations like losing is a
common challenge for children with ASD that contributes to their friendship difficulties (Hebron,
Hunphrey & Olfield, 2015). This concept is introduced with the third Level Three SAS computer
game virtual reality mission, and then rehearsed in session with the child playing a game with
their therapist and parent(s). The child is asked to practice being a losing champion when play-
ing games with family, classmates and other friends as their home mission for the week, provid-
ing them with opportunities to practice using their Relaxation Gadgets during ecologically valid
social experiences.
In Session Nine, the child is taught how to use their Relaxation Gadgets and other strategies
to cope with another social challenge often faced by children on the spectrum–​identifying, pre-
venting and managing bullying (Rowley et al., 2012). These skills are introduced to the child in
the final computer game virtual reality mission. The child learns clues to help them differentiate
friendly joking from mean teasing (e.g., whether the person says sorry or follows your instruc-
tions when you tell them to stop) and makes their own customized Bully Guard Body Armor
2
5

Figure 12.6  Summary of the D.E.C.O.D.E.R Problem-Solving Steps


Reprinted with the permission of Social Skills Training Pty Ltd (www.sst-​institute.net)
3
5
2

Conclusion 253

Code Card featuring stickers showing the bully blocking strategies that they intend to use. The
child is encouraged to put these skills into action during the week (with the help of their parent
mentor) as their home mission.
Session Ten, the final session of the program, involves the child and parent(s) learning how to
use a step-╉by-╉step problem-╉solving formula (the “D.E.C.O.D.E.R”) to change situations that are
causing them distress. As children with ASD often struggle to detect social problems when they
first arise due to their challenges with recognizing how others feel and knowing the implicit social
rules for different situations (Rowley et al., 2012), the formula focuses on helping them detect
social problems in the first instance based on their own internal emotional state and other people’s
facial expressions, voice tones, body postures and movements, what they say and do and the situ-
ation that they are in. For example, if a child accidentally says something that offends someone,
they need to detect this before they can determine the best way to respond. The formula also
focuses on the child developing a detailed plan for their chosen solution with adult support before
putting it into action, including planning and rehearsing the Relaxation Gadgets that they will use
to stay calm. A summary of the D.E.C.O.D.E.R steps is shown in Figure 12.6.
After the D.E.C.O.D.E.R formula is introduced in Session Ten, the child, therapist, and parent(s)
play a ball game to review the skills that have been taught in the SAS-╉OR program, plan how they
can use these skills to cope with future challenges and schedule weekly home review meetings
of session content. The child finishes the final level (Level four) of the SAS computer game, and
is presented with a graduation medal to reward them for their efforts throughout the program.
Finally, the child and parent(s) plan how the child can use the D.E.C.O.D.E.R steps to solve an
upcoming problem for their final home mission, and the parent(s) are given tips on how to sup-
port their child in continuing to develop their emotion regulation skills (e.g., gradually phasing
out the Home-╉School Diary, continuing to prompt and reward skill usage when needed with the
help of visual supports). As described, preliminary trial results of the SAS-╉OR intervention have
been encouraging, with parents and children reporting high program acceptability ratings, and
children on the spectrum showing improvements in their emotion regulation skills (Thomson,
Burnham, Riosa, & Weiss, 2015). Future research will inform how the user-╉friendliness and effec-
tiveness of the program can continue to be improved.

Conclusion
There is robust evidence indicating that we can address a common outcome of emotion dysregula-
tion in youth with ASD (i.e., anxiety), and growing interest in targeting core underlying emotion
regulation skills, primarily based on mindfulness and cognitive behavioral approaches. As we
progress toward building this evidence base, an improved understanding of a transdiagnostic
approach that is not solely focused on symptom reduction will likely emerge, and with it, more
effective interventions that can address a broader array of profiles and deficits. Understanding the
dynamic process of emotion regulation through the lens of ASD symptomatology, and in the face
of ASD-╉related stressors, is important to support individualized skill development and improved
outcomes.

References
Aldao, A., & Nolen-╉Hoeksema, S. (2010). Specificity of cognitive emotion regulation strategies:
A transdiagnostic examination. Behaviour Research and Therapy, 48(10), 974–╉983.
American Psychiatric Association (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Arlington, VA: American Psychiatric Publishing.
4
5
2

254 Emotion Regulation and Autism Spectrum Disorder

Attwood, T. (2004). Dylan is being teased. Exploring feelings: Cognitive behaviour therapy to manage anger.
Arlington, TX: Future Horizons Inc.
Barlow, D. H., Farchione, T. J., Fairholme, C. P., Ellard, K. K., Boisseau, C. L. Allen, L.B. & Ehrenreich-​
May, J. (2011). The unified protocol for transdiagnostic treatment of emotional disorders: Therapist guide.
New York: Oxford University Press.
Bebko, J. M., Schroeder, J. H., Weiss, J. A., Wells, K., McFee, K., & Goldstein, G. M. (2008). The face of
Autism research as reflected in the IMFAR looking glass. Research in Autism Spectrum Disorders, 2(3),
385–​394.
Ben Shalom, D., Mostofsky, S. H., Hazlett, R. L., Goldberg, M. C., Landa, R. J., Faran, Y., … Hoehn-​
Saric, R. (2006). Normal physiological emotions but differences in expression of conscious feelings in
children with high-​functioning autism. Journal of Autism and Developmental Disorders, 36, 395–​400.
Bennett, T., Szatmari, P., Bryson, S., Volden, J., Zwaigenbaum, L., Vaccarella, L., … & Boyle, M. (2008).
Differentiating autism and Asperger syndrome on the basis of language delay or impairment. Journal of
Autism and Developmental Disorders, 38(4), 616–​625.
Bishop-​Fitzpatrick, L., Minshew, N. J., Mazefsky, C. A., & Eack, S. M. (2016). Perception of Life as
Stressful, Not Biological Response to Stress, is Associated with Greater Social Disability in Adults with
Autism Spectrum Disorder, Journal of Autism and Developmental Disorders, Online first. doi:10.1007/​
s10803-​016-​2910-​6
Bird, G., & Cook, R. (2013). Mixed emotions: the contribution of alexithymia to the emotional symptoms
of autism. Translational Psychiatry, 3(7), e285.
Campbell-​Sills, L., & Barlow, D. H. (2007). Incorporating emotion regulation into conceptualizations and
treatments of anxiety and mood disorders. In J. Gross (Ed), Handbook of emotion regulation (pp. 542–​
559). New York, NY, US: Guilford Press.
Centers for Disease Control and Prevention (2007). Surveillance summaries. Morbidity & Mortality Weekly
Report, 56(SS-​1), 1–​28.
Centers for Disease Control and Prevention (2009). Prevalence of autism spectrum disorders: autism and
developmental disabilities monitoring network, United States, 2006. Morbidity & Mortality Weekly
Report Surveillance Summaries, 58(10), 1–​20.
Centers for Disease Control and Prevention (2014). Prevalence of autism spectrum disorders among
children aged 8 years: Autism and developmental disabilities monitoring network, 11 sites, United
States, 2010. Morbidity & Mortality Weekly Report Surveillance Summaries, 63(2), 1–​22.
Charman, T., Baird, G., Simonoff, E., Loucas, T., Chandler, S., Meldrum, D., & Pickles, A. (2007). Efficacy
of three screening instruments in the identification of autistic spectrum disorders. The British Journal of
Psychiatry, 191(6), 554–​559.
Chok, J. T., Demanche, J., Kennedy, A., & Studer, L. (2010). Utilizing physiological measures to facilitate
phobia treatment with individuals with autism and intellectual disability: A case study. Behavioral
Interventions, 25, 325–​337.
Church, B. A., Rice. C. L., Doygopoly, A., Lopata, C.J., Thomee, M. L., Nelson, A., & Mercado, E. III
(2015). Learning, plasticity, and atypical generalization in children with autism. Psychonomic Bulletin
and Review, 22(5), 1342–​1348. doi:10.3758/​s13423-​014-​0797-​9
Danial, J. T., & Wood, J. J. (2013). Cognitive behavioral therapy for children with autism: Review and
considerations for future research. Journal of Developmental and Behavioral Pediatrics, 34(9),
702–​715.
de Bruin, E. I., Blom, R., Smit, F. M., van Steensel, F. J., & Bögels, S. M. (2015). MYmind: Mindfulness
training for youngsters with autism spectrum disorders and their parents. Autism, 19(8), 906–​914.
DiGuiseppi, C., Hepburn, S., Davis, J. M., Fidler, D. J., Hartway, S., Lee, N. R., … & Robinson, C. (2010).
Screening for autism spectrum disorders in children with Down syndrome: Population prevalence and
screening test characteristics. Journal of Developmental and Behavioral Pediatrics, 31(3), 181–​191.
Dunn, W., & City, K. (1997). The impact of sensory processing abilities on the daily lives of young children
and their families: a conceptual model. Infants and Young Children, 9(4), 24–​35.
52

Conclusion 255

Gadow, K. D., DeVincent, C. J., & Pomeroy, J. (2006). ADHD symptom subtypes in children with pervasive
developmental disorder. Journal of Autism and Developmental Disorders, 36(2), 271–​283.
Granader, Y., Wallace, G. L., Hardy, K. K., Yerys, B. E., Lawson, R. A., Rosenthal, M., … & Kenworthy, L.
(2014). Characterizing the factor structure of parent reported executive function in autism spectrum
disorders: The impact of cognitive inflexibility. Journal of Autism and Developmental Disorders, 44(12),
3056–​3062.
Geurts, H. M., van den Bergh, S. F. W. M., & Ruzzano, L. (2014), Prepotent response inhibition and
interference control in autism spectrum disorders: Two meta-​analyses. Autism Research, 7, 407–​420.
doi: 10.1002/​aur.1369
Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. Handbook of Emotion
Regulation. Guilford Press, New York, NY.
Harms, M. B., Martin, A., & Wallace, G. L. (2010). Facial emotion recognition in autism spectrum
disorders: A review of behavioral and neuroimaging studies. Neuropsychology Review, 20(3), 290–​322.
Hebron, J., Humphrey, N., & Oldfield, J. (2015). Vulnerability to bullying of children with autism spectrum
conditions in mainstream education: a multi-​informant qualitative exploration. Journal of Research in
Special Educational Needs, 15(1), 185–​193. doi:10.1111/​1471-​3802.12108
Hill, E., Berthoz, S., & Frith, U. (2004) Brief report: Cognitive processing of own emotions in individuals
with autistic spectrum disorder and in their relatives. Journal of Autism and Developmental Disorders,
34, 229–​235.
Hus, V., Bishop, S., Gotham, K., Huerta, M. and Lord, C. (2013), Factors influencing scores on the
social responsiveness scale. Journal of Child Psychology and Psychiatry, 54, 216–​224. doi:10.1111/​
j.1469-​7610.2012.02589.x
Jahromi, L. B., Meek, S. E., & Ober-​Reynolds, S. (2012). Emotion regulation in the context of frustration
in children with high functioning autism and their typical peers. Journal of Child Psychology and
Psychiatry, 53, 1250–​1258. doi:10.1111/​j.1469-​7610.2012.02560.x
Jennett, H. K., & Hagopian, L. P. (2008). Identifying empirically supported treatments for phobic avoidance
in individuals with intellectual disabilities. Behavior Therapy, 39(2), 151–​161.
Kana, R. K., Patriquin, M. A., Black, B. S., Channell, M. M., & Wicker, B. (2015). Altered medial frontal
and superior temporal response to implicit processing of emotions in autism. Autism Research, 9(1)
55–​66. doi: 10.1002/​aur.1496
Kim, J. A., Szatmari, P., Bryson, S. E., Streiner, D. L., & Wilson, F. J. (2000). The prevalence of anxiety and
mood problems among children with autism and Asperger syndrome. Autism, 4(2), 117–​132.
King, M., & Bearman, P. (2009). Diagnostic change and the increased prevalence of autism. International
Journal of Epidemiology, 38(5), 1224–​1234.
Konstantareas, M. M., & Stewart, K. (2006). Affect regulation and temperament in children with autism
spectrum disorder. Journal of Autism and Developmental Disorders, 36, 143–​154. doi:10.1007/​
s10803-​005-​0051-​4
Lang, R., Mahoney, R., El Zein, F., Delaune, E., & Amidon, M. (2011). Evidence to practice: Treatment of
anxiety in individuals with autism spectrum disorders. Neuropsychiatric Disease and Treatment, 7, 27–​30.
Lartseva, A., Dijkstra, T., & Buitelaar, J. K. (2014). Emotional language processing in autism spectrum
disorders: A systematic review. Frontiers in Human Neuroscience, 8, 1–​24.
Lecavalier, L., Gadow, K. D., DeVincent, C. J., & Edwards, M. C. (2009). Validation of DSM-​IV
model of psychiatric syndromes in children with autism spectrum disorders. Journal of Autism and
Developmental Disorders, 39(2), 278–​289.
Leyfer, O. T., Folstein, S. E., Bacalman, S., Davis, N. O., Dinh, E., Morgan, J., … & Lainhart, J. E. (2006).
Comorbid psychiatric disorders in children with autism: Interview development and rates of disorders.
Journal of Autism and Developmental Disorders, 36(7), 849–​861.
Little, L. M., Ausderau, K., Sideris, J., & Baranek, G. T. (2015). Activity participation and sensory features
among children with autism spectrum disorders. Journal of Autism and Developmental Disorders, 45(9),
2981–​2990. doi: 10.1007/​s10803015-​2460-​3
6
5
2

256 Emotion Regulation and Autism Spectrum Disorder

Lord, C., & Bishop, S. L. (2015). Recent advances in autism research as reflected in DSM-​5 criteria for
autism spectrum disorder. Annual Review of Clinical Psychology, 11, 53–​70.
Lord, C., Petkova, E., Hus, V., Gan, W., Lu, F., Martin, D. M., … & Risi, S. (2012). A multisite study of the
clinical diagnosis of different autism spectrum disorders. Archives of General Psychiatry, 69(3), 306–​313.
Losh, M., & Capps, L. (2006). Understanding of emotional experience in autism: Insights from the personal
accounts of high-​functioning children with autism. Developmental Psychology, 42(5), 809–​818.
Lydon, S., Healy, O., O’Callaghan, O., Mulhern, T., & Holloway, J. (2014). A systematic review of the
treatment of fears and phobias among children with autism spectrum disorders. Review Journal of
Autism and Developmental Disorders, 2(2), 141–​154.
Lydon, S., Healy, O., Reed, P., Mulhern, T., Hughes, B. M., & Goodwin, M. S. (2014): A systematic review
of physiological reactivity to stimuli in autism. Developmental Neurorehabilitation, 1(21) 335–​355.
doi: 10.3109/​17518423.2014.971975
Mandell, D., & Lecavalier, L. (2014). Should we believe the Centers for Disease Control and Prevention’s
autism spectrum disorder prevalence estimates? Autism, 18(5), 482–​484.
Marcus, R. N., Owen, R., Manos, G., Mankoski, R., Kamen, L., McQuade, R. D., … & Findling, R.
L. (2011). Safety and tolerability of aripiprazole for irritability in pediatric patients with autistic
disorder: A 52-​week, open-​label, multicenter study. Journal of Clinical Psychiatry, 72(9), 1270–​1276.
Mazefsky, C. A., Borue, X., Day, T. N., & Minshew, N. J. (2014). Emotion regulation patterns in adolescents
with high-​functioning autism spectrum disorder: Comparison to typically developing adolescents and
association with psychiatric symptoms, Autism Research, 7(3), 344–​354.
Mazefsky, C. A., Herrington, J., Siegel, M., Scarpa, A., Maddox, B. B., Scahill, L., & White, S. W. (2013).
The role of emotion regulation in autism spectrum disorder. Journal of the American Academy of Child
& Adolescent Psychiatry, 52(7), 679–​688.
Mazefsky, C. A., Pelphrey, K. A., & Dahl, R. E. (2012). The need for a broader approach to emotion
regulation research in autism. Child Development Perspectives, 6(1), 92–​97.
Mazefsky, C. A., & White, S. W. (2014). Emotion regulation: Concepts & practice in autism spectrum
disorder. Child and Adolescent Psychiatric Clinics of North America, 23(1), 15–​24.
McCracken, J. T., McGough, J., Shah, B., Cronin, P., Hong, D., Aman, M. G., … & McMahon, D. (2002).
Risperidone in children with autism and serious behavioral problems. New England Journal of Medicine,
347(5), 314–​321.
Mullins, J. L., & Christian, L. (2001). The effects of progressive relaxation training on the disruptive
behavior of a boy with autism. Research in Developmental Disabilities, 22(6), 449–​462.
Odom, S. L., Thompson, J. L., Hedges, S., Boyd, B. A., Dykstra, J. R., Duda, M. A., … & Bord, A. (2014).
Technology-​aided interventions and instruction for adolescents with autism spectrum disorder. Journal
of Autism and Developmental Disorders, 1–​15. doi:10.1007/​s10803-​014-​2320-​6
Pitskel, N. B., Bolling, D. Z., Kaiser, M. D., Pelphrey, K. A., & Crowley, M. J. (2014). Neural systems
for cognitive reappraisal in children and adolescents with autism spectrum disorder. Developmental
Cognitive Neuroscience, 10, 117–​128.
Pouw, L. B., Rieffe, C., Stockmann, L., & Gadow, K. D. (2013). The link between emotion regulation, social
functioning, and depression in boys with ASD. Research in Autism Spectrum Disorders, 7(4), 549–​556.
Richey, J. A., Damiano, C. R., Sabatino, A., Rittenberg, A., Petty, C., Bizzell, J., … & Dichter, G. S.
(2015). Neural mechanisms of emotion regulation in autism spectrum disorder. Journal of Autism and
Developmental Disorders, 45(11), 1–​15.
Rieffe, C., Meerum Terwogt, M., & Kotronopoulou, K. (2007). Awareness of single and multiple
emotions in high-​functioning children with autism. Journal of Autism and Developmental Disorders, 37,
455–​465.
Rowley, E., Chandler, S., Baird, G., Simonoff, E., Pickles, A., Loucas, T., & Charman, T. (2012).
The experience of friendship, victimization and bullying in children with an autism spectrum
disorder: Associations with child characteristics and school placement. Research in Autism Spectrum
Disorders, 6(3), 1126–​1134. doi:10.1016/​j.rasd.2012.03.004
7
5
2

Conclusion 257

Samson, A. C., Hardan, A. Y., Podell, R. W., Phillips, J. M., & Gross, J. J. (2015). Emotion regulation in
children and adolescents with autism spectrum disorder. Autism Research, 8(1), 9–​18.
Samson, A. C., Huber, O., & Gross, J. J. (2012). Emotion regulation in Asperger’s syndrome and high-​
functioning autism. Emotion, 12(4), 659.
Samson, A. C., Phillips, J. M., Parker, K. J., Shah, S., Gross, J. J., & Hardan, A. Y. (2014). Emotion
dysregulation and the core features of autism spectrum disorder. Journal of Autism and Developmental
Disorders, 44(7), 1766–​1772.
Scarpa, A., & Reyes, N. M. (2011). Improving emotion regulation with CBT in young children with high
functioning autism spectrum disorders: A pilot study. Behavioral and Cognitive Psychotherapy, 39(4),
495–​500. doi: dx.doi.org/​10.1017/​S1352465811000063
Schmidt, J. D., Luiselli, J. K., Rue, H., & Whalley, K. (2013). Graduated exposure and positive
reinforcement to overcome setting and activity avoidance in an adolescent with autism. Behavior
Modification, 37, 128–​142.
Schoen, S. A., Miller, L. J., Brett-​Green, B., & Hepburn, S. L. (2008). Psychophysiology of children with
autism spectrum disorder. Research in Autism Spectrum Disorders, 2(3), 417–​429.
Schultz, D., Izard, C. E., & Bear, G. (2004). Children’s emotion processing: Relations to emotionality and
aggression. Development and Psychopathology, 16(2), 371–​388.
Selles, R. R., Arnold, E. B., Phares, V., Lewin, A. B., Murphy, T. K., & Storch, E. A. (2015). Cognitive-​
behavioral therapy for anxiety in youth with an autism spectrum disorder: A follow-​up study. Autism,
19(5), 613–​621.
Shabani, D. B., & Fisher, W. W. (2006). Stimulus fading and differential reinforcement for the treatment of
needle phobia in a youth with autism. Journal of Applied Behavior Analysis, 39(4), 449–​452.
Siegel, M., & Beaulieu, A. A. (2012). Psychotropic medications in children with autism spectrum
disorders: A systematic review and synthesis for evidence-​based practice. Journal of Autism and
Developmental Disorders, 42(8), 1592–​1605.
Silani, G., Bird, G., Brindley, R., Singer, T., Frith, C., & Frith, U. (2008). Levels of emotional awareness and
autism: An fMRI study. Social Neuroscience, 3(2), 97–​112.
Simonoff, E., Pickles, A., Charman, T., Chandler, S., Loucas, T., & Baird, G. (2008). Psychiatric disorders in
children with autism spectrum disorders: Prevalence, comorbidity, and associated factors in a population-​
derived sample. Journal of the American Academy of Child & Adolescent Psychiatry, 47(8), 921–​929.
Singh, N. N., Lancioni, G. E., Manikam, R., Winton, A. S., Singh, A. N., Singh, J., & Singh, A. D. (2011).
A mindfulness-​based strategy for self-​management of aggressive behavior in adolescents with autism.
Research in Autism Spectrum Disorders, 5(3), 1153–​1158.
Singh, N. N., Lancioni, G. E., Winton, A. S., Fisher, B. C., Wahler, R. G., Mcaleavey, K., … & Sabaawi,
M. (2006). Mindful parenting decreases aggression, noncompliance, and self-​injury in children with
autism. Journal of Emotional and Behavioral Disorders, 14(3), 169–​177.
Smith, S. M., & Vale, W. W. (2006). The role of the hypothalamic-​pituitary-​adrenal axis in neuroendocrine
responses to stress. Dialogues in Clinical Neuroscience, 8(4), 383–​395.
Sofronoff, K., Attwood, T., Hinton, S., & Levin, I. (2007). A randomized controlled trial of a cognitive
behavioural intervention for anger management in children diagnosed with Asperger syndrome.
Journal of Autism and Developmental Disorders, 37(7), 1203–​1214.
Spain, D., Sin, J., Chalder, T., Murphy, D., & Happe, F. (2015). Cognitive behaviour therapy for adults
with autism spectrum disorders and psychiatric co-​morbidity: A review. Research in Autism Spectrum
Disorders, 9, 151–​162.
Spek, A. A., van Ham, N. C., & Nyklíček, I. (2013). Mindfulness-​based therapy in adults with an autism
spectrum disorder: A randomized controlled trial. Research in Developmental Disabilities, 34(1),
246–​253.
Spencer, D., Marshall, J., Post, B., Kulakodlu, M., Newschaffer, C., Dennen, T., … & Jain, A. (2013).
Psychotropic medication use and polypharmacy in children with autism spectrum disorders. Pediatrics,
132(5), 833–​840.
8
5
2

258 Emotion Regulation and Autism Spectrum Disorder

Sukhodolsky, D. G., Bloch, M. H., Panza, K. E., & Reichow, B. (2013). Cognitive behavioral therapy for
anxiety in children with high-​functioning autism: A meta analysis. Pediatrics, 132(5), e1341–​e1350.
Thompson, R. A. (1994). Emotion regulation: A theme in search of definition. Monographs for the Society
for Research in Child Development, 59, 25–​52.
Thomson, K., Burnham Riosa, P., & Weiss, J. A. (2015). Brief report of preliminary outcomes of an
emotion regulation intervention for children with autism spectrum disorder. Journal of Autism and
Developmental Disorders, 45, 3487–​3495.
Totsika, V., Hastings, R. P., Emerson, E., Berridge, D. M., & Lancaster, G. A. (2011a). Behavior problems
at 5 years of age and maternal mental health in autism and intellectual disability. Journal of Abnormal
Child Psychology, 39(8), 1137–​1147.
Totsika, V., Hastings, R. P., Emerson, E., Lancaster, G. A. and Berridge, D. M. (2011b), A population-​
based investigation of behavioural and emotional problems and maternal mental health: Associations
with autism spectrum disorder and intellectual disability. Journal of Child Psychology and Psychiatry,
52(1), 91–​99. doi:10.1111/​j.1469-​7610.2010.02295.x
Turner, M. (1999). Annotation: Repetitive behaviour in autism: A review of psychological research. Journal
of Child Psychology and Psychiatry, 40(06), 839–​849.
Uljarevic, M., & Hamilton, A. (2013). Recognition of emotions in autism: A formal meta-​analysis. Journal
of Autism and Developmental Disorders, 43(7), 1517–​1526.
Ung, D., Selles, R., Small, B. J., & Storch, E. A. (2015). A systematic review and meta analysis of cognitive-​
behavioral therapy for anxiety in youth with high functioning autism spectrum disorders. Child
Psychiatry & Human Development, 1–​15.
Vasa, R. A., Carroll, L. M., Nozzolillo, A. A., Mahajan, R., Mazurek, M. O., Bennett, A. E., … & Bernal,
M. P. (2014). A systematic review of treatments for anxiety in youth with autism spectrum disorders.
Journal of Autism and Developmental Disorders, 44(12), 3215–​3229.
Weisbrot, D. M., Gadow, K. D., DeVincent, C. J., & Pomeroy, J. (2005). The presentation of anxiety in
children with pervasive developmental disorders. Journal of Child & Adolescent Psychopharmacology,
15(3), 477–​496.
Weiss, J. A. (2014). Transdiagnostic case conceptualization of emotional problems in youth with ASD: An
emotion regulation approach. Clinical Psychology: Science and Practice, 21(4), 331–​350.
Weitlauf, A. S., Gotham, K. O., Vehorn, A. C., & Warren, Z. E. (2014). Brief report: DSM-​5 “levels of
support:” a comment on discrepant conceptualizations of severity in ASD. Journal of Autism and
Developmental Disorders, 44(2), 471–​476.
Williams, D. L., Mazefsky, C. A., Walker, J. D., Minshew, N. J., & Goldstein, G. (2014). Associations
between conceptual reasoning, problem solving, and adaptive ability in high-​functioning autism.
Journal of Autism and Developmental Disorders, 44(11), 2908–​2920.
Wood, J. J., & Gadow, K. D. (2010). Exploring the nature and function of anxiety in youth with autism
spectrum disorders. Clinical Psychology: Science and Practice, 17(4), 281–​292.
Woodbury-​Smith, M., Klin, A., & Volkmar, F. (2005). Asperger’s syndrome: A comparison of clinical
diagnoses and those made according to the ICD-​10 and DSM-​IV. Journal of Autism and Developmental
Disorders, 35(2), 235–​240.
Wright, B., Clarke, N., Jordan, J. O., Young, A. W., Clarke, P., Miles, J., … & Williams, C. (2008). Emotion
recognition in faces and the use of visual context in young people with high-​functioning autism
spectrum disorders. Autism, 12(6), 607–​626.
9
5
2

Chapter 13

Emotion Dysregulation in Adolescents


with Borderline Personality Disorder
Carla Sharp & Timothy J. Trull

Borderline personality disorder


The achievement of independent emotion regulation without external regulation by caregivers
is considered a crucial milestone of adolescence (Steinberg et al., 2006). However, adolescence is
characterized by dramatic social-╉emotional developmental changes that include the creation of
an independent social network of stable friendships, the development of romantic relationships
while not neglecting maintaining closeness to family members, and honing the capacities required
for education and work tasks (Allen et al., 2006; Roisman, Masten, Coatsworth, & Tellegen, 2004).
These social demands coincide with significant functional and structural brain changes in brain
areas highly relevant to emotion regulation (Hare et al., 2008; Monk et al., 2003). For instance,
Monk et al. (2003) showed that the emotional content of social stimuli appeared to drive activa-
tion more strongly in the amygdala in adolescents than in adults. In other words, social-╉emotional
developmental tasks become salient at exactly the same time adolescents become more strongly
motivated by emotional content, thereby placing adolescents at risk for disorders of emotion reg-
ulation. One such disorder, which is often described as the quintessential disorder of emotion
regulation, is Borderline Personality Disorder (BPD).
The aim of this chapter is to shed light on emotion dysregulation in BPD in adolescents in terms
of its understanding and treatment. We begin by highlighting the main characteristics of BPD in
adolescents. Next, we discuss developmental theories of BPD that place emotion dysregulation
at its center. We then discuss definitional ambiguities for the constructs of emotion, emotion
regulation and emotion dysregulation to justify the use of a multi-╉component model of emotion
dysregulation in BPD (Carpenter & Trull, 2013). We then review the empirical literature on each
component of this model (emotional sensitivity, intense negative affect, inadequate emotion regu-
lation strategies and maladaptive regulation strategies), completing the chapter with a review of
the latest evidence-╉based intervention approaches for BPD, and providing an in-╉depth example of
one of these treatment approaches.

Borderline personality disorder in adolescents


DSM-╉5 criteria for BPD
In contrast to other disorders described in the fifth edition of the Diagnostic and Statistical
Manual of Psychiatric Disorders (American Psychiatric Association, 2013), the personality disor-
ders (PDs), which include BPD, are categorized in two sections. BPD is first categorized in Section
II of the DSM-╉5, alongside nine other personality disorders. Table 13.1 summarizes the diagnostic
criteria as defined in Section II.
0
6
2

260 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

Table 13.1╇ Diagnostic criteria of BPD as defined by the DSM-╉5 (Section II)

A pervasive pattern of instability of interpersonal relationships, self-╉image, and affects, and marked
impulsivity, beginning by early adulthood and present in a variety of context, as indicated by five (or
more) of the following:
1. Frantic efforts to avoid real or imagined abandonment. (Note: Do not include suicidal or
selfmutiliating behavior covered in Criterion 5).
2. A pattern of unstable and intense interpersonal relationships characterized by alternating between
extremes of idealization and devaluation.
3. Identity disturbance: markedly and persistently unstable self image or sense of self.
4. Impulsivity in at least two areas that are potentially self-╉damaging (e.g. spending, sex, substance
abuse, reckless driving, binge eating: (Note: Do not include suicidal or self mutilating behavior
covered in Criterion 5.)
5. Recurrent suicidal behavior, gestures, or threats, or self-╉mutilating behavior.
6. Affective instability die to a marked reactivity of mood (e.g., intense episodic dysphoria, irritability, or
anxiety usually lasting a few hours and only rarely more than a few days).
7. Chronic feelings of emptiness.
8. Inappropriate, intense anger or difficulty controlling anger (e.g., frequent displays of temper,
constant anger, recurrent physical fights).
9. Transient, stress-╉related paranoid ideation or severe dissociative symptoms.

While Section II represents the traditional categorical perspective that PDs are qualitatively
distinct clinical syndromes, an alternative to the categorical approach is articulated in Section III
of the DSM-╉5. Section III contains “Conditions for Further Study” and includes those for which
scientific evidence was deemed unavailable to support widespread clinical use. PDs are concep-
tualized in Section III from a dimensional perspective, such that PDs represent maladaptive vari-
ances of personality traits that lie on a continuum from normal to abnormal. DSM-╉5 Section III
requires clinicians to consider two sets of criteria (Criteria A and B) in the assessment of BPD.
Criterion A requires judgment of the severity of problems in identity, self-╉direction, empathy, and
intimacy. Criterion B is used to situate an individual within a dimensional personality disorder
space, such that an individual’s functioning may be profiled across five PD trait domains (negative
affectivity, detachment, antagonism, disinhibition and psychoticism) and 25 PD facets (emotional
lability, anxiousness, separation insecurity, submissiveness, hostility, perseveration, depressivity,
suspiciousness, restricted affectivity, withdrawal, intimacy avoidance, anhedonia, manipulative-
ness, deceitfulness, grandiosity, attention-╉seeking, callousness, irresponsibility, impulsivity, dis-
tractibility, risk-╉taking, rigid perfectionism, unusual beliefs/╉experiences, eccentricity, cognitive/╉
perceptual dysregulation). The typical trait profile suggested for BPD requires moderate to greater
impairment in personality functioning manifested by difficulties in two of Criterion A features
(poorly developed identity, problems in self-╉direction, compromised empathy and interpersonal
hypersensitivity, and problems in intimacy). In addition, for Criterion B, high ratings on four
or more of the following seven pathological personality traits are required:  Emotional lability,
anxiousness, separation insecurity, depressivity, impulsivity, risk-╉taking, and hostility; of which at
least one must be impulsivity, risk taking, or hostility.

Applying DSM criteria to adolescents


There has been longstanding, general consensus (as early as the DSM-╉II; American Psychiatric
Association, 1968) that BPD symptoms usually first become apparent in adolescence (Chanen &
Kaess, 2012a; Shiner, 2009). Accordingly, Section II of the DSM-╉5 states that the diagnosis of BPD
may be applied to children or adolescents when “the individual’s particular maladaptive personal-
ity traits appear to be pervasive, persistent, and unlikely to be limited to a particular developmental
1
6
2

Borderline personality disorder in adolescents 261

stage or another mental disorder” (American Psychiatric Association, 2013, p. 647). In contrast
to the two years necessary for an adult PD to be diagnosed, only one year is necessary for child/╉
adolescent PD. Section III of the DSM-╉5 states that impairments in personality function are stable
over time and onset can be traced back to “at least adolescence or early adulthood” (American
Psychiatric Association, 2013, p. 762). In addition, the trait facets incorporated in Section III of
the DSM-╉5 mirror developmental findings from maladaptive personality trait frameworks (e.g.,
De Clercq, De Fruyt, et al., 2014; De Clercq, Decuyper, & De Caluwé, 2014). The ICD-╉11, and
national treatment guidelines for the U.K. (National Institute for Health and Clinical Excellence,
2009) and Australia (National Health and Medical Research Council, 2013) also “legitimize” the
diagnosis of BPD in adolescence. In all, current official classification systems for mental disorders
support the evaluation and diagnosis of BPD in adolescents.
Despite these advances, there has been a general reluctance among clinicians to diagnose
BPD in adolescence due to fear of stigma and concerns that personality is not stable in adoles-
cence. The five-╉fold increase in research on BPD in adolescents over the last ten years (Sharp
& Tackett, 2014) is addressing some of these fears, and researchers agree that BPD is a valid
and reliable diagnosis in adolescents (Chanen & Kaess, 2012b; Chanen & McCutcheon, 2013a;
Miller, Muehlenkamp, & Jacobson, 2008; Sharp & Fonagy, 2015; Sharp & Kalpakci, 2015; Sharp
& Kim, 2015; Stepp, 2012). For example, the extant literature supports the construct of adoles-
cent BPD in terms of its clinical description (Chanen & Kaess, 2012b; Fossati, 2014), correlates
and causes (e.g. Carlson, Egeland, & Sroufe, 2009; Sharp et al., 2011), studies that delimitate the
disorder from other related syndromes (e.g. Chanen, Jovev, & Jackson, 2007), follow-╉up studies
that demonstrate a prototypical course and outcome of the symptoms (e.g. Bornovalova, Hicks,
Iacono, & McGue, 2009; Chanen et al., 2004; Cohen et al., 2008), and twin/╉family studies that
aim to identify a genetic basis of the biological phenomena associated with adolescent BPD
(Distel et al., 2008).

Summary
BPD is a valid and reliable disorder in adolescence and can be assessed categorically using DSM-╉
5 Section II, and dimensionally using DSM-╉5 Section III. Prevalence studies have shown that
adolescent BPD occurs at rates around 1% (Michonski, Sharp, Steinberg, & Zanarini, 2013) to 3%
(Zanarini et al., 2011; Johnson, Cohen, Kasen, Skodol, & Oldham, 2008) in community samples.
In clinical samples, rates are 11% in outpatients (Chanen et al., 2004), 33% (Ha, Balderas, Zanarini,
Oldham, & Sharp, in press) and 43–╉49% in inpatients (Levy et al., 1999). BPD is, therefore, not
a transient condition of adolescence, but should be diagnosed and treated to prevent youngsters
from a lifelong trajectory of increasingly severe psychopathology.

Emotion dysregulation and BPD: Developmental theories


As evident in Table 13.1, BPD is characterized by dysregulation in several functional domains
including behavior, identity, interpersonal relationships, cognition and emotion (Carpenter &
Trull, 2013). Of these, dysregulation in emotion is identified as a central mechanism in all develop-
mental theories of BPD. Linehan’s biosocial theory (Linehan, 1993) and recent extensions thereof
(Crowell, Beauchaine, & Linehan, 2009; Selby, Kranzler, & Panza, 2014), most prominently places
emotion dysregulation at the core of borderline pathology and describe dysregulation in other
domains as secondary. Linehan (1993) posits that children with a genetically-╉based, emotion-
ally sensitive, and reactive temperament are at an elevated risk of developing BPD when reared
in an invalidating (e.g., neglectful, abusive, and/╉or dismissive) early family environment. These
children are suggested to experience higher levels of negative affect across contexts and situations,
but instead of being matched with a family environment that scaffolds the child in regulating
these intense, negative emotions, the environment fails to impart adequate emotion management
2
6

262 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

skills, so the individual often resorts to short-╉term avoidance strategies (e.g., self-╉harm, impulsive
behavior, etc.) when experiencing unpleasant internal states.
Crowell and colleagues recently extended Linehan’s biosocial theory to develop a developmen-
tal psychopathology model with a specific focus on trait impulsivity. In their model, biologically
determined negative affectivity and high emotional sensitivity interact with trait impulsivity
and parental factors (invalidation and ineffective parenting) early on in development to confer
increased risk for impulse control deficits as development progresses. Impulse control deficits are
then further reinforced by the same parenting factors over time, culminating in BPD.
In another extension of Linehan’s biosocial theory, Selby and colleagues suggest, in their
Emotional Cascade Model (Selby & Joiner, 2009), that rumination (and catastrophization as
future-╉oriented rumination) potentiates the magnitude of biologically determined negative affect
which, in sequence, amplifies the level of rumination, initiating a vicious, self-╉perpetuating cas-
cade of negative emotions. This biologically determined deficit in emotional functioning interacts
with an invalidating family environment over time, from which a full syndrome of BPD emerges.
Fonagy and colleagues’ developmental theory of BPD (Fonagy, Gergely, Jurist, & Target, 2002;
Fonagy & Luyten, 2009; Sharp & Fonagy, 2008) describes emotion dysregulation as a core interac-
tive component with developing mentalizing capacity which, in the context of disrupted attach-
ment relationships, may foster poor self-╉other differentiation culminating in BPD over time.
This model of BPD is firmly rooted in the developmental psychology of emotion regulation.
Specifically, secure attachment and mentalizing capacity evolve when a parent communicates
contingent, marked, and ostensive cues to an infant/╉child. The brains of infants are presumed to
be hard-╉wired to preferentially attend to these cues (Fonagy et al., 2002; Kim, in press). Marked
communication (Fonagy, Gergely, & Target, 2007) refers to communication where a parent
understands the infant’s internal state, while concurrently signaling that the parent’s expression
of emotion concerns the infant, not the parent him/╉herself. The expression of emotion is marked
by modifying (e.g., exaggerating or slowing down) the display of the child’s affect, such that the
parent’s emotional expression resembles, but also modulates the child’s emotion simultaneously.
Ostensive communicative cues (Csibra, 2010; Gergely Csibra & Gergely, 2011) refer to the process
of calling attention to what the parent is about to communicate, for instance by making direct eye
contact with the child while calling the child by name, and/╉or speaking with a “motherese” into-
nation. These ostensive cues signal to the child that the parent’s emotion expression concerns the
child and is of importance. In all then, it is not essential that the adult is perfectly accurate every
time that he/╉she guesses what might be going on in the mind of the child; the point is that the
adult is genuinely interested in the child´s mind and this enables the child to develop a separate
sense of self and adequate emotion regulation capacity (Sadler et al., 2006). If these developmental
processes fail, a child is at risk for developing inadequate mentalizing, emotion dysregulation and
disturbed self-╉other processing—╉in short, BPD.

Summary
All developmental theories of BPD include a strong focus on emotion dysregulation. Moreover, all
theories emphasize the family and/╉or attachment context as highly relevant to the child’s develop-
ing emotion regulation capacity. Theories furthermore converge to suggest a reciprocal relation
between emotion dysregulation and other domains of functioning.

Emotion dysregulation and BPD


While the presence of emotion dysregulation in BPD is undeniable, research in the field contin-
ues to be plagued by definitional ambiguities (Bloch, Moran, & Kring, 2010; Carpenter & Trull,
2013). It is essential for researchers to commit to a particular working definition in this regard, as
3
6
2

Borderline personality disorder in adolescents 263

neglecting to do so may lead to a vague and non-​specific discussion of problems in emotion regu-
lation related to specific disorders. Given the developmental nature of this volume, and that this
chapter is concerned with emotion dysregulation as it relates to BPD, we will draw on literature
in developmental psychology, developmental psychopathology, and the psychology of emotion to
set the definitional parameters for the remaining discussion.
While several definitions of emotion have been put forward, they all converge on the central
idea that emotions have evolutionary utility—​that is, they prepare us biologically to appraise and
respond rapidly and flexibly to situations in service of our survival. Consistent with this view,
emotions are defined in the developmental psychology literature as “appraisal-​action readiness
stances, a fluid and complex progression of orienting toward the ongoing stream of experience”
(Cole, Martin, & Dennis, 2004, p.  320). In this sense, emotions are always context-​dependent,
although this context may include both the external and internal world. From a developmental
psychology perspective, and highly relevant to BPD, the family environment and the caregiving
relationship is regarded as the most relevant context influencing young children’s emotions. As
children mature into adolescence, peer-​and romantic-​relationships emerge as additional contexts
of high emotional salience (Cole et al., 2004). While most psychiatric disorders are characterized
by disturbances in emotions, emotions in and of themselves are not pathological. However, too
much or too little emotion may be indicative of pathology; we will return to how this bears on the
definition of emotion dysregulation later in this section.
Emotion regulation is defined in developmental psychology as either regulating or regulated
(Cole et  al., 2004). Emotion as regulating refers to instances during which emotion regulates
another system or another person (e.g., a child’s sadness leads to a mother picking her up). Emotion
as regulated refers to a change in a particular emotion (e.g., a mother picking up a child makes
the child calm down). For the purposes of the current chapter, we will focus solely on emotion as
regulated. Consistent with the concept of emotion as regulated, is Gross’s (Gross, 1988) definition
of emotion regulation as the processes by which individuals (or context) influence the type, tim-
ing, experience and expression of emotion. This definition was expanded by Gross and Thompson
(2007) to define emotion regulation as the automatic or controlled, conscious or unconscious pro-
cesses by which emotions in self and/​or others are influenced. Gross (1998) defines five regulat-
ing “processes” which include situation selection, situation modification, attentional deployment,
cognitive change, and response modulation. This definition implies that in order to effectively
regulate one’s emotions, one has to have available a set of skills to adopt in a particular situation
(e.g. when upset, an individual may know that turning to a loved one for support would help calm
him/​her down), as well as implementing the skill appropriately (e.g., suppressing one’s anger when
expressing it would lead to negative consequences).
Keeping in mind the definitions of emotion and emotion regulation discussed above, emotion
dysregulation denotes instances where emotion regulation processes are derailed, or, put differ-
ently, the inability to flexibly enhance or suppress emotional expression in accord with situational
demands (Bloch et al., 2010; Bonanno, Papa, Lalande, Westphal, & Coifman, 2004). Because too
much or too little emotion may be indicative of pathology, emotion dysregulation, within a devel-
opmental psychopathology framework, includes not only problems in emotion regulation, but
also affect dysfunction (Cicchetti, Ackerman, & Izard, 1995). The inclusion of affect dysfunction
in the definition of emotion dysregulation has meant that a wide variety of constructs are stud-
ied under the umbrella of emotion dysregulation in BPD (Carpenter & Trull, 2013) including
emotional sensitivity, emotion reactivity, affectivity lability, prolonged emotional responses and
emotional intensity, to name a few. For conceptual clarity, it has been suggested that emotion dys-
regulation be viewed as a process, consisting of many interactive components, rather than an end-​
state (Werner & Gross, 2010). To this end, Carpenter and Trull have developed a multi-​component
4
6
2

264 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

Emotion
sensitivity

Stimulus

Heightened and
labile negative
affect

Inadequate Maladaptive
appropriate regulation
regulation strategies strategies

Emotion dysregulation
consequences

Figure 13.1╇ Multi-╉component model of emotion dysregulation in BPD


Note. Multi-╉component model of emotion dysregulation in BPD (Carpenter and Trull, 2013) Individuals with BPD are
theorized to be sensitive to emotional stimuli from birth. Experiencing a negatively valenced stimulus (or interpreting a
stimulus in a negative way) in the environment leads to increases in negative affect and affective instability. Heightened
and unstable negative affect both makes it difficult to learn and to employ appropriate emotion regulation strategies and
leads to an increase in maladaptive and impulsive regulation strategies. Emotion dysregulation consequences occur as a
result, which, in turn, reinforce vigilance toward negatively valenced stimuli in the environment.
Reproduced from Current Psychiatry Reports, Components of Emotion Dysregulation in Borderline Personality
Disorder: A Review, 15 (1), p. 335, doi:10.1007/╉s11920-╉012-╉0335-╉2, Ryan W. Carpenter, Timothy J. Trull, Copyright ©
2012, Springer Science + Business Media New York. With permission of Springer.

model of emotion dysregulation relevant to BPD (see Figure 13.1). In this model, the experience
or subjective perception of a negatively-╉valenced stimulus in the environment leads to increases in
negative affect and affective instability. Heightened and unstable negative affect, in turn, impedes
the use of appropriate and effective regulation strategies, instead leading to increases in the use of
maladaptive strategies. The emotion dysregulation consequences that occur, as a result, reinforce
emotion sensitivity for negatively-╉valenced stimuli in the environment that maintains a vicious,
self-╉perpetuating cycle from which a full syndrome of BPD emerges. While this model is unmis-
takably rooted within Linehan’s (1993) biosocial theory of BPD, here, we infuse the model with
ideas from Fonagy and co-╉worker’s attachment-╉based theory of BPD, to suggest that it is most
often attachment-╉and relationship-╉based events that will provide the most evocative stimuli for
the initiation of this multi-╉component process of emotion dysregulation.

Adolescent BPD: Empirical evidence


Based on the depth, reach, and influence of developmental theories of BPD that place emotion
dysregulation at its center, the lack of prospective, longitudinal research to test the interactive
5
6
2

Borderline personality disorder in adolescents 265

and causal effects of affective dysfunction and emotion regulation problems on the development
of BPD is surprising (Matusiewicz, Weaverling, & Lejeuz, 2014). While this research is lacking,
there is a growing body of cross-╉sectional research focused on emotion dysregulation among
adolescents with BPD. While the latter does not provide a test of the causal relations inherent in
developmental models, it does provide an important starting point that can guide future work
(Matusiewicz et al., 2014). In this section, we use the Carpenter and Trull (2013) model of emo-
tion dysregulation in BPD to organize the empirical literature for each component of the model.

Emotional sensitivity
Emotional sensitivity is defined as heightened emotional reactivity to social and non-╉social stim-
uli (Carpenter & Trull, 2013); or, emotional sensitivity may be defined as the tendency to have
emotional responses to low-╉intensity stimuli (Matusiewicz et al., 2014). While findings are mixed,
several studies have demonstrated heightened emotional sensitivity in adults (see Carpenter &
Trull, 2013; Daros, Zakzanis, & Ruocco, 2013 for a review). Specifically, Daros et al. (2013) con-
cluded, in a recent meta-╉analytic review, that patients with BPD have a sensitivity for rejection-╉
related stimuli (captured in facial expressions of anger and disgust), which interferes with their
capacity to adequately regulate their emotions.
Two studies have investigated emotional sensitivity in adolescents, operationalized as atten-
tional bias to emotional stimuli. Jovev et al (2012) used a modified dot probe task in 21 subjects
between the ages of 15–╉24, who met three or more criteria of BPD, compared to 20 healthy con-
trols. The aim of the task was to assess whether emotion cues in facial stimuli interfered with a
simple discrimination task. Results showed that youth with borderline features had an attentional
bias for fearful faces that reflected difficulty in disengaging attention from threatening informa-
tion during the preconscious stages of attention. Similarly, Von Ceumern-╉Lindenstjerna et  al.
(2010) demonstrated a correlation between current mood and attentional bias to negative faces,
suggesting an inability to disengage attention from negative facial expressions during attentional
maintenance when in negative mood. Together, these findings suggest a diminished capacity for
affect regulation in the presence of negatively-╉valenced social stimuli.
Another way to operationalize emotional sensitivity is to evaluate whether individuals with
BPD accurately identify emotional expressions at earlier stages of expression (i.e., lower thresh-
olds of facial expressivity across all emotional valences). Findings in adolescents, like those in
adults, are mixed. Jovev et al. (2011) used a facial morphing task in which faces morph from neu-
tral to each of the six basic emotional expressions. No evidence of heightened sensitivity to emo-
tional facial expressions was found in the BPD group compared to the community control group.
Using a similar face morphing task, Robin et al. (2012) demonstrated that adolescents with BPD
were less sensitive to facial expressions of anger and happiness, i.e., they required more intense
facial expressions than control participants to correctly identify these two emotions. However,
they did not exhibit any deficit in recognizing fully expressed emotions.
A third way to operationalize emotional sensitivity is through evaluating the valence and inten-
sity of emotional reactions to aversive social or interpersonal events, in particular situations dur-
ing which there are perceived or real rejection and/╉or invalidation. While several social rejection/╉
invalidation studies of BPD have been conducted in college-╉age young adults (Ruocco et al., 2010;
Tragesser, Lippman, Trull, & Barrett, 2008; Woodberry, Gallo, & Nock, 2008), only one study has
included adolescents (Lawrence, Chanen, & Allen, 2011). This study examined the effect of social
exclusion, with the use of a Cyberball task, upon mood in a sample of young people (aged 15–╉24)
presenting for treatment early in the course of BPD, as compared with a healthy control group.
Cyberball is an experimental task designed to assess the effects upon mood of being excluded or
ignored without explanation in a social context (Williams & Jarvis, 2006). Results showed that
62

266 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

ostracism did not selectively induce negative mood in adolescents with BPD; nor did borderline
adolescents show more difficulty in regulating their mood back to baseline; however, the BPD
group rated their mood as more intense across all mood states and across time compared to the
control group.
While more research is clearly necessary to further examine emotional sensitivity in adoles-
cents with BPD, two conclusions can be drawn from the emotional sensitivity literature thus
far. First, there seems to be preliminary evidence in support of emotional sensitivity among this
group in the form of a “negativity bias” manifested as hyper-​responsiveness (hypersensitivity) to
negative emotions like anger and fear. This bias may not be specific to social-​emotional stimuli
as several studies (see von Ceumern-​Lindenstjerna et al.) have demonstrated negative biases in
borderline patients for non-​social stimuli. Therefore, it may be that the negative bias for social
stimuli is part of this general bias toward negative emotion. This proposed hypervigilance for
negative emotion (or emotion in general according to Frick et al., 2012) is thought to associate
with reduced amygdala volume and enhanced amygdala responding to emotional stimuli, such as
negative facial expressions, coupled with regulatory deficits of the orbital and prefrontal cortices
(Domes, Schulze, & Herpertz, 2009; Frick et al., 2012). Indeed, three neuroimaging studies uti-
lizing adult samples have explicitly investigated neural responses to emotion recognition in BPD
and have confirmed this hypothesis. Donegan et al. (2003) showed that borderline patients dem-
onstrated significantly greater left amygdala activation to the facial expressions of emotion (vs.
a fixation point) compared to healthy control subjects. (Minzenberg, Fan, New, Tang, & Siever,
2007) found that borderline patients exhibited changes in fronto-​limbic activity in the process-
ing of fear stimuli, with exaggerated amygdala response and impaired emotion-​modulation of
anterior cingulate cortex (ACC) activity. Similarly, Frick et  al. (2012) demonstrated stronger
activation of the amygdala in response to affective pictures, regardless of valence, compared to
healthy controls.
Second, while research in adolescents is still lacking, research in adults suggest that more com-
plex emotion recognition tasks more consistently distinguish BPD from non-​BPD groups. For
instance, in the Minzenberg, Poole, and Vinogradov (2006) study, where facial, prosodic (the
aspect of speech that communicates meaning by variation in stress and pitch independent of lexi-
cal and syntactic content) and integrated facial/​prosodic stimuli were used, borderline patients
showed no problems with isolated facial or prosodic emotion, but instead demonstrated deficits
in higher order integration of social information. Similarly, Dyck et al. (2009) investigated the
ability of individuals with BPD to recognize negative and neutral emotions in both timed and
untimed trials. They found that individuals with BPD were significantly impaired in their recogni-
tion when the task was timed; however, no such difficulty was noted when the participants were
not timed. Thus, the participants with BPD were significantly impaired when under time pressure
and were less able to correctly judge negative or neutral affect in a hasty manner. It is possible
therefore, that borderline patients have emotion recognition deficits when tasks require the inte-
gration of different modes of processing (emotion recognition and speed of response), or when
tasks are presented in the context of heightened emotional arousal (Dixon-​Gordon, Chapman,
Lovasz, & Walters, 2011b).
Crucial for future research, in this regard, is the inclusion of psychiatric control groups, as
studies typically compare adolescents with BPD with healthy controls. The specificity of emo-
tional sensitivity to BPD, beyond mere “caseness” or neuroticism is, therefore, not clear. Moreover,
the use of psychophysiology and neuroimaging to assess emotional sensitivity beyond subjective
self-​report in adolescents is completely absent and there is an urgent need for biologically-​based
studies. Finally, it is highly probable that attachment-​relevant interpersonal situations will evoke
stronger emotional reactions than more general social contexts. For instance, stimuli that include
7
6
2

Borderline personality disorder in adolescents 267

the faces or other identifying characteristics of actual attachment figures would increase emo-
tional salience in theoretically relevant ways.

Intense negative affect
As formulated by Carpenter and Trull (2013), the second component in the emotion dysregula-
tion process in BPD is the experience of intense, negative and labile affect. Intense, negative and
labile affect is seen as a direct result of emotional sensitivity to subtle events that may seem benign
to the casual observer, but which can cause rapid change in mood to an individual with BPD.
Typically, Ecological Momentary Assessment (EMA) methods, by which a research participant
repeatedly reports on symptoms, affect, behavior, and cognitions close in time to experience and
in the participants’ natural environment (Stone & Shiffman, 1994), provide the richest data in this
regard as it can track context-╉dependent (i.e., ecologically valid) valence, intensity and moment-╉
by-╉moment change in emotion, although trait-╉based approaches have also been used (Solhan,
Trull, Jahng, & Wood, 2009). In adults, EMA studies have generally supported greater negative
affective lability in BPD (see Nica & Links, 2009 for a review).
Only one study has used EMA methodology to assess negative affect in the context of adoles-
cent BPD (Scott et al., 2015). The study assessed the covariation of daily experiences of shame
and anger-╉related affects/╉hostile irritability and borderline symptoms in a community sample of
adolescent girls while they were going about their daily lives. Results generally supported the
hypothesized associations between shame and anger-╉related affects in those with greater border-
line features, such that, over the course of one week, shame (but not guilt) was associated with
greater hostile irritability, but only in girls with high levels of borderline symptoms.

Inadequate emotion regulation strategies


The third component of emotion dysregulation in BPD, as articulated by Carpenter and Trull
(2013), is a deficit in appropriate emotion regulation strategies. Here, the focus is on a lack of
adaptive strategies as opposed to the use of maladaptive strategies. In Gross’s (1998) definitional
terms discussed earlier, this deficit relates to the unavailability or access to skills, rather than the
capacity to implement skills appropriately. Both the deficit in appropriate emotion regulation
strategies and the use of maladaptive strategies are crucial components of the emotion dysregu-
lation model of BPD, because they provide clear, malleable treatment targets. In other words,
while it is counter-╉productive to tell an individual with BPD that their emotions are too intense
or that they are overly sensitive (in fact, this will reinforce their experience of invalidation which
originally contributed to the development of the disorder), learning emotion regulation skills to
manage intense, negative and labile emotions is a feasible alternative, which we return to in the
section on evidence-╉based intervention approaches.
An important first step to adequate emotion regulation is the accurate identification and dif-
ferentiation of one’s emotions. Adult patients with BPD have demonstrated difficulties in identify-
ing, differentiating and labeling emotions (Coifman, Berenson, Rafaeli, & Downey, 2012; Leible
& Snell, 2004; Suvak et al., 2011; Tomko, Lane, Pronove, Treloar, Brown, Solhan, Wood, & Trull,
in press). Adults with BPD have also been shown to have problems in distress tolerance suggest-
ing a lack of coping strategies to manage negative affect (Bornovalova, Matusiewicz, & Rojas,
2011; Gratz, Tull, Baruch, Bornovalova, & Lejuez, 2008). Borderline patients also report limited
access to emotion regulation strategies (Salsman & Linehan, 2012) as well as general difficulties
in employing emotion regulation strategies (Glenn & Klonsky, 2009; Gratz et al., 2008). Studies
such as these, that have used self-╉report to evaluate inadequate emotion regulation strategies, have
typically used the Difficulties in Emotion Regulation Scale (DERS; Gratz & Roemer, 2004) which
has become a popular assessment tool given its brevity and low cost. The DERS is a 36-╉item
8
6
2

268 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

measure and subjects are asked to rate the frequency of each statement using a Likert-╉type scale
ranging from 1 = “Almost Never” to 5 = “Almost Always.” A total score of emotion dysregulation is
derived by summing all responses (indicating greater difficulty) with additional emotion regula-
tion domains assessed for including 1) awareness and understanding of emotions; 2) acceptance
of emotions; 3) the ability to engage in goal-╉directed behavior, and refrain from impulsive behav-
ior when experiencing negative emotions; 4)  access to emotion regulation strategies perceived
as effective; and 5) the flexible use of situationally appropriate strategies to modulate emotional
responses.
To our knowledge, there are no studies, in adolescents with BPD, using experimental mea-
sures to evaluate identification and differentiation of their own emotion. There are also no avail-
able studies of distress tolerance in adolescents with BPD. There are, however, a few studies that
have utilized the DERS in adolescent samples. In a sample of inpatient adolescents, Sharp, Ha,
Michonski, Venta, and Carbonne (2012) showed that adolescents who met DSM-╉IV defined
criteria for BPD evidenced higher total DERS scores compared to adolescents not meeting cri-
teria for BPD. In addition, DERS total scores correlated positively with a self-╉report measure
of borderline features. In another study, difficulties in emotion regulation were shown to relate
to social-╉cognitive (mentalizing) capacity (Sharp et  al., 2011). Moreover, difficulties in emo-
tion regulation mediated the relation between impairment in social cognition and borderline
traits. In a study contrasting difficulties in emotion regulation strategies (DERS), with the use
of positive emotion regulation strategies, as measured by the Cognitive Emotion Regulation
Questionnaire, (CERQ; Garnefski, Kraaij, & Spinhoven, 2002) along with assessments of mater-
nal and paternal attachment security, Kim, Sharp, and Carbone (2014) showed that difficulties
in emotion regulation strategies and the use of positive emotion regulation strategies were dif-
ferentially implicated in the link between attachment insecurity and BPD features. Attachment
security functioned as a buffer against adolescent BPD by enhancing positive emotion regula-
tion strategies, while difficulties in emotion regulation strategies served to dilute the protec-
tive effect of attachment and positive regulation strategies, culminating in clinically significant
levels of borderline traits.
In all, two conclusions can be drawn from the above literature. First, like their adult counter-
parts, adolescents with BPD seem to experience a similar lack of emotion regulation strategies.
Second, this impairment appears to relate to attachment insecurity and also affect functioning
in other relationship-╉relevant domains, like social cognition. As with the other components of
Carpenter and Trull’s (2013) model of emotion dysregulation in BPD, more research is clearly
needed in this area. Research in adults highlight the need for considering the interaction between
components of emotion dysregulation, which should also be a goal of research in adolescent BPD.
For instance, research focusing on inadequate emotion regulation strategies in BPD would be sig-
nificantly enhanced if intense, negative affect is routinely assessed and controlled for in studies. In
so doing, one can begin to parse out the validity of different components of emotion dysregulation
to arrive at a more nuanced model of emotion dysregulation in BPD.

Maladaptive emotion regulation strategies


The fourth and final component of Carpenter and Trull’s (2013) emotion dysregulation model of
BPD is perhaps the component most closely associated with the concept of maladaptive emotion
regulation strategies. These are easily observable behaviors in individuals with BPD and include
problems in substance use and other impulsive behaviors like aggression towards others or self-╉
directed aggression (Carpenter & Trull, 2013), most notably, self-╉injurious behavior (SIB), with
several studies demonstrating that SIB serves an emotion regulation function in adults with BPD
(Klonsky, 2007).
9
6
2

Borderline personality disorder in adolescents 269

Maladaptive emotion regulation strategies may also include unobservable maladaptive cogni-
tive strategies that are employed to help manage intense and negative emotion. For instance, con-
sistent with the Emotional Cascades Model discussed above, adults with BPD have been shown
to engage in intense rumination, thereby increasing the magnitude of the negative affect that
caused the rumination in the first place, culminating in dysregulated behavior in order to distract
from rumination (Selby, Anestis, & Joiner, 2008). Individuals with BPD have also been shown
to engage in experiential avoidance (EA) (Dixon-╉Gordon, Chapman, Lovasz, & Walters, 2011a;
Gratz, Rosenthal, Tull, Lejuez, & Gunderson, 2006; Iverson, Follette, Pistorello, & Fruzzetti, 2012).
EA is defined as an “unwillingness to remain in contact with uncomfortable private events (e.g.,
thoughts, emotions, sensations, memories, urges)” that often manifests in behaviors that serve
to avoid unpleasant experiences (Hayes, Wilson, Gifford, Follette, & Strosahl, 1996, p.  1154).
Typical EA behaviors include thought suppression, denial, self-╉distraction, substance abuse, and
self-╉injury. While these behaviors alleviate distress in the short-╉term, avoidance of unpleasant
thoughts and sensations actually increases the likelihood of experiencing them again in the future,
elevating physiological arousal and distress (Chawla & Ostafin, 2007). This sets into motion a
vicious cycle of using more avoidance-╉based strategies, thereby thwarting healthy and effective
emotion regulation.
While studies are generally lacking in adolescents on the use of maladaptive emotion regulation
strategies, there is an emerging literature for SIB as an emotion regulation strategy in adoles-
cents with BPD. Consistent with Crowell and colleagues’ developmental psychopathology model
of BPD, SIB and BPD appear to co-╉occur in adolescence (see Gratz, Dixon-╉Gordon, and Tull,
2014 for a review). Adolescents with a history of deliberate self-╉harm also report higher levels of
overall emotion dysregulation on the DERS and a specific impairment in access to effective emo-
tion regulation strategies (Perez, Venta, Garnaat, & Sharp, 2012). Studies in adolescents have also
shown that adolescents who engage in SIB exhibit reduced respiratory sinus arrhythmia (RSA)
at baseline, greater RSA reactivity during negative mood induction, and attenuated peripheral
serotonin levels (Crowell et al., 2005). These studies suggest that SIB may serve a similar emotion
regulation function for adolescents with BPD as suggested for adults.
Similarly, at least two studies have demonstrated EA is associated with BPD in adolescents.
In a community sample of 881 adolescents (Sharp, Kalpakci, Mellick, Venta, & Temple, 2014) a
prospective relation between EA & BPD was demonstrated and, measured one year after baseline,
controlling for symptoms of anxiety and depression. In adolescent inpatients with BPD, Schramm,
Venta, and Sharp (2013) found that EA made a significant and independent contribution to the
variance in borderline features, while partially mediating the relation between difficulties in emo-
tion regulation and borderline features. The results of these studies were interpreted in the context
of a mentalization-╉based account of BPD (Fonagy & Luyten, 2009) in which the capacity to be
open and curious about one’s own mental states, without becoming distressed by them or trying to
control them (that is, EA), comes about in the context of secure attachment with primary caregiv-
ers. Indeed, in another study, we have shown that disorganized attachment predicted EA, which
in turn predicted the capacity to accurately assess mental states in others (Vanwoerden, Kalpakci,
& Sharp, 2015).

Summary
A major limitation of the emerging research on emotion dysregulation in BPD (beyond the mere
lack thereof), is the fact that research is not particularly developmentally sensitive and relies heav-
ily on self-╉report. Most of the emotion dysregulation measures are downward extensions of adult
measures and generally, few studies have employed experimental paradigms of emotion dysregu-
lation. Importantly, few studies have adopted a prospective design and it is unclear to what extent
0
7
2

270 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

emotion dysregulation is normative in adolescence. Despite these limitations, there is enough


evidence in support of each of the components of Carpenter and Trull’s (2013) multicomponent
model of BPD to warrant further research in adolescence. There is also enough research to justify
intervention approaches that focus explicitly on addressing problems of emotion regulation in
adolescents with BPD. It is to this literature that we turn to next.

Evidence-╉based intervention for BPD in adolescents


There are six intervention programs described in the literature, of which three have an evidence
base (defined here as an intervention for which a randomized-╉control trial [RCT] was conducted).
Below, we describe all six interventions and indicate the strength of evidence in support of each.
Perhaps most explicitly associated with targeting emotion dysregulation in BPD is Dialectical
behavior therapy (DBT; Linehan, 1993). DBT incorporates a focus on change and acceptance; spe-
cifically, to target difficulties in emotional dysregulation, distress tolerance, and interpersonal dif-
ficulties in BPD. An RCT was conducted in Norway with adolescents with non-╉suicidal self-╉injury
(NSSI) and at least two DSM-╉IV BPD criteria (plus the self-╉destructive criterion or at least DSM-╉
IV BPD criterion plus at least two subthreshold level criteria) randomized to DBT or Enhanced
Usual Care (EUC) (Mehlum et al., 2014). Results demonstrated significant decreases in self-╉harm
in the DBT group, but not in EUC, with differences emerging in the last third of the trial period.
An RCT was also conducted in New Zealand in adolescents with a history of NSSI and suicide
attempts (Cooney et al., 2012). Although this study was not focused on BPD specifically, DBT did
not show any improvements at 6 months above TAU for the DBT group.
Systems Training for Emotional Predictability and Problem Solving (STEPPS; Blum et al., 2008),
is a 20-╉week, manual-╉based, group treatment program for outpatients with BPD that combines
cognitive-╉behavioral elements and skills training with a systems component specifically targeting
emotion dysregulation in BPD. In fact, STEPPS refrains from using the word BPD when working
with patients and prefer the term “emotional sensitivity.” While STEPPS has an evidence base in
adults, it is yet to be evaluated in an RCT. Recently, STEPPS-╉A was evaluated in a small-╉scale study
in the United Kingdom in adolescents with BPD, and has shown strong potential for this popula-
tion (Harvey, Blum, Black, Burgess, & Henley-╉Cragg, 2014).
Emotion Regulation Training (ERT) (Schuppert et al., 2009b) is an adaptation of STEPPS in the
Netherlands for which two RCTs have been conducted in adolescents. It also includes additional
elements of Cognitive Behavior Therapy and DBT. The first RCT showed no benefit of ERT post-╉
treatment (Schuppert et al., 2009a), while the second showed significant improvement for adoles-
cents in the ERT condition at a six-╉month follow-╉up (Schuppert et al., 2012).
While DBT, STEPPS and ERT are rooted firmly in the tradition of behaviorism, two evidence-╉
based intervention approaches have emerged with psychodynamic roots. Neither of these treat-
ment approaches were designed to directly target emotion dysregulation and impulsivity, but they
both increase emotion regulation and behavioral control indirectly. Cognitive analytic therapy
(CAT; Ryle & Kerr, 2002), developed in Australia used in the Helping Young People Early (HYPE)
program (Chanen, Jackson, et al., 2009; Chanen & McCutcheon, 2013b; Chanen, McCutcheon,
et al., 2009), was the first individual therapy to be tested in an RCT for adolescent BPD and was
evaluated in the context of an early intervention program. HYPE is a comprehensive and inte-
grated indicated prevention and early intervention program for youth (15–╉25 years of age); it
includes adolescents who meet two or more BPD criteria, plus a childhood risk factor. More
recently, inclusion criteria for HYPE have been specified as meeting three BPD criteria with no risk
factors (Chanen, McCutcheon, & Kerr, 2014). HYPE includes both a service model and individual
therapy, and incorporates the principles of CAT into both components. CAT is time-╉limited and
1
7
2

Borderline personality disorder in adolescents 271

transdiagnostic, integrating elements of psychoanalytic object relations theory and cognitive psy-
chology. Compared to treatment as usual, CAT has demonstrated effectiveness and more rapid
recovery, although differences were not as marked at two-╉year follow-╉up (Chanen, Jackson, et al.,
2009). The CAT model is currently being disseminated in Europe.
Mentalization-╉based treatment (MBT; Bateman & Fonagy, 2009) shares many common features
with CAT (Bateman, Ryle, Fonagy, & Kerr, 2007), and has been adapted for use in adolescents.
This therapy assumes that the development of BPD in adolescence and its treatment is grounded
in a phase-╉specific compromise in the capacity to mentalize that occurs during adolescence
(Fonagy, Rossouw, et al., 2014). MBT for adolescents (MBT-╉A), which incorporates monthly ses-
sions of MBT for families (MBT-╉F) has been shown to be effective in an RCT in a sample of self-╉
harming adolescents (most of whom met criteria for BPD; Rossouw & Fonagy, 2012). MBT-╉A was
more effective than treatment as usual in reducing self-╉harm and depression. This superiority was
explained by improved mentalization and reduced attachment avoidance, and reflected improve-
ment in emergent BPD symptoms and traits.
Finally, transference-╉focused psychotherapy (TFP; Clarkin et al., 2001) has been adapted for use
in adolescents. TFP is based on contemporary psychoanalytic object relations theory as developed
by Kernberg. TFP-╉A is a manualized psychodynamic treatment for borderline adolescents deliv-
ered in individual sessions, ideally twice a week but not less often than once a week (Normandin,
Ensink, Yeomans, & Kernberg, 2014). Although commonly used with adolescents with BPD, TFP-╉
A has not yet been evaluated in an RCT, but also shows potential for indirectly affecting emotion
dysregulation through the process of increasing self integration as therapy progresses.

An in-depth look at MBT-A: Adolescents with BPD


Central to MBT-╉A is the construct of mentalizing, which refers to the capacity to reflect on own
and others’ minds in the context of relationships, in order to make sense of ourselves and our rela-
tionships. MBT-╉A proposes that adolescence is the point at which vulnerabilities, resulting from
early developmental difficulties, are exacerbated by neurodevelopmental changes, weakening
mentalizing and mentalizing-╉mediated affect regulation, and by intense psychosocial and devel-
opmental pressures that place greater demands on the capacity to represent the self and regulate
affect (Fonagy, Rosssouw, et al., 2014). This combination of factors creates the conditions for the
symptomatic expression of BPD. MBT-╉A is therefore very much rooted in attachment theory and
against this background, the primary aim of MBT-╉A is to help young people and their families
improve their awareness of their own mental states and the mental states of others by enhancing
their capacity to mentalize within the attachment relationship. The emphasis is on improving
their understanding of the mental states and processes that drive behavior and relational patterns.
Treatment is divided up into four phases, all of which are derived from the original MBT model
for adults (Bateman & Fonagy, 2004). Below, we summarize each of the phases briefly, but a more
detailed description can be found in Fonagy, Rosssouw, et al. (2014).

Assessment
Intervention begins with a two-╉week assessment period that includes all members of the family
and focuses on the evaluation of psychiatric symptoms through observations, interview and
standardized measures. The aim is firstly, to identify conditions that may require adjunctive
treatments (such as medication), to highlight any comorbidities, and to make the therapist
aware of any psychiatric conditions that can impair the ability to mentalize. In addition, assess-
ment also aims to fully characterize the mentalizing capacity, cognitive, executive function and
emotional regulation of the adolescent, as well as the general mentalizing capacity of family
2
7

272 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

members. The assessment of mentalizing in BPD patients will likely show that the adolescent
is able to mentalize, but does so intermittently; therefore, in highly charged situations, often in
the context of a family assessment session, the adolescent may show a temporary inability to
recognize the feelings and experiences of others resulting in hypermentalization (that is, the
over-╉attribution of mental states to others). Assessment tools that may aid clinical assessment
of mentalizing capacity in the adolescent include the Reflective Functioning Questionnaire for
Youth (Ha, Sharp, Ensink, Fonagy, & Cirino, 2013), which was recently validated and provides
an adequate measure of self-╉report mentalizing capacity in adolescents. A questionnaire-╉based
measure of hypermentalizing is currently being evaluated for its’ validity, but has shown prelim-
inary promise (Sharp, 2015). The Movie Task for the Assessment of Social Cognition (Dziobek
et al., 2006) has been used in inpatient adolescent settings (Sharp et al., 2009) and has shown
sensitivity to treatment outcome (Sharp et al., 2013). These measures provide valuable informa-
tion for a mentalization-╉based case formulation and may also be important in tracking change
and outcome.

Initial phase
The assessment phase is followed by the initial phase which consists of two sessions. First, two
parallel sessions with the adolescent and the family are carried out to share the mentalization-╉
based formulation. The aim is to make the adolescent and the family feel understood, and to use
the formulation to plan treatment. During these sessions, a crisis plan is developed which identi-
fies any triggers of emotional outbursts and/╉or impulsive behavior, including self-╉harm. In addi-
tion, a treatment contract is developed which sets out the duration of treatment and commitment
required from all those participating; it explains the importance of everyone’s engagement and the
process of working together in the therapy.
The family formulation session is followed by a psychoeducation session, which may be deliv-
ered to the individual family or in a group format. This aims to help the family understand that
behavior has meaning, that feelings arise in a relational context, and that people have a powerful
emotional impact on one another. Psychoeducation may involve informal discussion with the
family, using examples from everyday life, or in multifamily groups it may make use of group
discussion, role-╉play and videos.

Middle phase
The middle phase of MBT-╉A can be seen as the remediation and rehabilitation phase of ther-
apy, and lasts nine to ten months. It aims to enhance mentalization in the adolescent and family
through the development of mentalizing skills (i.e. active reflection on the mind of self and oth-
ers). This phase also aims to help the adolescent and family gain better emotion regulation and
impulse control (as dysregulation and impulsivity undermines the development and use of men-
talizing ability). MBT-╉A sessions are unstructured and focus on the young person’s current and
recent interpersonal experiences, while maintaining a constant focus on the mental states likely to
have been evoked by these experiences. The main tool of the therapist is the “mentalizing stance”
which is defined as an open, curious attitude towards the client. In addition, the therapist uses a
number of specific techniques that include supportive and empathic interventions, clarification
and elaboration techniques, basic mentalizing techniques, transference techniques, and inter-
pretive mentalizing techniques. In general, interventions are simple, “soundbite” interventions
that do not require excessive processing competencies on the part of the young person (Fonagy,
Rosssouw, et al., 2014). They are affect-╉focused and current (e.g., love, desire, hurt, catastrophe,
excitement), as these domains are most accessible for the construction of subjective states. To
facilitate accessibility, the therapist often uses his/╉her own mind as a model; not in the sense of
3
7
2

Conclusion 273

self-╉disclosure, but as a normalizing influence suggesting to the young person how the therapist
may feel or may think in the context the young person presents.

Final phase
The final phase of MBT-╉A addresses separation issues along with managing anticipated challenges
in a mentalizing manner. It aims to increase the adolescent’s independence and responsibility, and
consolidate relational stability and a sense of mastery (as opposed to helplessness or passivity) in
the adolescent and his/╉her family. In addition, a coping plan is created for the family, setting out
what to do in the future if difficulties return. The final phase of MBT-╉A lasts for approximately
two months and commonly includes a tapering-╉off of sessions at the end. Some families also find
it helpful to return for one final family session a few months afterwards.
The discussion of MBT-╉A above was necessarily brief and readers are referred to the adult man-
ual for treatment of BPD (Bateman & Fonagy, 2006) or Fonagy, Rosssouw, et al. (2014) for a more
detailed discussion of MBT-╉A for BPD in adolescents.

Conclusion
The aim of this chapter was to provide an overview of the construct of emotion dysregulation
in the context of BPD in adolescents. Our review has demonstrated that BPD has been seen as
the quintessential disorder of emotion regulation. While these conceptualizations have strong
theoretical and clinical foundations, there is room for more empirical research to further support
these ideas. In this regard, we identify two important goals for further research in adolescents.
First, research should be guided by a process-╉oriented and multi-╉component model of emotion
dysregulation in order to assess the complex interactions involved in emotion dysregulation.
Failing to do so will result in a piecemeal and potentially clinically meaningless understanding
of emotion dysregulation in BPD. Second, it is important to study BPD in the context of other
psychopathology. High comorbidity between BPD and other disorders has led authors to inves-
tigate the location of BPD within the latent structure of psychopathology in general. It is impor-
tant that these methods are combined with experimental approaches to emotion dysregulation,
where feasible, to further harness the transdiagnostic potential of emotion dysregulation and its
treatment.

References
Allen, J. P., Insabella, G., Porter, M. R., Smith, F. D., Land, D., & Phillips, N. (2006). A social-╉interactional
model of the development of depressive symptoms in adolescence. Journal of Consulting and Clinical
Psychology, 74(1), 55–╉65. doi:10.1037/╉0022-╉006x.74.1.55
American Psychiatric Association. (1968). Diagnostic and statistical manual of mental disorders (2nd ed.).
Washington, DC: American Psychiatric Association.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.).
Washington, DC: American Psychiatric Association.
Bateman, A., & Fonagy, P. (2004). Psychotherapy for borderline personality disorder: Mentalization-╉based
treatment. Oxford: Oxford University Press.
Bateman, A., & Fonagy, P. (2009). Randomized controlled trial of outpatient mentalization-╉based treatment
versus structured clinical management for borderline personality disorder. [Comparative Study
Randomized Controlled Trial Research Support, Non-╉U.S. Gov’t]. American Journal of Psychiatry,
166(12), 1355–╉1364. doi:0.1176/╉appi.ajp.2009.09040539
Bateman, A. W., & Fonagy, P. (2006). Mentalization based treatment for borderline personality
disorder: A practical guide. Oxford, UK: Oxford University Press.
4
7
2

274 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

Bateman, A. W., Ryle, A., Fonagy, P., & Kerr, I. B. (2007). Psychotherapy for borderline personality
disorder: Mentalization based therapy and cognitive analytic therapy compared. International Review of
Psychiatry, 19(1), 51–​62. doi: 10.1080/​09540260601109422
Bloch, L., Moran, E. K., & Kring, A. M. (2010). On the need for conceptual clarity in emotion regulation
research on psychopathology. In A. M. Kring & D. M. Sloan (Eds.), Emotion regulation and
psychopathology (pp. 88–​104). New York: Guilford Press.
Blum, N., St John, D., Pfohl, B., Stuart, S., McCormick, B., Allen, J., … Black, D. W. (2008). Systems
Training for Emotional Predictability and Problem Solving (STEPPS) for outpatients with borderline
personality disorder: A randomized controlled trial and 1-​year follow-​up. [Comparative Study
Randomized Controlled Trial Research Support, N.I.H., Extramural]. American Journal of Psychiatry,
165(4), 468–​478. doi: 10.1176/​appi.ajp.2007.07071079
Bonanno, G. A., Papa, A., Lalande, K., Westphal, M., & Coifman, K. (2004). The importance of being
flexible: the ability to both enhance and suppress emotional expression predicts long-​term adjustment.
Psychological Science, 15(7), 482–​487. doi:10.1111/​j.0956-​7976.2004.00705.x
Bornovalova, M. A., Hicks, B. M., Iacono, W. G., & McGue, M. (2009). Stability, change, and heritability
of borderline personality disorder traits from adolescence to adulthood: A longitudinal twin study.
[Research Support, N.I.H., Extramural Twin Study]. Development and Psychopathology, 21(4), 1335–​
1353. doi:10.1017/​S0954579409990186
Bornovalova, M. A., Matusiewicz, A., & Rojas, E. (2011). Distress tolerance moderates the relationship
between negative affect intensity with borderline personality disorder levels. Comprehensive Psychiatry,
52(6), 744–​753. doi:10.1016/​j.comppsych.2010.11.005
Carlson, E. A., Egeland, B., & Sroufe, L. A. (2009). A prospective investigation of the development of
borderline personality symptoms. Dev Psychopathol, 21(4), 1311–​1334.
Carpenter, R. W., & Trull, T. J. (2013). Components of Emotion Dysregulation in Borderline Personality
Disorder: A Review. Current Psychiatry Reports, 15(1). doi:Artn 2 Doi 10.1007/​S11920-​012-​0335-​2
Chanen, A. M., Jackson, H. J., McCutcheon, L. K., Jovev, M., Dudgeon, P., Yuen, H. P., … McGorry, P. D.
(2009). Early intervention for adolescents with borderline personality disorder: Quasi-​experimental
comparison with treatment as usual. [Comparative Study Research Support, Non-​U.S. Gov’t].
Australian and New Zealand Journal of Psychiatry, 43(5), 397–​408. doi:10.1080/​00048670902817711
Chanen, A. M., Jackson, H. J., McGorry, P. D., Allot, K. A., Clarkson, V., & Yuen, H. P. (2004). Two-​year
stability of personality disorder in older adolescent outpatients. [Research Support, Non-​U.S. Gov’t].
Journal of Personality Disorders, 18(6), 526–​541. doi: 10.1521/​pedi.18.6.526.54798
Chanen, A. M., Jovev, M., & Jackson, H. J. (2007). Adaptive functioning and psychiatric symptoms in
adolescents with borderline personality disorder. J Clin Psychiatry, 68(2), 297–​306.
Chanen, A. M., & Kaess, M. (2012a). Developmental pathways to borderline personality disorder.
[Research Support, Non-​U.S. Gov’t Review]. Current Psychiatry Reports, 14(1), 45–​53. doi:10.1007/​
s11920-​011-​0242-​y
Chanen, A. M., & Kaess, M. (2012b). Developmental pathways to borderline personality disorder.
[Research Support, Non-​U.S. Gov’t Review]. Curr Psychiatry Reports, 14(1), 45–​53. doi:10.1007/​
s11920-​011-​0242-​y
Chanen, A. M., & McCutcheon, L. (2013a). Prevention and early intervention for borderline personality
disorder: current status and recent evidence. [Research Support, Non-​U.S. Gov’t]. Br J Psychiatry Suppl,
54, s24–​s29. doi:10.1192/​bjp.bp.112.119180
Chanen, A. M., & McCutcheon, L. (2013b). Prevention and early intervention for borderline personality
disorder: Current status and recent evidence. [Research Support, Non-​U.S. Gov’t]. British Journal of
Psychiatry Supplement, 54, s24–​s29. doi:10.1192/​bjp.bp.112.119180
Chanen, A. M., McCutcheon, L., & Kerr, I. B. (2014). HYPE: A cognitive analytic therapy-​based prevention
and early intervention program for borderline personality disorder. In C. Sharp & J. L. Tackett (Eds.),
Handbook of borderline personality disorder in children and adolescents (pp. 261–​383). New York,
NY: Springer.
5
7
2

Conclusion 275

Chanen, A. M., McCutcheon, L. K., Germano, D., Nistico, H., Jackson, H. J., & McGorry, P. D. (2009).
The HYPE Clinic: An early intervention service for borderline personality disorder. [Case Reports
Research Support, Non-​U.S. Gov’t]. Journal of Psychiatric Practice, 15(3), 163–​172. doi:10.1097/​
01.pra.0000351876.51098.f0
Chawla, N., & Ostafin, B. (2007). Experiential avoidance as a functional dimensional approach to
psychopathology: an empirical review. [Comparative Study Review]. J Clin Psychol, 63(9), 871–​890.
doi: 10.1002/​jclp.20400
Cicchetti, D., Ackerman, B. P., & Izard, C. E. (1995). Emotions and emotion regulation in developmental
psychopathology. Development and Psychopathology, 7(1), 1–​10.
Clarkin, J. F., Foelsch, P. A., Levy, K. N., Hull, J. W., Delaney, J. C., & Kernberg, O. F. (2001).
The development of a psychodynamic treatment for patients with borderline personality
disorder: A preliminary study of behavioral change. Journal of Personality Disorders, 15(6), 487–​495.
doi:10.1521/​pedi.15.6.487.19190
Cohen, P., Chen, H., Gordon, K., Johnson, J., Brook, J., & Kasen, S. (2008). Socioeconomic background
and the developmental course of schizotypal and borderline personality disorder symptoms. [Research
Support, N.I.H., Extramural Research Support, Non-​U.S. Gov’t]. Development and Psychopathology,
20(2), 633–​650. doi: 10.1017/​S095457940800031X
Coifman, K. G., Berenson, K. R., Rafaeli, E., & Downey, G. (2012). From Negative to Positive and Back
Again: Polarized Affective and Relational Experience in Borderline Personality Disorder. Journal of
Abnormal Psychology, 121(3), 668–​679. doi:10.1037/​A0028502
Cole, P. M., Martin, S. E., & Dennis, T. A. (2004). Emotion regulation as a scientific
construct: Methodological challenges and directions for child development research. Child
Development, 75(2), 317–​333. doi:10.1111/​j.1467-​8624.2004.00673.x
Cooney, E., Davis, K., Thompson, P., Wharewera-​Mika, J., Stewart, J., & Miller, A. L. (2012). Feasibility
of comparing dialectical behavior therapy with treatment as usual for suicidal and self-​injuring
adolescents: Follow-​up data from a small randomized controlled trial. Paper presented at the Association
of Behavioral and Cognitive Therapies 46th Annual Convention, National Harbor, MD.
Crowell, S. E., Beauchaine, T. P., & Linehan, M. M. (2009). A Biosocial Developmental Model of
Borderline Personality: Elaborating and Extending Linehan’s Theory. Psychological Bulletin, 135(3),
495–​510. doi:10.1037/​A0015616
Crowell, S. E., Beauchaine, T. P., McCauley, E., Smith, C. J., Stevens, A. L., & Sylvers, P. (2005).
Psychological, autonomic, and serotonergic correlates of parasuicide among adolescent girls.
Development and Psychopathology, 17(4), 1105–​1127. doi:10.1017/​S0954579405050522
Csibra, G. (2010). Recognizing communicative intentions in infancy. Mind and Language, 25,
141–​168.
Csibra, G., & Gergely, G. (2011). Natural pedagogy as evolutionary adaptation. Philosophical Transactions
of the Royal Society B: Biological Sciences, 366(1567), 1149–​1157. doi:10.1098/​rstb.2010.0319
Daros, A. R., Zakzanis, K. K., & Ruocco, A. C. (2013). Facial emotion recognition in borderline personality
disorder. [Meta-​Analysis Research Support, Non-​U.S. Gov’t]. Psychol Med, 43(9), 1953–​1963.
doi: 10.1017/​S0033291712002607
De Clercq, B., De Fruyt, F., De Bolle, M., Van Hiel, A., Markon, K. E., & Krueger, R. F. (2014). The
Hierarchical Structure and Construct Validity of the PID-​5 Trait Measure in Adolescence. Journal of
Personality, 82(2), 158–​169. doi: 10.1111/​Jopy.12042
De Clercq, B., Decuyper, M., & De Caluwé, E. (2014). Developmental manifestations of Borderline
Personality Pathology from and age-​specific dimensional personality disorder trait framework. In C.
Sharp & J. L. Tackett (Eds.), Handbook of Borderline Personality Disorder in Children and Adolescents
(pp. 81–​94). New York: Springer.
Distel, M. A., Trull, T. J., Derom, C. A., Thiery, E. W., Grimmer, M. A., Martin, N. G., … Boomsma, D. I.
(2008). Heritability of borderline personality disorder features is similar across three countries. Psychol
Med, 38(9), 1219–​1229.
6
7
2

276 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

Dixon-​Gordon, K. L., Chapman, A. L., Lovasz, N., & Walters, K. (2011a). Too Upset to Think: The
Interplay of Borderline Personality Features, Negative Emotions, and Social Problem Solving in
the Laboratory. Personality Disorders-​Theory Research and Treatment, 2(4), 243–​260. doi:10.1037/​
A0021799
Dixon-​Gordon, K. L., Chapman, A. L., Lovasz, N., & Walters, K. (2011b). Too upset to think: the interplay
of borderline personality features, negative emotions, and social problem solving in the laboratory.
Personal Disord, 2(4), 243–​260. doi:10.1037/​a0021799
Domes, G., Schulze, L., & Herpertz, S. C. (2009). Emotion Recognition in Borderline Personality
Disorder—​a Review of the Literature. Journal of Personality Disorders, 23(1), 6–​19.
Donegan, N. H., Sanislow, C. A., Blumberg, H. P., Fulbright, R. K., Lacadie, C., Skudlarski, P., … Wexler,
B. E. (2003). Amygdala hyperreactivity in borderline personality disorder: implications for emotional
dysregulation. Biol Psychiatry, 54(11), 1284–​1293. doi:S000632230300636X [pii]
Dziobek, I., Fleck, S., Kalbe, E., Rogers, K., Hassenstab, J., Brand, M., … Convit, A. (2006). Introducing
MASC: a movie for the assessment of social cognition. J Autism Dev Disord, 36(5), 623–​636.
doi:10.1007/​s10803-​006-​0107-​0
Fonagy, P., Gergely, G., Jurist, E. L., & Target, M. (2002). Affect regulation, mentalization, and the
development of self. New York: Other Press.
Fonagy, P., Gergely, G., & Target, M. (2007). The parent-​infant dyad and the construction of the subjective
self. [Review]. Journal of Child Psychology and Psychiatry and Allied Disciplines, 48(3-​4), 288–​328.
doi:10.1111/​j.1469-​7610.2007.01727.x
Fonagy, P., & Luyten, P. (2009). A developmental, mentalization-​based approach to the understanding and
treatment of borderline personality disorder. Dev Psychopathol, 21(4), 1355–​1381.
Fonagy, P., Rossouw, T., Sharp, C., Bateman, A., Allison, L., & Farrar, C. (2014). Mentalization-​based
treatment for adolescents with borderline traits. In C. Sharp & J. L. Tackett (Eds.), Handbook of
borderline personality disorder in children and adolescents (pp. 313–​332). New York, NY: Springer.
Fonagy, P., Rosssouw, T., Sharp, C., Bateman, A., Allisson, L., & Farrar, C. (2014). Mentalization-​based
treatment for adolescents with borderline traits. In C. Sharp & J. L. TAckett (Eds.), The handbook of
borderline personality disorder in children and adolescents. New York: Springer.
Fossati, A. (2014). Borderline Personality Disorder in Adolescence: Phenomenology and Construct Validity.
In C. Sharp & J. Tackett (Eds.), Handbook of borderline personality disorder in children and adolescents.
New York: Springer.
Frick, C., Lang, S., Kotchoubey, B., Sieswerda, S., Dinu-​Biringer, R., Berger, M., … Barnow, S. (2012).
Hypersensitivity in Borderline Personality Disorder during Mindreading. Plos One, 7(8). doi:ARTN
e41650 10.1371/​journal.pone.0041650
Garnefski, N., Kraaij, V., & Spinhoven, P. (2002). Manual for the use of the Cognitive Emotion Regulation
Questionnaire. Leiderdorp, The Netherlands: DATEC.
Glenn, C. R., & Klonsky, E. D. (2009). Emotion Dysregulation as a Core Feature of Borderline Personality
Disorder. Journal of Personality Disorders, 23(1), 20–​28.
Gratz, K. L., Dixon-​Gordon, K. L., & Tull, M. T. (2014). Self-​injurious behaviors in adolescents wtih
Borderline Personality Disorder. In C. Sharp & J. L. Tackett (Eds.), Handbook of Borderline Personality
Disorder in Children and Adolescents (pp. 195–​210). New York: Springer.
Gratz, K. L., & Roemer, L. (2004). Multidimensional assessment of emotion regulation and
dysregulation: Development, factor structure and initial validation of the Difficulties in Emotion
Regulation Scale. Journal of Psychopathology and Behavioral Assessment, 26(1), 41–​54.
Gratz, K. L., Rosenthal, M. Z., Tull, M. T., Lejuez, C. W., & Gunderson, J. G. (2006). An experimental
investigation of emotion dysregulation in borderline personality disorder. Journal of Abnormal
Psychology, 115(4), 850–​855. doi:10.1037/​0021-​843x.115.4.850
Gratz, K. L., Tull, M. T., Baruch, D. E., Bornovalova, M. A., & Lejuez, C. W. (2008). Factors associated
with co-​occurring borderline personality disorder among inner-​city substance users: the roles of
childhood maltreatment, negative affect intensity/​reactivity, and emotion dysregulation. Comprehensive
Psychiatry, 49(6), 603–​615. doi:10.1016/​j.comppsych.2008.04.005
72

Conclusion 277

Gross, J. J. (1988). Antecendent-​and response-​focused emotion regualtion: Divergent consequences for


experinece, expression, and physiology. Journal of Personality and Social Psychology, 74, 224–​237.
Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations In J. J. Gross (Ed.),
Handbook of emotion regulation (pp. 3–​24). New York: Guilford Press.
Ha, C., Balderas, J., Zanarini, M., Oldham, J., & Sharp, C. (in press). Axis I Comorbidity in Hospitalized
Adolescents with Borderline Personality Disorder. Journal of Clinical Psychiatry, 75(5), 457–​464.
Ha, C., Sharp, C., Ensink, K., Fonagy, P., & Cirino, P. (2013). The measurement of reflective function
in adolescents with and without borderline traits. J Adolesc, 36(6), 1215–​1223. doi:10.1016/​
j.adolescence.2013.09.008
Hare, T. A., Tottenham, N., Galvan, A., Voss, H. U., Glover, G. H., & Casey, B. J. (2008). Biological
substrates of emotional reactivity and regulation in adolescence during an emotional go-​nogo task.
Biological Psychiatry, 63(10), 927–​934. doi:10.1016/​j.biopsych.2008.03.015
Harvey, R., Blum, N., Black, D. W., Burgess, J., & Henley-​Cragg, P. (2014). Systems Training for Emotional
Predictability and Problem Solving (STEPPS). In C. Sharp & J. L. Tackett (Eds.), Handbook of child and
adolescent borderline personality disorder (pp. 415–​430). New York, NY: Springer.
Hayes, S. C., Wilson, K. G., Gifford, E. V., Follette, V. M., & Strosahl, K. (1996). Experiential avoidance
and behavioral disorders: A functional dimensional approach to diagnosis and treatment. Journal of
Consulting and Clinical Psychology, 64(6), 1152–​1168. doi:10.1037//​0022-​006x.64.6.1152
Iverson, K. M., Follette, V. M., Pistorello, J., & Fruzzetti, A. E. (2012). An investigation of experiential
avoidance, emotion dysregulation, and distress tolerance in young adult outpatients with borderline
personality disorder symptoms. [Research Support, N.I.H., Extramural]. Personal Disord, 3(4), 415–​
422. doi:10.1037/​a0023703
Johnson, J. G., Cohen, P., Kasen, S., Skodol, A. E., & Oldham, J. M. (2008). Cumulative prevalence of
personality disorders between adolescence and adulthood. Acta Psychiatrica Scandinavica, 118(5),
410–​413. doi: DOI 10.1111/​j.1600-​0447.2008.01231.x
Jovev, M., Chanen, A., Green, M., Cotton, S., Proffitt, T., Coltheart, M., & Jackson, H. (2011). Emotional
sensitivity in youth with borderline personality pathology. Psychiatry Research, 187(1–​2), 234–​240.
doi:10.1016/​j.psychres.2010.12.019
Kim, S., Sharp, C., & Carbone, C. (2014). The Protective Role of Attachment Security for Adolescent
Borderline Personality Disorder Features via Enhanced Positive Emotion Regulation Strategies.
Personality Disorders-​Theory Research and Treatment, 5(2), 125–​136. doi:10.1037/​Per0000038
Klonsky, E. D. (2007). The functions of deliberate self-​injury: a review of the evidence. Clin Psychol Rev,
27(2), 226–​239. doi:S0272-​7358(06)00096-​1 [pii] 10.1016/​j.cpr.2006.08.002
Lawrence, K. A., Chanen, A. M., & Allen, J. S. (2011). The Effect of Ostracism Upon Mood in Youth with
Borderline Personality Disorder. Journal of Personality Disorders, 25(5), 702–​714.
Leible, T. L., & Snell, W. E. (2004). Borderline personality disorder and multiple aspects of emotional
intelligence. Personality and Individual Differences, 37(2), 393–​404. doi:10.1016/​j.paid.2003.09.011
Levy, K. N., Becker, D. F., Grilo, C. M., Mattanah, J. J., Garnet, K. E., Quinlan, D. M., … McGlashan,
T. H. (1999). Concurrent and predictive validity of the personality disorder diagnosis in adolescent
inpatients. American Journal of Psychiatry, 156(10), 1522–​1528.
Linehan, M. M. (1993). Cognitive-​behavioral treatment of borderline personality disorder. New York: The
Guildford Press.
Matusiewicz, A., Weaverling, G., & Lejeuz, C. W. (2014). Emotion dysregulation among adolescents with
borderline personality disorder. In C. Sharp & J. L. Tackett (Eds.), Handbook of borderline personality
disorder in children and adolescents (pp. 177–​194). New York: Springer.
Mehlum, L., Tormoen, A. J., Ramberg, M., Haga, E., Diep, L. M., Laberg, S., … Groholt, B. (2014).
Dialectical behavior therapy for adolescents with repeated suicidal and self-​harming behavior: a
randomized trial. [Research Support, Non-​U.S. Gov’t]. J Am Acad Child Adolesc Psychiatry, 53(10),
1082–​1091. doi:10.1016/​j.jaac.2014.07.003
Michonski, J. D., Sharp, C., Steinberg, L., & Zanarini, M. C. (2013). An item response theory analysis of
the DSM-​IV borderline personality disorder criteria in a population-​based sample of 11-​to 12-​year-​old
8
7
2

278 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

children. [Evaluation Studies Research Support, N.I.H., Extramural Research Support, Non-​U.S. Gov’t].
Personality Disorders: Theory, Research, and Treatment, 4(1), 15–​22. doi:10.1037/​a0027948
Miller, A. L., Muehlenkamp, J. J., & Jacobson, C. M. (2008). Fact or fiction: Diagnosing borderline
personality disorder in adolescents. Clinical Psychology Review, 28(6), 969–​981. doi:10.1016/​
j.cpr.2008.02.004
Minzenberg, M. J., Fan, J., New, A. S., Tang, C. Y., & Siever, L. J. (2007). Fronto-​limbic dysfunction in
response to facial emotion in borderline personality disorder: An event-​related fMRI study. Psychiatry
Research-​Neuroimaging, 155(3), 231–​243. doi:10.1016/​j.pseychresns.2007.03.006
Minzenberg, M. J., Poole, J. H., & Vinogradov, S. (2006). Social-​emotion recognition in borderline
personality disorder. Compr Psychiatry, 47(6), 468–​474. doi:S0010-​440X(06)00050-​2 [pii] 10.1016/​
j.comppsych.2006.03.005
Monk, C. S., McClure, E. B., Nelson, E. E., Zarahn, E., Bilder, R. M., Leibenluft, E., … Pine, D. S.
(2003). Adolescent immaturity in attention-​related brain engagement to emotional facial expressions.
Neuroimage, 20(1), 420–​428. doi:10.1016/​S1053-​8119(03)00355-​0
National Health and Medical Research Council. (2013). Clinical practice guideline for the management
of borderline personality disorder 2012. Melbourne, Australia: National Health and Medical Research
Council.
National Institute for Health and Clinical Excellence. (2009). Borderline personality disorder: Treatment
and management. Clinical guideline 78. London, UK: National Institute for Health and Clinical
Excellence.
Nica, E. I., & Links, P. S. (2009). Affective instability in borderline personality disorder: experience
sampling findings. [Review]. Curr Psychiatry Rep, 11(1), 74–​81.
Normandin, L., Ensink, K., Yeomans, F., & Kernberg, O. F. (2014). Transference-​focused psychotherapy for
personality disorders in adolescence. In C. Sharp & J. L. Tackett (Eds.), Handbook of child and adolesent
borderline personality disorder (pp. 333–​359). New York, NY: Springer.
Perez, J., Venta, A., Garnaat, S., & Sharp, C. (2012). The Difficulties in Emotion Regulation Scale: Factor
Structure and Association with Nonsuicidal Self-​Injury in Adolescent Inpatients. Journal of
Psychopathology and Behavioral Assessment, 34(3), 393–​404. doi:10.1007/​s10862-​012-​9292-​7
Robin, M., Pham-​Scottez, A., Curt, F., Dugre-​Le Bigre, C., Speranza, M., Sapinho, D., … Kedia, G. (2012).
Decreased sensitivity to facial emotions in adolescents with Borderline Personality Disorder. Psychiatry
Research, 200(2) 417–​421.
Roisman, G. I., Masten, A. S., Coatsworth, J. D., & Tellegen, A. (2004). Salient and emerging
developmental tasks in the transition to adulthood. Child Development, 75(1), 123–​133.
Rossouw, T. I., & Fonagy, P. (2012). Mentalization-​Based Treatment for Self-​Harm in
Adolescents: A Randomized Controlled Trial. Journal of the American Academy of Child and Adolescent
Psychiatry, 51(12), 1304–​1313. doi: 10.1016/​i.jaac.2012.09.018
Ruocco, A. C., Medaglia, J. D., Tinker, J. R., Ayaz, H., Forman, E. M., Newman, C. F., … Chute, D. L.
(2010). Medial prefrontal cortex hyperactivation during social exclusion in borderline personality
disorder. Psychiatry Research-​Neuroimaging, 181(3), 233–​236. doi:10.1016/​j.pscychresns.2009.12.001
Ryle, A., & Kerr, I. B. (2002). Introducing cognitive analytic therapy: Principles and practice. Chichester,
UK: John Wiley and Sons.
Salsman, N. L., & Linehan, M. M. (2012). An Investigation of the Relationships among Negative Affect,
Difficulties in Emotion Regulation, and Features of Borderline Personality Disorder. Journal of
Psychopathology and Behavioral Assessment, 34(2), 260–​267. doi:10.1007/​s10862-​012-​9275-​8
Schramm, A. T., Yenta, A., & Sharp, C. (2013). The Role of Experiential Avoidance in the Association
Between Borderline Features and Emotion Regulation in Adolescents. Personality Disorders-​Theory
Research and Treatment, 4(2), 138–​144. doi:10.1037/​A0031389
Schuppert, H. M., Giesen-​Bloo, J., van Gemert, T. G., Wiersema, H. M., Minderaa, R. B., Emmelkamp, P.
M., & Nauta, M. H. (2009a). Effectiveness of an emotion regulation group training for adolescents—​a
9
7
2

Conclusion 279

randomized controlled pilot study. [Multicenter Study Randomized Controlled Trial Research Support,
Non-​U.S. Gov’t]. Clin Psychol Psychother, 16(6), 467–​478. doi: 10.1002/​cpp.637
Schuppert, H. M., Giesen-​Bloo, J., van Gemert, T. G., Wiersema, H. M., Minderaa, R. B., Emmelkamp, P.
M., & Nauta, M. H. (2009b). Effectiveness of an emotion regulation group training for adolescents—​a
randomized controlled pilot study. [Multicenter Study Randomized Controlled Trial Research Support,
Non-​U.S. Gov’t]. Clinical Psychology and Psychotherapy, 16(6), 467–​478. doi: 10.1002/​cpp.637
Schuppert, H. M., Timmerman, M. E., Bloo, J., van Gemert, T. G., Wiersema, H. M., Minderaa, R. B.,
… Nauta, M. H. (2012). Emotion regulation training for adolescents with borderline personality
disorder traits: a randomized controlled trial. [Randomized Controlled Trial Research Support,
Non-​U.S. Gov’t]. J Am Acad Child Adolesc Psychiatry, 51(12), 1314–​1323 e1312. doi: 10.1016/​
j.jaac.2012.09.002
Scott, L. N., Stepp, S. D., Hallquist, M. N., Whalen, D. J., Wright, A. G. C., & Pilkonis, P. A. (2015). Daily
Shame and Hostile Irritability in Adolescent Girls With Borderline Personality Disorder Symptoms.
Personality Disorders-​Theory Research and Treatment, 6(1), 53–​63. doi:10.1037/​Per0000107
Selby, E. A., Anestis, M. D., & Joiner, T. E. (2008). Understanding the relationship between emotional
and behavioral dysregulation: emotional cascades. Behav Res Ther, 46(5), 593–​611. doi:10.1016/​
j.brat.2008.02.002
Selby, E. A., & Joiner, T. E. (2009). Cascades of Emotion: The Emergence of Borderline Personality
Disorder From Emotional and Behavioral Dysregulation. Review of General Psychology, 13(3), 219–​229.
doi:10.1037/​A0015687
Selby, E. A., Kranzler, A., & Panza, E. (2014). Development of emotional cascades in borderline personality
disorder. In C. Sharp & J. L. Tackett (Eds.), Handbook of borderline personality disorder in children and
adolescents (pp. 159–​176). New York: Springer.
Sharp, C. (2015). Hypermentalizing in adolescent BPD compared to psychiatry and healthy
controls: Introducing a new measure. Paper presented at the Annual Meeting of the American
Psychiatric Association, Montreal.
Sharp, C., & Fonagy, P. (2008). Social cognition and attachment-​related disorders. In C. Sharp, P.
Fonagy & I. M. Goodyer (Eds.), Social cognition and developmental psychopathology. Oxford: Oxford
University Press.
Sharp, C., & Fonagy, P. (in press). Practitioner review: Emergent borderline personality disorder in
adolescence:—​Recent conceptualization, intervention, and implications for clinical practice. Journal of
Child Psychology and Psychiatry 56(12), 1266–​1288.
Sharp, C., Ha, C., Carbone, C., Kim, S., Perry, K., Williams, L., & Fonagy, P. (2013). Hypermentalizing in
adolescent inpatients: treatment effects and association with borderline traits. J Pers Disord, 27(1), 3–​18.
doi: 10.1521/​pedi.2013.27.1.3
Sharp, C., Ha, C., Michonski, J., Venta, A., & Carbonne, C. (2012). The diagnosis of Borderline Personality
Disorder in adolescents: Evidence in support of the CI-​BPD in a sample of adolescent inpatients.
Comprehensive Psychiatry, 53(6), 765–​774.
Sharp, C., & Kalpakci, A. (2015). If it looks like a duck and quacks like a duck: Evaluating the validity
of borderline personality disorder in adolescents. The Scandinavian Journal of Child and Adolescent
Psychology and Psychiatry, 3(1), 49–​62.
Sharp, C., Kalpakci, A., Mellick, W., Venta, A., & Temple, J. R. (2014). First evidence of a prospective
relation between avoidance of internal states and borderline personality disorder features in
adolescents. Eur Child Adolesc Psychiatry, doi: 10.1007/​s00787-​014-​0574-​3
Sharp, C., & Kim, S. (2015). Recent advances in the developmental aspects of borderline personality
disorder. Current Psychiatry Reviews, 17(21), 1–​9.
Sharp, C., Pane, H., Ha, C., Venta, A., Patel, A. B., Sturek, J., & Fonagy, P. (2011). Theory of Mind and
Emotion Regulation Difficulties in Adolescents With Borderline Traits. Journal of the American
Academy of Child and Adolescent Psychiatry, 50(6), 563–​573. doi:10.1016/​j.jaac.2011.01.017
0
8
2

280 Emotion Dysregulation in Adolescents with Borderline Personality Disorder

Sharp, C., & Tackett, J. L. (2014). An idea whose time has come. In C. Sharp & J. L. Tackett (Eds.),
Handbook of Borderline Personality Disorder in Children and Adolescents (pp. 3–​8). New York: Springer.
Shiner, R. L. (2009). The development of personality disorders: Perspectives from normal personality
development in childhood and adolescence. Development and Psychopathology, 21(3), 715–​734.
doi:10.1017/​S0954579409000406
Solhan, M. B., Trull, T. J., Jahng, S., & Wood, P. K. (2009). Clinical Assessment of Affective
Instability: Comparing EMA Indices, Questionnaire Reports, and Retrospective Recall. Psychological
Assessment, 21(3), 425–​436. doi:10.1037/​A0016869
Steinberg, L., Dahl, R. E., Keating, D., Kupfer, D. J., Masten, A., & Pine, D. (2006). The study of
developmental psychopathology in adolescence: Integrating affective neuorsicence with the study
of context. In D. Cicchetti & D. Cohen (Eds.), Developmental Psychopathology (Vol. 2, pp. 710–​741).
New York: Wiley.
Stepp, S. D. (2012). Development of Borderline Personality Disorder in Adolescence and Young
Adulthood: Introduction to the Special Section. Journal of Abnormal Child Psychology, 40(1), 1–​5.
doi:10.1007/​s10802-​011-​9594-​3
Stone, A. A., & Shiffman, S. (1994). Ecological momentary assessment (EMA) in behavioral medicine.
Annual Behavioral Medicine 16(3), 199–​202.
Suvak, M. K., Litz, B. T., Sloan, D. M., Zanarini, M. C., Barrett, L. F., & Hofmann, S. G. (2011). Emotional
Granularity and Borderline Personality Disorder. Journal of Abnormal Psychology, 120(2), 414–​426. doi
10.1037/​A0021808
Tomko, R. L., Lane, S. P., Pronove, L. M., Treloar, H. R., Brown, W. C., Solhan, M. B., Wood, P. K., &
Trull, T. J. (in press). Undifferentiated Negative Affect and Impulsivity in Borderline Personality
Disorder: A Momentary Perspective. Journal of Abnormal Psychology, 124(3), 740–​753.
Tragesser, S. L., Lippman, L. G., Trull, T. J., & Barrett, K. C. (2008). Borderline personality disorder
features and cognitive, emotional, and predicted behavioral reactions to teasing. Journal of Research in
Personality, 42(6), 1512–​1523. doi:10.1016/​j.jrp.2008.07.003
Vanwoerden, S., Kalpakci, A. H., & Sharp, C. (2015). Experiential avoidance mediates the link between
maternal attachment style and theory of mind. [Research Support, Non-​U.S. Gov’t]. Compr Psychiatry,
57, 117–​124. doi: 10.1016/​j.comppsych.2014.11.015
von Ceumern-​Lindenstjerna, I. A., Brunner, R., Parzer, P., Mundt, C., Fiedler, P., & Resch, F.
Initial orienting to emotional faces in female adolescents with borderline personality disorder.
Psychopathology, 43(2), 79–​87. doi: 000274176 [pii] 10.1159/​000274176
Werner, K., & Gross, J. J. (2010). Emotion regulation and psychopatology: A conceptual framework. In A.
M. Kring & D. M. Sloan (Eds.), Emotion regulation and psychopathology: A transidagnostic approach to
etiology and treatment (pp. 13–​37). New York: Guilford Press.
Williams, K. D., & Jarvis, B. (2006). Cyberball: A program for use in research on interpersonal ostracism
and acceptance. Behavior Research Methods, 38(1), 174–​180. doi:10.3758/​Bf03192765
Woodberry, K. A., Gallo, K. P., & Nock, M. K. (2008). An experimental pilot study of response to
invalidation in young women with features of borderline personality disorder. Psychiatry Research,
157(1–​3), 169–​180. doi: 10.1016/​j.psychres.2007.06.007
Zanarini, M. C., Horwood, J., Wolke, D., Waylen, A., Fitzmaurice, G., & Grant, B. F. (2011). Prevalence of
Dsm-​Iv Borderline Personality Disorder in Two Community Samples: 6,330 English 11-​Year-​Olds and
34,653 American Adults. Journal of Personality Disorders, 25(5), 607–​619.
1
8
2

Chapter 14

Emotion Regulation in Severe


Irritability and Disruptive Mood
Dysregulation Disorder
Katharina Kircanski, Ellen Leibenluft,
& Melissa A. Brotman

Disruptive mood dysregulation disorder


By definition, dysfunction in normative emotion regulation is central to disruptive mood dys-
regulation disorder (DMDD), a new diagnosis in the Diagnostic and Statistical Manual of Mental
Disorders, fifth edition (DSM-╉5; American Psychiatric Association [APA], 2013). The hallmark
characteristic of DMDD is chronic, severe, and functionally impairing irritability. Here, irritabil-
ity refers to a propensity toward anger and is specifically operationalized as: 1) Recurrent severe
temper outbursts occurring at least three times per week, which are out of proportion to the
situation and inconsistent with developmental level; and 2) persistently irritable or angry mood
between outbursts, for most of the day, nearly every day. These core symptoms must be present
for at least one year across at least two of three settings (i.e., home, school, peers) and must begin
before age ten years; however, the diagnosis of DMDD cannot be made before age six years.
Given the recency with which DMDD was established as a formal diagnosis, data on its preva-
lence, clinical correlates, and outcomes are limited. In the first and largest epidemiological study
of the disorder, three-╉month prevalence estimates ranged from 0.8–╉3.3% across three community
samples of children and adolescents ages two to 17 years, notwithstanding the exclusion of chil-
dren younger than six years of age by the DSM-╉5 criteria (Copeland, Angold, Costello, & Egger,
2013). More recently, the three-╉month prevalence rate in a sample of six-╉year-╉olds (N = 462) was
reported to be 8.2% (Dougherty et al., 2014). No sex differences in prevalence rates were found in
either study. Importantly, these investigations also found DMDD to be associated with significant
functional impairment, particularly social role impairment, high rates of comorbid emotional
and behavioral disorders (Copeland et al., 2013; Dougherty et al., 2014), and high rates of service
use (e.g., mental health, school system; Copeland et al., 2013). Irritability itself in adolescence has
been associated with risk for suicidality in adulthood (Pickles et al., 2010). Therefore, the current
state of knowledge situates DMDD as a severe mood disorder that has serious consequences for
the health and daily functioning of those affected.

Severe mood dysregulation


The DSM-╉5 formulation of DMDD relied strongly on the syndrome of severe mood dysregulation
(SMD), operationalized by Leibenluft and colleagues (2003) for research purposes to examine
if chronic, severe irritability was a developmental presentation of bipolar disorder (reviewed in
2
8

282 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Leibenluft, 2011; Stringaris & Taylor, 2015). Unlike DMDD, SMD also included hyperarousal
symptoms such as insomnia, agitation, and distractibility, and required an onset of both chronic
irritability and hyperarousal symptoms before age 12 years. Consistent with prevalence estimates
of DMDD (Copeland et al., 2013), the lifetime prevalence of SMD was estimated to be 3.3% in a
community sample of children ages nine to 19 years (Brotman et al., 2006).
In a series of studies, SMD was compared to narrow phenotype bipolar disorder (NP-╉BD; i.e.,
a history of strictly defined, distinct episode(s) of euphoric mania/╉hypomania) in youth with
respect to pathophysiology, family history, and longitudinal outcomes. Evidence generally did not
support the conceptualization of SMD as a phenotype of bipolar disorder (see Leibenluft, 2011).
In fact, SMD (Brotman et al., 2006) and chronic irritability (Savage et al., 2015; Stringaris, Cohen,
Pine, & Leibenluft, 2009)  in children and adolescents were found to predict the development
of unipolar depressive and anxiety disorders, but not bipolar disorder, in adulthood. Additive
genetic factors were found to explain a substantial portion of the relation between irritability and
anxiety/╉depression (Savage et al., 2015). In addition, risk for future manic/╉hypomanic episodes
was documented to be 50 times higher in youth who met criteria for NP-╉BD than in youth with
SMD (Stringaris et al., 2010). Finally, youth who met criteria for SMD were significantly less likely
than those with NP-╉BD to have a parent diagnosed with bipolar disorder (Brotman et al., 2007).
Thus, although some youth with NP-╉BD may exhibit irritability while euthymic and/╉or increases
in irritability during mood episodes, evidence strongly suggests that chronic, severe irritability
is not a developmental precursor of bipolar disorder. Further, any history of manic/╉hypomanic
episodes is exclusionary for the diagnosis of DMDD, such that NP-╉BD and DMDD cannot be
comorbid.

Distinctions from associated conditions


DMDD is placed in the mood disorders section of the DSM-╉5 in recognition of its primary affec-
tive features and clinical associations with unipolar depression and anxiety, and to further dif-
ferentiate DMDD from other disorders (e.g., oppositional defiant disorder [ODD]). However, the
symptom criteria for DMDD and SMD do exhibit areas of overlap with those for ODD and atten-
tion deficit hyperactivity disorder (ADHD) (APA, 2013), and a sizeable proportion of youth with
DMDD or SMD meet concurrent criteria for ODD, ADHD, and/╉or conduct disorder (Brotman
et al., 2006; Dougherty et al., 2014; Roy et al., 2013). Supporting the conceptualization of irritabil-
ity as distinct from the full ODD syndrome, research by Stringaris and Goodman (2009ab) docu-
mented that the “irritable” dimension of ODD symptoms (in contrast to the “headstrong” and
“hurtful” dimensions) was the only feature of ODD to exhibit concurrent and prospective associa-
tions with major depressive disorder (MDD) and generalized anxiety disorder (GAD), consistent
with the pattern of associations observed for SMD. In addition, the range of clinical severity that
is seen in ODD does not correspond to the necessarily high level of severity that defines DMDD
(Leibenluft, 2011). Therefore, DSM-╉5 specifies that the DMDD diagnosis “trumps” ODD, such
that ODD should not be diagnosed in the context of DMDD. With respect to ADHD, whereas
SMD included hyperarousal symptoms in its definition, DMDD does not. Therefore, clinically
significant symptoms of hyperarousal, impulsivity, and/╉or inattention that accompany DMDD
should be designated as a comorbid diagnosis of ADHD when these features are present. Finally,
irritability is a symptom criterion of both GAD and MDD (APA, 2013). Irritability within the
syndrome of GAD must be associated with several other symptoms, including the cardinal fea-
ture of excessive and pervasive worry, which are not part of the criteria for DMDD or SMD; fur-
ther, irritability can present in GAD at varying levels of severity. Although DMDD can co-╉occur
with depression in children and adolescents (Copeland et al., 2013; Dougherty et al., 2014), the
3
8
2

The role of emotion (dys)regulation 283

expression of irritability in MDD is episodic by definition, and the diagnosis of DMDD should not
be made if irritability occurs only within the context of a depressive episode (APA, 2013).

The role of emotion (dys)regulation


General empirical approach
Research on the pathophysiological correlates of irritability has begun in the past decade. As
described earlier, the majority of empirical work in this area has focused on SMD and initially
examined whether youth with SMD could be distinguished from those with NP-╉BD. Further stud-
ies also have compared youth who met criteria for SMD to youth with other disorders and/╉or
youth with no history of psychiatric disorder. Integrating behavioral and biological metrics, this
research has aimed to identify the behavioral anomalies and dysfunctional neural circuitry that
differentiate SMD from both healthy functioning and other clinical syndromes, with the ultimate
goal of pinpointing biomarkers to facilitate diagnosis and intervention (Leibenluft, 2011). DMDD
is now the focus of ongoing and concerted research. Currently, it is unknown the extent to which
findings for SMD will generalize to DMDD; however, the vast majority of research participants
with SMD would have met diagnostic criteria for DMDD, suggesting that the pattern of findings
would be similar.
The field of affective neuroscience views the construct of emotion regulation broadly, as “any
process that maintains, accentuates, or attenuates emotional responses” (Davidson, Jackson, &
Kalin, 2000, p.  903). In this context, brain-╉based methodologies such as functional magnetic
resonance imaging (fMRI) have been used to characterize the neural mechanisms mediating
emotional processing as individuals with SMD engage in behaviors relevant to their condition.
In line with the knowledge that different neural pathways may mediate the same behavioral
deficits observed on tasks (Wilkinson & Halligan, 2004), neuroimaging techniques have been
particularly effective in elucidating neural circuit mechanisms of SMD that appear to differ from
those of NP-╉BD, even when behavioral impairments on the paradigms are similar across these
two groups.

Working model of pathologic irritability


Based on this research to date, we review below three domains of emotion regulation that have
been implicated in the pathophysiology of SMD: decreased context-╉sensitive regulation; dys-
regulated attention-╉emotion interactions; and misinterpretation of social-╉emotional stimuli. For
each domain, we describe the correlates of SMD with respect to 1) cognitive/╉behavioral impair-
ments and 2) neural function; we then integrate the results from these two levels of analysis in
summarizing each domain. Finally, we end our review with the most recent findings of pref-
erential processing of threat in SMD, which link with several findings in the other domains to
motivate new lines of ongoing research. Together, these contributing psychological processes
and neural circuits can be organized within an overarching systems neuroscience framework
for pathologic irritability, which both derives from the existing empirical literature and gener-
ates additional hypotheses to be tested (see Figure 14.1; revised from Leibenluft, 2011, Figure
2). Critically, underlying this framework is the conceptualization of irritability as a low (relative
to normative) threshold for experiencing negative affect in the context of frustration, in which
frustration refers to the emotional response when goal attainment is blocked or an action does
not produce the expected outcome (Blair, 2010; Leibenluft, 2011; Leibenluft et al., 2003). This is
consistent with the observed emphasis on anger as included in the clinical definition of DMDD
(APA, 2013).
4
8
2

284 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Amplification of Frustration

Dysregulated
attention-emotion
interactions

Decreased
Threshold

Misinterpretation
of social-emotional
stimuli

Blocked goal Increased irritability


Frustration
attainment Behavioral dyscontrol

Decreased
context-sensitive
Increased regulation
probability

Figure 14.1╇ Working etiopathological model of emotion dysregulation in chronic irritability, based


on reviewed empirical findings
Reproduced from Ellen Leibenluft, Severe Mood Dysregulation, Irritability, and the Diagnostic Boundaries of Bipolar
Disorder in Youths, The American Journal of Psychiatry, pp. 129–╉142, Figure 2 doi.org/╉10.1176/╉appi.ajp.2010.10050766
© American Psychiatric Association, 2011, with permission.

Decreased context-╉sensitive regulation


The human environment involves ever-╉changing situations and contingencies, and “context-╉
sensitive regulation” (Ochsner, 2008; p. 53), or the ability to detect and flexibly shift behaviors
in accord with such changes, promotes adaptive social functioning and emotion regulation. This
construct can be measured in the laboratory using response reversal tasks. In such tasks, over
a series of trials, participants initially learn to associate a particular stimulus (e.g., a geometric
shape on a computer screen) with the experience of reward (e.g., winning points by selecting that
shape among others presented simultaneously). Subsequently, reversal trials occur in which the
previously-╉rewarded stimulus is no longer followed by reward. Instead, participants must learn a
new set of contingencies (e.g., selecting a different geometric shape to continue obtaining points).
The task can be graded by utilizing a range of difficulty of contingency changes (e.g., using a
previously-╉irrelevant aspect of the stimuli as a new basis on which to reward responses), so that
correct responses on difficult trials require not only cognitive flexibility, but also the ability to
attend selectively and ignore irrelevant information. Impairments in mechanisms mediating these
abilities could increase the likelihood of experiencing blocked goal attainment, or of encounter-
ing situations in which outcomes are different than expected, causing more frequent or persistent
experiences of frustration (Blair, 2010; Leibenluft, 2011; Figure 14.1).

Behavioral findings
In the first investigation of reversal learning in SMD, Dickstein and colleagues (2007) compared
children and adolescents who met criteria for SMD to youth who met criteria for NP-╉BD and
5
8
2

The role of emotion (dys)regulation 285

non-╉psychiatric control participants. Compared to controls, participants with SMD and NP-╉BD
both made more errors on the difficult response reversal trials. However, the two clinical groups
did not differ from one another on these trials, suggesting that impairment in cognitive flexibility
is common to both syndromes. Dickstein et al. (2010) further examined the diagnostic specific-
ity of response reversal deficits by comparing groups of participants who met criteria for SMD,
NP-╉BD, unipolar depression, and anxiety disorders, as well as healthy controls. The probabilistic
response reversal task that was used in this study included both reward and punishment (point
loss) contingencies. In contrast to the earlier findings, participants with SMD did not differ sig-
nificantly from controls (or other clinical groups) on the primary performance metrics that were
assessed. However, the effect size associated with the comparison between SMD and control par-
ticipants was medium-╉to-╉large, so that the lack of a significant group difference may have reflected
insufficient statistical power. In secondary analyses, youth with SMD and NP-╉BD both exhibited
greater difficulty than did controls in using reward and punishment expectancies (i.e., learning on
previous trials) to facilitate adaptive responding to this task.

Neural findings
Adleman and colleagues (2011) used fMRI to assess the neural correlates of response reversal
learning in children and adolescents with SMD, NP-╉BD, and no history of psychiatric disorder.
The paradigm employed in the scanner was similar to that used by Dickstein et al. (2010) and had
been adapted previously for fMRI in a study of youth with ODD and conduct disorder (Finger
et al., 2008). Behaviorally, participants with SMD performed more poorly than did both the NP-╉
BD and control groups throughout the task, suggestive of a generalized deficit in contingency
learning. With respect to patterns of neural activation, on incorrect versus correct trials, both the
SMD and NP-╉BD groups failed to exhibit an increase in activity in the caudate nucleus that was
shown by healthy comparison youth. The caudate, a component of the striatum, supports motor
learning to enable behavioral adjustment following errors (Packard & Knowlton, 2002). Thus,
findings for this region directly implicate difficulty learning from errors in both SMD and NP-╉BD.
Interestingly, however, only the SMD group exhibited dysfunction in frontal activity when com-
pared to both the NP-╉BD and healthy comparison groups. Specifically, youth with SMD exhib-
ited hypoactivation in the inferior frontal gyrus, which supports the resolution of error-╉related
conflict through motor response selection and mediates key sub-╉processes, such as maintenance
of attention and cognitive representation of goals and contingencies (Budhani, Marsh, Pine, &
Blair, 2007). Activity in other regions, including frontal and cerebellar regions, was also found
to distinguish the SMD group and further implicated dysfunction in detecting and adapting to
errors in SMD.

Summary of findings
Therefore, across both behavioral and neural metrics, youth with SMD show impairments in the
ability to detect and flexibly respond to changing environmental contingencies of reward and
punishment. Whereas the behavioral deficits appear to be shared with NP-╉BD, the two conditions
differ in their mediating neural circuitry. In particular, SMD has been characterized by a unique
and pervasive pattern of frontostriatal dysfunction in the context of reversal learning, which may
increase the probability of experiencing blocked goal attainment or other unexpected outcomes
and heighten the resultant frustration response (Blair, 2010; Leibenluft, 2011).

Dysregulated attention-╉emotion interactions


As noted earlier, adaptive functioning requires the ability to selectively attend to motivation-
ally salient information in the environment while ignoring irrelevant information (Desimone
& Duncan, 1995). In addition, the recruitment of attentional strategies, such as flexibly shifting
6
8
2

286 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

attention toward or away from particular stimuli, can facilitate emotion regulation (Posner &
Rothbart, 1998) and behavioral self-╉control (Mischel, Shoda, & Rodriguez, 1989). Such norma-
tive attention-╉emotion interactions appear to be disrupted in children and adolescents with SMD,
especially in the context of frustration (see Figure 14.1). One commonly used task in this domain
is the affective Posner paradigm, which experimentally manipulates demands on participants’
attention and elicits frustration through the use of rigged performance feedback (Perez-╉Edgar &
Fox, 2005). On each trial, participants view two display frames on opposing sides of the computer
screen. Next, a cue (e.g., blue illumination) appears inside one of the frames, followed quickly by
a target in one of the two locations. Participants are instructed to indicate its location as quickly
and accurately as possible. A cue validity effect is typically observed, i.e., responses are relatively
slower on “invalid” trials, when the cue and target are presented in opposing locations. The affec-
tive adaptation of this task includes an initial baseline phase in which participants complete the
trials as described; a second phase in which participants win or lose money based on their perfor-
mance; and a critical third phase entailing blocked goal attainment in which participants receive
noncontingent negative performance feedback that results in monetary loss on a substantial pro-
portion of trials (i.e., frustration trials).

Behavioral findings
The affective Posner task has been conducted with SMD samples in three separate investigations,
all of which included both behavioral and neural assessments (Deveney et al., 2013; Rich et al.,
2007, 2011). Behaviorally, children and adolescents with SMD were found to differ consistently
from non-╉psychiatric comparison participants in their affective responses to the frustration trials.
As expected, SMD participants reported higher levels of arousal (Rich et al., 2007), unhappiness
(Rich et al., 2011), and frustration (Deveney et al., 2013) than did the comparison participants.
In the two studies that also included participants with NP-╉BD, these youth were similarly char-
acterized by heightened negative affect in response to frustration (Rich et al., 2007, 2011). With
respect to task performance, the pattern of findings is less consistent. In one study, participants
with both SMD and NP-╉BD were less accurate in their responses than healthy comparison youth
(Rich et al., 2007); however, another study did not document any group differences in behavioral
performance (Rich et al., 2011). Further, two findings were unique to SMD: These youth reported
greater arousal in response to negative performance feedback than did both NP-╉BD and healthy
comparison participants (Rich et al., 2011), and SMD youth were slower than healthy comparison
youth in responding to invalid trials during the frustration phase (Deveney et al., 2013). This lat-
ter effect, which we will return to in the neural findings, suggests decreased flexibility in spatial
attention in the context of frustration.

Neural findings
An electroencephalographic investigation by Rich and colleagues (2007) compared youth with
SMD, NP-╉BD, and no history of disorder in their event-╉related potentials (ERPs) to targets pre-
sented within the affective Posner paradigm. Despite similarities in the behavioral responses of
SMD and NP-╉BD participants in this study, ERP profiles diverged between these two syndromes.
Most notably, across all phases of the task, the SMD group exhibited lower N1 and P1 amplitudes
than did both the NP-╉BD and healthy comparison groups. The N1 and P1 components occur very
quickly following the presentation of a stimulus and reflect initial attentional orienting. Similar
reductions in N1/╉P1 amplitude have been reported for ADHD (Jonkman et al., 2000) and sug-
gest a generalized deficit in attentional orienting in SMD. In a follow-╉up study, Rich et al. (2011)
used magnetoencephalography to evaluate the neural responses of SMD, NP-╉BD, and typically-╉
developing participants to the performance feedback portion of the frustration trials. When
7
8
2

The role of emotion (dys)regulation 287

contrasting responses to negative feedback versus positive feedback, participants with SMD
uniquely showed heightened activity in the anterior cingulate cortex and medial frontal gyrus
relative to healthy comparison participants. Previous research has documented activity in these
brain structures to be associated with frustration (Moadab, Gilbert, Dishion, & Tucker, 2010),
which is consistent with the elevated arousal that was reported by SMD participants in response
to negative feedback in this study. Moreover, these results provide direct neurobiological evidence
for heightened frustration to blocked goal attainment in SMD.
Deveney and colleagues (2013) further explored the neural circuitry mediating frustration by
adapting the affective Posner task for fMRI with SMD and healthy comparison participants. In
response to frustration trials on which participants received negative feedback, SMD youth were
characterized by widespread neural deactivations, both relative to comparison youth in this con-
dition (i.e., between-╉group differences) and relative to their own responses to positive feedback
(i.e., within-╉group differences). In particular, SMD youth showed deactivations in the left and
right striatum, left amygdala, posterior cingulate cortex, and parietal cortex. Striatal deactiva-
tion is known to occur frequently in conjunction with negative prediction error, or conditions
in which an outcome is worse than expected (Schultz, Dayan, & Montague, 1997; Schultz, 2010).
Extrapolating from this knowledge and consistent with the response reversal findings, greater
deactivation in this region suggests that SMD youth experience blocked goal attainment as more
unexpected and/╉or unpleasant than do healthy youth. Deactivations in the amygdala and poste-
rior cingulate are more difficult to interpret, although as we review in the next section, Thomas
et  al. (2013) similarly found in SMD attenuated amygdala and posterior cingulate activity to
high levels of facial anger intensity. Finally, the parietal cortex plays a key role in spatial attention
(Corbetta, Kincade, Ollinger, McAvoy, & Shulman, 2000), and deactivation in this area suggests
that frustration reduces attentional flexibility for youth with SMD, which may have contributed to
their slowed performance on invalid trials when frustrated.
Summary of findings
The affective Posner paradigm has generated both behavioral and neural data to support the pro-
posed pathophysiology of SMD. In response to blocked goal attainment, children and adolescents
with SMD show heightened activity in brain structures associated with frustration and, corre-
spondingly, they report elevated levels of negative affective arousal. In addition, unique striatal
and parietal cortex dysfunctions suggest that, relative to healthy comparison youth, youth with
SMD experience blocked goal attainment as more unexpected and have limited attentional flex-
ibility to regulate their frustration.

Misinterpretation of social-╉emotional stimuli


As a third domain of emotion regulation, motivationally salient cues in one’s environment are
frequently social, i.e., others’ emotional expressions that directly or indirectly signal information
about current circumstances, such as an angry expression communicating hostility or a fearful
expression communicating the presence of a threat. The ability to identify such cues promotes
self-╉regulation and social adaptation (Ochsner, 2008). This construct has been translated to the
laboratory using a variety of experimental paradigms. Most of these tasks involve the presenta-
tion of controlled visual stimuli (e.g., facial expressions of emotions) over a series of trials in
order to characterize behavioral and neural responses as a function of varying stimulus char-
acteristics (e.g., the specific emotion expressed, the intensity of the emotion). Misinterpretation
of such social-╉emotional information is another process that has been found to broadly charac-
terize youth with SMD and is postulated to amplify their frustration (see Figure 14.1). Perhaps
especially for children and adolescents, the experience of blocked goal attainment often occurs
82

288 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

during social interactions (e.g., being told by one’s parents to stop playing a video game or another
preferred activity). Misinterpretations of others’ emotion cues in the setting of these interactions
(e.g., failing to identify the parent’s angry expression or, conversely, mislabeling the parent’s neu-
tral expression as angry) might exacerbate frustration in response to the precipitating events.

Behavioral findings
Guyer and colleagues (2007) conducted the first study of face emotion labeling in SMD, compar-
ing these youth to NP-╉BD, anxiety and/╉or unipolar depressive disorders, ADHD and/╉or conduct
disorder, and non-╉psychiatric comparison youth. Participants viewed child and adult faces dis-
playing angry, fearful, sad, or happy emotions. On each trial, participants indicated the emotion
being expressed. The SMD and NP-╉BD groups both made more identification errors than did all
other groups, and did not differ from one another. Further, in both SMD and NP-╉BD this impair-
ment was generalized across child and adult faces and across all emotions. In a subsequent eye
tracking study, Kim et al. (2013) replicated this broad deficit in face emotion labeling in SMD and
NP-╉BD relative to healthy comparison participants. However, whereas NP-╉BD was characterized
by reduced visual attention to the eyes in the facial stimuli, the pattern of visual attention to the
eyes in youth with SMD fell in between that in NP-╉BD and comparison participants. These data
suggest that, in SMD, the observed deficit in face emotion labeling is not fully explained by insuf-
ficient visual attention to emotion-╉relevant cues in others’ eyes.
In a more fine-╉grained examination of face emotion labeling in SMD, NP-╉BD, and healthy com-
parison participants, Rich et al. (2008) assessed the intensity of expression that was required to
label emotions correctly. Each trial began with the presentation of a neutral facial expression,
which gradually increased in emotional intensity to result in anger, fear, disgust, sadness, happi-
ness, or surprise. Participants were instructed to indicate, once they were aware, the specific emo-
tion being expressed by each face, and they were allowed to change their responses as each trial
progressed further. Across most emotions, both SMD and NP-╉BD participants required greater
intensity of facial expressions than did the comparison participants in order to label the emo-
tions correctly. In addition, within the SMD group, greater impairment in face emotion identifica-
tion was associated with poorer family functioning, including general relationships with family
members and typical interactions surrounding family rules, chores, and solving problems. Finally,
Deveney and colleagues (2012) examined children’s ability to label nonverbal emotional cues con-
veyed vocally. Extending the earlier findings for emotions presented visually, both the SMD and
NP-╉BD groups were found to perform more poorly than a healthy comparison group in identify-
ing the emotion in others’ speech.

Neural findings
Brotman and colleagues (2010) used an fMRI paradigm to investigate functional activity of the
amygdala to facial expressions in youth with SMD, NP-╉BD, ADHD, and no history of disorder. On
each trial participants used a rating scale to respond to one of four questions about the face being
presented. Importantly, some responses engaged attention to its emotional aspects (e.g., “How
afraid are you of this face?”) while other responses engaged attention to its non-╉emotional aspects
(e.g., “How wide is the nose?”). Behaviorally, participants with both SMD and NP-╉BD reported
being more afraid of the neutral faces than did ADHD and healthy participants. However, amyg-
dala responses to the neutral faces distinguished SMD and NP-╉BD. Specifically, participants with
SMD exhibited hyperactivity in the left amygdala when rating the nose width of the faces (i.e., dur-
ing implicit emotional processing), but hypoactivity in the left amygdala when rating their fear of
the faces (i.e., during explicit emotional processing). This response pattern was not shown in the
NP-╉BD, ADHD, or healthy comparison group. The amygdala is critical to detecting the emotional
9
8
2

The role of emotion (dys)regulation 289

salience of environmental stimuli, particularly in terms of threat value (Davis, 1992). Thus, in
direct contrast to the findings for SMD, greater amygdala activity is normative when attending to
the emotional versus non-​emotional aspects of a stimulus. Similarly aberrant findings of amyg-
dala hypoactivity to face emotion have been reported in children diagnosed with MDD (Beesdo
et al., 2009; Thomas et al., 2001), which is intriguing in light of the reviewed cross-​sectional and
longitudinal associations between SMD/​DMDD and unipolar depression (Brotman et al., 2006;
Copeland et al., 2013; Dougherty et al., 2014; Savage et al., 2015; Stringaris et al., 2009).
The brain-​based assessment of face emotion processing in SMD was recently advanced through
three programmatic studies by Thomas and colleagues (2012, 2013, 2014), all of which included
participants diagnosed with SMD, NP-​BD, and no history of disorder. First, Thomas et al. (2012)
examined the modulation of neural activity to faces that varied systematically along intensity
gradients of neutral to angry expressions and neutral to happy expressions; participants processed
faces both implicitly and explicitly in this task. In response to increasing intensity of anger expres-
sions, typically-​developing participants exhibited parametric increases in activity in the amygdala
and posterior cingulate cortex, the latter of which similarly activates to emotional stimuli and
participates in motivation-​driven allocation of spatial attention (Mohanty, Gitelman, Small, &
Mesulam, 2008). Both SMD and NP-​BD participants, however, failed to modulate amygdala or
posterior cingulate activity in this manner, indicating impoverished neural responsivity to higher
levels of anger intensity. With respect to increasing intensity of happy expressions, SMD partici-
pants exhibited a distinct frontoparietal pattern that entailed low initial activity (i.e., in response
to neutral faces) coupled with increasing activity in several regions that implement attention
(Kanwisher & Wojciulik, 2000), face processing (Kanwisher, 2000), and emotion processing in
a social context (Beer & Ochsner, 2006). Considered with the behavioral deficits in face emo-
tion labeling that typify SMD, increasing frontoparietal activity to happiness intensity may reflect
greater effort required to correctly identify others’ happy expressions.
Thomas et al. (2013) focused on implicit emotional processing of angry, fearful, and neutral
facial expressions, utilizing a common fMRI paradigm in which participants indicated the gender
of each of a series of faces. Across all emotion types, participants with both SMD and NP-​BD
exhibited hyperactivity in the right amygdala (i.e., increased activity during implicit emotional
processing); this finding for SMD is notably similar to that in Brotman et al. (2010). In response
to fearful expressions, only participants with SMD showed deactivation in several medial brain
regions that comprise a “default mode network” (Raichle & Snyder, 2007), which, among several
formulated functions, monitors information about one’s internal state. Extrapolating from cur-
rent knowledge of this network, the findings suggest that youth with SMD may not appropriately
monitor interoceptive cues in response to fearful stimuli to facilitate their identification.
Thomas et  al. (2014) integrated the assessment of automatic face emotion processing that
occurs outside of one’s conscious awareness. The authors used a backwards masking paradigm in
which, on a portion of trials (“non-​aware” trials), facial expressions were presented subliminally.
On other trials the faces were presented quickly but supraliminally, or of sufficient duration for
awareness (“aware” trials). Youth with both SMD and NP-​BD exhibited greater activity in occipi-
tal regions during “non-​aware” than “aware” trials of all face types, which was the opposite of the
occipital response pattern shown in healthy comparison participants and implicates disruption
in basic ventral visual stream functions (e.g., object recognition) in these two clinical syndromes
(Goodale & Milner, 1992). However, collapsing across “aware” and “non-​aware” trials of angry
faces, only youth with SMD showed elevated activity in several regions that support higher-​order
face processing and social cognition, including the posterior cingulate cortex, superior temporal
gyrus, and middle occipital gyrus (Allison, Puce, & McCarthy, 2000; Gallagher & Frith, 2003).
Finally, a recent replication study by Tseng and colleagues (2016) compared youth with SMD to
0
9
2

290 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

healthy comparison participants in the neural correlates of masked and unmasked face emotion
processing. Again, when viewing angry faces, youth with SMD exhibited hyperactivity relative to
comparison youth in several brain regions associated with face and emotion processing, such as
the parahippocampal gyrus and superior temporal gyrus. Hyperreactivity of these structures to
both subliminal and rapid supraliminal anger expressions suggests heightened neural sensitivity
toward anger.

The threat system: A new research domain in DMDD


Finally, two recent studies found youth with SMD or DMDD to preferentially process threaten-
ing information (Hommer et al., 2014; Stoddard et al., 2016). This tendency may increase their
likelihood of perceiving a goal as being blocked, further fueling frustration and problematic social
interactions. In the first investigation, Hommer and colleagues (2014) found youth with SMD
to demonstrate heightened attention toward threat. The authors used a dot-╉probe paradigm in
which participants were presented simultaneously with one emotional facial expression (angry or
happy) and one neutral expression over a series of trials. These face pairs were followed immedi-
ately by a probe (asterisk) in one of the two locations where a face had been, and participants were
instructed to indicate the location of the probe as quickly and accurately as possible. Biased alloca-
tion of visual attention is indexed using latencies to respond to the probe. Participants with SMD
exhibited a greater attentional bias toward angry faces, but not happy faces, compared to healthy
comparison participants. Intriguingly, this same threat-╉relevant attentional bias has been docu-
mented reliably in youth with anxiety disorders (Waters, Henry, Mogg, Bradley, & Pine, 2010).
Importantly, however, the significant difference between SMD and healthy participants was not
driven by concurrent anxiety (or depression) within the SMD group.
Most recently, Stoddard et  al. (2016, Study 1)  demonstrated that DMDD is associated with
a hostile interpretation bias, or elevated tendency to interpret ambiguous facial expressions as
angry. In this paradigm, participants viewed and classified as either “happy” or “angry” a series
of facial expressions that varied along a continuum of happy to angry and were presented in
random order. Critically, faces in the middle of the continuum were ambiguous in expression, as
they represented morphed images of both happy and angry faces. Youth with DMDD were more
likely than were healthy comparison participants to classify these ambiguous faces as angry. Taken
together, these recent findings indicate that chronic, severe irritability is associated with both
enhanced attention to, and interpretation of, threat in face emotion stimuli.
In this context, it is intriguing to consider neurobiological formulations of an overarching
threat-╉response system, a circuit that includes the amygdala, hypothalamus, and periaqueductal
gray and is posited to mediate the emotions of both anger and fear (Blair, 2010; Johansen, Tarpley,
LeDoux, & Blair, 2010; Panksepp, 2006). The concept of threat system dysregulation in SMD also
integrates some of the reviewed findings for other domains in which these youth reported height-
ened fear of neutral faces (Brotman et al., 2006) and showed elevated neural sensitivity to anger
expressions that were presented subliminally or rapidly (Thomas et al., 2014; Tseng et al., 2016).
Further, threat system dysregulation may underlie the concurrent and prospective associations
between SMD/╉DMDD and anxiety disorders (Brotman et al., 2006; Copeland, Shanahan, Egger,
Angold, & Costello, 2014; Savage et al., 2015; Stringaris et al., 2009). For all of these reasons, the
threat system is a focus of ongoing pathophysiological research in DMDD.

Summary of findings
In sum, SMD is characterized by extensive perturbations in face emotion processing.
Behaviorally, youth with SMD show broad deficits in identifying others’ emotional expressions
1
9
2

PSYCHOTHERAPEUTIC INTERVENTIONS 291

and detecting subtle expressions. NP-╉BD shares these behavioral impairments and evidence for
basic neural dysfunction in the ventral visual stream; however, other neural disruptions appear
unique to SMD. Although the precise findings have varied across studies (and these response
patterns may not generalize to non-╉facial emotional stimuli; see Rich et  al., 2007, 2010), an
overall pattern emerges in which youth with SMD seem to miscalibrate neural activity to faces
with respect to the appropriate depth of emotional processing (implicit versus explicit) and
the intensity or salience of emotional expression (neutral versus highly expressive). In addi-
tion, recent findings indicate that irritable youth preferentially process threatening face emo-
tions (Hommer et al., 2014; Stoddard et al., 2016). In the setting of everyday family and social
interactions, such miscalibrations may result in inappropriate responses, such as anger, toward
others.

Evidenced-╉based interventions
State of the literature
Currently, there are no well-╉established, evidence-╉based treatments specifically developed for
SMD or DMDD. However, a number of psychotherapeutic and psychopharmacological interven-
tions have been developed for related clinical syndromes (e.g., disruptive behavior disorders) or
selected symptoms that are common in SMD/╉DMDD (e.g., aggression, noncompliant behavior).
Below, we review these interventions and their evidence base with respect to the clinical symp-
toms or syndromes for which they were tested. Given the independent contributions of chronic,
severe irritability in youth to adverse outcomes, and the dearth of empirically-╉supported treat-
ments, there is a great need for the development of novel therapeutic approaches (Leibenluft,
2011; Stoddard et  al., 2016; Waxmonsky et  al., 2013). At the end of this section, we highlight
several novel approaches with promising initial findings.

Psychotherapeutic interventions
Numerous psychotherapeutic interventions have been tested for disruptive behavior and conduct
problems (Weisz, Jensen-╉Doss, & Hawley, 2006). Extant approaches fall into two general catego-
ries: Parent management training and cognitive-╉behavioral therapy. These may be delivered as
stand-╉alone treatments or in combination with one another. In this section, we briefly highlight
these interventions. In the final section of the chapter, we describe in greater detail two therapeu-
tic methods that may specifically target emotion dysregulation in SMD/╉DMDD.

Parent management training


Parent management training (PMT) is a type of behavioral treatment in which parents learn to
no longer engage in dysfunctional interactions with their children. The goal is to reduce child
disruptive behaviors, chiefly noncompliance and aggression, as manifested prototypically in
ODD and conduct disorder (reviewed in Kazdin, 2010; Sukhodolsky, Smith, McCauley, Ibrahim,
& Piasecka, 2016). Given its strong evidence base developed over several decades of research,
PMT has achieved the status of a “well-╉established” treatment (Chambless et al., 1996, 1998) for
disruptive behavior disorders (Eyberg, Nelson, & Boggs, 2008). Indeed, group-╉based PMT for
parents of children ages three to 11 years with oppositional defiant disorder and conduct dis-
order is included in the United Kingdom’s National Institute for Health and Clinical Excellence
official guidelines. However, the immediate benefits reported by parents may not generalize to
other contexts (e.g., with teachers) and may be difficult to sustain over the long term (e.g., one
year) (Pilling et al., 2013).
2
9

292 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Typically conducted with the parent(s) only, PMT techniques are grounded in parent-╉child
interaction research by Patterson and colleagues (Patterson, 1982; Patterson, DeBaryshe, &
Ramsey, 1989; Patterson, Reid, & Dishion, 1992). This work draws heavily on operant condi-
tioning (i.e., instrumental learning), in which behavior is shown to be shaped by the positive
and negative outcomes that follow it. Specific PMT techniques include instructing parents
in the use of positive reinforcement, selective attention and ignoring, and mild negative con-
sequences (reviewed in Kazdin, 2010; Sukhodolsky et al., 2016). Numerous clinician protocols
are available, including Kazdin’s (2005) 12-╉session and Barkley’s (2013) ten-╉session manuals. In
addition, a number of broader treatment programs for childhood disruptive behavior disorders
have a parent training component (e.g., Incredible Years, Webster-╉Stratton & Reid, 2010; Parent-╉
Child Interaction Therapy, Zisser & Eyberg, 2010; Positive Parenting Program/╉Triple-╉P, Sanders,
1999). A  large meta-╉analysis reported that, across numerous treatment packages that entailed
some degree of parent training, factors associated with the largest effect sizes included a focus on
enhancing parental consistency, positive interactions, and affective communication, and the use
of in-╉session practice with the child (Kaminski, Valle, Filene, & Boyle, 2008). Efforts are underway
to adapt PMT to disruptive behavior in the context of other diagnoses, such as autism spectrum
disorders (reviewed in Sukhodolsky et al., 2016).

Cognitive-╉behavioral therapy
In contrast to PMT, cognitive-╉behavioral therapy (CBT) is conducted directly with the child, aim-
ing to help the child learn skills to reduce maladaptive responses to everyday situations (reviewed
in Kazdin, 2010; Sukhodolsky et  al., 2016). Several CBT protocols have been developed for
school-╉age children and adolescents with disruptive behavior. Evidence supports these treatments
as “probably efficacious” (Eyberg et al., 2008), consistent with a meta-╉analysis by Sukhodolsky and
colleagues (2004) that reported a medium effect size of these interventions on outcomes.
CBT involves both in-╉session therapeutic techniques and out-╉of-╉session practice by the child,
often with monitoring of practice by the parent. The skills taught in CBT for disruptive behavior
draw on several theories, but most centrally the work of Dodge and colleagues regarding chil-
dren’s social information processing (Crick & Dodge, 1994; Dodge, 1980, 2003). In this frame-
work, cognitive processing of social cues comprises a series of steps: Encoding the cues of others,
interpreting these cues, searching for a response, selecting a response, and engaging in the chosen
response. Dysfunction in one or more of these processes is posited to underlie anger and aggres-
sive responding. Notably, the first two steps of this model (implicating impairments in encoding
and interpreting social cues) align with the findings, reviewed above, for misinterpretation of
social-╉emotional stimuli in SMD/╉DMDD, particularly in terms of heightened sensitivity to threat
cues (Stoddard et al., 2016; Thomas et al., 2014; Tseng et al., 2016). CBT emphasizes more adap-
tive ways to think about, and respond to, social situations that elicit anger and aggression. Specific
CBT protocols, among others, include Kazdin’s (2010) 20-╉to 25-╉session Problem-╉Solving Skills
Training (PSST), Sukhodolsky and Scahill’s (2012) ten-╉session treatment, and Lochman and col-
leagues’ (2010) group-╉based Anger Control Training. In the meta-╉analysis of CBT for disruptive
behavior, clinician techniques associated with the largest effect sizes were training in social skills,
modeling appropriate behaviors, providing direct feedback to the child, and assigning homework
to be done outside of the session (Sukhodolsky et al., 2004).

Psychopharmacological interventions
As described earlier, clinical and pathophysiological research has strongly suggested that SMD
is not a developmental presentation of bipolar disorder. This work has important treatment
3
9
2

Psychotherapeutic interventions 293

implications, particularly with respect to psychopharmacological interventions. First-╉line treat-


ments for NP-╉BD include mood stabilizers and atypical antipsychotics, which have considerable
side effect burden (Correll et al., 2009). Indeed, a major impetus of the research on SMD was con-
cern over the increasing use of these medications to treat nonepisodic irritability (Olfson, Blanco,
Liu, Moreno, & Laje, 2006). Dickstein et al. (2009) investigated the use of lithium to treat youth
with SMD in a small, six-╉week double-╉blind placebo-╉controlled trial. There was no significant
advantage of lithium over placebo, consistent with findings that the pathophysiology of SMD is
distinct from that of NP-╉BD. Divalproex, an antiepileptic medication that also may be used as a
mood stabilizer in NP-╉BD, has received slightly stronger support to target irritability and aggres-
sion, especially in youth with ADHD, although it has not been tested for SMD or DMDD directly
(Blader, Schooler, Jensen, Pliszka, & Kafantaris, 2009; Donovan et  al., 2000; Steiner, Petersen,
Saxena, Ford, & Matthews, 2003).
Further, whereas stimulants and selective serotonin reuptake inhibitors (SSRIs) are contraindi-
cated for NP-╉BD, these medications might be effective to treat SMD (or DMDD with co-╉occurring
ADHD) as the syndrome is conceptualized as irritability and hyperarousal symptoms that are lon-
gitudinally and genetically associated with unipolar depression and anxiety (Savage et al., 2015).
Two such treatment trials are underway to test the efficacy of stimulant plus SSRI versus stimulant
plus placebo (clinicaltrials.gov identifiers NCT00794040, NCT01714310). These studies reflect
the growing attention of the field to chronic, severe irritability as a broader clinical phenotype
(Hulvershorn, Fosselman, Dickstein, & Janicak, 2012ab). Previous studies of stimulant medication
for ADHD reported a decrease in aggression as a secondary outcome. A meta-╉analysis of these
studies indicated a medium-╉to-╉large effect size on reductions in aggressive behavior (Connor,
Glatt, Lopez, Jackson, & Melloni, 2002).
Finally, it is notable that irritability in the context of autism spectrum disorders has been the
symptom target in previous medication trials. As a result of such trials, risperidone and aripipra-
zole, second-╉generation antipsychotics, received FDA indication for the treatment of irritability
in autism (Marcus et al., 2009; McCracken et al., 2002).

Development of novel interventions


The review of extant treatment options makes clear the need for novel, pathophysiologically-╉
driven interventions specifically designed for DMDD. Several investigations have reported
promising preliminary findings. First, Stoddard et  al. (2016, Study 3)  furthered their inves-
tigation of hostile interpretation bias in DMDD by testing the efficacy of a computer-╉based
program to train away this bias. In an open trial, 14 youth with DMDD viewed a series of facial
expressions that were visually morphed along a continuum of happy to angry. As in the origi-
nal paradigm (Study 1), participants classified each face as either happy or angry. However, in
response to ambiguous faces in the middle of the continuum, participants were given feedback
from the computer. Specifically, participants were trained to classify faces that they originally
identified as “angry” as “happy.” Thus, youth with DMDD learned to make less hostile and more
benign interpretations of the ambiguous faces. Training over the course of four days was associ-
ated with a decrease in irritability symptoms and changes in lateral orbitofrontal activation to
happy versus angry faces. Together with another published trial in youth at-╉risk for aggressive
behavior (Penton-╉Voak et  al., 2013), these results provide the justification for a randomized
controlled trial of interpretation bias training in DMDD that is currently being conducted by
our group.
Second, Miller and colleagues (2015) have adapted empirically-╉supported interpersonal ther-
apy (IPT) for mood disorders to adolescents with SMD, based on the significant social-╉emotional
4
9
2

294 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

impairments documented in this population. A recent proof-╉of-╉concept study indicated excel-


lent participant attendance, high participant satisfaction, and overall improvement in core SMD
symptoms. A randomized controlled trial is presently being conducted by this research group.
Third, Waxmonsky and colleagues (2013) have developed a novel group therapy for youth with
ADHD and co-╉occurring SMD. The multimodal treatment incorporates elements of both PMT
and CBT. In a pilot trial, all 7 participants also were taking stimulant medication. Reductions in
mood lability, depressive symptoms, and overall child functioning were observed, although the
effect sizes for changes in disruptive behavior were not as large.

Selected psychotherapeutic techniques


Therapeutics through the lens of emotion regulation
As discussed earlier, the pathophysiology of chronic, severe irritability involves several disrup-
tions in normative emotion regulation. First, children and adolescents with SMD are less able
to regulate their behavior to changing environmental contingencies of reward and punishment
(Dickstein et al., 2007, 2010), and they exhibit pervasive frontostriatal dysfunction in this context
relative to healthy comparison youth (Adleman et al., 2011). Second, when an expected reward is
omitted, youth with SMD show hyperactivity in brain regions associated with frustration (Rich
et al., 2011), and hypoactivity in brain regions that mediate spatial attention, suggesting lim-
ited attentional flexibility to regulate frustration (Deveney et al., 2013). Third, youth with SMD
or DMDD miscalibrate their neural and behavioral responses to social-╉emotional stimuli, and
are especially sensitive to threatening facial expressions (Hommer et al., 2014; Stoddard et al.,
2016; Thomas et al., 2014; Tseng et al., 2016), which may lower their threshold for anger in social
situations.
Given that pathophysiological research on SMD began only in the past decade, it is not surpris-
ing that interventions specifically indicated for SMD/╉DMDD remain under development. Rather,
it is promising that several lines of psychotherapeutic and psychopharmacological research are
already underway. Notwithstanding future advances in treatment, it may be useful to consider
extant therapeutic techniques and how they might help to address the deficits in emotion regula-
tion that characterize chronic, severe irritability. Here, we examine two different psychotherapeu-
tic techniques through this lens.

Labeling irritable emotion


The “Feeling Thermometer” is one of the first tools introduced to patients in many CBT protocols
for disruptive behavior (e.g., Sukhodolsky & Scahill, 2012, p. 30), as in CBT packages for other
disorders (e.g., anxiety disorders; Kendall & Hedtke, 2006). In this technique, the clinician visu-
ally presents a thermometer or scale that has a simple numerical range (e.g., zero to ten) and
describes that the scale can be used to rate the intensity of angry, frustrated, or irritable feelings.
Through clinician prompts and discussion, the child learns to label his emotions verbally and
assign numerical values to different intensities of emotion that are elicited in different situations.
For example, a child may assign a “five” out of ten (moderate anger, irritability, or frustration) to a
scenario in which his parent asks him to stop playing a video game and come to the dinner table.
Some youth may be able to generate such scenarios independently, whereas others may require
the clinician to provide examples. The Feeling Thermometer also is useful to rate emotional inten-
sity when the child is actively experiencing irritability or anger in session.
With regard to emotion regulation, a growing literature supports the neural and behav-
ioral benefits of verbalizing negative emotions (reviewed in Lieberman, 2011). Neuroimaging
5
9
2

Selected psychotherapeutic techniques 295

studies document that assigning verbal labels to emotionally evocative images, versus viewing
those images, serves to decrease activity in the amygdala and increase activity in the right
ventrolateral prefrontal cortex (Lieberman et al., 2007; Tabibnia, Lieberman, & Craske, 2008).
Moreover, clinical analog studies have shown that labeling one’s emotions during exposure to
a feared stimulus reduces physiological arousal relative to a variety of comparison conditions
(Kircanski, Lieberman, & Craske, 2012; Niles, Craske, Lieberman, & Hur, 2015). Given the
emotion dysregulation and aberrant patterns of prefrontal and amygdala activity that char-
acterize SMD and DMDD (Adleman et al., 2011; Brotman et al., 2010; Thomas et al., 2014;
Tseng et  al., 2016), it is possible that verbally labeling irritable emotion may be helpful for
these youth. In addition, Feeling Thermometer ratings can be used to measure the success of
other emotion regulation strategies (e.g., cognitive restructuring) to decrease their irritability,
anger, or frustration.

Consistency in parental contingencies for behavior


As reviewed earlier, PMT teaches parents to be more consistent in delivering positive and mild
negative consequences for child behaviors, in order to shape these behaviors through instrumen-
tal learning. Typically, the clinician introduces specific therapeutic techniques in a prescribed
order; for detailed descriptions, see Kazdin (2005) and Barkley (2013). In general, however, early
sessions focus on enhancing positive reinforcement for adaptive and desired behaviors. For exam-
ple, in the case of a child asked to stop playing a video game and come to the dinner table, the
parent may positively attend to, praise, and/╉or provide a secondary reward (e.g., token or point)
in response to the child complying with this request. In later sessions, parents learn to administer
mild negative parental consequences or time-╉out from reinforcement in response to maladaptive
behaviors. For instance, if the child were to respond argumentatively to the request to come to
the dinner table, the parent may ignore this behavior or fail to provide the secondary reward. The
clinician works with the parent(s) intensively to achieve greater consistency in their responses
while minimizing negative, punitive reactions to the child.
Researchers have long recognized the influence of parental attention and contingencies on
children’s emotional and behavioral regulation (reviewed in Morris, Silk, Steinberg, Myers, &
Robinson, 2007). Given the deficits in instrumental learning that have been documented in
youth with SMD (Adleman et al., 2011; Dicksten et al., 2007, 2010), these deficits may be exac-
erbated by parental inconsistency in positive and negative consequences for behavior. Further,
as SMD and DMDD have been associated with the preferential processing of threat stimuli,
these youth may be particularly sensitive to their parents’ negative emotional expressions and
punitive consequences (Hommer et al., 2014; Stoddard et al., 2016). Indeed, Patterson and col-
leagues’ early research on the “coercive family process” (Patterson, 1982) in youth with conduct
problems demonstrated that parents’ negative, punitive reactions to child disruptive behaviors
tended to increase rather than decrease the frequency of problem behaviors. Additional studies
by Eisenberg and Fabes (1994) and Snyder et al. (2003) have shown that dismissing and punitive
parental responses to child expressions of anger are associated with anger amplification and poor
behavioral coping (e.g., escaping from the situation). Therefore, enhancing parental consistency
and minimizing punitive parenting through positive parent training techniques may be useful
components of treatment. Finally, it is important in this context to recall the heightened neural
and affective responses to blocked goal attainment that characterize SMD (Deveney et al., 2013;
Rich et al., 2007, 2011); failure to receive an expected reward from parents may be particularly
frustrating for these youth. This potential to elevate frustration will be an important area for
future work to address in adapting parent training techniques to chronic, severe irritability.
6
9
2

296 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Summary and need for continued research


In sum, it can be useful to consider how therapeutic techniques that are currently available to
clinicians may help to target pathophysiological processes in SMD/╉DMDD. Comprehensive treat-
ments for these patients likely will integrate well-╉conceptualized tools from extant interventions
with novel advances tailored to the mechanisms of chronic irritability. At present, DMDD is a
severe and impairing psychiatric disorder without any validated psychotherapeutic or psycho-
pharmalogical interventions. Many open, yet tractable, questions concerning its pathophysiology
and treatment remain, with the opportunity for investigators to make new and clinically meaning-
ful contributions.

References
Adleman, N. E., Kayser, R., Dickstein, D., Blair, R. J. R., Pine, D., & Leibenluft, E. (2011). Neural
correlates of reversal learning in severe mood dysregulation and pediatric bipolar disorder. Journal of
the American Academy of Child and Adolescent Psychiatry, 50(11), 1173–╉1185.
Allison, T., Puce, A., & McCarthy, G. (2000). Social perception from visual cues: role of the STS region.
Trends in Cognitive Sciences, 4(7), 267–╉278.
American Psychiatric Association. (2013). Diagnostic and Statistical Manual of Mental Disorders (5th ed.).
Washington, DC: American Psychiatric Publishing.
Barkley, R. A. (2013). Defiant children: A clinician’s manual for assessment and parent training. New York,
NY: Guilford Press.
Beer, J. S., & Ochsner, K. N. (2006). Social cognition: a multi level analysis. Brain Research, 1079(1),
98–╉105.
Beesdo, K., Lau, J. Y. F., Guyer, A. E., McClure-╉Tone, E. B., Monk, C. S., Nelson, E. E., … Pine, D. S.
(2009). Common and distinct amygdala-╉function perturbations in depressed vs anxious adolescents.
Archives of General Psychiatry, 66(3), 275–╉285.
Blader, J. C., Schooler, N. R., Jensen, P. S., Pliszka, S. R., & Kafantaris, V. (2009). Adjunctive divalproex
versus placebo for children with ADHD and aggression refractory to stimulant monotherapy. American
Journal of Psychiatry, 166(12), 1392–╉1401.
Blair, R. J. R. (2010). Psychopathy, frustration, and reactive aggression: the role of ventromedial prefrontal
cortex. British Journal of Psychology, 101(3), 383–╉399.
Brotman, M. A., Kassem, L., Reising, M. M., Guyer, A. E., Dickstein, D. P., Rich, B. A., … Leibenluft,
E. (2007). Parental diagnoses in youth with narrow phenotype bipolar disorder or severe mood
dysregulation. American Journal of Psychiatry, 164(8), 1238–╉1241.
Brotman, M. A., Rich, B. A., Guyer, A. E., Lunsford, J. R., Horsey, S. E., Reising, M. M., … Leibenluft, E.
(2010). Amygdala activation during emotion processing of neutral faces in children with severe mood
dysregulation versus ADHD or bipolar disorder. American Journal of Psychiatry, 167(1), 61–╉69.
Brotman, M. A., Schmajuk, M., Rich, B. A., Dickstein, D. P., Guyer, A. E., Costello, E. J., … Leibenluft,
E. (2006). Prevalence, clinical correlates, and longitudinal course of severe mood dysregulation in
children. Biological Psychiatry, 60(9), 991–╉997.
Budhani, S., Marsh, A. A., Pine, D. S., & Blair, R. J. R. (2007). Neural correlates of response
reversal: considering acquisition. NeuroImage, 34(4), 1754–╉1765.
Chambless, D. L., Baker, M. J., Baucom, D., Beutler, L. E., Calhoun, K. S., Crits-╉Christoph, P., … Woody,
S. R. (1998). Update on empirically validated therapies: II. The Clinical Psychologist, 51(1), 3–╉16.
Chambless, D. L., Sanderson, W. C., Shoham, V., Bennet Johnson, S., Pope, K. S., Crits-╉Cristoph, P., …
McCurry, S. (1996). An update on empirically validated therapies. The Clinical Psychologist, 49(2), 5–╉18.
Connor, D. F., Glatt, S. J., Lopez, I. D., Jackson, D., & Melloni, R. H. (2002). Psychopharmacology and
aggression. I: a meta-╉analysis of stimulant effects on overt/╉covert aggression-╉related behaviors in
ADHD. Journal of the American Academy of Child and Adolescent Psychiatry, 41(3), 253–╉261.
7
9
2

Selected psychotherapeutic techniques 297

Copeland, W. E., Angold, A., Costello, E. J., & Egger, H. (2013). Prevalence, comorbidity, and correlates
of DSM-​5 proposed disruptive mood dysregulation disorder. American Journal of Psychiatry, 170(2),
173–​179.
Copeland, W. E., Shanahan, L., Egger, H., Angold, A., & Costello, E. J. (2014). Adult diagnostic and
functional outcomes of DSM-​5 disruptive mood dysregulation disorder. American Journal of Psychiatry,
171(6), 668–​674.
Corbetta, M., Kincade, J. M., Ollinger, J. M., McAvoy, M. P., & Shulman, G. L. (2000). Voluntary orienting
is dissociated from target detection in human posterior parietal cortex. Nature Neuroscience, 3(3),
292–​297.
Correll, C. U., Manu, P., Olshanskiy, V., Napolitano, B., Kane, J. M., & Malhotra, A. K. (2009).
Cardiometabolic risk of second-​generation antipsychotic medications during first-​time use in children
and adolescents. JAMA, 302(16), 1765–​1773.
Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social information-​processing
mechanisms in children’s social adjustment. Psychological Bulletin, 115(1), 74–​101.
Davidson, R. J., Jackson, D. C., & Kalin, N. H. (2000). Emotion, plasticity, context, and
regulation: perspectives from affective neuroscience. Psychological Bulletin, 126(6), 890–​909.
Davis, M. (1992). The role of the amygdala in fear and anxiety. Annual Review of Neuroscience, 15(1),
353–​375.
Desimone, R., & Duncan, J. (1995). Neural mechanisms of selective visual attention. Annual Review of
Neuroscience, 18(1), 193–​222.
Deveney, C. M., Brotman, M. A., Decker, A. M., Pine, D. S., & Leibenluft, E. (2012). Affective prosody
labeling in youths with bipolar disorder or severe mood dysregulation. Journal of Child Psychology and
Psychiatry, 53(3), 262–​270
Deveney, C. M., Connolly, M. E., Haring, C. T., Bones, B. L., Reynolds, R. C., Kim, P., … Leibenluft,
E. (2013). Neural mechanisms of frustration in chronically irritable children. American Journal of
Psychiatry, 170(10), 1186–​1194.
Dickstein, D. P., Finger, E. C., Brotman, M. A., Rich, B. A., Pine, D. S., Blair, J. R., & Leibenluft, E. (2010).
Impaired probabilistic reversal learning in youths with mood and anxiety disorders. Psychological
Medicine, 40(7), 1089–​1100.
Dickstein, D. P., Nelson, E. E., McClure, E. B., Grimley, M. E., Knopf, L., Brotman, M. A., … Leibenluft,
E. (2007). Cognitive flexibility in phenotypes of pediatric bipolar disorder. Journal of the American
Academy of Child and Adolescent Psychiatry, 46(3), 341–​355.
Dickstein, D. P., Towbin, K. E., Van Der Veen, J. W., Rich, B. A., Brotman, M. A., Knopf, L., … Leibenluft,
E. (2009). Randomized double-​blind placebo-​controlled trial of lithium in youths with severe mood
dysregulation. Journal of Child and Adolescent Psychopharmacology, 19(1), 61–​73.
Dodge, K. A. (1980). Social cognition and children’s aggressive behavior. Child Development, 51(1), 162–​170.
Dodge, K.A. (2003). Do social information processing patterns mediate aggressive behavior? In B. Lahey, T.
Moffitt, & A. Caspi (Eds.), Causes of conduct disorder and juvenile delinquency (pp. 254–​274). New York,
NY: Guilford Press.
Donovan, S. J., Stewart, J. W., Nunes, E. V., Quitkin, F. M., Parides, M., Daniel, W., … Klein, D. F. (2000).
Divalproex treatment for youth with explosive temper and mood lability: a double-​blind, placebo-​
controlled crossover design. American Journal of Psychiatry, 157(5), 818–​820.
Dougherty, L. R., Smith, V. C., Bufferd, S. J., Carlson, G. A., Stringaris, A., Leibenluft, E., & Klein, D. N.
(2014). DSM-​5 disruptive mood dysregulation disorder: correlates and predictors in young children.
Psychological Medicine, 44(11), 2339–​2350.
Eisenberg, N., & Fabes, R. (1994). Mothers’ reactions to children’s negative emotions: relations to children’s
temperament and anger behavior. Merrill-​Palmer Quarterly, 40(1), 138–​156.
Eyberg, S. M., Nelson, M. M., & Boggs, S. R. (2008). Evidence-​based psychosocial treatments for children
and adolescents with disruptive behavior. Journal of Clinical Child & Adolescent Psychology, 37(1),
215–​237.
8
9
2

298 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Finger, E. C., Marsh, A. A., Mitchell, D. G., Reid, M. E., Sims, C., Budhani, S., … Blair, J. R. (2008).
Abnormal ventromedial prefrontal cortex function in children with psychopathic traits during reversal
learning. Archives of General Psychiatry, 65(5), 586–​594.
Gallagher, H. L., & Frith, C. D. (2003). Functional imaging of “theory of mind.” Trends in Cognitive
Sciences, 7(2), 77–​83.
Goodale, M. A., & Milner, A. D. (1992). Separate visual pathways for perception and action. Trends in
Neurosciences, 15(1), 20–​25.
Guyer, A. E., McClure, E. B., Adler, A. D., Brotman, M. A., Rich, B. A., Kimes, A. S., … Leibenluft, E.
(2007). Specificity of facial expression labeling deficits in childhood psychopathology. Journal of Child
Psychology and Psychiatry, 48(9), 863–​871.
Hommer, R. E., Meyer, A., Stoddard, J., Connolly, M. E., Mogg, K., Bradley, B. P., … Brotman, M. A. (2014).
Attention bias to threat faces in severe mood dysregulation. Depression and Anxiety, 31(7), 559–​565.
Hulvershorn, L. A., Fosselman, D. D., Dickstein, D. P., & Janicak, P. G. (2012a). Psychopharmacology of
nonepisodic irritability, aggression, and mood swings in children and adolescents: Part I. Stimulants
and antidepressants. Psychopharmacology Review, 47(1), 1.
Hulvershorn, L. A., Fosselman, D. D., Dickstein, D. P., & Janicak, P. G. (2012b). Psychopharmacology
of nonepisodic irritability, aggression, and mood swings in children and adolescents: Part II.
Antipsychotics, antimanic agents, and alpha agonists. Psychopharmacology Review, 47(1), 9.
Johansen, J. P., Tarpley, J. W., LeDoux, J. E., & Blair, H. T. (2010). Neural substrates for expectation-​
modulated fear learning in the amygdala and periaqueductal gray. Nature Neuroscience, 13(8),
979–​986.
Jonkman, L. M., Kemner, C., Verbaten, M. N., Van Engeland, H., Camfferman, G., Buitelaar, J. K., &
Koelega, H. S. (2000). Attentional capacity, a probe ERP study: differences between children with
attention-​deficit hyperactivity disorder and normal control children and effects of methylphenidate.
Psychophysiology, 37(3), 334–​346.
Kaminski, J. W., Valle, L. A., Filene, J. H., & Boyle, C. L. (2008). A meta-​analytic review of components
associated with parent training program effectiveness. Journal of Abnormal Child Psychology, 36(4),
567–​589.
Kanwisher, N. (2000). Domain specificity in face perception. Nature Neuroscience, 3(8), 759–​763.
Kanwisher, N., & Wojciulik, E. (2000). Visual attention: insights from brain imaging. Nature Reviews
Neuroscience, 1(2), 91–​100.
Kazdin, A. E. (2010). Problem-​solving skills training and parent management training for oppositional
defiant disorder and conduct disorder. In A. E. Kazdin & J. R. Weisz (Eds.), Evidence-​based
psychotherapies for children and adolescents (2nd ed.) (pp. 211–​226). New York, NY: Guilford Press.
Kendall, P. C., & Hedtke (2006). Cognitive behavioral therapy for anxious children: therapist manual (3rd
ed.) Ardmore: PA: Workbook Publishing.
Kim, P., Arizpe, J., Rosen, B. H., Razdan, V., Haring, C. T., Jenkins, S. E., … Leibenluft, E. (2013).
Impaired fixation to eyes during facial emotion labelling in children with bipolar disorder or severe
mood dysregulation. Journal of Psychiatry & Neuroscience: JPN, 38(6), 407–​416.
Kircanski, K., Lieberman, M. D., & Craske, M. G. (2012). Feelings into words: contributions of language to
exposure therapy. Psychological Science, 23(10), 1086–​1091.
Leibenluft, E. (2011). Severe mood dysregulation, irritability, and the diagnostic boundaries of bipolar
disorder in youths. American Journal of Psychiatry, 168(2), 129–​142.
Leibenluft, E., Charney, D. S., Towbin, K. E., Bhangoo, R. K., & Pine, D. S. (2003). Defining clinical
phenotypes of juvenile mania. American Journal of Psychiatry, 160(3), 430–​437.
Lieberman, M. D. (2011). Why symbolic processing of affect can disrupt negative affect: Social cognitive
and affective neuroscience investigations. In A. Todorov, S. Fiske, & D. Prentice (Eds.), Social
neuroscience: Toward understanding the underpinnings of the social mind (pp. 188–​209). Oxford,
England: Oxford University Press.
92

Selected psychotherapeutic techniques 299

Lieberman, M. D., Eisenberger, N. I., Crockett, M. J., Tom, S. M., Pfeifer, J. H., & Way, B. M. (2007).
Putting feelings into words: affect labeling disrupts amygdala activity in response to affective stimuli.
Psychological Science, 18(5), 421–​428.
Lochman, J. E., Boxmeyer, C. L., Powell, N. P., Barry, T. D., & Pardini, D. A. (2010). In A. E. Kazdin & J.
R. Weisz (Eds.), Evidence-​based psychotherapies for children and adolescents (2nd ed.) (pp. 227–​242).
New York, NY: Guilford Press.
Marcus, R. N., Owen, R., Kamen, L., Manos, G., McQuade, R. D., Carson, W. H., & Aman, M. G. (2009).
A placebo-​controlled, fixed-​dose study of aripiprazole in children and adolescents with irritability
associated with autistic disorder. Journal of the American Academy of Child and Adolescent Psychiatry,
48(11), 1110–​1119.
McCracken, J. T., McGough, J., Shah, B., Cronin, P., Hong, D., Aman, M. G., … Research Units on
Pediatric Psychopharmacology Autism Network. (2002). Risperidone in children with autism and
serious behavioral problems. The New England Journal of Medicine, 347(5), 314–​321.
Miller, L., Hlastala, S., Mufson, M., Leibenluft, E., & Riddle, M. (2015). Interpersonal psychotherapy for
adolescents with mood and behavior dysregulation: Proof of concept pilot study. Poster presented at the
1st Congress on Pediatric Irritability and Dysregulation, Burlington, VT.
Mischel, W., Shoda, Y., & Rodriguez, M. I. (1989). Delay of gratification in children. Science, 244(4907),
933–​938.
Moadab, I., Gilbert, T., Dishion, T. J., & Tucker, D. M. (2010). Frontolimbic activity in a frustrating
task: covariation between patterns of coping and individual differences in externalizing and
internalizing symptoms. Development and Psychopathology, 22(2), 391–​404.
Mohanty, A., Gitelman, D. R., Small, D. M., & Mesulam, M. M. (2008). The spatial attention network
interacts with limbic and monoaminergic systems to modulate motivation-​induced attention shifts.
Cerebral Cortex (New York, N.Y.: 1991), 18(11), 2604–​2613.
Morris, A. S., Silk, J. S., Steinberg, L., Myers, S. S., & Robinson, L. R. (2007). The role of the family context
in the development of emotion regulation. Social Development, 16(2), 361–​388.
Niles, A. N., Craske, M. G., Lieberman, M. D., & Hur, C. (2015). Affect labeling enhances exposure
effectiveness for public speaking anxiety. Behaviour Research and Therapy, 68, 27–​36.
Ochsner, K. N. (2008). The social-​emotional processing stream: five core constructs and their translational
potential for schizophrenia and beyond. Biological Psychiatry, 64(1), 48–​61.
Olfson, M., Blanco, C., Liu, L., Moreno, C., & Laje, G. (2006). National trends in the outpatient
treatment of children and adolescents with antipsychotic drugs. Archives of General Psychiatry, 63(6),
679–​685.
Packard, M. G., & Knowlton, B. J. (2002). Learning and memory functions of the basal ganglia. Annual
Review of Neuroscience, 25, 563–​593.
Panksepp, J. (2006). Emotional endophenotypes in evolutionary psychiatry. Progress in Neuro-​
Psychopharmacology & Biological Psychiatry, 30(5), 774–​784.
Patterson, G. R. (1982). Coercive family process. Eugene: OR: Castalia.
Patterson, G. R., DeBaryshe, B. D., & Ramsey, E. (1989). A developmental perspective on antisocial
behavior. The American Psychologist, 44(2), 329–​335.
Patterson, G. R., Reid, J. B., & Dishion, T. J. (1992). Antisocial boys. Eugene: OR: Castalia.
Penton-​Voak, I. S., Thomas, J., Gage, S. H., McMurran, M., McDonald, S., & Munafò, M. R. (2013).
Increasing recognition of happiness in ambiguous facial expressions reduces anger and aggressive
behavior. Psychological Science, 24(5), 688–​697.
Perez-​Edgar, K., & Fox, N. A. (2005). A behavioral and electrophysiological study of children’s
selective attention under neutral and affective conditions. Journal of Cognition and Development, 6(1),
89–​118.
Pickles, A., Aglan, A., Collishaw, S., Messer, J., Rutter, M., & Maughan, B. (2010). Predictors of suicidality
across the life span: The Isle of Wight study. Psychological Medicine, 40(09), 1453–​1466.
03

300 Emotion Regulation in Severe Irritability and Disruptive Mood Dysregulation Disorder

Pilling, S., Gould, N., Whittington, C., Taylor, C., Scott, S., & Guideline Development Group. (2013).
Recognition, intervention, and management of antisocial behaviour and conduct disorders in children
and young people: summary of NICE-​SCIE guidance. BMJ, 346, f1298.
Posner, M. I., & Rothbart, M. K. (1998). Attention, self-​regulation and consciousness. Philosophical
Transactions of the Royal Society of London. Series B, Biological Sciences, 353(1377), 1915–​1927.
Raichle, M. E., & Snyder, A. Z. (2007). A default mode of brain function: a brief history of an evolving idea.
NeuroImage, 37(4), 1083–​1090; discussion 1097–​1099.
Rich, B. A., Brotman, M. A., Dickstein, D. P., Mitchell, D. G. V., Blair, R. J. R., & Leibenluft, E. (2010).
Deficits in attention to emotional stimuli distinguish youth with severe mood dysregulation from youth
with bipolar disorder. Journal of Abnormal Child Psychology, 38(5), 695–​706.
Rich, B. A., Carver, F. W., Holroyd, T., Rosen, H. R., Mendoza, J. K., Cornwell, B. R., … Leibenluft, E.
(2011). Different neural pathways to negative affect in youth with pediatric bipolar disorder and severe
mood dysregulation. Journal of Psychiatric Research, 45(10), 1283–​1294.
Rich, B. A., Grimley, M. E., Schmajuk, M., Blair, K. S., Blair, R. J. R., & Leibenluft, E. (2008). Face emotion
labeling deficits in children with bipolar disorder and severe mood dysregulation. Development and
Psychopathology, 20(2), 529–​546.
Rich, B. A., Schmajuk, M., Perez-​Edgar, K. E., Fox, N. A., Pine, D. S., & Leibenluft, E. (2007). Different
psychophysiological and behavioral responses elicited by frustration in pediatric bipolar disorder and
severe mood dysregulation. American Journal of Psychiatry, 164(2), 309–​317.
Roy, A. K., Klein, R. G., Angelosante, A., Bar-​Haim, Y., Leibenluft, E., Hulvershorn, L., … Spindel, C.
(2013). Clinical features of young children referred for impairing temper outbursts. Journal of Child and
Adolescent Psychopharmacology, 23(9), 588–​596.
Sanders, M. R. (1999). Triple P-​Positive Parenting Program: towards an empirically validated multilevel
parenting and family support strategy for the prevention of behavior and emotional problems in
children. Clinical Child and Family Psychology Review, 2(2), 71–​90.
Savage, J., Verhulst, B., Copeland, W., Althoff, R. R., Lichtenstein, P., & Roberson-​Nay, R. (2015). A
genetically informed study of the longitudinal relation between irritability and anxious/​depressed
symptoms. Journal of the American Academy of Child and Adolescent Psychiatry, 54(5), 377–​384.
Schultz, W. (2010). Dopamine signals for reward value and risk: basic and recent data. Behavioral and Brain
Functions, 6, 24.
Schultz, W., Dayan, P., & Montague, P. R. (1997). A neural substrate of prediction and reward. Science,
275(5306), 1593–​1599.
Snyder, J., Stoolmiller, M., Wilson, M., & Yamamoto, M. (2003). Child anger regulation, parental responses
to children’s anger displays, and early child antisocial behavior. Social Development, 12(3), 335–​360.
Steiner, H., Petersen, M. L., Saxena, K., Ford, S., & Matthews, Z. (2003). Divalproex sodium for the
treatment of conduct disorder: a randomized controlled clinical trial. The Journal of Clinical Psychiatry,
64(10), 1183–​1191.
Stoddard, J., Sharif-​Askary, B., Harkins, E. A., Frank, H. R., Brotman, M. A., Penton-​Voak. A. S., …
Leibenluft, E. (2016). An open pilot study of training hostile interpretation bias to treat disruptive
mood dysregulation disorder. Journal of Child and Adolescent Psychopharmacology, 26(1), 49–​57.
Stringaris, A., Baroni, A., Haimm, C., Brotman, M., Lowe, C. H., Myers, F., … Leibenluft, E. (2010).
Pediatric bipolar disorder versus severe mood dysregulation: risk for manic episodes on follow-​up.
Journal of the American Academy of Child and Adolescent Psychiatry, 49(4), 397–​405.
Stringaris, A., Cohen, P., Pine, D. S., & Leibenluft, E. (2009). Adult outcomes of youth irritability: a 20-​year
prospective community-​based study. American Journal of Psychiatry, 166(9), 1048–​1054.
Stringaris, A., & Goodman, R. (2009a). Longitudinal outcome of youth oppositionality: irritable,
headstrong, and hurtful behaviors have distinctive predictions. Journal of the American Academy of
Child and Adolescent Psychiatry, 48(4), 404–​412.
Stringaris, A., & Goodman, R. (2009b). Three dimensions of oppositionality in youth. Journal of Child
Psychology and Psychiatry, 50(3), 216–​223.
1
0
3

Selected psychotherapeutic techniques 301

Stringaris, A., & Taylor, E. (2015). Disruptive mood: irritability in children and adolescents. Oxford,
England: Oxford University Press.
Sukhodolsky, D. G., Kassinove, H., & Gorman, B. S. (2004). Cognitive-​behavioral therapy for anger in
children and adolescents: a meta-​analysis. Aggression and Violent Behavior, 9(3), 247–​269
Sukhodolsky, D. G., & Scahill, L. (2012). Cognitive-​behavioral therapy for anger and aggression in children.
New York, NY: Guilford Press.
Sukhodolsky, D. G., Smith, S. D., McCauley, S. A., Ibrahim, K., & Piasecka, J. B. (2016). Behavioral
interventions for anger, irritability and aggression in children and adolescents. Journal of Child and
Adolescent Psychopharmacology, 26(1), 58–​64.
Tabibnia, G., Lieberman, M. D., & Craske, M. G. (2008). The lasting effect of words on feelings: words may
facilitate exposure effects to threatening images. Emotion, 8(3), 307–​317.
Thomas, K. M., Drevets, W. C., Dahl, R. E., Ryan, N. D., Birmaher, B., Eccard, C. H., … Casey, B. J.
(2001). Amygdala response to fearful faces in anxious and depressed children. Archives of General
Psychiatry, 58(11), 1057–​1063.
Thomas, L. A., Brotman, M. A., Bones, B. L., Chen, G., Rosen, B. H., Pine, D. S., & Leibenluft, E. (2014).
Neural circuitry of masked emotional face processing in youth with bipolar disorder, severe mood
dysregulation, and healthy volunteers. Developmental Cognitive Neuroscience, 8, 110–​120.
Thomas, L. A., Brotman, M. A., Muhrer, E. J., Rosen, B. H., Bones, B. L., Reynolds, R. C., … Leibenluft, E.
(2012). Parametric modulation of neural activity by emotion in youth with bipolar disorder, youth with
severe mood dysregulation, and healthy volunteers. Archives of General Psychiatry, 69(12), 1257–​1266.
Thomas, L. A., Kim, P., Bones, B. L., Hinton, K. E., Milch, H. S., Reynolds, R. C., … Leibenluft, E. (2013).
Elevated amygdala responses to emotional faces in youths with chronic irritability or bipolar disorder.
NeuroImage: Clinical, 2, 637–​645.
Tseng, W.-​L., Thomas, L. A., Harkins, E., Pine, D. S., Leibenluft, E., & Brotman, M. A. (2016). Neural
correlates of masked and unmasked face emotion processing in youth with severe mood dysregulation.
Social Cognitive and Affective Neuroscience, 11(1), 78–​88.
Waters, A. M., Henry, J., Mogg, K., Bradley, B. P., & Pine, D. S. (2010). Attentional bias towards angry
faces in childhood anxiety disorders. Journal of Behavior Therapy and Experimental Psychiatry, 41(2),
158–​164.
Waxmonsky, J. G., Wymbs, F. A., Pariseau, M. E., Belin, P. J., Waschbusch, D. A., Babocsai, L., … Pelham,
W. E. (2013). A novel group therapy for children with ADHD and severe mood dysregulation. Journal
of Attention Disorders, 17(6), 527–​541.
Webster-​Stratton, C., & Reid, M. J. (2010). The Incredible Years parents, teachers and children training
series: a multifaceted treatment approach for young children with conduct disorders. In A. E. Kazdin &
J. R. Weisz (Eds.), Evidence-​based psychotherapies for children and adolescents (2nd ed.) (pp. 194–​2010).
New York, NY: Guilford Press.
Weisz, J. R., Jensen-​Doss, A., & Hawley, K. M. (2006). Evidence-​based youth psychotherapies versus usual
clinical care: a meta-​analysis of direct comparisons. The American Psychologist, 61(7), 671–​689.
Wilkinson, D., & Halligan, P. (2004). The relevance of behavioural measures for functional-​imaging studies
of cognition. Nature Reviews Neuroscience, 5(1), 67–​73.
Zisser, A., & Eyberg, S. M. (2010). Parent-​child interaction therapy for oppositional children. In A. E.
Kazdin & J. R. Weisz (Eds.), Evidence-​based psychotherapies for children and adolescents (2nd ed.) (pp.
179–​193). New York, NY: Guilford Press.
2
0
3
3
0

Part III

Emotion Regulation
in Specific Behavior/​
Population
4
0
3
5
0
3

Chapter 15

Children of Abuse and Neglect


Faye Riley, Anna Bokszczanin, & Cecilia A. Essau

Abuse and neglect


Violence against children dates back to ancient civilisations, with infanticide, child abuse and
cruelty routinely noted in ancient Greece, Rome, Egypt and China (Ten, Bensel, Rheinberger, &
Radbill, 1997). For example, in ancient Rome a child was viewed as the possession of the father,
who had the legal right to sell, mutilate or kill their child at will. It was also forbidden to rear
deformed children; thus, making infanticide or abandonment the only option (Zigler & Hall,
1989). Throughout history, children have also been forced to work in poor conditions and histori-
cal accounts have emerged of neglect and sexual abuse (Krug, Mercy, Dahlberg, & Zwi, 2002).
However, the issue of child protection remained largely unacknowledged until 1874 with the
landmark case of nine-╉year-╉old Mary Ellen Wilson in the United States, who had been severely
abused and neglected by her guardian. This case gained widespread media attention and led to
the acknowledgement of child abuse and the founding of the New York Society for the Prevention
of Cruelty to Children, believed to be the first child protective agency in the world (Myers, 2008).
Even with the knowledge of such abuse and the existence of charitable groups concerned with
the protection of children, the issue did not receive widespread attention until 1962, with the
publication of “The battered child syndrome” (Kempe, Silverman, Steele, Droegemueller, & Silver,
1962). The term “battered child syndrome” refers to serious physical abuse in children and it was
after this seminal work that parents and caregivers began to be held responsible for these injuries
(Kempe et al., 1962). Kempe and colleagues further asserted that healthcare professionals were
responsible for the protection of children and preventing continued abuse, and made recommen-
dations on how to identify and report maltreatment.

Definition
Following the publication of this seminal volume, awareness of other forms of abuse such as child
sexual abuse and neglect also increased, along with an increased understanding of the physical,
emotional and behavioral consequences of all forms of abuse. However, it was not until 1999 that
the World Health Organisation Consultation on Child Abuse Prevention gave a definition of child
abuse as, “all forms of physical and/╉or emotional ill-╉treatment, sexual abuse, neglect or negligent
treatment, or commercial or other exploitation of children, resulting in actual or potential harm
to a child’s health, survival, development, or dignity in the context of a relationship of responsibil-
ity, trust or power.”
Forms of child abuse and neglect are generally recognised to fall into four main categories
(Higgins & McCabe, 2001). Firstly, physical abuse which involves hitting, shaking, throwing, poi-
soning, burning or scalding, suffocating, or otherwise causing the child actual physical harm or
creating the potential for harm. The second form is sexual abuse which involves forcing or entic-
ing a child to take part in sexual activities, not necessarily involving a high level of violence,
6
0
3

306 Children of Abuse and Neglect

irrespective of whether or not the child is cognizant of what is happening to them. The third
from is emotional abuse, which is defined as the persistent emotional maltreatment of a child
and includes the failure of a caregiver to provide an appropriate and supportive environment. It
includes acts that have an adverse effect on the emotional health and development of a child, such
as denigration, ridicule, threats, intimidation, discrimination, rejection and other nonphysical
forms of hostile treatment. The final form is neglect, which has been defined as the persistent
failure to meet a child’s basic physical and/╉or psychological needs and is likely to result in the
serious impairment of the child’s health or development. Neglect may result when a caregiver fails
to provide adequate food, clothing and shelter (including exclusion from the home or abandon-
ment), fails to protect a child from physical and emotional harm or danger, fails to respond to a
child’s basic emotional needs, does not provide adequate supervision, or does not ensure access to
appropriate medical care or treatment.

Prevalence
Child maltreatment is complex and difficult to study, consequently current estimates of preva-
lence can vary widely depending on a variety of factors such as differing legal and cultural defi-
nitions of child maltreatment used between countries; the type of child maltreatment studied;
the population studied and how the sample was recruited; the methods of research used; the
breadth and quality of official statistics and the extent of mandatory reporting; the breadth and
quality of population-╉based self-╉report surveys collected from victims, parents or caregivers
(Finklehor, 1994).
Prevalence estimates of abuse and neglect in the child population vary considerably depending
on the country of study. For example, in a review of international studies, Radford et al. (2011)
showed prevalence rates for different forms of abuse ranging from 1.8% to 34% for physical vio-
lence; 1.1% to 32% for sexual abuse; 5.4% to 37.5% for emotional abuse; and 6% to 41.5% for
neglect. Prevalence rates can also vary depending on the age of the population studied. In the UK,
a National Society for the Prevention of Cruelty to Children (NSPCC) prevalence study found
that 2.5% of children under 11 years of age and 6% of 11 to 17 year olds had experienced mal-
treatment by a parent or caregiver in the past year. For lifetime rates, it was found that one in 17
(5.9%) children under 11 years and one in five 11–╉17 year olds (18.6%) had experienced severe
maltreatment during childhood.
A further issue is that abuse and neglect are often hidden from view and therefore large numbers
of cases are not detected, reported or recorded, even where mandatory reporting exists (Theodore
& Runyan, 1999). Consequently, official statistics often reveal little about the true rates of child
abuse. There are currently over 50,000 children identified that are in need of protection from
abuse in the UK, yet the NSPCC estimates that for every child identified as needing protection
from abuse, another eight may be suffering abuse (Harker et al., 2013). In part this may be because
children do not disclose what is happening to them out of fear of the repercussions or due to fear
that they will not be believed. In addition, some victims may be too young to realise that what is
happening to them is wrong, or abuse may not be officially reported even if other family members
or adults know about it. Moreover, in many countries, there are no legal or social systems with
specific responsibility for recording, or responding to, reports of child abuse and neglect (Bross,
Miyoshi, Miyoshi, & Krugman, 2000).
Therefore, prevalence rate estimates gathered through research studies and official statistics in
particular, may often only represent the “tip of the iceberg” in terms of the extent of child abuse
and neglect. However, such data are useful to gain a picture of the general patterns of abuse and
have shown a global trend that has significant, deleterious consequences for public health (Krug
et al., 2002).
7
0
3

RISK FACTORS FOR ABUSE AND NEGLECT 307

Assessment
Early detection and intervention of abuse is important in limiting the damage done to the devel-
opment of the child. When early intervention does not occur, approximately one in three children
will suffer continued abuse (Seifert et  al., 2010). Child abuse and neglect is rarely detected or
prevented before hospitalisation, most likely due to limited contact with non-╉family members and
because violent incidents often occur within or around the family, in “a circle of trust” (Finkelhor,
1994). It is estimated that 2% to 10% of children who visit hospital emergency departments are
victims of child abuse and neglect (Holt, Buckley, & Whelan, 2008; Hussey, Chang, & Kotch,
2006; Palazzi, de Girolamo, & Liverani, 2005). Therefore, hospital staff may be the first contact
and opportunity for physical abuse to be identified. However, recognising maltreated children
in the everyday routine of an emergency department is a major challenge and detection rates of
child abuse remain low, often going undetected by both clinical and nursing staff (Gilbert, Kemp,
et  al., 2009; Louwers et  al., 2012; Oral, Blum, & Johnson, 2003). This is influenced by factors
such as knowledge, training, attitude and the experience of health care professionals, and avail-
able resources for referral to name a few (Flaherty et al., 2008; Fraser, Mathews, Walsh, Chen, &
Dunne, 2010; Jones et al., 2008).
Due to limitations associated with identifying abuse at emergency departments, other screen-
ing assessments attempt to facilitate the detection of abuse before hospitalisation. For example, the
“Tool for identifying families at risk of or with already established infant and toddler abuse and
neglect problems” (INTOVIAN Tool), is a five-╉item checklist addressing risk indicators for physi-
cal and/╉or psychological violence, neglect, and disordered/╉abusive relationship patterns between
the child and caregiver. This tool aims to prevent violence, or to break the cycle of violence in the
family before children suffer more severe maltreatment. The tool can be used by professionals as
part of routine observations in children’s centers and nurseries, or by medical staff during routine
health examinations. Staff who have been trained in education and who work in nurseries and
children’s centers will likely be in a better position to detect signs of neglect, because it is more
long-╉term; whereas, staff in health and social services will be better placed to identify child abuse
(Essau, 2015). Professionals can use the aforementioned tool to assess the quality of the caregiver
and child interaction; levels of affection shown; and the psychological involvement of both the
caregiver and child, in order to uncover any potential issues regarding the caregiver-╉child rela-
tionship, which may be indicative of maltreatment.

Risk factors for abuse and neglect


A variety of theories and models have been developed to explain the occurrence of child abuse.
The most widely adopted explanatory model is the ecological model. This risk factor approach
assumes that there is no single pathway to abuse and that risk factors occur across multiple devel-
opmental domains or levels of a child’s social ecology (Bronfenbrenner, 1988). As applied to child
abuse and neglect, the ecological model proposes a number of factors that contribute to the risk
of maltreatment, covering characteristics of the individual child, those of the caregiver or family,
parent-╉child relationships, neighbourhood characteristics, and societal factors.

Child factors
A number of child characteristics have been associated with greater likelihood of being abused
or neglected, which include age; being perceived as problematic; or having a disability or ill-
ness. In terms of age, the risk of maltreatment is greater for children under four years of age
or during adolescence. Fatal cases of physical abuse are more common among young infants
(Damashek, Nelson, & Bonner, 2013; Klevens & Leeb, 2010), who are particularly vulnerable due
8
0
3

308 Children of Abuse and Neglect

to their dependency, small size, and inability to defend themselves. In contrast, rates of sexual
abuse tend to rise after the onset of puberty, with the highest rates occurring during adolescence
(Finklehor, 1994).
Caregivers who perceive their children as having more problem behaviors are more likely to
physically abuse them (During & McMahon, 1991; Whipple & Webster-╉Stratton, 1991). In addi-
tion, difficult child temperament (as perceived by the parent) has been specifically associated with
emotional neglect (Harrington, Black, Starr, & Dubowitz, 1998). Such behavior and perceptions
may strain the parent-╉child relationship, increasing the risk of maltreatment.
Finally, it has been documented that children with disabilities are 1.8 times more likely to be
neglected, 1.6 times more likely to be physically abused, and 2.2 times more likely to be sexually
abused than are children without disabilities (Sullivan & Cork, 1996). It may be that due to the
demands of raising a child with a disability caregivers become overwhelmed and respond with
irritability, inconsistency, or punitive discipline. Furthermore, disabled children may be unre-
sponsive or have limited ability to interact with, or show affection to their caregivers, which conse-
quently may interfere with attachment and bonding (Hibbard & Desch, 2007). Compounding the
issue further, children with disabilities may be less able to protect themselves and are also highly
dependent on adults for their safety and well-╉being, meaning they may be particularly vulnerable
to abuse or neglect.

Caregiver or family factors


Research has linked certain characteristics of the caregiver and the wider family environment
to child abuse and neglect. Such factors may compromise parenting, resulting in maltreatment.
These factors include: The family structure and resources; family size and household composition;
personality and behavioral characteristics of the caregiver; stress; lack of social support; failure
to bond with and nurture the child; being maltreated themselves as a child; misusing alcohol or
drugs; experiencing financial difficulties; and family breakdown or conflict.
In terms of family structure and resources, physically abusive parents are more likely to be
young, single, poor, unemployed and less educated than non-╉abusing parents. For example, being
in a single parent household increases the risk of child neglect by 87% (Connell-╉Carrick, 2003)
and single mothers are three times more likely to report using harsh physical discipline when com-
pared with mothers in two-╉parent families (Straus & Gelles, 1990). For single parents there may be
less time to accomplish household tasks, including child supervision, spending quality time with
children and earning sufficient money to adequately provide for them (DePanfilis, 2006).
With relation to family size and household composition, households of a larger size have been
shown to present a greater risk of child abuse (Damashek et al., 2013). In part this may be due to
unstable family environments, where the make-╉up of the household frequently changes as mem-
bers move in and out (Dubowitz & Black, 2001). Furthermore, children living in households with
adults who are not related to them, have 50 times greater risk of fatal injury as children living with
two biological parents or a single parent (Schnitzer & Ewigman, 2005).
Personality and behavioral characteristics of parents may also influence child abuse. Parents
who are more likely to abuse their children tend to have lower self-╉esteem and reduced mental
well-╉being. For example, research has shown that physically abusive parents have more negative
self-╉perceptions than parents who are not physically abusive (Christensen et al., 1994) and have
greater general emotional distress and unhappiness (Caliso & Milner, 1992; Milner & Robertson,
1990). These characteristics may compromise their parenting capabilities and may be associated
with disrupted social relationships, an inability to cope with stress and difficulty in forming social
support systems (Krug et al., 2002).
9
0
3

Risk factors for abuse and neglect 309

In relation to stress, it is believed that stress resulting from job changes, loss of income, health
problems or other stressors can exacerbate characteristics in the family, such as hostility, anxi-
ety, or depression, which in turn might increase levels of family conflict and child maltreatment
(Goldman, Salus, Wolcott, & Kennedy, 2003). Indeed, abusive parents reported more stressful
life events (Coohey & Braun, 1997)  and scored higher on perceived daily stress (Williamson,
Borduin, & Howe, 1991), compared to non-​abusive parents. Physically abusive mothers also per-
ceive themselves as using more inefficient coping strategies when faced with stress (Cantos, Neale,
O’Leary, & Gaines, 1997). While families who are coping with such problems may also lack the
time or emotional capacity to provide for the basic needs of their children, resulting in neglect
(DePanfilis, 2006).
Social support can also play a role in the occurrence of child abuse. Abusive mothers report
receiving less social support, compared to non-​abusive mothers (Chan, 1994) and the support
that is received, seems to be weaker with abusive mothers reporting they receive less emotional
resources (e.g., listening, decision-​making, companionship) from their social networks (Coohey &
Braun, 1997). A lack of social support may mean parents or caregivers have less of a support net-
work to act as alternative caregivers, or to provide additional support to the caregiver or the child.
A further contributing factor is failure to bond with and nurture the child. Abusive parents
show greater irritation and annoyance in response to their children’s moods and behavior and
have been shown to be more controlling and hostile, and less supportive, affectionate, playful and
responsive to their children (Bardi & Borgognini-​Tarli, 2001; National Research Council, 1993).
Furthermore, abusive mothers have greater negative expectations of their children (Larrance &
Twentyman, 1983) and are less likely to blame themselves for failed mother–​child interactions,
and give less credit to their children for successful mother–​child interactions than other mothers
(Bradley & Peters, 1991). Such characteristics can impact upon the successful formation of the
parent-​child relationship and thus, increase the risk for abuse.
The caregivers’ own childhood can also serve as a risk factor, specifically if they have been mal-
treated themselves as a child. Research has shown that abusive mothers are more likely to report
having been physically victimized as children by their parents (Coohey & Braun, 1997) and expe-
riencing corporal punishment as a teen has also been shown to be a significant predictor of per-
petrating severe child abuse (Ross, 1996). Additionally, neglectful mothers are three times more
likely to have been abused in childhood than mothers who did not neglect their children (Zuravin
& DiBlasio, 1996). This can be explained by social learning theory, which suggests that children’s
behavior is largely shaped by their parents via modelling and schedules of reinforcement such that
exposure to abusive and maltreating parents during their own childhood encourages the caregiver
to believe such behaviors are acceptable and effective, leading them to incorporate these into their
own parenting styles as adults (Dodge, Bates, & Pettit, 1990).
Substance abuse can also be a contributing factor. Research has demonstrated that physically
abusive mothers are more likely to have a history of drug problems (Whipple & Webster-​Stratton,
1991) and, children whose parents abused alcohol and other drugs are more than four times more
likely to be neglected than children whose parents did not (Jaudes, Ekwo, & Van Voorhis, 1995).
Such substance abuse may limit the ability of the caregiver to provide adequate care for their
children or to make appropriate decisions regarding their welfare (Slack et al., 2011), particularly
when intoxicated, with one study reporting that 65% of maltreated children who had parents with
substance abuse problems were maltreated while the parent was intoxicated (Donohue, 2004).
Experiencing financial difficulties in the family has also been found to be a strong predictor of
child neglect (Slack et al., 2011), with a lack of familial financial resources having serious negative
consequences on the ability of the caregiver to meet even the most basic needs of their child. For
0
1
3

310 Children of Abuse and Neglect

example, poverty can affect the ability of the caregiver to provide adequate supervision (e.g., can-
not afford child care), housing, nutrition, medical care, clothing, and safety.
Finally, family breakdown and conflict can be a further contributing factor. Straus and Gelles
(1990) found that parents with higher rates of inter-╉parental verbal aggression were more likely
to perpetrate severe child abuse. Additionally, physically abused adolescents are more likely to
come from families experiencing high levels of family stress, and these adolescents perceived their
families as being less adaptive and cohesive (Williamson et al., 1991). Family life with such stress
and conflict may be so disorganized and hostile that caregivers are unable to meet the basic needs
of their children on a consistent basis (Hornor, 2014).

Community factors
A number of community characteristics may increase the risk of child maltreatment. These factors
include poverty; community disorganisation; and poor social cohesion. In relation to poverty, this
has been shown to adversely affect children through its impact on parental behavior and the avail-
ability of community resources (McLoyd, 1990). Communities with high levels of poverty tend
to have weaker physical and social infrastructures and fewer resources and amenities in place to
prevent and detect child abuse, such as accessible health care, social services, and affordable child
care, which can reduce the ability of caregivers to appropriately care for their children (Corcoran
& Nichols-╉Casebolt, 2004). Community disorganization can be a further risk factor. Rates of child
abuse and neglect are higher in communities characterized by high levels of unemployment, high
population turnover and high availability of alcohol and drugs (DePanfilis, 2002; Gillham et al.,
1998). Finally, children living in communities that lack social cohesion and solidarity have been
found to be at greater risk of abuse (Runyan et al., 1998), due to the lack of positive informal and
formal support systems for families (Cash & Wilke, 2003).

Societal factors
A range of society level factors are considered to have important influences on child maltreat-
ment. Many of these broader cultural and social factors can influence how caregivers treat their
children. These include factors such as the presence of policies and programs to prevent child
abuse and neglect, and the responsiveness of the criminal justice system. In addition, cultural
norms and the cultural definitions of generally accepted child-╉rearing principles, may contribute
to this issue. Specifically, social and cultural norms that promote or tolerate violence towards oth-
ers and support the use of corporal punishment or certain cultural practices that may be viewed
as abusive or neglectful to the larger society (e.g. genital mutilation). Moreover, social, economic,
health and education policies that lead to poor living standards, or to socioeconomic instability,
may also be a contributing factor to child abuse.

Outcomes
The consequences of child abuse and neglect can vary widely. Physical injuries and, in extreme
cases, death are direct consequences; however, there are also a variety of psychological and behav-
ioral outcomes. Various aspects of the maltreatment situation, such as duration, type and sever-
ity of abuse, can directly impact the development of negative outcomes (Manly, Cicchetti, &
Barnett, 1994).
Chronically maltreated children appear to be at a high-╉risk of developing clinical levels of
psychological problems (Ethier, Lemelin, & Lacharité, 2004). Research has shown that there
is a particularly strong association between childhood abuse and mood and anxiety disorders,
including depression, bipolar disorder, generalized anxiety disorder, panic disorder, phobias, and
13

Emotion regulation 311

posttraumatic stress disorder (PTSD) (Gilbert, Widom, et al., 2009; Heim & Nemeroff, 2001; Hill,
2003; Katerndahl, Burge, & Kellogg, 2005; Kendler et al., 2000; Molnar, Buka, & Kessler, 2001). In
addition, traumatic experiences early in life are associated with other psychological conditions,
such as schizophrenia, reactive attachment disorder, eating disorders, and personality disorders
(Ackard & Neumark-╉Sztainer, 2003; Kaplan & Klinetob, 2000; Noll, Horowitz, Bonanno, Trickett,
& Putnam, 2003; Zeanah et al., 2004).
Heim, Shugart, Craighead, and Nemeroff (2010) conducted a meta-╉analysis of 124 studies
that investigated the relationship between child physical abuse, emotional abuse, or neglect and
various health outcomes. It was found that emotionally abused children were three times more
likely to develop depression than non-╉abused individuals. Physically abused and neglected chil-
dren also had a higher risk of developing a depressive disorder. Cicchetti and Rogosch (1997)
examined the level of adaptation of school-╉aged maltreated children who were evaluated over
a three-╉year period. Over this period, the maltreated children exhibited more externalising and
internalising behavior problems, less prosocial behavior, greater symptoms of depression, and
more withdrawn behavior than the non-╉maltreated children. This study confirms continuity in
the difficulties experienced by abused children. In addition, research has shown children who
experience maltreatment are also at increased risk for substance misuse, engagement in high-╉risk
sexual behaviors, delinquency and poor academic performance (Dubowitz, 2009; Felitti et  al.,
1998; Lansford et al., 2007).
As well as causing immediate harm and later childhood impairment, abuse and neglect can
have many long-╉term consequences that endure well into adulthood. A  longitudinal study
revealed that as many as 80% of young adults who had been abused as a child met the diag-
nostic criteria for at least one psychiatric disorder at age 21 (Silverman, Reinherz, & Giaconia,
1996). Adult psychological difficulties most frequently associated with child abuse include
depression, anxiety, PTSD, low self-╉esteem, poor adjustment, criminal behavior, risky sexual
behavior and substance misuse (Afifi et al., 2008; Fergusson, Boden, & Horwood, 2008; Huang
et al., 2011; Widom, Marmorstein, & White, 2006; Wilson & Widom, 2011). Moreover, chil-
dren who experience maltreatment are also at increased risk for adverse health effects and
certain chronic diseases as adults, including heart disease, cancer, chronic lung disease, liver
disease, obesity, high blood pressure and high cholesterol (Danese et  al., 2009; Felitti et  al.,
1998; Springer, Sheridan, Kuo, & Carnes, 2007). A further long-╉term consequence of maltreat-
ment is the heightened likelihood of abusing or neglecting one’s own children (Thornberry &
Henry, 2013), which, therefore, contributes to the establishment of an intergenerational cycle
of neglect and abuse.

Emotion regulation
Emotion regulation is generally defined as the internal and external processes by which the indi-
vidual manages the occurrence, intensity, and expression of emotions to reach goals or situational
demands (Cicchetti & Howes, 1991; Eisenberg & Morris, 2002; Thompson, 1994). The childhood
years are thought to be a critical period for the development of emotion regulation skills, and suf-
fering abuse and neglect during this time can interfere with the acquisition of these skills (Shields
& Cicchetti, 1998). Maltreated children can experience conflicting feelings and impulses arising
from being neglected and/╉or harmed by an adult who is often also an attachment figure. Abusive
and neglectful caregivers also likely fail to engage in behaviors that enable children to develop
optimal emotion regulation strategies. Consequently, maltreated children often exhibit a range of
emotion regulation deficits, with nearly 80% of maltreated children displaying dysregulated emo-
tion patterns, compared with only 36% of non-╉maltreated children (Maughan & Cicchetti, 2002).
2
1
3

312 Children of Abuse and Neglect

In the next section, we discuss how abuse and neglect can lead to the development of emotion
regulation problems, how these problems manifest in abused children and how these problems
can heighten the. risk of continued abuse and exacerbate negative outcomes and psychopathology.

Development of emotion regulation problems


Caregiver influence
Individual differences in emotion regulation emerge from the quality of familial relationships
and from caregiver socialisation of appropriate emotion related behaviors, through processes
such as modelling, reinforcement, discipline, displays of positive or negative affect and sensitiv-
ity (Calkins, 1994; Morris, Silk, Steinberg, Myers, & Robinson, 2007; Thompson & Meyer, 2007;
Valiente & Eisenberg, 2006). Caregiver behavior can influence the child’s internal emotional expe-
riences, emotional reactivity, and modulate their emotional arousal (van der Kolk & Fisler, 1994).
From the perspective of attachment theory, securely attached children are able to use care-
givers effectively to help regulate their emotions (Bowlby, 1969, 1982). However, in abusive and
neglectful environments characterized by inconsistency and hostility, caregivers lack certain char-
acteristics thought to be important to the development of their children’s emotion management
skills (Howes, Cicchetti, Toth, & Rogosch, 2000). In particular, abusive caregivers socialize emo-
tion management skills differently than non-╉abusive caregivers, by providing less support and
acknowledgement in response to their children’s emotions, showing lower levels of emotional
expression in the parent-╉child relationship, and providing less emotion-╉related information
exchange (Bousha & Twentyman, 1984; Camras, Sachs-╉Alter, & Ribordy, 1996; Gaudin, Polansky,
Kilpatrick, & Shilton, 1996; Shipman, Zeman, Penza, & Champion, 2000; Shipman & Zeman,
2001). Such parenting practices are imbedded in the caregivers’ own attitudes towards emotion
and have been found to correlate with children’s successful regulation of emotion (Gottman,
Katz, & Hooven, 1996; Lunkenheimer, Shields, & Cortina, 2007). It is also believed that abusive
caregivers tend to be more socially isolated, and in turn isolate their children from interactions
with others, resulting in fewer opportunities for the child to engage in emotional communication
(Salzinger, Feldman, Hammer, & Rosario, 1993).
Abusive caregivers have additionally been found to show less positive and more negative emo-
tion, than non-╉abusive caregivers (Bugental & Others, 1989; Kavanagh, Youngblade, Reid, &
Fagot, 1988) and as such, abused children are exposed to frequent negative emotional experiences
including anger, frustration, reactivity, and irritability (Alessandri, 1991; Shields & Cicchetti,
1998). This can place the child in a state of near constant arousal and vigilance. Such overwhelming
emotional arousal can foster conditions that undermine the child’s ability to process and manage
emotions effectively (Briere & Jordan, 2009; Maughan & Cicchetti, 2002) and may later manifest
in problematic behavior, such as aggression or hypervigilance (Greenberg, Speltz, & Deklyen,
1993). For example, abused children may be more sensitive to anger from their abuser and fearful
of those around them, because this could help them identify threat quickly and potentially avoid
additional abuse (Masten et al., 2008).
In other cases, the emotional regulatory problems of maltreated children arise from the incon-
sistent emotional demands expected from the caregiver, which means a child can have difficulty
predicting the consequences of their behavior (Dadds & Salmon, 2003). Thus, the same emo-
tional signals from the child that can elicit nurturance from non-╉abusive caregivers may result
in hostile or angry reactions from an abusive caregiver. Consequently, children may learn that it
is unacceptable or even dangerous to discuss their feelings and emotions, particularly negative
ones (Beeghly & Cicchetti, 1994) and accordingly develop strategies of emotion regulation that
3
1

Development of emotion regulation problems 313

attempt to accommodate these conditions. Emotionally abusive environments in which children


are told their emotional reactions are bad or inappropriate, or in which negative emotions are
punished, ignored, or met with hostility, may elicit patterns of avoidance, suppression, and harm-
ful attempts at managing emotions (Krause, Mendelson, & Lynch, 2003). For example, Shipman,
Edwards, Brown, Swisher, and Jennings (2005) found support for this from the perspective of
neglected children. Neglected children were more likely to inhibit the expression of negative emo-
tion to their mothers and expected less support and more punishment or conflict from mothers
in response to displays of negative emotion. Children who anticipate non-╉supportive or disparag-
ing responses to their emotional displays are then likely to restrict emotional expressiveness as a
learned response (Cole, Zahn-╉Waxler, & Smith, 1994).
In summary, certain parenting practices and behavior associated with abuse and neglect may
interfere with the normal acquisition of emotional understanding and emotion regulation skills
in maltreated children.

Neurobiology
It is probable that the emotional difficulties displayed by maltreated children are not only a reflec-
tion of caregiver influence, but also affected by changes in neurobiological structure and function-
ing that occur as a result of abuse or neglect (Cicchetti & Tucker, 1994). These changes can reduce
a child’s capacity to regulate affective states and modulate behavioral responses to stressors, and
thus, may influence emotion regulation processes.
Maltreatment in childhood appears to be associated with the reorganisation of neural circuits in
ways that alter the processing of emotional information, which may underlie the emotion regula-
tion deficits reported in maltreated children (Pollak, 2008). For instance, youth with histories of
early-╉life abuse and trauma display greater amygdala activity to threatening cues and emotional
conflict, showing a hypervigilance to threat stimuli and reduced ability to regulate emotional
processing (Marusak, Martin, Etkin, & Thomason, 2015; McCrory et al., 2011). Engagement of
the amygdala is thought to support preferential processing to emotional stimuli (Vuilleumier,
Armony, Driver, & Dolan, 2001), in order to allow potential threats to be rapidly detected and
evaluated (LeDoux, 1996). Thus, the alterations in the neural systems of abused children alter
how they monitor the environment for threatening information. Pollak, Klorman, Thatcher, and
Cicchetti (2001) found that the event-╉related potential amplitude responses of maltreated chil-
dren exceeded those of non-╉maltreated children in response to angry facial expressions, but not to
fearful or happy expressions, reflecting a response bias for angry stimuli. These results suggest that
the abusive experiences encountered by these children during their development may enhance
the memory of salient stimuli, due to the stored mental representations that have been associated
with that stimulus over time. As such, maltreated children are particularly sensitive to, and quick
to detect, anger, as they develop an association between this emotion and abusive behavior from
their caregiver.
Furthermore, there is substantial evidence that amygdala reactivity is under inhibitory control
of medial prefrontal regions (Ochsner & Gross, 2005), particularly Pregenual anterior cingulate
cortex (pgACC) (Maier & di Pellegrino, 2012). Marusak et al. (2015) found an absence of nega-
tive regulation-╉related amygdala–╉pgACC connectivity in maltreated youth, indicating an absence
of effective inhibitory control. Maltreated children have also been found to exhibit dysregulation
of the hypothalamic-╉pituitary-╉adrenal axis following social interactions, signifying the impaired
ability to cope with stressors and negative emotions (Tarullo & Gunnar, 2006).
Collectively, these neurobiological findings imply abuse and neglect is associated with a simul-
taneous heightened sensitivity to conflicting emotional information and a lack of regulatory
4
1
3

314 Children of Abuse and Neglect

control over emotion processing. Thus, the combination of these factors is likely to limit the abil-
ity of the maltreated child to master appropriate emotional skills.

Expression of emotion regulation deficits


The emotion regulatory deficits associated with abuse and neglect manifest through a range
of expressions and behaviors, including the poor understanding and recognition of emotions
(Cicchetti & Curtis, 2005; Shipman et al., 2005), greater negative affect (Gaensbauer, 1982), fewer
adaptive emotion regulation skills (e.g., situational appropriateness of affective displays, empathy,
and emotional self-╉awareness) (Shipman et al., 2005), heightened vigilance to threatening stimuli
(Pollak, Vardi, Putzer Bechner, & Curtin, 2005; Rieder & Cicchetti, 1988), and poor regulation of
emotions (Maughan & Cicchetti, 2002; Shipman et al., 2007). Behaviorally, maltreated children
exhibit less self-╉control and social competence (Shields, Cicchetti, & Ryan, 1994) and more irrita-
bility, reactivity and anger (Alessandri, 1991).
The atypical emotional development of maltreated children can become apparent from the first
few months of life. Early-╉life abuse appears to be a particular risk for emotion regulation problems.
Children who experienced maltreatment during early-╉life showed significantly reduced emotion
regulation compared to non-╉maltreated children; whereas, children who experienced a later onset
did not (Kim & Cicchetti, 2010). Early-╉life abuse and neglect has been associated with high levels
of negativity and anger in toddlers, and a lack of self-╉control in pre-╉schoolers (Erickson, Egeland,
& Pianta, 1989). For example, Gaensbauer, Mrazek, and Harmon (1981) found that maltreated
infants as young as three months of age displayed higher rates of fearfulness, anger, and sad-
ness during mother-╉infant interactions, and expressed a reduced range of emotions and increased
duration of negative affect, compared with non-╉maltreated controls. Similarly, maltreated chil-
dren aged one to three years, exhibited more anger and less positive affect compared to non-╉
maltreated children (Robinson et  al., 2008). Furthermore, Macfie, Cicchetti, and Toth (2001)
studied dissociation in a sample of maltreated and non-╉maltreated preschoolers and found that
physically abused, sexually abused, and neglected children all demonstrated greater dissociation
than non-╉maltreated children.
In primary school-╉age children, Shields and Cicchetti (1998) found that abuse was associated
with regulatory deficits characterized by emotional lability and negativity. In particular, physi-
cally abused children were more likely than their non-╉maltreated peers to be rated by counsellors
as expressing greater emotional negativity when transitioning from one activity to the next (i.e.,
becoming anxious, angry, or distressed) and as having more difficulty recovering emotionally
from exposure to stressful events. Similarly, school-╉aged girls with sexual abuse histories have
been found to display regulation difficulties. When compared to non-╉maltreated peers, these
abused children showed limited emotional awareness, a decreased capacity to regulate their emo-
tions appropriately and greater affective lability (Shipman et al., 2000).
Older children and adolescents with histories of abuse, or currently suffering abuse, often
express emotion regulation problems through high levels of arousal, greater reactive aggres-
sion, and have more difficulty regulating this response, when exposed to potentially threaten-
ing situations, such as interpersonal anger and conflict (Cummings, Hennessy, Rabideau, &
Cicchetti, 2008; Shackman & Pollak, 2014; Shields et al., 1994). Reactive aggression is emotion-
ally driven and associated with a high degree of sympathetic arousal and angry reactivity, and is
thought to be motivated by a desire to protect oneself from real or perceived threat. Physically
abused adolescents in particular, seem especially prone to exhibiting this form of aggression,
which can lead to poor social competence and difficulty with peer relationships (Shields &
Cicchetti, 1998).
5
1
3

Propagation and maintenance of outcomes 315

Additionally, different forms of neglect and abuse have been associated with differential expres-
sions of poor emotion regulation. For example, Pollak, Cicchetti, Hornung, and Reed (2000)
reported that neglected children, who often suffer from an extremely limited emotional envi-
ronment, had more difficulty in identifying and distinguishing between emotions, compared to
physically abused or non-╉maltreated children. Furthermore, a history of neglect has been found
to relate more strongly to a bias toward sad facial expressions; whereas, physical abuse has been
linked to a response bias to angry stimuli and a greater hostile attribution bias in ambiguous social
situations (Pollak et al., 2000, 2005; Pollak & Sinha, 2002; Shackman, Shackman, & Pollak, 2007).
This suggests physically abused children feel an exaggerated need to defend themselves from
perceived threats, which is reflected in their emotional processing (Dodge et  al., 1990; Rieder
& Cicchetti, 1989). Pollak et al. (2005) suggested that in physically abusive home environments
children learn to associate anger with threat of harm and therefore, become better prepared at
identifying threats. In contrast, neglect is typically associated with an emotionally impoverished
environment, with few opportunities for meaningful social interactions. If children are deprived
of interactive emotional experiences with others, their capacity to tolerate intense emotional
states may be underdeveloped, which can manifest in problems discriminating between emotions.
When examining additional types of maltreatment, childhood sexual abuse has been associated
with lower impulse control, whereas emotional abuse has been linked to impulsivity and problems
with behaving in accordance with desired goals (Oshri, Sutton, Clay-╉Warner, & Miller, 2015).
In summary, emotion regulation deficits manifest in various ways, including difficulties in iden-
tifying, understanding, expressing and regulating emotions, and can be dependent on the devel-
opmental stage of the child and the type of abuse experienced.

Propagation and maintenance of outcomes


The maladaptive emotion regulation strategies caused by abuse and neglect may leave children
vulnerable to other risks, and has been suggested as one reason why maltreated children are more
likely to develop clinical symptomatology and behavioral problems, both during childhood and in
adulthood. Impairment in emotion regulation is, therefore, both a negative consequence, and an
important mechanism connecting the experience of childhood abuse with subsequent difficulties
(Choi, Choi, Gim, Park, & Park, 2014).

Psychopathology
As noted above, child abuse and neglect have been associated with internalising disorders includ-
ing anxiety, depression, and PTSD (Cannon, Bonomi, Anderson, Rivara, & Thompson, 2010;
Gilbert, Widom, et  al., 2009; Kaufman, Plotsky, Nemeroff, & Charney, 2000; Springer et  al.,
2007), along with greater externalising problems, such as aggressive behavior and psychopathy
(Bernstein & Watson, 1997; Kolla et al., 2013; Lang, Klinteberg, & Alm, 2002; Weiler & Widom,
1996). Evidence has also indicated that many such outcomes associated with childhood abuse are
characterized by deficits in the processing and regulation of emotion (Burns, Jackson, & Harding,
2010; Etkin & Wager, 2007). Thus, child maltreatment may be linked with negative outcomes via
ineffective emotion regulation strategies, which may lead to a domino effect of deficits, resulting in
maladaptive functioning during childhood (Eberhart, Auerbach, Bigda-╉Peyton, & Abela, 2011).
Internalising problems, which can result from child abuse, have been linked with emotional
processing and regulation deficits. For example, enhanced anxiety symptoms are associated
with poorer emotional awareness and perception (Bradley, Mogg, White, Groom, & Bono,
1999; Eisenberg et al., 2001) and deficits in the regulation of emotions (Suveg, Morelen, Brewer,
& Thomassin, 2010; Suveg & Zeman, 2004). In addition, PTSD has also been associated with a
6
1
3

316 Children of Abuse and Neglect

lack of emotional clarity and acceptance, difficulty engaging in goal-​directed behavior when dis-
tressed, and an attentional bias towards trauma related stimuli (Buckley, Blanchard, & Trammell
Neill, 2000; Cloitre, Miranda, Stovall-​McClough, & Han, 2005; Tull, Barrett, McMillan, & Roemer,
2007). Similarly, depression is also characterized by deficits in regulating emotions (Joormann,
Siemer, & Gotlib, 2007), perceiving emotion in others (Stuhrmann, Suslow, & Dannlowski, 2011),
the inability to support oneself when experiencing negative emotions (Berking et al., 2011) and to
modify negative emotions (Ehring, Fischer, Schnülle, Bösterling, & Tuschen-​Caffier, 2008; Kassel,
Bornovalova, & Mehta, 2007). Finally, substance misuse is widely theorized as an effort to regulate
or avoid negative emotions (Baker, Piper, McCarthy, Majeskie, & Fiore, 2004; Wupperman et al.,
2012). For example, negative affect has been shown to predict increases in desire to drink and
drinking levels in individuals treated for alcohol dependence (Birch et al., 2004; Falk, Yi, & Hilton,
2008; Gamble et al., 2010; Sinha et al., 2009) and deficits in emotion regulation skills have been
shown to predict relapse during and after cognitive–​behavioral therapy for dependence (Berking
et al., 2011).
A number of studies have also reported emotional deficits in individuals with externalising
symptoms (Blair, Peschardt, Budhani, Mitchell, & Pine, 2006; Eisenberg et al., 2001; Hill, Degnan,
Calkins, & Keane, 2006), including deficits in empathy (Blair, 1995), experiencing emotion (Blair
et al., 2006; Frick, Lilienfeld, Ellis, Loney, & Silverthorn, 1999), and identifying emotional expres-
sions (Blair et  al., 2004; Iria & Barbosa, 2009; Pham & Philippot, 2010). For example, antiso-
cial behavior has been linked with deficits in perceiving negative emotions in facial expressions
and an inability to distinguish between emotions (Blair, Colledge, Murray, & Mitchell, 2001).
Additionally, psychopathic traits reflect greater emotional desensitisation, and an inability to
empathize or respond to the emotional needs of others (Weiler & Widom, 1996).
The emotion regulation problems that characterize children of abuse and these disorders seem
to be important for understanding linkages between maltreatment and maladjustment. The
emotion-​based deficits resulting from the abusive behavior of a caregiver and neurobiological
changes related to abuse, may in turn influence the development and maintenance of maladap-
tive psychological and behavioral outcomes (Dodge, 1991). This is supported by Kim-​Spoon,
Cicchetti, and Rogosch (2013) who showed that poor emotion regulation predicted a subsequent
increase in internalising symptomatology. Early maltreatment was associated with high emotion
lability at age seven, which contributed to poor emotion regulation at age eight, which in turn was
predictive of increases in internalising symptomatology, from age eight to nine.
In addition, numerous studies have established that the relationship between childhood abuse
and various psychological symptoms is mediated by impairments in emotion regulation (Choi
et al., 2014; Schwartz & Proctor, 2000; Shields & Cicchetti, 2001). For example, in women with
a history of childhood abuse, emotion regulation difficulties mediated the relationship between
abuse and current post-​traumatic symptomology (PTS) (Burns et  al., 2010; Choi & Oh, 2014;
Stevens et  al., 2013). Further to this, the relationship between specific dimensions of emotion
regulation and PTS was investigated among undergraduates with a history of trauma exposure
during childhood (Tull et  al., 2007). After controlling for negative affect, three dimensions of
emotion regulation—​difficulties in impulse-​control, diminished access to effective emotion regu-
lation strategies, and a lack of emotional clarity—​remained significant predictors of PTS symp-
tom severity. These findings provide support for the notion that difficulties in emotion regulation
could indirectly influence the maintenance of trauma symptoms. Specifically, abused or neglected
children were more likely to experience difficulties in emotion regulation, which then served to
exacerbate trauma symptoms. In turn, increased PTS resulted in increased physiological arousal,
maintaining the cycle of dysregulation, as increased arousal is harder to regulate (Tull et al., 2007).
7
1
3

Adaptive outcomes of emotional regulation deficits 317

Interpersonal problems
Emotion regulation also appears to be important for understanding linkages between abuse and
neglect and subsequent problems forming relationships. Maltreated children are at greater risk of
unpopularity among their peers (Anthonysamy & Zimmer-╉Gembeck, 2007; Bolger, Patterson, &
Kupersmidt, 1998; Rogosch, Cicchetti, & Aber, 1995), and are more likely to be rejected by peers,
not only on a single occasion, but also across multiple years from second through seventh grade
(Bolger & Patterson, 2001). Evidence has indicated that children’s deficits in emotional under-
standing mediated the link between earlier abuse and later interpersonal functioning and rejec-
tion by peers (Briere & Rickards, 2007; Cloitre et al., 2005; Rogosch, Cicchetti, & Aber, 1995).
Shipman et  al. (2005) found neglected children used avoidance strategies when responding
to other’s emotion distress. Such strategies reflect an emphasis on self-╉reliance when handling
emotional distress, as opposed to seeking help and support. For example, when asked how they
would respond to negative emotional displays in others, neglected children frequently indicated
that they would ignore or remove themselves from the situation. In contrast, non-╉maltreated
children generally indicated that they would provide the other child with assistance or support.
These avoidance strategies utilized by neglected children may hinder their ability to form and
maintain positive interpersonal relationships and have profoundly negative consequences for self-╉
regulation, which could add to the continuation of their abuse in contexts outside the home. This
is consistent with research, which demonstrates that emotional understanding is related to peer
acceptance (Cassidy & Parke, 1991; Denham et al., 1990; Underwood, 1997).
Children who have difficulty managing their negative emotions are more likely to become dis-
ruptive, impulsive and reactively aggressive in social interactions, leading to lower acceptance and
more rejection by peers (Maszk, Eisenberg, & Guthrie, 1999). Peer rejection can then place abused
children at risk of subsequent adjustment problems, including internalising and externalising dis-
orders (Kupersmidt & Coie, 1990; Ladd & Troop-╉Gordon, 2003). Kim and Cicchetti (2010) found
support for this mechanism. In this study early experiences of abuse and neglect were related
to emotion dysregulation, which placed these children at a greater risk of peer rejection, which
then contributed to internalising and externalising symptomology one year later, after controlling
for initial symptomology. Kim and Cicchetti (2010) argued that in the absence of positive peer
interactions and relationships, abused children may become more vulnerable to stress, which can
manifest as internalising or externalising disorders. Maladaptive emotion regulation, therefore,
may impede children’s abilities to establish positive peer relationships due to an underdevelop-
ment of certain traits, such as perspective-╉taking and empathy, which are vital to the development
of social competence.

Adaptive outcomes of emotional regulation deficits


Although these patterns of emotion regulation are typically viewed as maladaptive and may
result in impaired functioning, in neglectful and abusive situations these strategies may serve cer-
tain adaptive functions that enable children to protect themselves from further harm (Rogosch,
Cicchetti, Shields, & Toth, 1995). For example, in a home environment in which the child’s emo-
tional expressions elicit an aversive or punitive reaction from the caregiver, a child may suppress
their future emotions in an attempt to avoid further harm and hostile reactions. Shipman et al.
(2005) found that although neglected children did not understand emotion in a manner con-
sistent with cultural norms, their understanding may be adaptive within the neglectful context.
Therefore, in response to questions regarding self-╉awareness of emotional experience, neglected
children tended to deny experiencing negative emotions (e.g., “I never feel mad at any one”). This
8
1
3

318 Children of Abuse and Neglect

decreased self-╉awareness may help neglected children cope with a home environment which fails
to support emotional expressiveness.
Maltreated children’s hypervigilance to potential threats and overregulation of distress responses
may also assist in reducing rates of abuse. For children who have experienced abuse, displays of
anger in their environment are the strongest predictors of threat and therefore a selective attention
to threat-╉related (i.e., angry) stimuli at the expense of attention to other emotional cues would
be adaptive, as children are quickly able to detect signs of anger and remove themselves from a
potentially dangerous situation (Pollak et al., 2005). Additionally, managing the intense emotions
arising from physical abuse may also include cognitive strategies that enable the child to maintain
a sense of control over their circumstances.
However, although emotional withdrawal and hypervigilance during conflict can result in
short-╉term relief for the abused child, it may also heighten the long-╉term risk of future abuse both
in the home and in other contexts. Thus, immediate goals may conflict with longer-╉term goals
and the emotional strategies developed may be unsuccessful at accomplishing both, leaving the
child vulnerable to further risks. In settings outside the home, these deviant strategies of emotion
regulation may simultaneously create other problems, such as social difficulties and poor peer
relations (Cicchetti & Schneider-╉Rosen, 1986). Ultimately, Thompson and Calkins (1996) suggest
that although children may adapt by trying to cope with the emotional demands of abuse and
neglect, there are likely to be no optimal approaches to emotion regulation available to them and
efforts are likely to result in a problematic mixture of adaptive and maladaptive outcomes.
In summary, emotion regulation strategies developed in the context of abusive homes may
serve specific protective functions, while concurrently resulting in maladaptive outcomes, with
evidence suggesting these emotional deficits play a role in the development and maintenance of
psychopathology and interpersonal difficulties. In turn, such problems can make a child more
vulnerable to future abuse and victimisation both in and outside the home.

Implications for intervention
The body of research on emotion regulation and child abuse and neglect has important implica-
tions for the development and utilisation of intervention programs targeted at abusive caregiv-
ers and their children. Difficulties with emotion regulation have become an important target of
clinical interventions in maltreated children (Cloitre, Koenen, Cohen, & Han, 2002) and many
approaches to individual psychotherapy with victims of abuse focus on improving emotion regu-
lation skills (Leahy, Tirch, & Napolitano, 2011; Paivio & Laurent, 2001). These skills play a vital
role in attaining successful social interactions and psychological adjustment and, therefore, may
be a source of resiliency related to maltreatment and psychopathology (Chang, Schwartz, Dodge,
& McBride-╉Chang, 2003; Shields et al., 1994). If intervention programs can help abused children
to develop such resiliency this may reduce the emotional, psychological and social problems asso-
ciated with neglect and abuse.
Several empirically supported group treatments emphasize building emotion regulation and
interpersonal skills in the aftermath of abuse; these include Dialectical Behavior Therapy (Robins,
Schmidt, & Linehan, 2004), Seeking Safety (Najavits, 2002), Trauma Adaptive Recovery Group
Education and Therapy, (Ford & Russo, 2006), and Skills Training in Affect and Interpersonal
Regulation (STAIR)/╉Prolonged Exposure (Cloitre et al., 2002). STAIR is a cognitive–╉behavioral
treatment that targets the development of emotion management and interpersonal skills. Each
session focuses on a different deficit understood within the context of the experience of child
abuse:  1)  labelling and identifying emotions, 2)  managing emotions, 3)  distress tolerance,
4) acceptance of emotions and enhanced experiencing of positive emotions, 5) identification of
9
1
3

Conclusion 319

trauma based interpersonal schemas and their enactment in day-╉to-╉day life, 6) identification of
conflict between trauma-╉generated feelings and current interpersonal goals, 7) role plays related
to issues of power and control, and 8)  role plays highlighting the presence and expression of
emotion, related to developing flexibility in interpersonal situations involving power differentials.
This program was found to be valuable for the emotional development of women with a history
of child abuse, who showed significant reductions in negative mood regulation, anger expression
and interpersonal skills deficits, when compared to those on the wait-╉list (Cloitre et al., 2002).
However, an unmet need remains for the more extensive use of emotion regulation skills train-
ing in the context of child abuse and neglect (Hernandez, Nesman, Mowery, Acevedo-╉Polakovich,
& Callejas, 2009). In particular, policies and programs that focus on parenting behaviors are likely
to be beneficial. A large body of research suggests caregiver behavior can have a positive influ-
ence on children’s emotional development. Morris et al. (2011) examined specific parenting prac-
tices associated with children’s emotion management, assessing whether particular practices were
linked to successful emotion management. Findings indicated that certain practices were more
effective at improving children’s emotion regulation. Specifically, redirecting attention away from,
and cognitively reframing emotions with the child, were the most effective strategies used by
mothers to help children manage the expression of negative emotions, and were associated with
less expressed anger and sadness. Further parenting practices which have been linked to improved
emotion regulation skills include directing positive emotion and behaviors toward children, help-
ing children to label and discuss emotions, soothing children’s reactions through appropriate
physical contact and encouraging activities such as reading or drawing to reduce arousal (Calkins
& Hill, 2007; Eisenberg et al., 2001). Such practices should therefore be utilized in interventions
targeted at abusive caregivers.
Furthermore, school-╉based interventions that focus on promoting social emotional compe-
tence among children have been shown to be effective in fostering emotion regulation develop-
ment. For example, the Promoting Alternative Thinking Strategies (PATHS) curriculum has been
successful in reducing internalising problems among primary school students by teaching them to
identify, understand, and discuss their emotions (Greenberg, Kusche, Cook, & Quamma, 1995).
In the PATHS program improvements were seen in the range of emotion related vocabulary, flu-
ency in discussing emotional experiences, efficacy beliefs regarding the management of emotions,
and the developmental understanding of emotions. Such training could potentially be applied to
equip abused children with the emotional skills they fail to develop in abusive homes.

Conclusion
Through processes of caregiver socialisation and neurobiological changes, child abuse and neglect
can lead to a range of emotional problems during childhood, which can endure into adulthood.
These emotion regulation deficits are manifested throughout development in a variety of ways,
including difficulties in identifying, understanding, expressing and regulating emotions and
hypervigilance. Although the altered emotion regulation strategies developed by abused and
neglected children do have certain adaptive qualities, which allows the child to avoid further
emotional or physical harm by quickly removing themselves from conflict, evidence indicates that
such strategies also lead to a range of maladaptive consequences. Via mechanisms of emotion reg-
ulation, abuse can lead to internalising and externalising problems, and interpersonal problems,
including peer rejection. The presence of such psychological, behavioral and social problems can
result in the maintenance of abuse.
The extensive range of evidence, which indicates how critical emotion regulation problems are
in the relationship between child abuse and negative outcomes, has led to a range of interventions
0
2
3

320 Children of Abuse and Neglect

which target emotion regulation skill building. There is potential for further usage of such
approaches in interventions to train caregivers in the importance of socialisation to develop adap-
tive emotion regulation strategies in children, and for at-╉risk and abused children to build resil-
ience through improved emotional development.

References
Ackard, D. M., & Neumark-╉Sztainer, D. (2003). Multiple sexual victimizations among adolescent boys
and girls: prevalence and associations with eating behaviors and psychological health. Journal of Child
Sexual Abuse, 12(1), 17–╉37. doi:10.1300/╉J070v12n01_╉02
Afifi, T. O., Enns, M. W., Cox, B. J., Asmundson, G. J. G., Stein, M. B., & Sareen, J. (2008). Population
attributable fractions of psychiatric disorders and suicide ideation and attempts associated with
adverse childhood experiences. American Journal of Public Health, 98(5), 946–╉952. doi:10.2105/╉
AJPH.2007.120253
Alessandri, S. M. (1991). Play and social behavior in maltreated preschoolers. Development and
Psychopathology, 3(2), 191–╉206. doi:10.1017/╉S0954579400000079
Anthonysamy, A., & Zimmer-╉Gembeck, M. J. (2007). Peer status and behaviors of maltreated children
and their classmates in the early years of school. Child Abuse & Neglect, 31(9), 971–╉991. doi: 10.1016/╉
j.chiabu.2007.04.004
Baker, T. B., Piper, M. E., McCarthy, D. E., Majeskie, M. R., & Fiore, M. C. (2004). Addiction motivation
reformulated: an affective processing model of negative reinforcement. Psychological Review, 111(1),
33–╉51. doi:10.1037/╉0033-╉295X.111.1.33
Bardi, M., & Borgognini-╉Tarli, S. M. (2001). A survey on parent-╉child conflict resolution: intrafamily
violence in Italy. Child Abuse & Neglect, 25(6), 839–╉853. doi:10.1016/╉S0145-╉2134(01)00242-╉3
Beeghly, M., & Cicchetti, D. (1994). Child maltreatment, attachment, and the self system: Emergence of
an internal state lexicon in toddlers at high social risk. Development and Psychopathology, 6(1), 5–╉30.
doi:10.1017/╉S095457940000585X
Berking, M., Margraf, M., Ebert, D., Wupperman, P., Hofmann, S. G., & Junghanns, K. (2011). Deficits in
emotion-╉regulation skills predict alcohol use during and after cognitive-╉behavioral therapy for alcohol
dependence. Journal of Consulting and Clinical Psychology, 79(3), 307–╉318. doi:10.1037/╉a0023421
Bernstein, J. Y., & Watson, M. W. (1997). Children Who Are Targets of Bullying: A Victim Pattern. Journal
of Interpersonal Violence, 12(4), 483–╉498. doi:10.1177/╉088626097012004001
Birch, C. D., Stewart, S. H., Wall, A.-╉M., McKee, S. A., Eisnor, S. J., & Theakston, J. A. (2004). Mood-╉
induced increases in alcohol expectancy strength in internally motivated drinkers. Psychology of
Addictive Behaviors╯: Journal of the Society of Psychologists in Adictive Behaviors, 18(3), 231–╉238.
doi: 10.1037/╉0893-╉164X.18.3.231
Blair, R. (1995). A cognitive developmental approach to morality: investigating the psychopath. Cognition,
57(1), 1–╉29. doi:10.1016/╉0010-╉0277(95)00676-╉P
Blair, R. J. R., Colledge, E., Murray, L., & Mitchell, D. G. V. (2001). A Selective Impairment in the
Processing of Sad and Fearful Expressions in Children with Psychopathic Tendencies. Journal of
Abnormal Child Psychology, 29(6), 491–╉498. doi:10.1023/╉A:1012225108281
Blair, R. J. R., Mitchell, D. G. V., Peschardt, K. S., Colledge, E., Leonard, R. A., Shine, J. H., … Perrett, D.
I. (2004). Reduced sensitivity to others’ fearful expressions in psychopathic individuals. Personality and
Individual Differences, 37(6), 1111–╉1122. doi: 10.1016/╉j.paid.2003.10.008
Blair, R. J. R., Peschardt, K. S., Budhani, S., Mitchell, D. G. V, & Pine, D. S. (2006). The development
of psychopathy. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 47(3-╉4), 262–╉276.
doi:10.1111/╉j.1469-╉7610.2006.01596.x
Bolger, K. E., & Patterson, C. J. (2001). Developmental Pathways from Child Maltreatment to Peer
Rejection. Child Development, 72(2), 549–╉568. doi:10.1111/╉1467-╉8624.00296
1
2
3

Conclusion 321

Bolger, K. E., Patterson, C. J., & Kupersmidt, J. B. (1998). Peer Relationships and Self-​Esteem among
Children Who Have Been Maltreated. Child Development, 69(4), 1171–​1197. doi 10.1111/​j.1467-​
8624.1998.tb06166.x
Bousha, D. M., & Twentyman, C. T. (1984). Mother–​child interactional style in abuse, neglect, and control
groups: Naturalistic observations in the home. Journal of Abnormal Psychology, 93(1), 106–​114.
Bowlby, J. (1969). Attachment: Vol 1. Attachment and loss. London: Hogarth Press.
Bowlby, J. (1982). Attachment and loss: retrospect and prospect. The American Journal of Orthopsychiatry,
52(4), 664–​678. Retrieved from http://​europepmc.org/​abstract/​med/​7148988
Bradley, B. P., Mogg, K., White, J., Groom, C., & Bono, J. (1999). Attentional bias for emotional faces
in generalized anxiety disorder. British Journal of Clinical Psychology, 38(3), 267–​278. doi:10.1348/​
014466599162845
Bradley, E. J., & Peters, R. D. (1991). Physically abusive and nonabusive mothers’ perceptions of parenting
and child behavior. American Journal of Orthopsychiatry, 61(3), 455–​460. doi:10.1037/​h0079263
Briere, J., & Jordan, C. E. (2009). Childhood maltreatment, intervening variables, and adult psychological
difficulties in women: an overview. Trauma, Violence & Abuse, 10(4), 375–​388. doi:10.1177/​
1524838009339757
Briere, J., & Rickards, S. (2007). Self-​awareness, affect regulation, and relatedness: differential sequels of
childhood versus adult victimization experiences. The Journal of Nervous and Mental Disease, 195(6),
497–​503. doi:10.1097/​NMD.0b013e31803044e2
Buckley, T., Blanchard, E. B., & Trammell Neill, W. (2000). Information processing and ptsd
A review of the empirical literature. Clinical Psychology Review, 20(8), 1041–​1065. doi:10.1016/​
S0272-​7358(99)00030-​6
Bugental, D. B., & Others, A. (1989). Caregiver Beliefs and Dysphoric Affect Directed to Difficult Children.
Developmental Psychology, 26(4), 631–​638. Retrieved from http://​eric.ed.gov/​?id=EJ415413
Burns, E. E., Jackson, J. L., & Harding, H. G. (2010). Child Maltreatment, Emotion Regulation, and
Posttraumatic Stress: The Impact of Emotional Abuse. Journal of Aggression, Maltreatment & Trauma.
Retrieved from http://​www.tandfonline.com/​doi/​abs/​10.1080/​10926771.2010.522947#.VbeQ4bNVhBc
Caliso, J. A., & Milner, J. S. (1992). Childhood history of abuse and child abuse screening. Child Abuse &
Neglect, 16(5), 647–​659. doi: 10.1016/​0145-​2134(92)90103-​X
Calkins, S. D. (1994). Origins and outcomes of individual differences in emotion regulation. Monographs of
the Society for Research in Child Development, 59(2-​3), 53–​72. doi:10.1111/​j.1540-​5834.1994.tb01277.x
Camras, L. A., Sachs-​Alter, E., & Ribordy, S. C. (1996). Emotion understanding in maltreated
children: Recognition of facial expressions and integration with other emotion cues. In M. W.
Lewis, Michael; Sullivan (Ed.), Emotional development in atypical children (pp. 203–​225). Hillsdale,
NJ: Lawrence Erlbaum Associates, Inc.
Cannon, E. A., Bonomi, A. E., Anderson, M. L., Rivara, F. P., & Thompson, R. S. (2010). Adult Health
and Relationship Outcomes Among Women With Abuse Experiences During Childhood. Violence and
Victims, 25(3), 291–​305. doi:10.1891/​0886-​6708.25.3.291
Cantos, A. L., Neale, J. M., O’Leary, K. D., & Gaines, R. W. (1997). Assessment of coping strategies of child
abusing mothers. Child Abuse & Neglect, 21(7), 631–​636. doi:10.1016/​S0145-​2134(97)00022-​7
Cash, S. J., & Wilke, D. J. (2003). An ecological model of maternal substance abuse and child neglect: issues,
analyses, and recommendations. The American Journal of Orthopsychiatry, 73(4), 392–​404. doi:10.1037/​
0002-​9432.73.4.392
Chan, Y. C. (1994). Parenting stress and social support of mothers who physically abuse their children in
Hong Kong. Child Abuse & Neglect, 18(3), 261–​269. doi:10.1016/​0145-​2134(94)90110-​4
Chang, L., Schwartz, D., Dodge, K. A., & McBride-​Chang, C. (2003). Harsh parenting in relation to
child emotion regulation and aggression. Journal of Family Psychology, 17(4), 598–​606. doi:10.1037/​
0893-​3200.17.4.598
23

322 Children of Abuse and Neglect

Choi, J. Y., Choi, Y. M., Gim, M. S., Park, J. H., & Park, S. H. (2014). The effects of childhood abuse on
symptom complexity in a clinical sample: mediating effects of emotion regulation difficulties. Child
Abuse & Neglect, 38(8), 1313–​1319. doi:10.1016/​j.chiabu.2014.04.016
Christensen, M. J., Brayden, R. M., Dietrich, M. S., McLaughlin, F. J., Sherrod, K. B., & Altemeier, W.
A. (1994). The prospective assessment of self-​concept in neglectful and physically abusive low income
mothers. Child Abuse & Neglect, 18(3), 225–​232. doi: 10.1016/​0145-​2134(94)90107-​4
Cicchetti, D., & Curtis, W. J. (2005). An event-​related potential study of the processing of affective facial
expressions in young children who experienced maltreatment during the first year of life. Development
and Psychopathology, 17(3), 641–​677. doi:10.1017/​S0954579405050315
Cicchetti, D., & Howes, P. W. (1991). Developmental psychopathology in the context of the
family : illustrations from the study of child maltreatment. Canadian Journal of Behavioural Science,
23(3), 257–​281. Retrieved from http://​psycnet.apa.org/​journals/​cbs/​23/​3/​257/​
Cicchetti, D., & Rogosch, F. A. (1997). The role of self-​organization in the promotion of resilience in
maltreated children. Development and Psychopathology, 9(4), 797–​815. Retrieved from http://​journals.
cambridge.org/​abstract_​S0954579497001442
Cicchetti, D., & Tucker, D. (1994). Development and self-​regulatory structures of the mind. Development
and Psychopathology, 6(4), 533–​549. doi: 10.1017/​S0954579400004673
Cloitre, M., Koenen, K. C., Cohen, L. R., & Han, H. (2002). Skills training in affective and interpersonal
regulation followed by exposure: A phase-​based treatment for PTSD related to childhood abuse. Journal
of Consulting and Clinical Psychology, 70(5), 1067–​1074.
Cloitre, M., Miranda, R., Stovall-​McClough, K., & Han, H. (2005). Beyond PTSD: Emotion regulation
and interpersonal problems as predictors of functional impairment in survivors of childhood abuse.
Behavior Therapy, 36(2), 119–​124. doi:10.1016/​S0005-​7894(05)80060-​7
Cole, P. M., Zahn-​Waxler, C., & Smith, K. D. (1994). Expressive Control during a Disappointment:
Variations Related to Preschoolers’ Behavior Problems. Developmental Psychology, 30(6), 835–​846.
Retrieved from http://​eric.ed.gov/​?id=EJ498080
Connell-​Carrick, K. (2003). A Critical Review of the Empirical Literature: Identifying Correlates of Child
Neglect. Child and Adolescent Social Work Journal, 20(5), 389–​425. doi:10.1023/​A:1026099913845
Coohey, C., & Braun, N. (1997). Toward an integrated framework for understanding child physical abuse.
Child Abuse & Neglect, 21(11), 1081–​1094. doi:10.1016/​S0145-​2134(97)00067-​7
Corcoran, J., & Nichols-​Casebolt, A. (2004). Risk and Resilience Ecological Framework for Assessment
and Goal Formulation. Child and Adolescent Social Work Journal, 21(3), 211–​235. doi:10.1023/​
B:CASW.0000028453.79719.65
Cummings, E. M., Hennessy, K. D., Rabideau, G. J., & Cicchetti, D. (2008). Responses of physically abused
boys to interadult anger involving their mothers. Development and Psychopathology, 6(01), 31–​41.
doi:10.1017/​S0954579400005861
Dadds, M. R., & Salmon, K. (2003). Punishment Insensitivity and Parenting: Temperament and Learning
as Interacting Risks for Antisocial Behavior. Clinical Child and Family Psychology Review, 6(2), 69–​86.
doi:10.1023/​A:1023762009877
Damashek, A., Nelson, M. M., & Bonner, B. L. (2013). Fatal child maltreatment: characteristics of
deaths from physical abuse versus neglect. Child Abuse & Neglect, 37(10), 735–​44. doi:10.1016/​
j.chiabu.2013.04.014
Danese, A., Moffitt, T. E., Harrington, H., Milne, B. J., Polanczyk, G., Pariante, C. M., … Caspi, A. (2009).
Adverse childhood experiences and adult risk factors for age-​related disease: depression, inflammation,
and clustering of metabolic risk markers. Archives of Pediatrics & Adolescent Medicine, 163(12), 1135–​
1143. doi:10.1001/​archpediatrics.2009.214
Dodge, K. A. (1991). The structure and function of reactive and proactive aggression. In D. J. Pepler &
K. H. Rubin (Eds.), The development and treatment of childhood aggression (pp. 201–​218). Hillsdale,
NJ: Lawrence Erlbaum Associates, Inc.
3
2

Conclusion 323

Dodge, K., Bates, J., & Pettit, G. (1990). Mechanisms in the cycle of violence. Science, 250(4988), 1678–​
1683. doi:10.1126/​science.2270481
Donohue, B. (2004). Coexisting child neglect and drug abuse in young mothers: specific recommendations
for treatment based on a review of the outcome literature. Behavior Modification, 28(2), 206–​233.
doi: 10.1177/​0145445503259486
Dubowitz, H. (2009). Tackling child neglect: a role for pediatricians. Pediatric Clinics of North America,
56(2), 363–​378. doi:10.1016/​j.pcl.2009.01.003
During, S. M., & McMahon, R. J. (1991). Recognition of Emotional Facial Expressions by Abusive
Mothers and Their Children. Journal of Clinical Child Psychology, 20(2), 132–​139. doi: 10.1207/​
s15374424jccp2002_​4
Eberhart, N. K., Auerbach, R. P., Bigda-​Peyton, J., & Abela, J. R. Z. (2011). Maladaptive Schemas and
Depression: Tests of Stress Generation and Diathesis-​Stress Models. Journal of Social and Clinical
Psychology, 30(1), 75–​104. doi: 10.1521/​jscp.2011.30.1.75
Ehring, T., Fischer, S., Schnülle, J., Bösterling, A., & Tuschen-​Caffier, B. (2008). Characteristics of emotion
regulation in recovered depressed versus never depressed individuals. Personality and Individual
Differences, 44(7), 1574–​1584. doi:10.1016/​j.paid.2008.01.013
Eisenberg, N., Cumberland, A., Spinrad, T. L., Fabes, R. A., Shepard, S. A., Reiser, M., … Guthrie, I. K.
(2001). The Relations of Regulation and Emotionality to Children’s Externalizing and Internalizing
Problem Behavior. Child Development, 72(4), 1112–​1134. doi:10.1111/​1467-​8624.00337
Eisenberg, N., & Morris, A. S. (2002). Children’s emotion-​related regulation. Advances in Child
Development and Behavior, 30, 189–​229. Retrieved from http://​europepmc.org/​abstract/​med/​12402675
Erickson, M. F., Egeland, B., & Pianta, R. (1989). The effects of maltreatment on the development of young
children. In D. Cicchetti & V. Carlson (Eds.), Theory and research on the causes and consequences of
child abuse and neglect (pp. 647–​684). New York: Cambridge University Press.
Ethier, L. S., Lemelin, J.-​P., & Lacharité, C. (2004). A longitudinal study of the effects of chronic
maltreatment on children’s behavioral and emotional problems. Child Abuse & Neglect, 28(12), 1265–​
1278. doi:10.1016/​j.chiabu.2004.07.006
Etkin, A., & Wager, T. D. (2007). Functional neuroimaging of anxiety: a meta-​analysis of emotional
processing in PTSD, social anxiety disorder, and specific phobia. The American Journal of Psychiatry,
164(10), 1476–​1488. doi:10.1176/​appi.ajp.2007.07030504
Falk, D. E., Yi, H.-​Y., & Hilton, M. E. (2008). Age of onset and temporal sequencing of lifetime DSM-​IV
alcohol use disorders relative to comorbid mood and anxiety disorders. Drug and Alcohol Dependence,
94(1-​3), 234–​245. doi:10.1016/​j.drugalcdep.2007.11.022
Felitti, V. J., Anda, R. F., Nordenberg, D., Williamson, D. F., Spitz, A. M., Edwards, V., … Marks, J.
S. (1998). Relationship of Childhood Abuse and Household Dysfunction to Many of the Leading
Causes of Death in Adults. American Journal of Preventive Medicine, 14(4), 245–​258. doi:10.1016/​
S0749-​3797(98)00017-​8
Fergusson, D. M., Boden, J. M., & Horwood, L. J. (2008). Exposure to childhood sexual and physical
abuse and adjustment in early adulthood. Child Abuse & Neglect, 32(6), 607–​619. doi:10.1016/​
j.chiabu.2006.12.018
Finkelhor, D. (1994). The international epidemiology of child sexual abuse. Child Abuse & Neglect, 18(5),
409–​417. doi:10.1016/​0145-​2134(94)90026-​4
Flaherty, E. G., Sege, R. D., Griffith, J., Price, L. L., Wasserman, R., Slora, E., … Binns, H. J. (2008). From
suspicion of physical child abuse to reporting: primary care clinician decision-​making. Pediatrics,
122(3), 611–​619. doi:10.1542/​peds.2007-​2311
Ford, J. D., & Russo, E. (2006). Trauma-​focused, present-​centered, emotional self-​regulation approach to
integrated treatment for posttraumatic stress and addiction: trauma adaptive recovery group education
and therapy (TARGET). American Journal of Psychotherapy, 60(4), 335–​355. Retrieved from http://​
www.ncbi.nlm.nih.gov/​pubmed/​17340945
4
2
3

324 Children of Abuse and Neglect

Fraser, J. A., Mathews, B., Walsh, K., Chen, L., & Dunne, M. (2010). Factors influencing child abuse and
neglect recognition and reporting by nurses: a multivariate analysis. International Journal of Nursing
Studies, 47(2), 146–​153. doi:10.1016/​j.ijnurstu.2009.05.015
Frick, P. J., Lilienfeld, S. O., Ellis, M., Loney, B., & Silverthorn, P. (1999). The Association between Anxiety
and Psychopathy Dimensions in Children. Journal of Abnormal Child Psychology, 27(5), 383–​392.
doi:10.1023/​A:1021928018403
Gaensbauer, T. J. (1982). Regulation of Emotional Expression in Infants from Two Contrasting Caretaking
Environments. Journal of the American Academy of Child Psychiatry, 21(2), 163–​170. doi:10.1016/​
S0002-​7138(09)60915-​8
Gamble, S. A., Conner, K. R., Talbot, N. L., Yu, Q., Tu, X. M., & Connors, G. J. (2010). Effects of
pretreatment and posttreatment depressive symptoms on alcohol consumption following treatment in
Project MATCH. Journal of Studies on Alcohol and Drugs, 71(1), 71–​77. Retrieved from http://​www.
pubmedcentral.nih.gov/​articlerender.fcgi?artid=2815065&tool=pmcentrez&rendertype=abstract
Gaudin, J. M., Polansky, N. A., Kilpatrick, A. C., & Shilton, P. (1996). Family functioning in neglectful
families. Child Abuse & Neglect, 20(4), 363–​377. doi:10.1016/​0145-​2134(96)00005-​1
Gilbert, R., Kemp, A., Thoburn, J., Sidebotham, P., Radford, L., Glaser, D., & Macmillan, H. L. (2009).
Recognising and responding to child maltreatment. Lancet (London, England), 373(9658), 167–​180.
doi:10.1016/​S0140-​6736(08)61707-​9
Gilbert, R., Widom, C. S., Browne, K., Fergusson, D., Webb, E., & Janson, S. (2009). Burden and
consequences of child maltreatment in high-​income countries. Lancet, 373(9657), 68–​81. doi:10.1016/​
S0140-​6736(08)61706-​7
Gillham, B., Tanner, G., Cheyne, B., Freeman, I., Rooney, M., & Lambie, A. (1998). Unemployment rates,
single parent density, and indices of child poverty: Their relationship to different categories of child
abuse and neglect. Child Abuse & Neglect, 22(2), 79–​90. doi:10.1016/​S0145-​2134(97)00134-​8
Goldman, J., Salus, M. K., Wolcott, D., & Kennedy, K. Y. (2003). A Coordinated Response to Child
Abuse and Neglect: The Foundation for Practice. Child Abuse and Neglect User Manual Series.
Washington, DC: US Department of health and human services. Retrieved from http://​eric.ed.gov/​
?id=ED474857
Gottman, J. M., Katz, L. F., & Hooven, C. (1996). Parental meta-​emotion philosophy and the emotional life
of families: Theoretical models and preliminary data. Journal of Family Psychology, 10(3), 243–​268.
Greenberg, M. T., Kusche, C. A., Cook, E. T., & Quamma, J. P. (1995). Promoting emotional competence
in school-​aged children: The effects of the PATHS curriculum. Development and Psychopathology, 7(1),
117–​136.
Greenberg, M. T., Speltz, M. L., & Deklyen, M. (1993). The role of attachment in the early development
of disruptive behavior problems. Development and Psychopathology, 5(1-​2), 191. doi:10.1017/​
S095457940000434X
Harrington, D., Black, M. M., Starr, R. H., & Dubowitz, H. (1998). Child neglect: Relation to child
temperament and family context. American Journal of Orthopsychiatry, 68(1), 108–​116. doi:10.1037/​
h0080275
Heim, C., & Nemeroff, C. B. (2001). The role of childhood trauma in the neurobiology of mood and
anxiety disorders: preclinical and clinical studies. Biological Psychiatry, 49(12), 1023–​1039. doi:10.1016/​
S0006-​3223(01)01157-​X
Heim, C., Shugart, M., Craighead, W. E., & Nemeroff, C. B. (2010). Neurobiological and psychiatric
consequences of child abuse and neglect. Developmental Psychobiology, 52(7), 671–​690. doi:10.1002/​
dev.20494
Hernandez, M., Nesman, T., Mowery, D., Acevedo-​Polakovich, I. D., & Callejas, L. M. (2009). Cultural
competence: a literature review and conceptual model for mental health services. Psychiatric Services
(Washington, D.C.), 60(8), 1046–​1050. doi:10.1176/​appi.ps.60.8.1046
Hibbard, R. A., & Desch, L. W. (2007). Maltreatment of children with disabilities. Pediatrics, 119(5), 1018–​
1025. doi:10.1542/​peds.2007-​0565
5
2
3

Conclusion 325

Higgins, D. J., & McCabe, M. P. (2001). Multiple forms of child abuse and neglect: adult retrospective
reports. Aggression and Violent Behavior, 6(6), 547–​578. doi:10.1016/​S1359-​1789(00)00030-​6
Hill, A. L., Degnan, K. A., Calkins, S. D., & Keane, S. P. (2006). Profiles of externalizing behavior problems
for boys and girls across preschool: the roles of emotion regulation and inattention. Developmental
Psychology, 42(5), 913–​928. doi:10.1037/​0012-​1649.42.5.913
Holt, S., Buckley, H., & Whelan, S. (2008). The impact of exposure to domestic violence on children
and young people: a review of the literature. Child Abuse & Neglect, 32(8), 797–​810. doi:10.1016/​
j.chiabu.2008.02.004
Hornor, G. (2014). Child neglect: assessment and intervention. Journal of Pediatric Health Care : Official
Publication of National Association of Pediatric Nurse Associates & Practitioners, 28(2), 186–​192; quiz
193–​4. doi:10.1016/​j.pedhc.2013.10.002
Howes, P. W., Cicchetti, D., Toth, S. L., & Rogosch, F. A. (2000). Affective, organizational, and relational
characteristics of maltreating families: A system’s perspective. Journal of Family Psychology, 14(1),
95–​110. doi:10.1037//​0893-​3200.14.1.95
Huang, S., Trapido, E., Fleming, L., Arheart, K., Crandall, L., French, M., … Prado, G. (2011). The
long-​term effects of childhood maltreatment experiences on subsequent illicit drug use and
drug-​related problems in young adulthood. Addictive Behaviors, 36(1-​2), 95–​102. doi:10.1016/​
j.addbeh.2010.09.001
Hussey, J. M., Chang, J. J., & Kotch, J. B. (2006). Child maltreatment in the United States: prevalence, risk
factors, and adolescent health consequences. Pediatrics, 118(3), 933–​942. doi:10.1542/​peds.2005-​2452
Iria, C., & Barbosa, F. (2009). Perception of facial expressions of fear : comparative research with criminal
and non-​criminal psychopaths. The Journal of Forensic Psychiatry & Psychology, 20(1), 66–​73. Retrieved
from http://​www.tandfonline.com/​doi/​abs/​10.1080/​14789940802214218
Jaudes, P. K., Ekwo, E., & Van Voorhis, J. (1995). Association of drug abuse and child abuse. Child Abuse &
Neglect, 19(9), 1065–​1075. doi:10.1016/​0145-​2134(95)00068-​J
Jones, R., Flaherty, E. G., Binns, H. J., Price, L. L., Slora, E., Abney, D., … Sege, R. D. (2008). Clinicians’
description of factors influencing their reporting of suspected child abuse: report of the Child
Abuse Reporting Experience Study Research Group. Pediatrics, 122(2), 259–​266. doi:10.1542/​
peds.2007-​2312
Joormann, J., Siemer, M., & Gotlib, I. H. (2007). Mood regulation in depression: Differential effects of
distraction and recall of happy memories on sad mood. Journal of Abnormal Psychology, 116(3), 484–​
490. doi:10.1037/​0021-​843X.116.3.484
Kaplan, M. J., & Klinetob, N. A. (2000). Childhood Emotional Trauma and Chronic Posttraumatic Stress
Disorder in Adult Outpatients with Treatment-​Resistant Depression. The Journal of Nervous and Mental
Disease, 188(9), 596–​601. doi:10.1097/​00005053-​200009000-​00006
Kassel, J. D., Bornovalova, M., & Mehta, N. (2007). Generalized expectancies for negative mood regulation
predict change in anxiety and depression among college students. Behaviour Research and Therapy,
45(5), 939–​950. doi:10.1016/​j.brat.2006.07.014
Katerndahl, D., Burge, S., & Kellogg, N. (2005). Predictors of Development of Adult Psychopathology in
Female Victims of Childhood Sexual Abuse. The Journal of Nervous and Mental Disease, 193(4),
258–​264. doi:10.1097/​01.nmd.0000158362.16452.2e
Kaufman, J., Plotsky, P. M., Nemeroff, C. B., & Charney, D. S. (2000). Effects of early adverse experiences
on brain structure and function: clinical implications. Biological Psychiatry, 48(8), 778–​790.
doi:10.1016/​S0006-​3223(00)00998-​7
Kavanagh, K. A., Youngblade, L., Reid, J. B., & Fagot, B. I. (1988). Interactions Between Children and
Abusive Versus Control Parents. Journal of Clinical Child Psychology, 17(2), 137–​142. Retrieved from
http://​www.tandfonline.com/​doi/​abs/​10.1207/​s15374424jccp1702_​5#.Vd72fSVVhBc
Kendler, K. S., Bulik, C. M., Silberg, J., Hettema, J. M., Myers, J., & Prescott, C. A. (2000). Childhood
Sexual Abuse and Adult Psychiatric and Substance Use Disorders in Women. Archives of General
Psychiatry, 57(10), 953. doi:10.1001/​archpsyc.57.10.953
6
2
3

326 Children of Abuse and Neglect

Kim, J., & Cicchetti, D. (2010). Longitudinal pathways linking child maltreatment, emotion regulation, peer
relations, and psychopathology. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 51(6),
706–​716. doi:10.1111/​j.1469-​7610.2009.02202.x
Kim-​Spoon, J., Cicchetti, D., & Rogosch, F. A. (2013). A longitudinal study of emotion regulation, emotion
lability-​negativity, and internalizing symptomatology in maltreated and nonmaltreated children. Child
Development, 84(2), 512–​527. doi:10.1111/​j.1467-​8624.2012.01857.x
Klevens, J., & Leeb, R. T. (2010). Child maltreatment fatalities in children under 5: Findings from the
National Violence Death Reporting System. Child Abuse & Neglect, 34(4), 262–​266. doi:10.1016/​
j.chiabu.2009.07.005
Kolla, N. J., Malcolm, C., Attard, S., Arenovich, T., Blackwood, N., & Hodgins, S. (2013). Childhood
maltreatment and aggressive behaviour in violent offenders with psychopathy. Canadian Journal of
Psychiatry. Revue Canadienne de Psychiatrie, 58(8), 487–​494. Retrieved from http://​europepmc.org/​
abstract/​med/​23972111
Krause, E. D., Mendelson, T., & Lynch, T. R. (2003). Childhood emotional invalidation and adult
psychological distress: the mediating role of emotional inhibition. Child Abuse & Neglect, 27(2), 199–​
213. doi:10.1016/​S0145-​2134(02)00536-​7/​
Krug, E. G., Mercy, J. A., Dahlberg, L. L., & Zwi, A. B. (2002). The world report on violence and health.
Lancet, 360(9339), 1083–​1088. doi:10.1016/​S0140-​6736(02)11133-​0
Kupersmidt, J. B., & Coie, J. D. (1990). Preadolescent Peer Status, Aggression, and School Adjustment as
Predictors of Externalizing Problems in Adolescence. Child Development, 61(5), 1350. doi:10.2307/​1130747
Ladd, G. W., & Troop-​Gordon, W. (2003). The Role of Chronic Peer Difficulties in the Development of
Children’s Psychological Adjustment Problems. Child Development, 74(5), 1344–​1367. doi:10.1111/​
1467-​8624.00611
Lang, S., af Klinteberg, B., & Alm, P.-​O. (2002). Adult psychopathy and violent behavior in males with early
neglect and abuse. Acta Psychiatrica Scandinavica, 106(s412), 93–​100. doi:10.1034/​j.1600-​0447.106.
s412.20.x
Lansford, J. E., Miller-​Johnson, S., Berlin, L. J., Dodge, K. A., Bates, J. E., & Pettit, G. S. (2007). Early
physical abuse and later violent delinquency: a prospective longitudinal study. Child Maltreatment,
12(3), 233–​245. doi:10.1177/​1077559507301841
Larrance, D. T., & Twentyman, C. T. (1983). Maternal attributions and child abuse. Journal of Abnormal
Psychology, 92(4), 449–​457. doi:10.1037//​0021-​843X.92.4.449
Louwers, E. C. F. M., Korfage, I. J., Affourtit, M. J., Scheewe, D. J. H., van de Merwe, M. H., Vooijs-​
Moulaert, A.-​F. S. R., … de Koning, H. J. (2012). Effects of systematic screening and detection of child
abuse in emergency departments. Pediatrics, 130(3), 457–​464. doi:10.1542/​peds.2011-​3527
Lunkenheimer, E. S., Shields, A. M., & Cortina, K. S. (2007). Parental Emotion Coaching and Dismissing
in Family Interaction. Social Development, 16(2), 232–​248. doi:10.1111/​j.1467-​9507.2007.00382.x
Macfie, J., Cicchetti, D., & Toth, S. L. (2001). Dissociation in maltreated versus nonmaltreated preschool-​
aged children. Child Abuse & Neglect, 25(9), 1253–​1267. doi:10.1016/​S0145-​2134(01)00266-​6
Maier, M. E., & di Pellegrino, G. (2012). Impaired conflict adaptation in an emotional task context
following rostral anterior cingulate cortex lesions in humans. Journal of Cognitive Neuroscience, 24(10),
2070–​2079. doi:10.1162/​jocn_​a_​00266
Manly, J. T., Cicchetti, D., & Barnett, D. (1994). The impact of subtype, frequency, chronicity, and severity
of child maltreatment on social competence and behavior problems. Development and Psychopathology,
6(1), 121–​143. doi:10.1017/​S0954579400005915
Marusak, H. A., Martin, K. R., Etkin, A., & Thomason, M. E. (2015). Childhood trauma exposure
disrupts the automatic regulation of emotional processing. Neuropsychopharmacology : Official
Publication of the American College of Neuropsychopharmacology, 40(5), 1250–​1258. doi:10.1038/​
npp.2014.311
Masten, C. L., Guyer, A. E., Hodgdon, H. B., McClure, E. B., Charney, D. S., Ernst, M., … Monk, C. S.
(2008). Recognition of facial emotions among maltreated children with high rates of post-​traumatic
stress disorder. Child Abuse & Neglect, 32(1), 139–​153. doi:10.1016/​j.chiabu.2007.09.006
7
2
3

Conclusion 327

Maszk, P., Eisenberg, N., & Guthrie, I. K. (1999). Relations of Children’s Social Status to Their Emotionality
and Regulation: A Short-​Term Longitudinal Study. Merrill-​Palmer Quarterly, 45(3), 468–​492. Retrieved
from http://​eric.ed.gov/​?id=EJ599896
Maughan, A., & Cicchetti, D. (2002). Impact of Child Maltreatment and Interadult Violence on Children’s
Emotion Regulation Abilities and Socioemotional Adjustment. Child Development, 73(5), 1525–​1542.
doi:10.1111/​1467-​8624.00488
McLoyd, V. C. (1990). The Impact of Economic Hardship on Black Families and Children: Psychological
Distress, Parenting, and Socioemotional Development. Child Development, 61(2), 311–​346.
doi:10.1111/​j.1467-​8624.1990.tb02781.x
Milner, J. S., & Robertson, K. R. (1990). Comparison of Physical Child Abusers, Intrafamilial Sexual
Child Abusers, and Child Neglecters. Journal of Interpersonal Violence, 5(1), 37–​48. doi:10.1177/​
088626090005001003
Molnar, B. E., Buka, S. L., & Kessler, R. C. (2001). Child sexual abuse and subsequent
psychopathology: results from the National Comorbidity Survey. American Journal of Public Health,
91(5), 753–​760. Retrieved from http://​www.pubmedcentral.nih.gov/​articlerender.fcgi?artid=1446666&t
ool=pmcentrez&rendertype=abstract
Morris, A. S., Silk, J. S., Morris, M. D. S., Steinberg, L., Aucoin, K. J., & Keyes, A. W. (2011). The
influence of mother-​child emotion regulation strategies on children’s expression of anger and sadness.
Developmental Psychology, 47(1), 213–​225. doi:10.1037/​a0021021
Morris, A. S., Silk, J. S., Steinberg, L., Myers, S. S., & Robinson, L. R. (2007). The Role of the Family
Context in the Development of Emotion Regulation. Social Development (Oxford, England), 16(2),
361–​388. doi:10.1111/​j.1467-​9507.2007.00389.x
Noll, J. G., Horowitz, L. A., Bonanno, G. A., Trickett, P. K., & Putnam, F. W. (2003). Revictimization and
self-​harm in females who experienced childhood sexual abuse: results from a prospective study. Journal
of Interpersonal Violence, 18(12), 1452–​1471. doi:10.1177/​0886260503258035
Ochsner, K. N., & Gross, J. J. (2005). The cognitive control of emotion. Trends in Cognitive Sciences, 9(5),
242–​249. doi:10.1016/​j.tics.2005.03.010
Oral, R., Blum, K. L., & Johnson, C. (2003). Fractures in young children: are physicians in the emergency
department and orthopedic clinics adequately screening for possible abuse? Pediatric Emergency Care,
19(3), 148–​153. doi:10.1097/​01.pec.0000081234.20228.33
Oshri, A., Sutton, T. E., Clay-​Warner, J., & Miller, J. D. (2015). Child maltreatment types and risk
behaviors: Associations with attachment style and emotion regulation dimensions. Personality and
Individual Differences, 73, 127–​133. doi:10.1016/​j.paid.2014.09.015
Palazzi, S., de Girolamo, G., & Liverani, T. (2005). Observational study of suspected maltreatment
in Italian paediatric emergency departments. Archives of Disease in Childhood, 90(4), 406–​410.
doi:10.1136/​adc.2003.040790
Pham, T. H., & Philippot, P. (2010). Decoding of facial expression of emotion in criminal psychopaths.
Journal of Personality Disorders, 24(4), 445–​459. doi:10.1521/​pedi.2010.24.4.445
Pollak, S. D. (2008). Mechanisms Linking Early Experience and the Emergence of Emotions: Illustrations
From the Study of Maltreated Children. Current Directions in Psychological Science, 17(6), 370–​375.
doi:10.1111/​j.1467-​8721.2008.00608.x
Pollak, S. D., Cicchetti, D., Hornung, K., & Reed, A. (2000). Recognizing emotion in faces: Developmental
effects of child abuse and neglect. Developmental Psychology, 36(5), 679–​688. doi:10.1037/​
0012-​1649.36.5.679
Pollak, S. D., Klorman, R., Thatcher, J. E., & Cicchetti, D. (2001). P3b reflects maltreated children’s reactions
to facial displays of emotion. Psychophysiology, 38(2), 267–​274. doi:10.1111/​1469-​8986.3820267
Pollak, S. D., & Sinha, P. (2002). Effects of early experience on children’s recognition of facial displays of
emotion. Developmental Psychology, 38(5), 784–​791. doi:10.1037//​0012-​1649.38.5.784
Pollak, S. D., Vardi, S., Putzer Bechner, A. M., & Curtin, J. J. (2005). Physically abused children’s
regulation of attention in response to hostility. Child Development, 76(5), 968–​977. doi:10.1111/​
j.1467-​8624.2005.00890.x
8
2
3

328 Children of Abuse and Neglect

Radford, L., Corral, S., Bradley, C., Fisher, H., Bassett, C., Howat, N., & Collishaw, S. (2011).
Child Abuse and Neglect in the UK Today. London: National Society for the Prevention of Cruelty
to Children. Retrieved from http://​clok.uclan.ac.uk/​6022/​1/​child_​abuse_​neglect_​research_​PDF_​
wdf84181.pdf
Rieder, C., & Cicchetti, D. (1989). Organizational Perspective on Cognitive Control Functioning and
Cognitive-​Affective Balance in Maltreated Children. Developmental Psychology, 25(3), 382–​393.
Retrieved from http://​eric.ed.gov/​?id=EJ392566
Robinson, L. R., Morris, A. S., Heller, S. S., Scheeringa, M. S., Boris, N. W., & Smyke, A. T. (2008).
Relations Between Emotion Regulation, Parenting, and Psychopathology in Young Maltreated
Children in Out of Home Care. Journal of Child and Family Studies, 18(4), 421–​434. doi:10.1007/​
s10826-​008-​9246-​6
Rogosch, F. A., Cicchetti, D., & Aber, J. L. (1995). The role of child maltreatment in early deviations in
cognitive and affective processing abilities and later peer relationship problems. Development and
Psychopathology, 7(4), 591–​609. doi:10.1017/​S0954579400006738
Rogosch, F. A., Cicchetti, D., Shields, A., & Toth, S. L. (1995). Parenting dysfunction in child maltreatment.
In M. H. Bornstein (Ed.), Handbook of parenting, Vol. 4: Applied and practical (pp. 127–​159). Hillsdale,
NJ: Lawrence Erlbaum Associates, Inc.
Ross, S. M. (1996). Risk of physical abuse to children of spouse abusing parents. Child Abuse & Neglect,
20(7), 589–​598. doi:10.1016/​0145-​2134(96)00046-​4
Runyan, D. K., Hunter, W. M., Socolar, R. R. S., Amaya-​Jackson, L., English, D., Landsverk, J., …
Mathew, R. M. (1998). Children Who Prosper in Unfavorable Environments: The Relationship
to Social Capital. Pediatrics, 101(1), 12–​18. Retrieved from http://​pediatrics.aappublications.org/​
content/​101/​1/​12.short
Salzinger, S., Feldman, R. S., Hammer, M., & Rosario, M. (1993). The Effects of Physical Abuse on
Children’s Social Relationships. Child Development, 64(1), 169–​187. doi:10.1111/​j.1467-​8624.1993.
tb02902.x
Schnitzer, P. G., & Ewigman, B. G. (2005). Child deaths resulting from inflicted injuries:
household risk factors and perpetrator characteristics. Pediatrics, 116(5), e687–​e693. doi:10.1542/​
peds.2005-​0296
Schwartz, D., & Proctor, L. J. (2000). Community violence exposure and children’s social adjustment in
the school peer group: The mediating roles of emotion regulation and social cognition. Journal of
Consulting and Clinical Psychology, 68(4), 670–​683. doi:10.1037/​0022-​006X.68.4.670
Seifert, D., Krohn, J., Larson, M., Lambe, A., Püschel, K., & Kurth, H. (2010). Violence against
children: further evidence suggesting a relationship between burns, scalds, and the additional injuries.
International Journal of Legal Medicine, 124(1), 49–​54. doi:10.1007/​s00414-​009-​0347-​6
Shackman, J. E., & Pollak, S. D. (2014). Impact of physical maltreatment on the regulation of negative
affect and aggression. Development and Psychopathology, 26(4 Pt 1), 1021–​1033. doi:10.1017/​
S0954579414000546
Shackman, J. E., Shackman, A. J., & Pollak, S. D. (2007). Physical abuse amplifies attention to threat
and increases anxiety in children. Emotion (Washington, D.C.), 7(4), 838–​852. doi:10.1037/​
1528-​3542.7.4.838
Shields, A., & Cicchetti, D. (1998). Reactive aggression among maltreated children: the contributions of
attention and emotion dysregulation. Journal of Clinical Child Psychology, 27(4), 381–​395. doi:10.1207/​
s15374424jccp2704_​2
Shields, A., & Cicchetti, D. (2001). Parental maltreatment and emotion dysregulation as risk factors for
bullying and victimization in middle childhood. Journal of Clinical Child Psychology, 30(3), 349–​363.
doi:10.1207/​S15374424JCCP3003_​7
Shields, A. M., Cicchetti, D., & Ryan, R. M. (1994). The development of emotional and behavioral
self-​regulation and social competence among maltreated school-​age children. Development and
Psychopathology, 6(1), 57–​75. doi:10.1017/​S0954579400005885
9
2
3

Conclusion 329

Shipman, K., Edwards, A., Brown, A., Swisher, L., & Jennings, E. (2005). Managing emotion in a
maltreating context: a pilot study examining child neglect. Child Abuse & Neglect, 29(9), 1015–​1029.
doi:10.1016/​j.chiabu.2005.01.006
Shipman, K. L., Schneider, R., Fitzgerald, M. M., Sims, C., Swisher, L., & Edwards, A. (2007). Maternal
Emotion Socialization in Maltreating and Non-​maltreating Families: Implications for Children’s
Emotion Regulation. Social Development, 16(2), 268–​285. doi:10.1111/​j.1467-​9507.2007.00384.x
Shipman, K. L., & Zeman, J. (2001). Socialization of children’s emotion regulation in mother-​child dyads: a
developmental psychopathology perspective. Development and Psychopathology, 13(2), 317–​336.
Retrieved from http://​europepmc.org/​abstract/​med/​11393649
Shipman, K., Zeman, J., Penza, S., & Champion, K. (2000). Emotion management skills in sexually
maltreated and nonmaltreated girls: a developmental psychopathology perspective. Development and
Psychopathology, 12(1), 47–​62. Retrieved from http://​europepmc.org/​abstract/​med/​10774595
Silverman, A. B., Reinherz, H. Z., & Giaconia, R. M. (1996). The long-​term sequelae of child and
adolescent abuse: A longitudinal community study. Child Abuse & Neglect, 20(8), 709–​723. doi:10.1016/​
0145-​2134(96)00059-​2
Sinha, R., Fox, H. C., Hong, K. A., Bergquist, K., Bhagwagar, Z., & Siedlarz, K. M. (2009). Enhanced
negative emotion and alcohol craving, and altered physiological responses following stress and cue
exposure in alcohol dependent individuals. Neuropsychopharmacology : Official Publication of the
American College of Neuropsychopharmacology, 34(5), 1198–​1208. doi:10.1038/​npp.2008.78
Slack, K. S., Berger, L. M., DuMont, K., Yang, M.-​Y., Kim, B., Ehrhard-​Dietzel, S., & Holl, J. L. (2011).
Risk and protective factors for child neglect during early childhood: A cross-​study comparison.
Children and Youth Services Review, 33(8), 1354–​1363. doi:10.1016/​j.childyouth.2011.04.024
Springer, K. W., Sheridan, J., Kuo, D., & Carnes, M. (2007). Long-​term physical and mental health
consequences of childhood physical abuse: Results from a large population-​based sample of men and
women. Child Abuse & Neglect, 31(5), 517–​530. doi:10.1016/​j.chiabu.2007.01.003
Stevens, N. R., Gerhart, J., Goldsmith, R. E., Heath, N. M., Chesney, S. A., & Hobfoll, S. E. (2013).
Emotion regulation difficulties, low social support, and interpersonal violence mediate the link between
childhood abuse and posttraumatic stress symptoms. Behavior Therapy, 44(1), 152–​161. doi:10.1016/​
j.beth.2012.09.003
Stuhrmann, A., Suslow, T., & Dannlowski, U. (2011). Facial emotion processing in major depression: a
systematic review of neuroimaging findings. Biology of Mood & Anxiety Disorders, 1(10). doi:10.1186/​
2045-​5380-​1-​10
Suveg, C., Morelen, D., Brewer, G. A., & Thomassin, K. (2010). The Emotion Dysregulation Model of
Anxiety: a preliminary path analytic examination. Journal of Anxiety Disorders, 24(8), 924–​930.
doi:10.1016/​j.janxdis.2010.06.018
Suveg, C., & Zeman, J. (2004). Emotion regulation in children with anxiety disorders. Journal of
Clinical Child and Adolescent Psychology : The Official Journal for the Society of Clinical Child and
Adolescent Psychology, American Psychological Association, Division 53, 33(4), 750–​759. doi:10.1207/​
s15374424jccp3304_​10
Tarullo, A. R., & Gunnar, M. R. (2006). Child maltreatment and the developing HPA axis. Hormones and
Behavior, 50(4), 632–​639. doi:10.1016/​j.yhbeh.2006.06.010
Ten Bensel, R. W., Rheinberger, M. M., & Radbill, S. X. (1997). Children in a world of violence: The roots
of child maltreatment. In R. E. Helfer, R. Kempe, & D. Krugman (Eds.), The battered child (pp. 3–​28).
Chicago: University of Chicago Press.
Theodore, A. D., & Runyan, D. K. (1999). A Medical Research Agenda for Child Maltreatment: Negotiating
the Next Steps. Pediatrics, 104(1), 168–​177. Retrieved from http://​pediatrics.aappublications.org/​
content/​104/​Supplement_​1/​168.short
Thompson, R. A. (1994). Emotion regulation: A theme in search of definition. Monographs of the Society for
Research in Child Development, 59(2-​3), 25–​52. doi:10.1111/​j.1540-​5834.1994.tb01276.x
0
3

330 Children of Abuse and Neglect

Thompson, R. A., & Calkins, S. D. (1996). The double-​edged sword: Emotional regulation for children at
risk. Development and Psychopathology, 8(1), 163–​182.
Thompson, R. A., & Meyer, S. (2007). Socialization of Emotion Regulation in the Family. In J. J. Gross
(Ed.), Handbook of Emotion Regulation (pp. 249–​268). New York: Guilford Press.
Thornberry, T. P., & Henry, K. L. (2013). Intergenerational continuity in maltreatment. Journal of Abnormal
Child Psychology, 41(4), 555–​569. doi:10.1007/​s10802-​012-​9697-​5
Tull, M. T., Barrett, H. M., McMillan, E. S., & Roemer, L. (2007). A preliminary investigation of the
relationship between emotion regulation difficulties and posttraumatic stress symptoms. Behavior
Therapy, 38(3), 303–​313. doi:10.1016/​j.beth.2006.10.001
Valiente, C., & Eisenberg, N. (2006). Parenting and Children’s Adjustment: The Role of Children’s Emotion
Regulation. In J. N. Snyder, Douglas K; Simpson, Jeffry; Hughes (Ed.), Emotion regulation in couples and
families: Pathways to dysfunction and health (pp. 123–​142). Washington, DC: American Psychological
Association.
Van der Kolk, B. A., & Fisler, R. E. (1994). Childhood abuse and neglect and loss of self-​regulation. Bulletin
of the Menninger Clinic, 58(2), 145–​168. Retrieved from http://​europepmc.org/​abstract/​med/​7519094
Vuilleumier, P., Armony, J. L., Driver, J., & Dolan, R. J. (2001). Effects of Attention and Emotion on Face
Processing in the Human Brain. Neuron, 30(3), 829–​841. doi:10.1016/​S0896-​6273(01)00328-​2
Weiler, B. L., & Widom, C. S. (1996). Psychopathy and violent behaviour in abused and neglected young
adults. Criminal Behaviour and Mental Health, 6(3), 253–​271. doi:10.1002/​cbm.99
Whipple, E. E., & Webster-​Stratton, C. (1991). The role of parental stress in physically abusive families.
Child Abuse & Neglect, 15(3), 279–​291. doi:10.1016/​0145-​2134(91)90072-​L
Widom, C. S., Marmorstein, N. R., & White, H. R. (2006). Childhood victimization and illicit drug use in
middle adulthood. Psychology of Addictive Behaviors : Journal of the Society of Psychologists in Addictive
Behaviors, 20(4), 394–​403. doi:10.1037/​0893-​164X.20.4.394
Williamson, J. M., Borduin, C. M., & Howe, B. A. (1991). The ecology of adolescent maltreatment:
A multilevel examination of adolescent physical abuse, sexual abuse, and neglect. Journal of Consulting
and Clinical Psychology, 59(3), 449–​457. doi:10.1037//​0022-​006X.59.3.449
Wilson, H. W., & Widom, C. S. (2011). Pathways from childhood abuse and neglect to HIV-​risk sexual
behavior in middle adulthood. Journal of Consulting and Clinical Psychology, 79(2), 236–​246.
doi:10.1037/​a0022915
Wupperman, P., Marlatt, G. A., Cunningham, A., Bowen, S., Berking, M., Mulvihill-​Rivera, N., & Easton,
C. (2012). Mindfulness and modification therapy for behavioral dysregulation: results from a pilot
study targeting alcohol use and aggression in women. Journal of Clinical Psychology, 68(1), 50–​66.
doi:10.1002/​jclp.20830
Zeanah, C. H., Scheeringa, M., Boris, N. W., Heller, S. S., Smyke, A. T., & Trapani, J. (2004). Reactive
attachment disorder in maltreated toddlers. Child Abuse & Neglect, 28(8), 877–​888. doi:10.1016/​
j.chiabu.2004.01.010
Zuravin, S. J., & DiBlasio, F. A. (1996). The correlates of child physical abuse and neglect by adolescent
mothers. Journal of Family Violence, 11(2), 149–​166. doi:10.1007/​BF02336667
Chapter 16

Children of Divorce
Maria Caridad H. Tarroja, Ma. Araceli Balajadia-╉Alcala,
& Maria Aurora Assumpta D. Catipon

Effects of divorce
Parental separation and divorce significantly impacts children’s development. Many books and
articles have been written on the experiences of children whose parents separated and how the
parental separation affected them throughout their development, from childhood into adulthood.
The majority of studies report that compared to children from intact families, children of divorce
are at a higher risk for behavior and adjustment problems which may manifest in different facets
of their lives (e.g., academic performance, self-╉esteem, social relations, expression of emotions,
and psychological wellbeing) (Alubokin & Akyina, 2015; Boring, Velez, Sandler, Tein, & Horan,
2015; Schick, 2002). The terms parental separation and parental divorce are often interchanged,
perhaps because the impact of this event on children’s and families’ lives are similar. Regardless
of the term used, the importance of harmonious relationships between the divorcing/╉separating
couples remains imparative (Lee & Bax, 2000).
Studies on children of divorce show inconsistent results. On one hand, some studies suggest
the after-╉effects of divorce are short-╉term and that adjustment improves over time. On the other
hand, some longitudinal studies show that long-╉term effects on the children are undeniable and
are emphasized to bring attention to the phenomenology of children of divorce so that other
steps to preserve their best interests can be pursued. Likewise, findings from the 25 year study of
Wallerstein and Lewis (2004) demonstrate that the lives of children of divorce are forever changed
or transformed. Thus, divorce is not simply an acute stressor from which a child bounces back
because challenges throughout their life course and the shifting relationships therein, continu-
ously demand a series of adjustments from the child, who needs to deal with the many losses
associated with separation and divorce. Hence, the challenge is to develop interventions that may
enable children of divorce to maintain hope, develop a positive view of relationships, and carry
this perception into succeeding relationships they may have.
This chapter describes children of divorce, its prevalence, experiences, behavior and adjust-
ment issues, and interventions that have been identified in the literature that are helpful given
their needs. In addition, the role of emotion regulation in the manifestation of their psychological
problems and as a protective factor in their adjustment process, is discussed.

Prevalence of divorce
The Social Trends Institute (2012) reported that generally high rates of marriage in Asia and the
Middle East and low rates of cohabitation suggest that marriage is important in the life path
of adults, especially in these regions. However, it has less of a dominant role in Africa, the
Americas, Europe, and Oceania, where cohabitation or non-╉marriage is a common practice. Thus,
2
3

332 Children of Divorce

termination of partnerships, or separation of parents, is not reported and included in marriage


and divorce statistics. Excluding separations that happen in the context of cohabitation or non-╉
marriage, plus the non-╉practice of divorce in some countries, a crude divorce rate (percentage for
every population of 1,000 individuals) of 1.9% is seen across the globe (UN Demographic Report,
2012), with Europe and the Americas registering the highest rates (Divorcescience.org, 2015).
While crude divorce rates in excess of 1% have been predominant in Western countries, divorce
became comparatively common in most countries in 1985. Considerable variations in divorce
trends are still noted within regions. Currently, the divorce rates in most developed countries
exceed 2%. By 1995, the divorce rates in East Asian countries became similar to other industrial-
ized countries. Divorce rates, on the other hand, were noted to decline in 1995 in New Zealand,
Australia, and other countries in Europe and North America, which further narrowed the divorce-╉
rate gap between countries across the world (Social Trends Institute, 2012).
It was noted that the observed peaks in England and Wales in the early 1980s, 1990s, 2000s and
the increase in 2010 before the slight drop in 2011, coincided with the economic recession. These
statistics may suggest a tendency for marriages to succumb under financial pressure (Sedghi &
Rodgers, 2014).
Although 50% of marriages in the US end in divorce, the NY Times (December 2014) reported
a decline over the last three decades after its peak in the 1970s and early 1980s. Much of the trend
may be attributed to changing gender roles, especially among women, and the fact that fewer
people are getting married. Some of the other many reasons for the drop in divorce include later
marriages, birth control and the rise of so-╉called love marriages (referring to marrying for love as
opposed to the idea of arranged marriages).
Perhaps one of the biggest upheavals that divorce creates is the break-╉up of the household.
When a couple divorces or separates, the process also involves dividing up their lives. Children
unfortunately, also find themselves included in this procedure, regardless of their readiness to
do so. Living alternatively and equally with both parents after a parental separation has become
more common in the past 20  years for children in the Western world (Melli & Brown, 2008;
Spruijt & Duindam, 2010). Through most of twentieth-╉century America, mothers were routinely
granted sole custody of their children. Research from the University of Wisconsin–╉Madison sug-
gests a dramatic shift away from mother-╉only custody in favor of shared custody as mother-╉only
custody agreements dropped from 80% in 1986 to 74% in 1994. An analysis of nearly 10,000
Wisconsin divorce cases showed that the trend toward joint custody gained momentum since
then (i.e., mother-╉only custody decreased to 42% in 2008, while shared custody increased from
8% to 45%). On the other hand, father-╉only custody remains low and largely unchanged (Cancian,
2014). As shall be discussed in various sections of this chapter, there are several factors to consider
in examining the adjustment and well-╉being of children of divorce, including parent factors, child
factors, and environmental context. In addition, the role of family processes and functioning is
also highlighted in this section.

Impact of parental separation on children of divorce


Children’s reactions to their parents’ divorce or separation depend on several factors including
individual characteristics (i.e., temperament, personality characteristics, support network etc.),
parent characteristics, family processes and functioning, and the environmental context. Some
of these reactions may include internalizing behaviors (e.g., blame and guilt, depression or sad-
ness) or externalizing behaviors (e.g., anger and hostility). The majority of the research on the
differences between children of divorce and children from intact families has established that
children of divorce are at a higher risk for adjustment problems than children from intact fami-
lies (Schick, 2002). This is significant, as long term effects may manifest in different stages of
3

Impact of parental separation on children of divorce 333

development, especially in adulthood. Schick (2002) also found that significant differences in
levels of social withdrawal, behavior problems and academic difficulties occur between children
coming from intact families and children whose parents had been separated for a maximum of
two-​and-​a-​half years.
However, it is important to note that recent literature has shown that children who enjoy a
joint-​custody arrangement are generally better adjusted than those from single-​custody homes
(Bauserman, 2002). Thus, when marriage cannot be salvaged joint custody may be a more viable
option, in addition to other critical factors, such as individual difference, the parents maintain-
ing a positive relationship with healthy, transparent communication and ensuring the child has
a sound support network to name a few. Amidst the negative outcomes in children of divorce, it
is worthwhile to reiterate that a few studies indicated psychological problems of these children
existed even before marital separation, and may not be attributable to divorce per se (Furstenberg
& Teitler, 1994). Other findings have suggested that the negative effects of parental separation and
divorce had only a short term impact on children’s well-​being that was likely to disappear over
time (Allison & Furstenberg, 1989; Ambert, 1984). Moreover, a 2000 APA overview of the psy-
chological literature on the outcomes of divorce on children showed that the differences between
children of divorce and children from intact families is smaller and less pronounced than often-
times indicated i.e., the majority of children fall within the normal range of standard measures
of adjustment (Amato, 1994). Hetherington and Kelly (2002) concluded that 80% of the children
from divorced homes become reasonably well adjusted and insist that, this event is not a form of
developmental predestination. In sum, these studies (Bauserman, 2002; Schick, 2002) invite us to
re-​appraise the individual experience of each child so that the long term impact of divorce, not
just on the individual child, but on society as well, can be attenuated.
Some scholars have examined specific effects of divorce on various intra and interpersonal vari-
ables, a specific area of interest has been academic achievement. Kim (2011) found that children
of divorce, specifically those within the first to third grade levels, showed significant delays in
their math test scores during and after the divorce of their parents. No such setback was seen in
this group in reading test scores when compared with children from intact families. However, a
more recent article has demonstrated that reading test scores and particularly children’s reading
comprehension scores are affected pre-​and post-​divorce. The problems appear to precede divorce
and even follow it. Specifically, it was seen that reading test scores are most affected in seven to 14
year old children, two to four years prior to the family disruption. For reading comprehension,
negative effects tended to persist and in fact, even intensify with the passage of time (Arkes, 2015).
Whereas, some studies posit that poor academic achievement directly results from the divorce
experience (Alubokin & Akyina, 2015), others qualify, that social class or socio-​economic status
plays an important role in the eventual academic outcome of children from divorced families
(Grätz, 2015), as has been pointed out in an earlier section of this chapter. Children belonging to
higher social classes may have more financial resources and social support to overcome academic
problems brought about by the parents’ separation.
Potter (2010) found that elementary school children who experience parental divorce imme-
diately begin performing worse academically than their peers from intact families. According to
Schick (2002), inconsistent academic performance of children of divorce may be brought about
by the consumption of energy required to cope with the loss and re-​adjust to their new situation
(Schick, 2002). Gruber (2004) also showed that children exposed to unilateral divorce are less
educated by adulthood.
Another area that can be affected by parents’ separation is social relationships. For example,
in a 25  year follow-​up study conducted on a white, middle-​class, divorced population from
California, USA, Wallerstein and Lewis (2004) maintained that divorce appears to affect the social
relationships of people at virtually every stage of development, from childhood to adolescence to
4
3

334 Children of Divorce

adulthood. Instead of these children overcoming the impact of divorce when they reach adult-
hood, this is the time when they may be impacted significantly as long-​hidden emotional prob-
lems stemming from parental divorce may suddenly emerge during young adult life (Wallerstein
& Blakeslee, 1989). According to Wallerstein and Blakeslee (1989) without clear inner images of
stable relationships they may struggle more than adults from intact families, to attain love, sexual
intimacy, and lasting commitment. Thus, 60% of the participants in the 25-​year landmark study
had trouble with social relationships. They tended to expect failure in their lives, have a lingering
fear of loss due to childhood fears of abandonment, and experienced strong fear of change. In
addition, they feared betrayal and being alone, which was associated with self-​destructive choices
in relationships.
Relationships between the child and their parents were not just affected in the short-​term,
but also in the long-​term, especially for fathers. De Graaf and Fokkema (2007) found that cus-
tody arrangements during divorce affect subsequent contact of fathers with their adult children.
Bouchard and Doucet (2010) examined the transition into parenthood of adults who were chil-
dren of divorce using the family systems theory. They looked into interaction and dynamics
among family members in explaining how far-​fetching the impact of divorce is on relationships
within the family. They determined that in their sample of couples who experienced parental
divorce or separation growing up, and were expecting their first child, the quality of relation-
ship with the father continued to affect their relationship as adults. However, more importantly,
it was found that support coming from the gender-​matched parent (women with their mothers,
and men with their fathers) predicted the level of positive adjustment of couples as they went
through this significant life transition. Moreover, it is significant to note that a more recent article
has emerged examining the effect of parental separation/​divorce on relationships of children of
divorce in later life. Fergusson, McLeod and Horwood (2014) presented findings from a 30 year
longitudinal study of a birth cohort of over 1000 children in New Zealand, in 1977, stating that
parental divorces/​separations alone was not sufficient to account for relationship outcomes of
children coming from these families. Instead, there are a host of different contextual factors that
contribute to this outcome, including parent factors (history of illicit drug use and criminality,
conflict and violence between couples), child abuse (sexual, physical maltreatment), and fam-
ily socioeconomic status at birth of the child. The presence of conduct problems in childhood
(from ages seven to nine) was also one of the covariates mentioned. This latter factor may be
related to challenges that children have in regulating their emotions which may lead to behavioral
difficulties.
Parents who divorce may have fewer resources to effectively maintain the stability of social rela-
tionships that their child sorely needs, as they themselves have trouble maintaining close ties with
their own families of origin. This is known as the “negative divorce effect” which states that adult
children who are divorced have even less frequent contact with their parents compared to their
never-​married siblings (Kalmijn, 2014). This puts divorced individuals at risk for disconnection
from extended relatives at a time when they may need the support the most.
Children’s psychological well-​being is another facet that may be affected by the parents’ separa-
tion. For example, several researches say that children of divorced parents generally score lower
on indicators or measures of well-​being than children from intact families (Amato, 1991; Clarke-​
Stewart, Vandell, McCartney, Owen, & Booth, 2000; Kurtz, 1994; Weaver & Schofield, 2014).
Thus, compared to children in intact, two-​parent families, children of divorce show more symp-
toms of psychological maladjustment (e.g., exhibiting more conduct problems such as aggressive,
impulsive, and antisocial behaviors and more problems in their relationships with their moth-
ers and fathers (Amato, 1994; Hetherington, Stanley-​Hagan, 1999), lower academic achievement
(Amato, 1994; Hethereington & Stanley-​Hagan,1999; Wallerstein, 1991), more social difficulties
5
3

Impact of parental separation on children of divorce 335

and a reduced self-​concept (Amato, 1994; Kurtz & Derevensky, 1993). Children may regress, dis-
play anxiety and depressive symptoms, appear more irritable, become demanding and noncom-
pliant, and experience problems in social relationships and academic performance. Not only in
the US has the negative effect of divorce been demonstrated, but in Europe as well. Dronkers
(1999) found that the presence of increased parental conflict during and after divorce was strongly
linked to a lower sense of well-​being in secondary school pupils. Given this link between children
experiencing parental divorce and problems facing psychological maladjustment in children,
divorce has been identified as a stressor among health professionals as being potentially respon-
sible for maladjusted neuropsychological responses and for a decline in children’s physical health
(Nunes-​Costa, Lamela, Figueiredo, 2009) as well as for the experience of pain symptomatology
(Lee, 2000).
From the period of 1996 to 2009, a birth cohort of approximately 35,000 Chinese adolescents
was tracked for changes in self-​esteem. A decline in self-​esteem rates were seen in that decade,
and one of the key factors seemingly contributing to this was the increasing divorce rate, which
led to decreased familial connections. Findings like these demonstrate that the impact of divorce
can be particularly troubling especially as self-​esteem is generally found to be stable over time in
the general populace (Trzesniewski, Donnellan, & Robins, 2003).
Marquardt (2006) further noted that children of divorce experience early pressure to create
their own moral systems, as they cannot fully endorse the rules of two different households. Such
resulting behaviors and symptoms may indicate emotion suppression, which refers to consciously
inhibiting the ongoing expression of emotion-​related behavior (Gross & John, 2003), as the com-
mon form of emotion regulation used by children of divorce. Moreover, a repressive style may have
adverse effects not only on subjective well-​being, but also on physical health (John & Gross, 2004).
In an interview with over 400 divorced adults over 20 years who grew up in divorced homes,
Brooks (2010) found two dominant traits in most children of divorce, i.e., fear of abandonment
and fear of not being good enough to be loved. While adult children of divorce are consciously
aware of these fears, the underlying processes and underpinning beliefs often operate on a sub-
conscious level and it may be difficult to link back to limiting behavioral patterns. Brooks’ study
captured a pattern of four destructive relationship behaviors that both male and female adult
children of divorce engage in: 1) Trying to constantly please one’s partner and suppressing one’s
own needs for fear of rejection or in other words, being a martyr to please one’s partner to main-
tain the relationship, 2) looking for a relationship, sometimes taking on the first person who gives
attention and not wanting to be alone, 3) fixing a partner—​falling for a person who needs help
and love to fix all their problems, and being convinced that love is the solution to everything, 4)
shutting down emotionally for fear of getting divorced and settling for superficial relationships
with no real commitment.
Simons (2009) noted that children of divorcing parents tend to be good actors as they put on
different masks to fit into their parents’ different worlds. This is further noted by Marquardt’s
(2006), who discusses the split existence of children of divorce. The children reported that they
felt like different people with each of their parents and that they believe their parents were polar
opposites (even when they were not). In addition, they felt they needed to keep more secrets from
their parents than other kids do, and that they did not want to resemble one of their parents too
much, as this may lead to alienation from the other parent.
Development of coping skills may be influenced by the stability or the lack of it in parents’
relationship. At best, research has consistently shown that children from divorced families exhibit
less stereotyped sex behavior, greater maturity, and greater independence (Emery & Coiro, 1995).
Such positive characteristics may reflect the children’s coping attempts, which in turn, suggests
active regulation of emotions that allow them to function or compensate for the weaknesses
6
3

336 Children of Divorce

brought to the family system. However, the fact that they become vulnerable in situations that
they are not entirely ready for, e.g., when children leave the parental home earlier and have sex
at an earlier age (Amato & Keith, 1991) may suggest that emotions are not regulated enough for
self-╉protective purposes (Marquardt, 2005).

Mediating factors
Researchers agree that interparental conflict, rather than divorce or residential placement, is
the single most critical determining factors in children’s adjustment post-╉divorce. For example,
Reiter, Hjörleifsson, Breidablik, and Meland (2013) found that the rising prevalence of divorce
does not take away the emotional impact of it on adolescents. Rates of anxiety, depression, and
psychosomatic complaints continue to be present; however, the damaging effects of divorce can be
seen with those children who lose parental contact. The children who succeed after divorce appar-
ently are those who have parents who can communicate and work together. This indicates good
psychological health of the parents, and augurs well for the parent-╉child relationship. In addition
to this, contact and support from nonparental adults, such as extended family or support from
other authority figures, was cited as an additional factor in the adjustment of children of divorce
(Wolchik et al, 1993). With well managed interparental conflict, plus the continuing supportive
presence of the parents and extended family who are able to listen and value the emotions of these
children, the children can better learn to regulate themselves (i.e., they may possibly be aware of
their feelings, but not be overwhelmed by them).
Although divorce occurs in many cultural contexts, there is some evidence that social classes
have an effect on the impact of divorce on children. For example, the negative effect of parental
separation is concentrated among children from lower class families, but is less evident in higher
income families (Grätz, 2015). The decrease in parents income, often caused by having to support
two households, directly affects children’s basic needs, e.g., proper nutrition, clothing, and school
choices and extra-╉curricular activities. Some research has shown that while there are long-╉term
negative consequences of divorce, the consequences are more closely linked to fewer opportuni-
ties arising from lower educational attainment, rather than divorce per se (APA, 2004). Thus, if
educational opportunities are maintained amidst the divorce, the better the guidance the children
receive. This in turn can mitigate the after-╉effects of the experience of divorce, and facilitate chil-
dren’s ability to regulate their emotions. As a result, children of divorce can adjust better after their
parents’ separation.
Clarke-╉Stewart, Vandell, McCartney, Owen, and Booth (2000) indicate that the children’s psy-
chological development was not affected by parental separation per se. Instead stability of the
mother was noted, as psychological development was related to the mothers’ income, education,
ethnicity, childrearing beliefs, depressive symptoms, and behavior. Results further suggested that
what is most important for children in the years following divorce is not family structure or mari-
tal status per se, but family process such as family relationships, interaction, and communication.
An APA (2004) review of divorce literature summarized the following key factors that con-
tribute to healthy adjustment post-╉divorce. These include a healthy family system that enables
appropriate parenting, access to the non-╉custodial parent, custody arrangements, and low inter-
parental conflict. Appropriate parenting refers to the provision of emotional support, monitoring
of children’s activities, authoritative discipline, and maintenance of age-╉appropriate expectations.
Joint legal custody (which allows access to both parents, and shared decision making between
parents) is often associated with more frequent father-╉child visits, regular child support pay-
ments, and more satisfied and better adjusted children. For example, children of divorced families
with fathers who assist with homework, provide emotional support, and listen to their children’s
7
3

EMOTION REGULATION FOR CHILDREN AT RISK 337

problems have more positive academic achievement and fewer behavioral problems. Likewise,
children who alternate time between the parent’s respective homes after a separation experience
fewer psychosomatic problems than those living mostly or only with one parent (Bergström,
Fransson, Modin, Berlin, Gustafsson, & Hjern, 2015).
Such factors mentioned in this section allow children to adjust and cope better with the impact
of divorce. The process underlying adjustment, which will later be seen as bolstering the coping
capability of children dealing with divorce, is discussed in the next section.

Emotion regulation
Emotion regulation is critical in children’s development, cognitively, affectively, socially, and psy-
chologically. It has been found to be associated with social competence, cognitive performance,
and the management of stress (Thompson & Calkins, 1996). Studies have shown that children
who are able to effectively regulate their emotions perform well in school, relate positively with
others, and cope adaptively with life stressors (Barish, 2012).
According to Gross (2002), emotion regulation “refers to the processes by which we influence
which emotions we have, when we have them, and how we experience and express them” (p. 282).
He further explained that emotion regulation may involve maintaining, increasing, or decreasing
negative or positive emotions, occurs on a spectrum from the conscious to subconscious, and
is neither inherently good nor bad. A recent study on emotion regulation of mothers of young
children concluded that emotion regulation is composed of interrelated domains, which include
physiological, cognitive, and temperament aspects (Deater-╉Deckard, Li, & Bell, 2015). Hence, it
is important to consider these elements in understanding emotion regulation processes when
analyzing children of separated parents.

Emotion regulation for children at risk


There are many studies that have investigated the role of emotion regulation in adjustment and
behaviors of children at risk. Thompson and Calkins (1996) indicated that emotion regulation
can both foster resiliency and vulnerability among children at risk (e.g., children living with a
parent who has depression, or who had experienced domestic violence or who is temperamentally
inhibited).
Thompson and Calkins (1996) described emotional regulation for children at risk enumerating
several characteristic features which include 1) both intrinsic and extrinsic influences, i.e., both
self-╉regulated and managed by others; 2) serves goals for individuals which may be conflicting
at times; 3) includes monitoring, evaluating, and modifying emotional experiences; 4) manages
emotion by maintaining, heightening, or inhibiting emotional arousal; and 5) manages the tem-
poral and intensity characteristics of emotions.
More recent conceptualizations of emotional regulation appear simpler but still capture its
complex nature. For example, Fosco and Grych (2012) conceptualized it “as a balance between
adequate control of excessive or inappropriate expressions of emotions (e.g., losing control when
angry) and the ability to express one’s emotional needs, rather than suppressing emotions” (p. 566).
For Barish (2012), emotion regulation means thinking constructively about how to express and
manage one’s feelings. For children in distress, this implies recognizing and expressing their feel-
ings without being overwhelmed. For example, it is okay for children to feel angry without hurting
others, to feel discouraged without giving up, and to be afraid without feeling immobilized.
While it is generally true that the development of emotion regulation strategies can help in
the development of healthy personality characteristics, this path may not always be the case for
8
3

338 Children of Divorce

children who are in difficult circumstances or conditions of risk (Thompson et al., 1996). The
complex processes involved in emotion regulation may be more pronounced when dealing with
individuals who are at risk for displaying behavior problems due to their adverse circumstances
including children of divorced or separated parents. They further noted that there may be no opti-
mal means of regulating emotion (i.e., a child at risk might use a strategy that can result in both
protection and vulnerability). Likewise, under adverse circumstances, children may have more
difficulty regulating their emotions, not because they lack the strategies but because they employ
coping strategies that are most appropriate given their situations.

Influential factors
Emotional regulation is indeed an important element that impacts children’s adjustment to their
parents’ separation. A number of factors influence the development of emotion regulation, such as
the child’s age, temperament, neurophysiology, cognitive development, and social context are said
to influence emotion regulation (Morris, Silk, Steiberg, Myers, & Robinson, 2007).
A child’s ability for emotion regulation comes with neurological development, which is related
to a child’s chronological age. Thus, school-╉age children who experience the divorce at younger
ages are generally more likely to have problems as they lack the scaffolds for grasping the experi-
ence. However, infants and toddlers have little comprehension that a divorce has occurred and
thus, do not directly react to this change in their family set-╉up. The risks for this age group are
decreased interaction with the custodial parent and loss of contact with the noncustodial parent,
who can fade entirely from their lives. The experience can leave a tenuous and unstable model of
a parent, and a poor map for understanding and coping with emotions which is made possible by
stable primary relationships.
Thus, the child may benefit from frequent, short visits with the noncustodial parent that do not
disrupt the stable daily routine and secure attachment to the custodial parent (Thompson, 1998).
Wallerstein and Kelly’s (1980) observations, in which Piagetian influences are most notable, fur-
ther illustrates this. According to their observation, three-╉to five-╉year-╉old children commonly are
bewildered by divorce. Their limited cognitive capacity prevents them from fully understanding
its meaning and implications, leading to poor emotion regulation that results in unusual fantasies,
the fear of abandonment, emotional neediness, and aggressive acting out. The improved under-
standing of children six-╉to eight years old allows for greater acceptance of divorce; thus, grief
replaces denial. Adolescents have the most complete and abstract conception of the reasons for
their parents’ divorce, which Wallerstein and Kelly (1980) assert, facilitates their adjustment. They
suggest that perhaps the most difficult cognitive task for adolescents is to integrate the divorce
experience with their developing self.
In his paper, Gross (2002) described how reappraisal and suppression, as emotion regulation
strategies, impact emotion experience and behavioral expression. While reappraisal decreases
emotion experience and behavior expression, suppression decreases behavioral expression only
and not emotion experience. The latter impairs memory while the former has little to no impact on
memory (Gross, 2002). Similarly, a recent study claimed that recovery experiences depend on how
people were able to utilize emotion regulation strategies (Schraub, Turgut, Clavairoly, & Sonntag,
2013). Importantly, such resulting behaviors and symptoms indicate emotion suppression, which
refers to consciously inhibiting the ongoing expression of emotion-╉related behavior, as the com-
mon form of emotion regulation used by children of divorce. Moreover, a repressive style may have
adverse effects not only on subjective well-╉being, but also on physical health (John & Gross, 2004).
Another factor that has been shown to impact wellbeing outcomes following divorce is gender,
such that divorced children living with the same-╉sex parent showed fewer effects than children
9
3

Influential factors 339

living with the opposite-​sex parent (Kelly, 1998). Children living in single mother families have
attitudes that are more favorable to women, whereas, children living with single father families
have attitudes less favorable to women’s rights and gender equality (Prokic & Dronkers, 2009). In
addition, children living in single-​parent families tend to have less trust in societal institutions
than children from two-​parent families, and tend to have higher level of civic participation in
some countries. Such a relationship between the gender of children and the gender of the custo-
dial parent seem to point to a unique socialization process, which may include extensive cognitive
appraisal of gender roles, which is a core component of emotion regulation.
One important factor that can influence children’s ability to regulate their emotions is their
family, regardless of their individual and situational context. A tripartite model of familiar influ-
ence on emotional regulation has been described by Morris et al. (2007). In this model, emo-
tional regulation is developed through observational learning, modelling, and social referencing.
The second aspect of the model talks about how parenting practices pertaining to expression
and management of emotion can impact children’s emotion regulation. Thirdly, the emotional
climate of the family and its correlates such as parenting style, attachment relationship, marital
relationship and family expressiveness have been noted to impact emotion regulation. These
factors are important for all types of families. It is certainly significant to look into how these
different family variables and emotion regulation are affected with the change in family structure
following the separation of their parents. Following the model presented by Morris et al. (2007),
it is important to look at how each one can impact the emotion regulation of children of divorce
in particular.
Prior to the separation of their parents, children of divorce may have witnessed interparental
conflict or experienced domestic violence (Davies & Cummings, 1995). In the same research of
Davies and Cummings (1995), these stressors may trigger the development of problems in emo-
tion regulation, as they have observed them in their parents. They further noted that interparental
conflict may create negative affect for children, which may make them wish that their parents
separate to end the conflict both as a way of managing their parents’ relationship and of regulating
their own emotions.
Numerous studies have looked into the parental roles in children’s emotional regulation. In
a seminal review, Frankel, et  al., (2012) enumerated different ways that parents can help their
children regulate their emotions, namely modeling emotion, responding to children’s emotional
expression, teaching emotion regulation strategies, and motivating through rewards and punish-
ments. Another study highlighted the role of maternal emotional regulation on parenting behav-
ior and how the latter can influence children’s emotional regulation (Morelen, Shaffer, & Suveg,
2014). Taken together, the findings from these studies suggest the critical role of parenting prac-
tices and behaviors in emotion regulation which may mitigate the impact of the change in family
structure brought on due to the parents’ separation.
A variety of studies have established the important role of family functioning and process in
the development of emotion regulation among children, including those who are in vulnerable
situations. It is reported that their reactions may not necessarily be due to their parents’ separation
or divorce but as a result of reduced social support from the father and the children’s perception
of interparental conflict (Schick, 2002). However, this does not undermine the impact of parental
influences on emotion regulation, it is also essential to look into other parent and child factors
that can explain the differences in how children develop emotion regulation.
Relatedly, Fosco and Grych (2012) described the family context of emotion regulation and
explained how the family systems approach can explain the emotion regulation of children.
Rather than looking into the impact of parents on their children’s development and adjustment,
some examined inter-​parental relationships and family functioning and emotional regulation.
0
4
3

340 Children of Divorce

According to Clarke-╉Stewart et al. (2000), more than parental separation, children’s psychological
development is influenced by mother-╉related characteristics such as income, education, ethnicity,
childrearing beliefs, depressive symptoms, and behavior. On the other hand, Fosco and Grych
(2012) found important family variables such as, family climate, maternal warmth, sensitivity, and
interparental conflict as significant predictors of emotional regulation of children. It is important
to note, however, that the relationship between interparental conflict and internalizing problems
can be mediated by children’s perceived threat, i.e., negative parental conflict resolution styles may
have greater impact on internalizing problems because of the children’s perceived threat of their
current situation.
Hence, it is recommended that family processes and functioning are included in order to
have a more complete understanding of emotion regulation, especially emotional regulation of
children (Fosco & Grych, 2012). These family processes and how they impact emotion regula-
tion may help explain the behaviors and well-╉being of children whose parents are separated or
divorced.

Emotion regulation as a protective adjustment


When parents separate, children react and adjust differently. While parental separation generally
puts children more at risk or vulnerable to adjustment problems, studies have revealed that path-
ways to adjustment vary. Some of the key factors that appear to be different between children of
divorce and non-╉divorced parents are social anxiety and academic achievement (Schick, 2002).
Social anxiety is often related to shame and shyness which may be due to the children’s negative
perception of divorce and separation and these results highlighted the role of emotional factors in
understanding the behaviors of children of divorced parents in comparison with their peers from
intact families (Schick, 2002).
There are many studies that have investigated the role of emotion regulation in adjustment
behaviors of children at risk (Maughan & Cicchetti, 2002; Siffert & Schwarz, 2011; Thomson &
Calkins, 1996). Emotion regulation can be an important factor in the adjustment process of chil-
dren when their parents separate. Thomson and Calkins (1996) claimed that emotion regulation
can both foster resiliency and vulnerability among children at risk (e.g., children living with a par-
ent with depression, had experienced domestic violence or are temperamentally inhibited). Siffert
and Schwarz (2011) further validated the role of emotion regulation as a mediator between inter-
parental conflict and children’s internalizing and externalizing behaviors. Relatedly, Maughan and
Cicchetti (2002) determined, emotion regulation mediates the relationship between maltreatment
and children’s anxious/╉depressed symptoms.
Gross (2002) identified five points in his process model of emotion regulation, namely situ-
ation selection, situation modification, attentional deployment, cognitive change, and response
modulation. In this model, emotional regulation can be antecedent-╉focused (any of the first four
points) or response-╉focused (behavior change). Many studies examined reappraisal (cognitive
change) and suppression (response modulation) as emotion regulation strategies. According to
Gross (2002), reappraisal decreases negative emotional experience and increases positive emo-
tion experience. An example of reappraisal is when a child focuses on the positive aspects of their
parents separation, for example, less intrapersonal conflict.
On the other hand, suppression decreases positive emotion experience but has no impact on
negative emotional experience. Given the emotional experiences of children of divorce, those who
can reappraise their family situations may be able to handle the negative emotions brought about
by the parental separation more adaptively, than those who suppress their negative thoughts and
emotions.
1
4
3

INTERVENTIONS 341

Protective and risk factors and children’s adjustment


to parental separation
Aside from emotion regulation, there are other protective and risk factors that can impact the
psychological well-╉being and adjustment a child has to their parental separation. These factors
include personality traits and may be directly or indirectly related to emotion regulation. They
may be either minimized or exacerbated by emotion regulation. Aside from identifying individual
factors such as personality traits and demographic variables, it may be important to consider
familial and social factors.
There are certain personality traits that are found to moderate the impact of separation on
children’s adjustment and behaviors. For example, optimism was found to moderate the impact of
parental separation on the academic achievement of adolescents (Tetzner & Becker, 2015). In gen-
eral, optimism is directly related to academic achievement and the self-╉esteem of adolescents such
that the higher the level of optimism, the greater the achievement in school and the higher the
self-╉esteem. However, it is important to note that for children whose parents separated, optimism
does not diminish the impact of separation on self-╉esteem. In relation to resilience, social compe-
tence (Kliewer & Sandler, 1993) and absence of depression (Weaver & Schofield, 2014) were found
to be positively related with adaptive coping among children of divorce.
Other protective factors include parental and peer relationships (Landstedt, Hammarström, &
Winefield, 2015), and parental sensitivity (Weaver & Schofield, 2014). In the Northern Swedish
Cohort Study of Landstedt, Hammarström, and Winefield, (2015), the quality of parental and
peer relationships predicted mental (internalizing behaviors) and somatic (physical symptoms)
health even 26 years later. In relation to parental sensitivity the study of Weaver and Scoffed (2014)
was more direct in their findings stating that predivorce maternal sensitivity can serve as a pro-
tective factor of children of divorce such that children of divorce are likely to display behavior
problems when the mothers are less sensitive and more depressed.

Interventions
This section discusses the various intervention programs geared towards helping children of
divorce and their families. While not necessarily targeting emotion regulation, key aspects of
these programs nevertheless target important areas of children’s functioning that are associated
with cultivating adaptive emotion regulation. Broadly speaking, skills for approaching, diffusing,
or temporarily avoiding interparental conflict may point to adequate situation selection, while
components of attentional deployment and cognitive change may occur in skill building sessions
where children are taught to reappraise their situation, understand their divorce related thoughts
and feelings better, and find better ways of solving problems related to these that are developmen-
tally appropriate. Findings in the literature have demonstrated the impact of divorce on children,
thus, interventions that are geared towards targeting crucial areas of potential disruption and
reverse negative lifelong consequences are paramount. As emotion regulation involves multiple
components, interventions that are dynamic and multi-╉factorial are required. As the succeed-
ing sections will show, most programs focus on children’s emotions, ensuring they are given the
opportunity to express what they are experiencing, untangling the confusion brought about by
this experience in order to help them derive support from others who are going through the
same experience. However, there are also components of cognitive processing of the experience,
which places the experience in perspective and provides opportunities for correcting maladaptive
thought patterns and beliefs. Finally, skill building components are also embedded into many
interventions, which target things such as coping, social competence, and stress management.
2
4
3

342 Children of Divorce

Court-╉connected programs
Court-╉connected programs are spread throughout the United States (Pollet, 2009). According
to Pollet (2009), these programs intend to reduce the experience of emotional pain in children
whilst targeting better outcomes via a solution focused approach. Ten states in the US have court-╉
ordered or legislated parent-╉education programs for divorcing couples. On the other hand, 35
states have no such requirements. Basically, there are two streams of divorce programs: Those
that are offered by the state governments and those that are offered in the community, such as
through churches and schools. For example, in Alabama, a court judge may require attendance
of a child of divorcing parents to a four-╉hour program called “Families in transition” (for ages
six to 16). For this program there is a cost for attendance per child and there is also a parent
program that occurs simultaneously. Nevertheless, divorcing parties also have the option to file
a motion to waive their attendance to these sessions (p. 532 in Pollet, 2009). However, as previ-
ous literature and studies have shown, it may be in the best interest of the child for health care
providers and lawmakers to ensure the high rate of attendance in parenting sessions during and
after divorce.

School-╉based programs
A second category of interventions is those that are based in schools. Perhaps one of the most
studied of these is the Children of Divorce Intervention Program (CODIP), developed by
Pedro-╉Carrol and Cowen (1985) in the US, and now being run in various countries apart
from the US, such as Canada, New Zealand, and Australia. This program was designed with
five goals, which serve to lessen the impact of divorce, to provide a supportive environment
for children of divorce, to process divorce-╉related feelings, to clarify and promote the under-
standing of divorce-╉related concepts and misconceptions, to impart skills for problem-╉solv-
ing, and to foster positive self and family perceptions (Alpert-╉Gillis, Pedro-╉Carroll, & Cowen,
1989). Looking at the components of the program, crucial elements of emotion regulation are
addressed by its goals. For example, clarifying and promoting the understanding of divorce-╉
related concepts and potentially correcting misconceptions children may have surrounding
this can be seen to target the cognitive change component of emotion regulation. The CODIP
has been viewed as providing a multitude of advantages for children such as increasing their
healthy adjustment and reducing internalizing (anxiety, somatic symptoms) and externaliz-
ing (behavior) problems (Pedro-╉Carroll, 2005). In more recent years, the program has been
adapted for different age groups and countries. For instance, a feasibility study of the program
was undertaken for Dutch children in the kindergarten and first grade, with generally positive
results (Velderman, Cloostermans, & Pannebakker, 2014). Relatedly, the program was recently
piloted in two South African Schools for ten-╉to 14-╉year old boys, using an experimental
design; improvements in socio-╉emotional and behavioral functioning were reported (Botha
& Wild, 2013).
Another popular school-╉based intervention program provides support while also teaching chil-
dren skills for coping. Stolberg and Mahler (1994) conducted a study of 103 students in the third
to fifth grades, who had separated or divorced parents. They were assigned to one of three treat-
ment conditions (i.e., support; support and skill building; or support, skill building, transfer and
parent training procedures) with a no-╉treatment group serving as control. Those who entered the
support-╉alone condition experienced the most benefits, although the ones in the skill-╉building
conditions experienced long-╉lasting improvements in affect. Effective coping skills have been seen
in more recent literature to predict better emotion regulation skills in children going through
stressful events (Zalewski, Lengua, Wilson, Trancik, & Bazinet, 2011).
3
4

Programs targeting parenting skills 343

Community-╉based programs
A third category of interventions is community-╉based programs. One program that has received
empirical support is the Kids’ Turn program, which was evaluated by Cookston and Fung (2011).
This is a supportive, skills based program, the goals of which included providing emotional
support for children and their families coming from a variety of cultures in processing nega-
tive emotions, teaching skills for preventing children’s at-╉risk behaviors, and “demystifying” and
“de-╉stigmatizing” the process of separation. While the study by Cookston and Fung (2011) on
61 parents yielded no changes in parenting behaviors, improvements were reported in terms of
interparental conflict, topics that parents fight about, parental alienation behaviors, anxiety and
behavior, as well as children’s internalizing behaviors.
In light of what the literature has revealed from this chapter on the benefits of ongoing positive
parental relationships and child adjustment, this program may yield much benefit in offsetting the
long-╉term negative consequences of divorce on children.
Other interventions have targeted specific areas of thinking and feeling which may contribute
to emotion regulation in the long term. One study used expressive art in the form of music, to
successfully target children’s irrational beliefs about the divorce, which are thought to be related to
depressive symptoms (DeLucia-╉Waack & Gellman, 2007). Another study (Rossiter, 1988) targeted
pre-╉schoolers’ capacity to express their pain at their parents’ separation (which they may not have
the cognitive wherewithal to comprehend yet), and increased their adjustment to the separation
or divorce. A  component of parent feedback (telling them what their children need) was also
seen as germane to the program’s implementation. There are also significant relationships among
pretend play, creativity, and emotion regulation among children (Hoffman & Russ, 2012), indicat-
ing the potential use of play and storytelling in teaching children emotion regulation strategies
indirectly.

Programs targeting parenting skills


While most programs target the interventions to the children themselves, there have also been a
number of programs that focus on the parents, in as much as they are crucial in improving the
child’s adjustment after divorce.
A well-╉researched community based program that specifically targets parenting behaviors is
the New Beginnings Program (NBP) which is conducted largely in Arizona, USA. This program
was developed based on the theory that certain risk and protective factors may contribute to the
development or prevention of children’s mental health and social adjustment problems. These
factors include the quality of the parent-╉child relationship, effective discipline, father-╉child con-
tact, parental conflict, children’s coping, and their assessment of their stressors (Sandler, Wolchik,
Davis, Haine & Ayers, 2003 in Zhou, Sandler, Millsap, Wolchik, & Dawson-╉McClure, 2008), many
of which also directly affect how well children regulate their emotions and behaviors. The ini-
tial program specifically targeted mothers only, and looked at resilience developed in children
as a result of their mothers’ participation in this preventive parenting program as early findings
determined high maternal demoralization and low self-╉regulatory skills in children were seen to
contribute to poor program gains (Hike, Wolchik, Sandler, & Braver, 2002). The program was seen
to be effective as well in reducing problems in adolescents (e.g., substance use and risky sexual
behaviors) six years after program participation, especially for those children who were at high
risk for maladjustment at the outset of the program participation (baseline). Specifically, the qual-
ity of the mother-╉child relationships appeared to have a mediational effect on subsequent exter-
nalizing and internalizing behavior problems of adolescents (Velez, Wolchik, Tein, & Sandler,
2011; Zhou, Sandler, Millsap, Wolchik, & Dawson-╉McClure, 2008). In other words, those children
43

344 Children of Divorce

who are most at-╉risk for behavioral and emotional problems seem to benefit the most from these
programs. Importantly, the program has been shown to have a transformative affect, as it has been
linked to the intergenerational transmission of parenting behaviors (Mahrer, Winslow, Wolchik,
Tein, & Sandler, 2014). The longitudinal study showed that children whose parents participated
in the NBP program demonstrated a higher degree of warm parenting attitudes and lower levels
of harsh discipline attitudes when they themselves became parents. This seems to show that inter-
vention geared towards improving separated or divorced parents’ behaviors toward their children,
had long-╉lasting positive effects in terms of the children’s own parenting behaviors again, stressing
the crucial factor of adequate parenting in this time of crisis and change in the child’s life.
Additional programs have been cited in literature which target positive parenting skills to
improve child and adolescent behavior outcomes (Basson, 2013; Stallman & Sanders, 2014).
Parenting through Change is one example, which is associated with positive outcomes in the
literature. This program is designed to teach positive discipline and problem solving skills in par-
ents (Forgatch, 1994). A large randomized study of 238 divorcing mothers tested the efficacy of
this program. It determined effective parenting practices and teacher-╉reported school adjustment
in the treatment group although no direct effects on child outcomes were observed (Forgatch &
DeGarmo, 1999).

Online programs or website-╉based programs


In keeping with this era of the Internet, a final category of intervention programs for children
of divorce falls into the category of on-╉line delivery learning. A number of online programs are
available which are directed at improving children’s coping skills as well as helping parents going
through a divorce or separation in the service of improving parent-╉child relationships. One exam-
ple is the Children of Divorce-╉Coping with Divorce (Boring, Velez, Sandler, Tein, & Horan, 2015)
an online coping skills program designed to help children and adolescents deal with the divorce
or separation of their parents. 147 children and adolescents ranging in ages from 11 to 16 years
old took part in the study. The program was shown to be most effective for youth at the highest
risk for emotional problems.
A second program is the sandcastles program, which was designed for children ages six to
17 years old. This program recognizes that children go through a variety of negative emotions
in the aftermath of their parents’ separation or divorce; it has a three and a half hour—╉one time
format, targeting self-╉expression, improvement of self-╉concept, and better understanding of the
divorce overall. Similar to most of the other programs mentioned, program one offers a support-
ive atmosphere where children feel free to express themselves and gain knowledge that they are
not alone. Relatedly, Kids First Center and Kids First offer support for children and families going
through divorce, mitigating factors by offering a supportive atmosphere. Children of Divorce
Coping with Divorce also has a similar focus to the other programs advertized on websites, geared
for children. That is, they also promise to help normalize the divorce experience, help children
deal with and manage difficult emotions that go with this experience, and develop tools for cop-
ing, which appear to be skills that can help in emotion regulation (e.g., cognitive restructuring,
using relaxation, cognitive distraction).
In conclusion, studies on the impact of divorce and separation generally show that children go
through emotional stress that may make them at risk for behavior and adjustment problems in
the short-╉term following the divorce, while adjustment improves over time, with the majority of
children falling within the normal range. In addition, research has shown mediating and protec-
tive factors that can help children of divorce adjust and cope better with the new demands brought
about by their parents’ separation. Studies that compare children of divorce and those coming
5
4
3

Programs targeting parenting skills 345

from intact families show that significant differences between the two groups are not the result
of parental separation or divorce alone but also by other factors, such as interparental conflict,
family and environmental factors. It is important to look at family factors in understanding emo-
tion regulation of children, especially those who are considered to be at risk for adjustment and
behavior problems as a result of parental separation.
A crucial protective factor is the cultivation of adaptive emotion regulation competencies. The
literature on different factors influencing outcomes of divorce do seem to point to ER as the inter-
nal process underlying the capacity of a child to adjust to divorce and rise from it. In fact, the
majority of interventions mentioned in this chapter, while not directly referring to the term emo-
tion regulation in their targeted goals, nevertheless focus on common and crucial elements of this
construct. Moreover, it is well worth mentioning that emotion regulation is not only something
that the child learns from these programs, but from parents who undergo skills training in those
programs that offer it.
Family functioning and other family process variables more than family structure may also
mediate the impact of separation on adjustment and behaviors of children of divorce. Again here,
we stress the importance of lessened inter-​parental conflict, increase in contact with both parents,
especially with fathers, and in strengthening the ties even with extended relatives. This kind of
emotional support provides a “safety net” by which the child can come to rely in a time of poten-
tial upheaval in their lives. This safety net also provides the protective structure needed by the
child to develop strategies (which all seem to address emotion regulation) in dealing with the
challenges.
Thus, while there is enough evidence in terms of the effectiveness of certain programs to help
children of divorce adjust to their situations and cope with problems (Velez, et al, 2011), there may
be a need to understand the emotion regulation process in interventions that tap family systems,
in particular interparental conflicts.
“It takes a village to raise a child” they say. But what happens when this formerly tight village
(which the intact family and the extended networks that it weaves together represents) that the
child has come to rely on for security and comfort disintegrates? As this chapter has demon-
strated, the impact on the child’s life can be substantial, as their internal frame of reference is
greatly challenged with the onset of divorce. Interventions that help them adjust to their new
circumstance and help them regain some sense of control over their own lives are thus crucial.
Much has been said and written about the phenomenon of divorce and its effects on the major
players, particularly key family members. Likewise, sufficient attention and effort has been given
to children who are by products of these separation and divorce experiences. With all that has
been said and done, what else is left to be done? As has been shown by some well-​researched
intervention programs, incorporating ongoing research into the practice of mental health pro-
fessionals working with children and families of divorce generates much valuable and relevant
information that increases our knowledge of this complex issue as well as the most effective ways
to deal with it.
In terms of practice-​based research, other countries where divorce or separation occurs fre-
quently may do well to test the efficacy of the well-​researched programs that help alleviate the
potentially destructive aftermath of divorce on children and their families. For instance, we have
not been able to find evidence of such programs in other Asian and most European countries
where divorce or separation is socially sanctioned. The effects of divorce or separation in countries
where they are not yet approved but where the instance of annulment is on the rise (as in the case
of the Philippines), may benefit from such research on the impact and efficacy of intervention
programs as well.
6
4
3

346 Children of Divorce

Practitioners working with children and their families need to be able to approach the issue
of divorce with high levels of compassion and cognitive know-╉how regarding recent evidence in
this field. For example, it is recommended that focusing on the role of emotion regulation in the
adjustment of children of divorce or of separated parents be explored further. As highlighted by
Deater-╉Deckard et al. (2015), emotion regulation is best understood when all aspects are consid-
ered, namely the physiological, cognitive, and temperament domains.
Furthermore, in understanding emotion regulation of children of divorce, it is important to
involve the children themselves in research, to get their perspective not only the perspective of
their parents or the adults around them. One such research strategy is the use of Q methodol-
ogy which has been found to be useful in drawing out children’s experiences and emotions that
are going through difficult circumstances such as their parents’ divorce. (Ellingsen, Thorsen, &
Størksen, 2014). This methodology was undertaken as a form of participatory research with chil-
dren. The Q methodology which was developed by William Stephenson (in Ellingsen et al, 2014),
looks at emerging patterns of people’s feelings, beliefs about a certain topic, and determines the
degree of their agreement or disagreement with an issue or a point of view. Through the use of
subjective statements and visual images, Q methodology may enable researchers to elicit sensitive
feelings and issues from children participants in research. In the conduct of research on children
of divorce, it is important that all perspectives are taken into account, parents, peers, teachers and
the children of divorce themselves. Depending on their age and developmental stage, more appro-
priate methodology can be used aside from the usual survey and interview.

References
Alpert-╉Gillis, L., Pedro-╉Carroll, J., & Cowen, E. (1989). The children of divorce intervention
program: Development, interpretation, and evaluation of a program for young urban children. Journal
of Consulting and Clinical Psychology, 57, 583–╉589.
Allison, P., Furstenberg, F. (1989). How marital dissolution affects children: variations by age and sex.
Developmental Psychology, 25, 540–╉549.
Alubokin, B. & Akyina, K. O. (2015). Effects of divorce on the academic performance of some selected
public senior high school students in the Bolgatanga municipality of Ghana. International Journal of
Multidisciplinary Research and Development, 2, 375–╉381.
Amato, P. R. (2000). The consequences of divorce for adults and children. Journal of Marriage and the
Family, 62, 1269–╉1287.
Amato, P. R. (1994). Life-╉span adjustment of children to their parents’ divorce. Future Child, 4, 143–╉164.
Amato, P. R., Keith, B (1991). Parental Divorce and Adult Well-╉Being: A Meta-╉Analysis. Journal of Marriage
and Family, 53, 43–╉58.
Ambert, A. M. (1989). Ex-╉spouses and new spouses: A study of relationships. Greenwich, CT: JAI Press.
American Psychological Association (2004). An Overview of the Psychological Literature on the Effects of
Divorce on Children (May 2004). Retrieved from http://╉www.apa.org/╉about/╉gr/╉issues/╉cyf/╉divorce.aspx
Arkes, J. (2015) (abs). The temporal effects of divorces and separations on children’s academic achievement
and problem behavior. Journal of Divorce and Remarriage, 56, 25–╉42.
Barish, K. (2012). Emotions in Child Psychotherapy: An Integrative Framework. Oxford: Oxford
University Press.
Basson, W. (2013). Helping Divorced Parents to Benefit Adolescent Children: A Prospective Enrichment
Programme. Journal of Psychology in Africa, 23, 675–╉678.
Bauserman, R. (2002). (abs). Child adjustment in joint-╉custody versus sole-╉custody arrangements: A meta-╉
analytic review. Journal of Family Psychology, 16, 91–╉102.
Bergström, M., Fransson, E., Modin, B., Berlin, M., Gustafsson, P. & Hjern, A. (2015). Fifty moves a
year: is there an association between joint physical custody and psychosomatic problems in children?
Journal of Epidemiology and Community Health, 69, 769-╉774. doi:10.1136/╉jech-╉2014-╉205058
7
4
3

Programs targeting parenting skills 347

Boring, J., Velez, C, Sandler, I., Tein, J., & Horan, J. (2015). Children of Divorce-​Coping with
Divorce: A randomized control trial of an online prevention program for Youth experiencing parental
divorce. Journal of Consulting and Clinical Psychology, 83, 999–​1005.
Botha, C. & Wild, L. (2013). Evaluation of a school-​based intervention programme for South African
children of divorce. Journal of Child and Adolescent Mental Health, 25, 81–​91.
Bouchard, D. & Doucet, D. (2010). Parental Divorce and Couples’ Adjustment During the Transition to
Parenthood: The Role of Parent–​Adult Child Relationships. Journal of Family Issues, 20, 1–​21
Clarke-​Stewart, K., Vandell, D., McCartney, K., Owen, M., & Booth, C. (2000). Effects of parental
separation and divorce on very young children. Journal of Family Psychology, 14, 304–​326.
Cookston, J. & Fung, W. (2011). The Kids’ Turn Program Evaluation: Probing change within a community-​
based intervention for separating families. Family Court Review, 49, 348–​363.
Cummings, E. M. & Davies, P. T. (1994). Children and marital conflict: The impact of family dispute and
resolution. Journal of Child Psychology and Psychiatry, 35, 73–​122.
Davies, P. T., & Cummings, E. M. (1994). Marital conflict and child adjustment: An emotional security
hypothesis. Psychological Bulletin, 116, 387–​411.
De Graaf, P. M. & Fokkema, T. (2007). Contacts between divorced and non-​divorced parents and their adult
children in the Netherlands: An investment perspective. European Sociological Review, 23, 263–​277.
Deater-​Deckard, K., Li, M., & Bell, M. A. (2015). Multifaceted emotion regulation, stress and
affect of mothers of young children. Cognition and Emotion, 30(3), 444–​457. doi:10.1080/​
02699931.2015.1013087
DeLucia-​Waack, J., & Gellman, R. (2007). The efficacy of using music in children of divorce groups: Impact
on anxiety, depression and irrational beliefs about divorce. Group Dynamics: Theory, Research and
Practice, 11, 272–​282.
Divorcescience.org (2015). Updated World Divorce Statistics, 2012, 70 countries reporting. Retrieved
October 2015 from https://​divorcescience.files.wordpress.com/​2015/​01/​update-​world-​divorce-​rates-​
2012-​70-​countries.jpg
Dronkers, J. (1999). The effects of parental conflicts and divorce on the well-​being of pupils in Dutch
secondary education. European Sociological Review, 15, 195–​212.
Emery, R. (1999). Marriage, divorce, and children’s adjustment. Thousand Oaks, California 91320: Sage
Publications.
Ellingsen, I. T., Thorsen, A. A., & Størksen, I. (2014). Revealing children’s experiences and emotions
through Q methodology. Child Development Research, 22, 1–​9. doi:.org/​10.1155/​2014/​910529
Emery, R. E, & Coiro, M. J. (1995). Divorce: consequences for children. Pediatric Review, 16, 306–​310.
Fergusson, D. M., McLeod, G. F. H., & Horwood, L. M. (2014). Parental separation/​divorce in childhood
and partnership outcomes at age 30. Journal of Child Psychology and Psychiatry, 55, 352–​360.
Fosco, G. M., & Grych, J. H. (2012). Capturing the family context of emotion regulation: A family systems
model comparison approach. Journal of Family Issues, 34, 558–​578. doi:10.1177/​0192513X12445889
Frankel, L. A., Hughes, S. O., O’Connor, T. M., Power, T. G., Fisher, J. O., & Hazen, N. L. (2012). Parental
influences on children’s self-​regulation of energy intake: Insights from developmental literature on
emotion regulation. Journal of Obesity, 2012 1–​12. doi:10.1155/​2012/​327259
Furstenberg, F. F., & Cherlin, A. J. (1991). Divided families: what happens to children when parents part.
Cambridge: Harvard University Press.
Furstenberg, F. F., & Teitler, J. O. (1994). Reconsidering the effects of marital disruption: What happens to
children of divorce in early adulthood? Journal of Family Issues, 15, 173–​190.
Grätz, M. (2015). When growing up without a parent does not hurt: Parental separation and the
compensatory effect of social origin. European Sociological Review, 2015, 1–​12. doi:10.1093/​esr/​jcv057
Grillon, C., Quispe-​Escudero, D., Mathur, A., & Ernst, M. (2015). Mental fatigue impairs emotional
regulation. Emotion, 15, 383–​389. doi.org/​10.1037/​emo0000058
Gross, J. J. (2002). Emotion regulation: Affective, cognitive and social consequences. Psychophysiology, 39,
281–​291. doi:10.1017.S0048577201393198
8
4
3

348 Children of Divorce

Gruber, J. (2004). Is making divorce easier, bad for Children? The Long-​Run Implications of Unilateral
Divorce, Journal of Labor Economics, 22, 830.
Grych, J. H., & Fincham, F. D. (1990). Marital conflict and children’s adjustment: A cognitive contextual
framework. Psychological Bulletin, 267–​290.
Hetherington, E. M & Stanley-​Hagan, M. (1999). The adjustment of children with divorced parents: A risk
and resiliency perspective. Journal of Child Psychology and Psychiatry, 40, 129–​140.
Hike, K., Wolchik, S, Sandler, I. & Braver, S (2002). Predictors of Children’s Intervention-​Induced
Resilience in a Parenting Program for Divorced Mothers. Family Relations, 51, 121–​129.
Hoffman, J., & Russ, S. (2012). Pretend play, creativity and emotion regulation in children. Psychology of
Aesthetics, Creativity, and the Arts, 6, 175–​184. doi:10.1037/​a0026299
John, O. P. & Gross, J. J. (2004), Healthy and unhealthy emotion regulation: Personality processes,
individual differences, and life span development. Journal of Personality, 72, 1301–​1334. doi:10.1111/​
j.1467-​6494.2004.00298.x
Kalmijn, M. (2016). Children’s divorce and parent–​child Contact: A within-​family analysis of older
European parents. Journals of Gerontology, Series B: Psychological Sciences and Social Sciences, 71(2),
332–​343. doi:10.1093/​geronb/​gbu141
Kelly, J. B. (1998). Marital conflict, divorce and children’s adjustment. Child and Adolescent Psychiatry
Clinics of North America, 7, 259–​271.
Kim, H. S. (2011). Consequences of parental divorce for child development. American Sociological Review,
20, 1–​25
Kinard, E. M. & Reinherz, H. (1986). Effects of Marital Disruption on Children’s School Aptitude and
Achievement, Journal of Marriage and Family, 48, 289–​290.
Kliewer, W., & Sandler, I. (1993). Social competence and coping among children of divorce. American
Journal of Orthopsychiatric Association, 63, 432–​440.
Kurtz, L. (1994). Psychosocial coping resources in elementary school-​age children of divorce. American
Journal Orthopsychiatric, 64, 554–​563.
Lacey, R., Kumari, M. & McMunn, A. (2013). Parental separation in childhood and adult
inflammation: The importance of material and psychosocial pathways. Psychoneuroendocrinology, 38,
2476–​2484.
Landstedt, E., Hammarström, A., & Winefield, H. (2015). How well do parental and peer relationships in
adolescence predict health in adulthood? Scandinavian Journal of Public Health, 43, 460–​468.
Lee, B., & Bax, K. (2000). Children’s reactions to parental separation and divorce. Paediatric Child Health, 5,
217–​218.
Lewis, J. L., & Wallerstein, J. S. (1987). Methodological issues in longitudinal research on divorced families.
In J. P. Vincent (Ed.), Family Intervention, Assessment and Theory, 4, 121–​142. Greenwich, CT: JAI Press.
Mahrer, R., Winslow, E., Wolchik, S., Tein, J., & Sandler, I. (2014). Effects of a Preventive Parenting
Intervention for Divorced Families on the Intergenerational Transmission of Parenting Attitudes in
Young Adult Offspring. Child Development, 85(5), 2091–​2105
Marquardt, E. (2005). Between two worlds: The inner lives of children of divorce. NY: Three Rivers Press.
Maughan, A., & Cicchetti, D. (2002). Impact of child maltreatment and interadult violence on children’s
emotion regulation abilities and socioemotional adjustment. Child Development, 73, 1525–​1542.
Morelen, D., Shaffer, A., & Suveg, C. (2016). Maternal emotion regulation: Links to emotion parenting and
child emotion regulation. Journal of Family Issues, 37, 1891–​1916. doi:10.1177/​0192513X14546720
Morris, A. S., Silk, J. S., Steiberg, L., Myers, S. S., & Robinson, L. R. (2007). The role of the family
context in the development of emotion regulation. Social Development, 16, 361–​388. doi:10.1111/​
j.1467-​9507.2007.00389.x
Pedro-​Carroll, J. Fostering Resilience in the Aftermath of Divorce: The Role of Evidence-​Based Programs
for Children, 2005. Family Court Review, 43, 52–​64.
9
4
3

Programs targeting parenting skills 349

Pollet, S. (2009) A Nationwide Survey of Programs for Children of Divorcing and Separating parents.
Family Court Review, 47, 523–​543.
Potter, D. (2010). Psychosocial well-​being and the relationship between divorce and children’s academic
achievement, Journal of Marriage and the Family, 72, 941.
Prokic, T., Dronkers, J (2009). Parental Divorce and Attitudes about Society of their children. Paper
Presented at the Seventh Meeting of the European Network for the Sociological and Demographic
Study of Divorce, June 25–​26, Antwerp Belgium: European University Institute, Retrieved October 2015
from http://​apps.eui.eu/​Personal/​Dronkers/​English/​Prokic.PDF
Reiter, S. F., Hjörleifsson, S., Breidablik,H. J., Meland, E. (2013). Impact of divorce and loss of parental
contact on health complaints among adolescents. Journal of Public Health, 35, 278–​285.
Rossiter, A. B. (1988). A model for group intervention with preschool children experiencing separation and
divorce. American Journal of Orthopsychiatry, 58, 387–​396.
Sandler, I. N., Ayers, T. S., Wolchik, S. A., Tein, J., Kwok, & O. Haine, R. A. (2003). The Family
Bereavement Program: Efficacy evaluation of a theory-​based prevention program for parentally
bereaved children and adolescents. Journal of Consulting and Clinical Psychology, 71, 587–​600.
Schick, A. (2002). Behavioral and emotional differences between children of divorce and children from
intact families: Clinical significances and meditating process. Swiss Journal of Psychology, 61, 5–​14.
Schraub, E. M., Turgut, S., Clavairoly, V., & Sonntag, K. (2013). Emotion regulation as determinant of
recovery experiences and well-​being: A day-​level study. International Journal of Stress Management, 20,
309–​355.doi:10.1037/​a0034483
Sedghi, A & Rodgers, S. (2014). Divorce rates data, 1858 to now: how has it changed? The Guardian.
Retrived from http://​www.theguardian.com/​news/​datablog/​2010/​jan/​28/​divorce-​rates-​marriage-​ons
Siffert, A., & Schwarz, B. (2011). Parental conflict resolution styles and children’s adjustment: Children’s
appraisals and emotion regulation as mediators. The Journal of Genetic Psychology, 172, 21–​39.
Social Trends Institute (2012). Global Structure, Sustainable Demographic Dividend, What do marriage and
fertility have to do with the economy? Retrieved from http://​sustaindemographicdividend.org/​articles/​
international-​family-​indicators/​global-​family-​structure
Stallman, H., & Sanders, M. (2014) (abs). A randomized controlled trial of family transitions triple p: a
group-​administered parenting program to minimize the adverse effects of parental divorce on children.
Journal of Divorce & Remarriage, 55, 33–​48.
Stolberg, A. & Mahler, J. (1994). Enhancing treatment gains in a school-​based intervention for children of
divorce through skill training, parental involvement, and transfer procedures. Journal of Consulting and
Clinical Psychology, 62, 147–​156.
Tetzner, J., & Becker, M. (2015). How being an optimist makes a difference: The protective role of optimism
in adolescents’ adjustment to parental separation. Social Psychological and Personality Science, 6, 325–​
333. doi:10.1177/​1948550614559605
Thomson, R. A., & Calkins, S. D. (1996). The double-​edged sword: Emotion regulation in high risk
children. Development and Psychopathology, 8, 163–​182.
Trzesniewski, K. H., Donnellan, M. B. & Robins, R. W. (2003). Stability of self-​esteem across the life span.
Journal of Personality and Social Psychology, 84, 205–​220.
United Nations Statistical Division (UNSTAT) (2011). Divorces and crude divorce rates. Retrieved from http://​
unstats.un.org/​unsd/​demographic/​products/​dyb/​dyb2011/​Table25.pdf
Velderman, M., Cloostermans, A., & Pannebakker, F. (2014). Child adjustment in divorced families: Can
we successfully intervene with Dutch 4-​to 6-​year olds? Feasibility study Children of Divorce
Intervention Program (CODIP) in the Netherlands. TNO report, TNO/​CH 2014 R10697
Velez, C., Wolchik, S., Tein, J. & Sandler, I. (2011). Protecting Children from the Consequences of
Divorce: A Longitudinal Study of the Effects of Parenting on Children’s Coping Processes. Child
Development, 82, 244–​257. http://​www.child-​encyclopedia.com/​divorce-​and-​separation/​according-​
experts/​interventions-​help-​parents-​and-​children-​through/​-​separation
0
5
3

350 Children of Divorce

Wallerstein, J. S., & Blakeslee, S. (1989). Second chances: Men, women and children a decade after divorce.
Boston: Houghton Mifflin.
Wallerstein, J. S., & Kelly, J. B. (1980). Surviving the breakup: How parents and children cope with divorce.
New York: Basic Books.
Wallerstein, J. S., Lewis, J. M., & Blakeslee, S. (2002). The unexpected legacy of divorce: A 25-​year landmark
study. New York: Hyperion.
Weaver, J., & Schofield, T. (2014). Mediation and moderation of divorce effects on children’s behavior
problems. Journal of Family Psychology, 29, 39–​48.
Wolchik S. A., West S. G., Westover S., Sandler I. N., Martin A., Lustig J., et al (1993). The children of
divorce parenting intervention: Outcome evaluation of an empirically based program. American Journal
of Community Psychology, 21, 293–​331.
Zalewski, M., Lengua, L., Wilson, A., Trancik, A., & Bazinet, A. (2011). Associations of coping and
appraisal styles with emotion regulation during preadolescence. Journal of Experimental Child
Psychology, 110, 141–​158.
Zhou, Q., Sandler, I., Millsap, R. Wolchik, S., & Dawson-​McClure, S. (2008). Mother-​Child relationship
quality and effective discipline as mediators of the 6-​year effects of the New Beginnings Program for
Children from divorced families. Journal of Consulting and Clinical Psychology, 76, 579–​594.
1
5
3

Chapter 17

Children’s and Adolescents’ Emotion


Regulation in the Context of Parental
Incarceration
Janice Zeman & Danielle Dallaire

“I felt so sad. I was just crying. It just made my head hurt, my brain hurt, my
stomach hurt. It just got control of me. It got my mind twisted. I couldn’t
focus on anything else. A whole lot of days I couldn’t go to sleep without my
mum. I had some bad dreams, so my daddy gave me an invisible necklace. I
couldn’t live without her. It was like a curse. It was like prison.” Jasmine
(Zehr, 2011)

Parental incarceration
The United States (U.S.) incarcerates more individuals than any other country (Walmsley, 2013),
and approximately half of U.S.  prisoners are parents (Glaze & Maruschak, 2008). Accordingly,
millions of U.S.  children experience the collateral consequences of parental separation due to
incarceration. It has been documented that children and adolescents with incarcerated parents are
prone to a diverse array of maladaptive developmental outcomes, including aggressive and anti-
social behavior (Murray, Farrington, & Sekol, 2012), depression (Wilbur et al., 2007), attachment
insecurity (Poehlmann, 2005a), and diminished educational attainment (Haskins, 2014; Trice
& Brewster, 2004). Increasingly, evidence suggests that youth with a history of parental incar-
ceration also evidence detrimental effects on physical health, including health disease and other
chronic diseases (Gjelsvik, Dumont, & Nunn, 2013). What is strikingly absent from the literature
on parental incarceration is an investigation into the possible mediators and moderators of the
relationship between the experience of parental incarceration and the psychosocial and physical
health outcomes. One such factor may be emotion-╉related processes, such as emotion regulation.
The primary focus of this chapter is to review the literature examining emotion processes in
children and adolescents who have experienced parental incarceration with a specific focus on
emotion regulation. Background information on parental incarceration within the U.S. penal sys-
tem is presented first to provide the context from which to understand and interpret the research
regarding emotion processes. The first section thus provides information describing the variety
of contextual variables that characterize maternal and paternal incarceration and its sequaelae.
Second, we provide an overview of the negative effects of parental incarceration on children’s aca-
demic, social, and psychological functioning. Third, we discuss emotion processes and propose
two important factors that may explain, in part, why parental incarceration confers additional
risk for psychological maladjustment through:1) The disruption of the attachment relationship
between the incarcerated parent and the child, and 2) the development of poor emotion regulation
2
5
3

352 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

skills. We also discuss how adaptive emotion regulation skills may be a potential protective factor
that may help to lessen the impact of the negative outcomes associated with having an incarcer-
ated parent. We conclude by highlighting the limitations of the current research in this area, the
challenges inherent in conducting research in this area, and the implications and directions for
future research.

The context of parental incarceration


Overview of incarceration types and incidence
Incarceration is defined as confinement, typically within a jail or prison facility. There are some
important differences between prison and jail incarcerations, which may impact children’s experi-
ence of parental incarceration. Individuals incarcerated at jail facilities may be awaiting trial for
pending charges or serving a relatively short sentence, typically less than one year, though this
varies by state (Bureau of Justice Statistics, 2015). In contrast, prisons are long term facilities oper-
ated by the state or the federal government and typically hold felons and inmates with sentences
of more than one year (Bureau of Justice Statistics, 2015). Because jail stays can be relatively short,
many more individuals cycle in and out of jail systems in a typical year (Minton, 2013; Wagner &
Sakala, 2014). For example, although at any given time state and federal prisons house approxi-
mately one-╉and-╉a-╉half million individuals, jails typically only house about half that number. But,
throughout the course of a year, almost 12 million people will cycle through local jails, a num-
ber much higher than that of state and federal prisons (Minton, 2013; Wagner & Sakala, 2014).
Jail facilities, in comparison to federal or state prisons, are often located in closer proximity to
inmates’ homes and families making the possibility of visitation more likely. Additionally, because
jail stays are usually of a shorter duration and may occur several times over the course of a year,
they may cause considerably more family disruption than a single long-╉term prison sentence.
Although the U.S. represents about 4.4% of the world’s population, it houses around 22% of the
world’s prisoners (Walmsley, 2013). The extraordinarily high rates of incarceration in the U.S. are
unprecedented in the history of the U.S. as well as internationally (Travis, Western, & Redburn,
2014). There are a multitude of causes of mass incarceration (see Travis et al., 2014) including
harsher sentencing laws and mandatory minimum sentencing implementation. Surprisingly, rates
of incarceration have increased while crime rates have decreased (Bureau of Justice Statistics,
2015). Furthermore, the U.S. prison populations also show high levels of mental illness, particularly
since the mass closings of the state-╉run mental hospitals in the 1960s and 1970s (Harcourt, 2011).
Mass incarceration has resulted in the imprisonment of many parents. Over half of the nation’s
federal and state prisoners report being a parent of a child under the age of 18 (Glaze & Maruschak,
2008), and there are even more individuals incarcerated in local community jails. Parental incar-
ceration has far-╉reaching consequences, particularly on communities and the health and well-╉
being of American families and children (Hatzenbuehler, Keyes, Hamilton, Uddin, & Galea,
2015). In some urban communities, it is more likely that a young man will go to jail or prison
than to college (Justice Policy Institute, 2002). There are other impacts on communities as well,
including a lack of social and human capital (e.g., skilled workers, thriving schools community
infrastructure) and high rates of poverty (see Clear, 2008).

Maternal vs paternal incarceration


The empirical study of how parental incarceration impacts children and families has been a bur-
geoning research topic for the last decade (Murray et al., 2012). This growing research base has
yielded considerable insights into the family environment of children with incarcerated parents
3
5

The context of parental incarceration 353

and demonstrated the heterogeneous impact of parental incarceration on children (Turney &
Wildeman, 2015). In particular, when considering the impact of parental incarceration, it is
important to differentiate between incarceration in prison vs. jail as well as maternal vs. paternal
incarceration. Even the ground-​breaking Adverse Childhood Experiences study (see below) did
not differentiate between parental incarceration in jail or prison, or maternal or paternal impris-
onment, nor take into account the length of parental imprisonment (Anda et  al., 2001; Felitti
et al., 1998). Unfortunately, to date, no research that we are aware of has examined differences in
children’s experience of parental incarceration in prison or jail or how the length of incarceration
may impact child and family functioning. However, a growing body of research has examined the
experience of paternal vs. maternal incarceration.
Paternal incarceration is much more common than maternal incarceration (Glaze & Maruschak,
2008). The U.S. jail and prison population is predominately male and therefore, there are more
children impacted by paternal than maternal incarceration. However, maternal and paternal
incarceration is more likely to co-​occur when a mother is incarcerated (Dallaire, 2007). Thus,
many children impacted by a mother’s incarceration may be experiencing separation from both
their mother and their father. The heterogeneous impact of parental incarceration is well docu-
mented in the literature (Murray et al., 2012) and may be partly a reflection of the diversity of the
experience of paternal, maternal, or dual-​parental incarceration.
In comparison to incarcerated mothers, incarcerated fathers are less likely to report having lived
with the child prior to their incarceration and are less likely to remain in contact with their chil-
dren during their incarceration (Glaze & Maruschak, 2008). Conversely, most incarcerated moth-
ers report living with their child during the month prior to arrest and/​or incarceration and being
the child’s primary caregiver. According to incarcerated mothers’ report, their adult children are
two-​and-​a-​half times more likely to be incarcerated than adult children with incarcerated fathers
(Dallaire, 2007), and three times as likely to be incarcerated as adults, compared to children whose
mothers have never been incarcerated (Huebner & Gustafson, 2007).
According to the Bureau of Justice Statistics (Glaze & Maruschak, 2008), when fathers are incar-
cerated, their children likely remain in the care of their biological mother. When mothers are
incarcerated, their children are more likely to transition to the care of a grandparent or other
relative. Most children whose parents are incarcerated remain in familial care. However, approxi-
mately 3% of incarcerated parents report that their child is in the foster care system, with 11%
of incarcerated mothers, compared to only 2% of incarcerated men reporting that one or more
of their children are in state custody (Glaze & Maruschak, 2008). Thus, maternal incarceration,
because of the increased likelihood of severed or limited contact with both biological parents, may
serve as an additional or intensifying risk factor for these children.
As previous research has demonstrated (e.g., Sameroff, Bartko, Baldwin, Baldwin, & Siefer,
1998), as the number of risk factors in children’s lives accumulate, children evidence increased
likelihood of problematic outcomes, which may reflect the situation of children with an incarcer-
ated mother. Dallaire (2007) found that as the number of risks accumulate in the lives of adult
children of incarcerated parents, so too does their children’s risk for incarceration; the relation
was particularly pronounced for children of incarcerated mothers. As contextual (e.g., large fam-
ily size, single parenthood) and incarceration-​specific (e.g., familial incarceration, number of
previous incarcerations) risk factors increased, the likelihood of intergenerational incarceration
increased as well. Specifically, Dallaire (2007) reported that:
Of the 21% of mothers who reported that their adult child was incarcerated, two-​thirds of these moth-
ers had four or more risk factors. With 1 risk factor, 1% of mothers reported that their adult children
were incarcerated; with 4 risk factors, the percentage of mothers with adult children incarcerated was
4
5
3

354 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

almost 6%. There was a similar trend for fathers, such that of the 8.5% of fathers who reported that
their adult child was incarcerated, almost three-╉quarters of these fathers had four or more risk factors.
With 1 risk factor almost no (0.1%) fathers reported that their adult child was incarcerated, but with
4 risk factors the percentage of incarcerated fathers with adult children incarcerated was 2%. (p. 448).

General contextual risk vs. incarceration-╉specific risk


The negative effects of parental incarceration on children may be due, in part, to the fact that chil-
dren are exposed to contextual risks, including parental substance abuse and mental illness. James
and Glaze (2006) found that 25% of inmates in state and federal jails and prisons who participated
in a clinical, diagnostic interview were likely to have a history of mental health problems—╉a per-
centage much higher than the estimated rate of 11% in the general population. Children’s experi-
ence of other contextual risk factors, such as poverty, and substance abuse, when added to the
stress of parental incarceration, may help explain why children with incarcerated parents may be
at heightened risk for psychopathology and antisocial outcomes.
Research examining the family environments of children with incarcerated parents has dem-
onstrated that in comparison to children and youth whose parents are not involved with the
criminal justice system, children of an incarcerated parent are more likely to be exposed to
parental substance abuse, domestic violence, and poverty (Phillips, Burns, Wagner, & Barth,
2004). Specifically with regard to poverty, analyses of the Fragile Families and Child Well Being
(FFCWB) data set has demonstrated that children exposed to parental incarceration were more
likely to experience poverty and family instability than children with no history of parental
incarceration (Western & Wildeman, 2009). Mackintosh, Myers, and Kennon (2006) also found
that caregivers (who included mothers and grandmothers) reported clinically significant levels
of parenting stress.
To summarize, first, with more individuals incarcerated in the U.S. than elsewhere in the world,
more American children are separated from one or more parents because of parental incarcera-
tion in jail and prison facilities. Second, although more children are impacted by a father’s impris-
onment, maternal incarceration confers unique and specific risks and challenges for children and
families. Third, children impacted by parental incarceration face numerous risks in their caregiv-
ing environment, including parental mental illness, parental substance abuse, poverty, and harsh
caregiving.

Effects of parental incarceration on children


The importance of studying parental incarceration is validated through the landmark study which
identified key risk factors known as adverse child experiences (ACEs) that are thought to derail
children’s optimal developmental trajectories (Anda et al., 2001; Feletti et al., 1998). Parental incar-
ceration has been documented as an ACE along with general household dysfunction (i.e., parental
substance abuse, parental divorce/╉separation, parental mental illness), domestic violence, and the
experience of childhood abuse and neglect. Research indicates that ACEs are highly inter-╉related
and are associated with health problems and depression (Anda et al., 2001; Chapman et al., 2004;
Dube et al., 2001). In particular, parental incarceration has been demonstrated to co-╉occur with
other ACEs (Johnson & Waldfogel, 2002) and independently predict mental and physical health
problems in adulthood (Gjelsvik et al., 2013; Gjelsvik, Dumont, Nunn, & Rosen, 2014). Adverse
childhood experiences are thought to impair social and emotional development, leading to the
adoption of health-╉risk behaviors, which then predispose one to disease and other social prob-
lems (including incarceration) and potentially early death. The impact of parental incarceration
on psychopathology may vary based on the extent to which children’s emotion-╉related process are
affected by parental incarceration.
53

The context of parental incarceration 355

A body of research has indicated that numerous negative academic, social, and psychological
outcomes are present in children with an incarcerated parent. With respect to cognitive function-
ing, in a sample of children ages two-​and-​a-​half to seven-​and-​a-​half years of currently incarcer-
ated mothers, Poehlmann (2005b) found that 42.4% of children were characterized by sub average
cognitive functioning and delays as assessed by the Stanford-​Binet. However, these findings were
mediated by the quality of the home environment. That is, children who lived in environmental
contexts with higher numbers of contextual risk (e.g., low education, unemployment) had poorer
cognitive outcomes, but these outcomes were mediated by positive, safe, stimulating home envi-
ronments. Hanlon et al. (2005) interviewed families of 88, nine to 14 year old youth in Baltimore
City whose mothers were currently incarcerated and participating in a parenting program. The
results indicated that although the youth did not show significant indications of psychological
maladjustment, school behavior problems and associations with deviant peer groups were evi-
dent. The authors posited that the longer term, consistent care provided by caregivers may have
mitigated some of the anticipated negative outcomes.
Regarding psychological outcomes associated with being a child of an incarcerated parent the
findings are mixed. When children are first separated from parents because of incarceration they
show many emotions, including sadness, worry, confusion, anger, loneliness, and sleep problems
(Poehlmann, 2005a). However, over time children and families make adaptations and adjust-
ments, and the relation between parental incarceration and psychopathology is less clear. Given
the myriad of contextual risks in the caregiving environments of children of an incarcerated par-
ent, it is difficult to disentangle the contribution of parental incarceration over and above the
impact of other significant experiences to which these children have been exposed. Increasingly,
it appears that there might not be a direct link between the experience of parental incarcera-
tion and children’s psychopathology. A comprehensive meta-​analysis by Murray and colleagues
(2012) of 40 studies that examined the impact of parental incarceration in relation to a compari-
son group of youth who did not have experience with parental incarceration yielded two main
conclusions. First, the authors concluded that, “there are zero or only weak associations between
parental incarceration and children’s poor mental health, drug use, and educational performance”
(pp.  190–​191). Second, they concluded that, “children with incarcerated parents are at signifi-
cantly higher risk for antisocial behavior compared with their peers” (pp. 191).
These findings compliment work conducted by researchers analyzing the FFCWB data set,
which has indicated that the impact of parental incarceration may be negligible (Turney &
Wildeman, 2015). Other researchers too have documented themes of resilience and adaptation
among these children and youth (Miller, 2007; Siegel, 2007). It is well documented, however, that
children with incarcerated parents experience high levels of risk in their environment, and they
show high rates of antisocial behavior, delinquency, and have high rates of incarceration as youth
and young adults (Huebner & Gustafson, 2007; Murray & Farrington, 2005). For example, Murray
and Farrington (2005) analyzed the date from the Cambridge Study in Delinquent Development
to investigate the effects of various types of parent-​child separation including paternal incarcera-
tion. The results indicated that the experience of paternal imprisonment for boys prior to age ten
was related to future antisocial and delinquent outcomes, even up to age 40, even after controlling
for other risk factors. However, this study was conducted from 1953–​1964 in Great Britain and
the results may not replicate to present times and to the US In conclusion, based on the recent
meta-​analyses conducted by Murray and colleagues (2012), there appears to be no or little direct
link between parental incarceration and psychopathology.
Although there has been considerable research examining psychological outcomes for chil-
dren of incarcerated parents, very few researchers have studied emotion-​related processes in this
sample. Emotion regulation and coping abilities in particular may help explain the diversity of
6
5
3

356 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

outcomes in this population and act as an intervening variable, which can help better explain why
some children of parental incarceration demonstrate resilient outcomes; whereas, others evidence
significant adjustment problems. The following sections of this chapter provides an in depth focus
on the scant research that has examined the relations between the experience of parental incar-
ceration and attachment relationships given its foundational role in the development of emotional
processes, and the relation between parental incarceration and the development of children’s emo-
tion regulation skills.

Emotion processes
Attachment in children with an incarcerated parent
Attachment theory has been proposed as a useful theoretical framework to understand why chil-
dren with incarcerated parents may be an at-╉risk population, and also offers pathways to resil-
ience that can serve as junctures for intervention or prevention efforts. A secure mother-╉child
attachment relationship has been demonstrated to provide the necessary foundation for positive
emotional development that is integrally related to functioning in cognitive and social domains
(Cassidy, 1994; Sroufe, 2005). In particular, research has indicated that attachment security pro-
motes the development of key emotional processes such as emotional awareness and empathetic
responding in children (Kerns, 2008; Thompson, 2008). These emotion development skills are
foundational for the developmental of more advanced emotional competencies including emo-
tion regulation (Saarni, 1999). Infants and young children with incarcerated parents may not
form secure attachment relationships with the incarcerated parent which has significant implica-
tions for their risk of developing subsequent emotional and behavioral problems (Myers, Smarsh,
Amlund-╉Hagen, & Kennon, 1999; Parke & Clark-╉Stewart, 2001).
Poehlmann (2005a) studied attachment relationships in a sample of 54 children ages two-╉and-╉
a-╉half to seven-╉and-╉a-╉half years (M age = 56.8 months, 47% girls, 48% African American) with
a currently incarcerated mother. One of the eligibility requirements for participation was that
the mother had been the primary caregiver prior to incarceration. The average age of the child
when separated from his or her mother due to incarceration was 33 months. Multiple methods
of assessment and multiple informants (i.e., mothers, caregivers, children) were used to assess
attachment. The primary assessment of attachment relationship was based on caregiver and
mother report of the child’s initial two-╉week response to the maternal separation as well as on
children’s responses to an attachment story completion task in which children responded to four
increasingly stressful hypothetical events concerning parent-╉child relationships (Bretherton,
Ridgeway, & Cassidy, 1990). Results based on the story-╉completion task indicated that most
(63%) of the children were classified as having insecure relationships with mothers and caregiv-
ers; 54% of children had insecure relationships with both mothers and caregivers, and 28% had
secure relationships with both adults. Eighteen percent of children had one secure and one inse-
cure parental representation. Interestingly, rates of insecurity in the Poehlmann (2005a) study
were not significantly different than rates of insecurity in other high-╉risk populations (Cassidy
et al., 2007). In the Poehlmann (2005a) study, secure attachment relationships were more likely
when children lived in a stable caregiving situation, were older, and evidenced sad rather than
angry responses to the initial parental separation. These results indicate that parental incarcera-
tion negatively impacts children’s ability to form and maintain stable, secure attachment repre-
sentations with their incarcerated parent. To the extent that children with incarcerated parents
may be able to form and/╉or maintain a secure, organized attachment relationship with their
incarcerated parent or another caregiver, they may be protected from some of the other risks
associated with the experience of parental incarceration.
7
5
3

Emotion processes 357

Specific aspects of children’s interactions with an incarcerated parent in the context of incar-
ceration may stress the attachment and caregiving systems. In particular, attachment-​related
thoughts and feelings may be an important issue for these children and families in relation to the
child’s level and quality of contact with the incarcerated parent (see also Poehlmann, Dallaire,
Loper, & Shear, 2010). There is mounting evidence that children’s contact with an incarcerated
parent may activate a child’s attachment system and that lack of contact may increase feelings
of anger. Shlafer and Poehlmann (2010) examined attachment representations in 57 youth (M
age = 9.1 years, range:four–​15 years; 60% girls; 93% minority ethnicity/​race) with an incarcer-
ated parent (86% had incarcerated fathers). Children were interviewed about their relationships
with their caregivers (primarily mothers) and their incarcerated parent using the Inventory of
Parent and Peer Attachment Inventory (IPPA; Armsden, 1986; Armsden & Greenberg, 1987).
Caregivers provided information about their perceptions of their relationship with the child. All
families were involved in a mentoring program thus measures were administered at baseline and
six months later. When asked directly about their relationship with the incarcerated parent, 39%
provided no answer, 31% perceived the incarcerated parent negatively, and 41% reported a posi-
tive relationship. Interestingly, for children who were older than eight years old and had no con-
tact with the incarcerated parent, they reported feeling alienated and having negative feelings
towards that parent. Children in this study who did have contact with their incarcerated parent
reported significantly fewer feelings of anger and alienation toward the incarcerated parent. This
suggests the importance of maintaining and developing the attachment relationship that may have
been present prior to incarceration.
Dallaire, Ciccone, and Wilson (2012) compared attachment security in a sample of 24 children
(M age = 7.7 years, range: six to ten years; 42% girls; 58% African American) with an incarcerated
parent (71% had incarcerated fathers) to a sample of 20 children (M age = 8.6 years, range: six
to ten years; 30% girls; 85% African American) separated from a parent for another reason (e.g.,
divorce, abandonment, substance abuse recovery). The authors assessed attachment security
through the use of the Attachment Family Drawing task (Fury, Carlson, & Sroufe, 1997)  that
was then related to the child’s experience of contact and visitation with the incarcerated parent.
They found that greater contact and visitation with an incarcerated parent was associated with
more depictions of role-​reversal in their family drawings. Role reversal involved depictions of
the mother as vulnerable or less important in the relationship than the child. The authors inter-
preted this finding as suggesting that children with incarcerated parents perceived that their care-
giver needed to be protected and it was their responsibility to be the protector. This research also
examined specific dimensions in the children’s drawings with caregiver behaviors such as warmth
and hostility. Children of incarcerated parents’ perceptions of caregiver hostility was related to
increased global insecurity as depicted in the drawings, whereas, both child and caregiver per-
ceptions of stress were associated with more child global pathology as well as “bizarreness/​dis-
sociation” depicted in the drawings (p.  178). Interestingly, these findings were not seen in the
comparison group of children. This set of findings indicates that importance of considering the
effect of parental incarceration on children’s attachment relationship with both the caregiver and
the incarcerated parent. Not all children experience insecure attachment in this high risk context
and thus, future research needs to uncover what factors may help to strengthen the relationship
between the incarcerated parent and his or her child while incarcerated and also upon reunion
and reintegration into the family upon release.
In sum, attachment in the context of parental incarceration has been examined in two different
ways. Poehlmann (2005a) assessed children’s current representations of attachment with mothers
and caregivers and found high rates of insecurity. Other researchers have examined attachment-​
related thoughts and feelings in relation to children’s contact with their incarcerated parent.
8
5
3

358 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

Shlafer and Poehlmann (2010) found that lack of contact was associated with increased feelings
of anger and alienation with the incarcerated parent, and Dallaire and colleagues (2012) reported
that greater contact and visitation was associated with greater role reversal. These findings illus-
trate the heterogeneity of the impact of parental incarceration on children and their attachment-╉
related thoughts and feelings. Specifically, some children, particularly those with a non-╉residential
incarcerated father, may not show attachment-╉related difficulties as they may have developed an
attachment relationship with their mother or other caregiver. However, young children with an
incarcerated mother may be at-╉risk for attachment-╉difficulties as they are more likely to experi-
ence separation from a primary attachment figure. To further complicate matters, children who
wish to maintain an attachment relationship with their incarcerated parent via visitation and con-
tact may experience attachment-╉related distress following a visit or phone call with the incarcer-
ated parent. Although it is difficult to isolate the impact of parental incarceration on children’s
attachment relationships, attachment-╉related thoughts and feelings are important to consider, as
they are foundational to emotional development and coping. Attachment-╉focused research with
children of incarcerated parents has illustrated that they are at risk for developing insecure attach-
ments and, as discussed more in the next section, visitation and contact (or lack thereof) can also
impact the attachment system.

Emotional experiences in visitation


Only one study has examined children’s emotional responses to parental incarceration. One
of the unique situations for children with an incarcerated parent is the emotional arousal they
experience during the brief reunions when and if children are permitted to visit their incarcer-
ated parent (Poehlmann-╉Tynan et  al., 2015). Research to date has primarily focused on the
type (e.g., plexiglass barrier, video visitation, in person) and frequency of visitation, and has
relied on the adult perspective of the visit (Arditti, Lambert-╉Schute, & Joest, 2003). To address
this gap in the literature, Poehlmann-╉Tynan and colleagues (2015) observed the emotional and
behavioral reactions of 20 children (M age = 3.9 years, range: two to six years; 55% girls; 40%
African American) when they visited their incarcerated parent (90% were incarcerated fathers).
Children were accompanied by their caregiver (75% mothers, 25% grandmothers) to the jail
and visited with the incarcerated parent through a plexiglass barrier or by closed circuit tele-
vision (video visit). In both situations, only one person could talk to and hear the parent at a
time using a headset. The observer noted the children’s emotional reactions during the security
procedures, the wait time, and the visit with the incarcerated parent. Among other variables,
children’s affect (e.g., tired, fearful, sad, happy, angry, anxious), attachment behaviors toward
caregiver (e.g., staying in close proximity, clinging, hitting adult), and their emotions towards
the incarcerated parent (e.g., sad, happy, excited, confused, crying) were coded. The child’s
overall level of emotional lability was rated on a five-╉point scale. Behavioral measures were also
observed but are not discussed here given the focus of this chapter. Overall, the observations
indicated that most children exhibited two predominant emotions including happy and seri-
ous/╉somber facial expressions. The longer children visited their parent in jail, the more their
distress increased as did their proximity seeking behaviors with the caregiver. Further, there
was an increase in negative affect (i.e., fatigue, sadness, confusion, anger) over the course of
the visit, although the children were also observed to express happiness and loving behaviors
during the visit.
The authors concluded that the children experienced the visit as mildly stressful as indicated by
the increase in emotion dysregulation and the activation of the attachment system as evidenced
by increased contact seeking and contact maintenance over the course of the visit. Thus, although
9
5
3

Emotion processes 359

there were positive emotional moments during the visit, there were stressful, emotionally nega-
tive aspects as well. The findings from this research have important implications for interven-
tion and policy for child-╉friendly visitation to incarcerated parents. Further, from an emotional
development perspective, the visitation context provides caregivers with a valuable opportunity
to teach children how to manage emotional arousal in constructive ways. That is, drawing from
meta-╉emotion theory (Gottman, Katz, & Hooven, 1997), adopting an emotion coaching stance
in which instances of negative emotionality are viewed as “teachable moments” provides children
with the necessary support and scaffolding to experience, validate, and learn adaptive problem-╉
solving responses to manage negative emotional arousal. The use of emotion coaching strategies
has been demonstrated to result in positive psychosocial outcomes in low risk samples of children
(Lunkenheimer, Shields, & Cortina, 2007). Importantly, to be an effective emotion coach, care-
givers must be able to regulate their own negative emotionality, thereby providing the necessary
attention to the child’s emotional needs (Cassano & Zeman, 2010). However, the stress of the visit
not only affects children but also the caregivers who have their own set of emotional responses
to the brief reunion. Thus, managing their own emotions in order to be emotionally available to
their child is likely a challenging parenting task. Thus, future research needs to focus on devel-
oping ways to assist caregivers in becoming effective emotion coaches to help guide their child
through the myriad of emotional responses the child may experience when visiting the incarcer-
ated parent.

Emotion regulation in these children


Overview
The development of self-╉regulation, with emotion regulation as a key component (Sroufe, 2005),
is a critical skill that is integrally involved in children’s social (e.g., Denham, Bassett, & Wyatt,
2007), academic (e.g., Trentacosta & Izard, 2007), cognitive (e.g., Blair, 2002; Simonds, Kieras,
Rueda, & Rothbart, (2007), psychological (e.g., Zeman, Shipman, & Suveg, 2002), and physical
health functioning (e.g., Whitson & El Sheikh, 2003). Definitional clarity regarding the term
“emotion regulation” continues to be a topic of debate within the developmental psychology lit-
erature (Campos, Frankel, & Camras, 2004; Cole, Martin, & Dennis, 2004). We adopt Thompson’s
(1994) well-╉accepted definition which states that emotion regulation “consists of the extrinsic
and intrinsic processes responsible for monitoring, evaluating, and modifying emotional reac-
tions, especially their intensive and temporal features, to accomplish one’s goals” (pp.  27–╉28).
Importantly, the regulation of emotion involves multiple forms of management (e.g., suppression,
exaggeration, substitution of one emotion for another) that when implemented successfully, is
sensitive to the demands of the social context in which the emotion is evoked. This perspective
also acknowledges that emotion regulation efforts can originate from external sources from the
child such as when a parent may protect or shield a child from particularly emotionally intense
situations (e.g., watching a R-╉rated move when six years old) that may be too emotionally arous-
ing and exceed the child’s coping resources (Zeman, Cassano, Perry-╉Parrish, & Stegall, 2006). This
latter point is particularly germane for children experiencing parental incarceration, a point that
will be elaborated on more fully below. In contrast to emotion regulation, emotion dysregulation
is considered to be a core component of most aspects of psychopathology (e.g., Bradley, 2003,
Cicchetti, Ackerman, & Izard, 1995; Gross & Munoz, 1995; Keenan, 2000). Thus, understanding
the processes that promote the adaptive management of emotion in children is key, particularly
in high risk contexts in which emotion socializing agents such as parents, may be absent or poor
models and teachers of emotion regulation skills.
0
6
3

360 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

Emotion regulation is a critical competency to examine in children of parental incarceration as


they are exposed to a large number of stressors that likely elicit strong and frequent arousal of neg-
ative affect. For example, youth may feel sadness and anger at the loss of the parental figure in their
life due to incarceration and the resultant absences and support at key events or milestones in
their lives as well as in their day-╉to-╉day ordinary events. For more optimal outcomes, youth must
learn how to manage this emotional arousal in flexible ways that are sensitive to the demands
of their school, home, and peer environments. It may be that children and adolescents who are
capable of managing their emotions in constructive ways will be less susceptible to the negative
consequences of parental incarceration. Because emotion regulation is a foundational skill that
underlies competencies in other domains of functioning (e.g., social, cognitive), the benefits of
adaptive emotion regulation may lessen the negative impact of parental incarceration through
its indirect effects in these domains. That is, adaptive emotion regulation has been demonstrated
to buffer against psychosocial maladjustment (e.g., Zeman, Cassano, & Adrian, 2013), promote
academic success (Gumora & Arsenio, 2002; Izard et al., 2001), and is associated with constructive
peer relations and social competence (Legerski, Biggs, Greenhoot, & Sampilo, 2015 Perry-╉Parrish
& Zeman, 2011). As discussed previously, emotion regulation is posited to be a transdiagnostic
process that underlies many, if not all forms of psychopathology (Aldao, 2013; Kring & Sloan,
2010). Thus, learning adaptive emotion regulation skills may protect youth from the negative
effects of PI through its indirect relations to psychosocial and academic functioning. Although
the importance of studying emotion regulation in youth with an incarcerated parent is clear, little
research has been conducted on this topic, with five exceptions. Given the focus of this chapter on
emotion regulation processes, considerable detail about these five studies will be provided.

Emotion regulation and problem behaviors


Lotze, Ravidran, and Myers (2010) investigated the relations among poor emotion regulation,
the experience of shame and guilt, and how these variables predicted callous-╉unemotional
traits and problem behaviors. These relations were examined with a sample of 50 children (M
age = 9.8 years, range: six to 12 years; 62% girls; 55% African American) enrolled in a summer
camp for children with a currently incarcerated mother. To assess emotion regulation, children
answered questions assessing how they manage anger and frustration using two subscales from
the Early Adolescent Temperament Scale-╉Revised (Ellis & Rothbart, 1999). The camp counsel-
ors completed the Emotion Regulation Checklist (Shields & Cicchetti, 1997). Both reporters also
completed versions of questionnaires to assess callous-╉unemotional traits as well as internalizing
and externalizing behaviors. Only children completed the measures assessing shame and guilt.
Using regression analyses with age and gender entered in the first step, based on child report data,
poor regulation of anger and frustration, higher shame, and lower guilt predicted externalizing
behaviors, whereas, only poor emotion regulation predicted internalizing behaviors. Lower levels
of guilt predicted callous-╉unemotional traits. Based solely on adult report, low levels of adaptive
emotion regulation predicted internalizing behaviors and callous-╉unemotional traits, whereas,
emotional dysregulation predicted externalizing behaviors. When adult report was used to pre-
dict child reported outcomes, only emotional dysregulation predicted externalizing behaviors.
Gender and age were also significant predictors but because of the relatively small sample size,
they were not considered as moderators. The findings mirror those found in the literature using
community samples in which youth who have poor emotion regulation exhibit higher numbers
of both internalizing and externalizing behaviors (Chaplin, 2006; Supplee, Skuban, Shaw, & Prout,
2009; Zeman et al., 2002). The authors posit that the subset of children in this summer camp who
exhibited poor emotion regulation may have had poor models of effective emotion regulation at
home. This speculation was based on research which indicates that emotional understanding is
1
6
3

Emotion processes 361

less developed in children who reside in families with high levels of stress (Dunn & Brown, 1994;
Shipman & Zeman, 1999) and that experiencing a greater number of risk variables is associated
with poorer emotion regulation (Lengua, 2002).

Emotion regulation and children’s exposure to incarceration-╉related risk


Children of an incarcerated parent participated in a study examining the relation between socio-╉
demographic risk factors as well as incarceration-╉specific risk factors to emotion regulation
skills and psychological functioning outcomes (Dallaire & Wilson, 2010). A sample of 32 fami-
lies participated including children (M = 10.74, range: seven to 17 years; 47% boys, 60% African
American), an incarcerated parent (57% incarcerated mothers and 43% incarcerated fathers), and
the children’s caregivers (57% mothers). About half of the children (52%) were separated from
both parents. To assess emotion regulation, children completed the How I  Feel questionnaire
(Walden, Harris, & Catron, 2003) in which ten items were used to assess children’s perceptions
of how well they can control or change positive (i.e., happy, excited) and negative (i.e., scared,
mad, sad) emotions. Caregivers evaluated the child’s psychological functioning, whereas, caregiv-
ers and parents reported on incarceration-╉related risk events (e.g., exposure to parental crimi-
nal activity). The results specifically pertaining to emotion regulation indicated that children’s
exposure to incarceration-╉related events was associated with poorer emotion regulation, even
after controlling for children’s age and self-╉reported negative life events. Interestingly, emotion
regulation was not significantly correlated with measures of psychological maladjustment but this
may have been due to the relatively small sample size as the correlations were in the expected
direction, albeit non-╉significant, for externalizing types of problems. These findings illuminate
the important association between children’s exposure to incarceration-╉specific experiences that
are associated with poorer emotion regulation and increased psychological maladjustment above
socio-╉demographic risk factors.

Cross-sectional and longitudinal examination


The potentially protective effects of emotion regulation coping in children with incarcerated moth-
ers was examined using cross-╉sectional and longitudinal designs (Dallaire, Zeman, Borowski, &
Poon, 2014). At Time 1, 154 children (M age = 9.8 years, range: six to 13 years; 54% boys, 62%
African American), and their caregivers (62% grandparents) and incarcerated mothers partici-
pated. Children completed the Children’s Emotion Management Scales (Zeman, Shipman, &
Penza-╉Clyve, 2001; Zeman, Cassano, Suveg, & Shipman, 2010) to assess the adaptive regulation of
anger, sadness, and worry. Caregivers completed a measure of children’s internalizing and exter-
nalizing behaviors. Mothers reported on their children’s exposure to an index of incarceration-╉
specific risk experiences (ISRE; Dallaire, Zeman, & Thrash, 2015). At Time 1, regression analyses
indicated that anger regulation coping predicted fewer internalizing and externalizing behav-
ior problems. Interestingly, sadness and worry regulation coping, and a global emotion coping
scale (combination of anger, sadness, and worry coping scales) were not significant predictors of
psychological functioning. The ISRE also was a significant predictor of externalizing problems
but did not moderate the relation with anger regulation coping. The relations between emotion
regulation and psychological functioning were then examined at Time 2, 18 months later, with
83 families participating from the original sample (M child age = 11.3 years). During this time
period, 43% of children’s mothers had been re-╉arrested and 58% had been re-╉incarcerated. There
was a significant effect of Time 1 global emotion coping as well as anger regulation coping predict-
ing externalizing problems at Time 2 but no significant effect for internalizing problems. Maternal
recidivism predicted externalizing behavior problems and the effect for internalizing problems
approached significance.
2
6
3

362 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

These findings point to the potential protective effect of anger regulation in this high-╉risk
sample of children. That is, adaptive regulation of anger not only predicted fewer externalizing
difficulties concurrently but also longitudinally. It is interesting that the regulation of anger but
not the regulation of the internalizing types of emotions (i.e., sadness, worry) predicted fewer
behavioral problems. Understanding the mechanisms underlying how children of incarcerated
parents learn adaptive emotion regulation skills is of key importance as this information could be
used in prevention and intervention programs. That is, it would be important to examine whether
children with stronger emotion regulation skills have: 1) More secure attachment relationships
with their incarcerated parent and/╉or caregiver thus setting the stage for more adaptive emotional
development, 2) experienced more effective and adaptive emotion socialization efforts at home by
their parent(s) or caregivers, or in the community (e.g., teachers, religious figures, sports coaches),
3) have been exposed to fewer numbers of risk variables, and/╉or (d) other variables that could
account for the stronger anger regulation skills.

Observations of emotion regulation and psychological outcomes


Poon, Zeman, Dallaire, and Sanders (2012) used observational methods to evaluate children’s
emotion regulation skills. The sample was the same as described in Dallaire et al. (2015), except
that 15 children did not have complete data for the behavioral task due to technical glitches or
noncompliance. The final sample was comprised of 139 children with a currently incarcerated
mother (M age  =  9.4  years, range:  six to 13  years; 54% boys, 62% African American) with 76
children between the ages of six to nine years and 66 youth in the pre-╉adolescent stage (ages
ten to 13 years). Caregivers (n = 118) and 117 incarcerated mothers also participated. Mothers
completed the Children’s Emotion Management Scales (Zeman et al., 2001, 2010) to assess their
child’s inhibition, dysregulation, and adaptive coping with anger, sadness, and worry. Caregivers
and mothers reported on their child’s psychological functioning and children reported on their
symptoms of depression and anxiety. Children’s emotion regulation was assessed using the classic
disappointment display rule paradigm (Cole, 1986; Saarni, 1979, 1984). Toward the beginning of
the 60-╉minute research protocol, children were presented with eight gifts that they rank ordered
from desirable (e.g., colored markers) to undesirable (e.g., binder clip). Children were then told
that they would receive their preferred toy later in the session. Halfway through the session, the
child was presented with his or her second least preferred gift. The interviewer then “realized”
her mistake and gave the child his or her first-╉choice gift. The child was then debriefed. A secure
laptop with a webcam video-╉recorded children’s responses when they received the undesirable
and desirable gifts. The first ten seconds of their videotaped response were coded by two graduate
students.
Use of a display rule was defined as inhibiting the expression of a negative feeling (e.g., dis-
appointment, anger) by substituting a positive affect display (e.g., smiling, saying thank you).
Responses were coded as indicating the presence, absence, or an attempt to use a display rule. The
responses were also coded based on Saarni’s (1984) four categories of expressive behavior: Positive
(e.g., smiling, saying “thank you”), negative (e.g., knit brow, negative comments), tension (e.g., lip
biting, nervous blinking), and social monitoring responses (e.g., starting at researchers, mum-
bling “thank you”).
Regression analyses examining age, gender, and ethnicity differences yielded three main effects
such that older children, girls, and Caucasian children were more likely to utilize display rules,
although the ethnicity finding was marginally significant. Using analyses of variance, age, gender,
and ethnicity differences in the frequency of positive, negative, social monitoring, and tension-╉
related behaviors were examined. Two main effects emerged in which younger children and boys
exhibited higher frequencies of negative behaviors than older children and girls. A  marginally
3
6

Emotion processes 363

significant trend indicated that non-╉ Caucasian children exhibited more social monitoring
behaviors than Caucasian children. The use of display rules was examined in relation to other
emotion management strategies. Children who used more anger and sadness coping responses
were observed to use more display rules suggesting a higher level of emotional competence.
Interestingly, children who inhibited their anger, sadness, and worry were less likely to use a dis-
play rule. Although this may seem counter-╉intuitive, display rules are considered to be adaptive,
emotionally competent behaviors that are sensitive to the demands of the social context whereas
inhibition is a global strategy of suppression of negative emotion that has been linked to negative
outcomes (Gross & Levenson, 1997). Thus, the inverse relations found between inhibition and
display rule use found in this study are consistent with the emotion regulation literature.
Finally, display rule use was examined in relation to psychological functioning and found to be
significantly negatively correlated with social problems, and marginally negatively correlated with
externalizing symptomatology. The number of positive behaviors displayed (another indication
of display rule use) was significantly positively associated with self-╉reported anxiety, whereas, the
number of negative behaviors displayed was marginally positively correlated with self-╉reported
depressive symptoms. Taken together, the findings from this study are consistent with those found
in the literature that has used middle-╉class, Caucasian samples (Cole, 1986; McDowell & Parke,
2000; Saarni, 1984). However, the patterns of findings relating display rule use to behavioral prob-
lems was not as robust as indicated in other literature. Research indicates that norms for emo-
tional responsiveness are sensitive to cultural variations (Cole, Tamang, & Shrestha, 2006; Markus
& Kitayama, 1994). Thus, the violation of norms held in Caucasian, middle-╉class samples (e.g.,
smile when you get a present you do not like) may not yield the same psychological or behavioral
outcomes in non-╉Caucasian samples, particularly for those children who live in high-╉risk contexts
characterized by parental incarceration.

Maternal socialization of emotion regulation


In an effort to investigate one possible mechanism underlying the development of adaptive
emotion regulation skills, Zeman, Dallaire, and Borowski (2016) examined children’s percep-
tions of their incarcerated mother’s responses to their expressions of sadness and anger. Using
the sample described above in Dallaire et  al. (2015), 154 children, their currently incarcerated
mothers, and the children’s caregivers participated in individual interviews. To assess perceived
maternal socialization of emotion, children answered the question: “If your mom saw you look-
ing (sad, angry), what would she do?” Children’s responses were coded based on the socializa-
tion responses described in the Coping with Children’s Negative Emotions Scales (Fabes, Poulin,
Eisenberg, & Madden-╉Derdich, 2002). The two mostly highly endorsed categories for sadness
and for anger were used in the analyses including Problem-╉focused responses (e.g., help solve the
problem), Emotion-╉focused responses (e.g., help alleviate distress through comfort) for sadness,
and Problem-╉focused responses and Negative Reactions (e.g., combination of punitive, minimiza-
tion and neglect responses) for anger. Regarding outcomes, children reported depressive symp-
toms and their participation in risky, externalizing behaviors. Mothers completed the Emotion
Regulation Checklist (Shields & Cicchetti, 1997) to evaluate adaptive and maladaptive emotion
regulation, and a measure of their child’s psychological behavioral problems. Caregivers provided
information to derive the ISRE (Dallaire et al., 2015), which was used as a moderator.
The results indicated that at high levels of incarceration-╉specific risk, maternal socialization
responses did not operate in the same way as has been summarized in the literature that primarily
uses Caucasian, middle-╉class samples of children (e.g., Fabes et al., 2002). Specifically, for sadness,
at high levels of incarceration-╉specific risk, maternal emotion-╉focused responses predicted poorer
emotion regulation, more emotional lability, more depressive symptoms, and more psychological
4
6
3

364 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

problems. Interestingly, there were no significant findings for anger. Taken together, these find-
ings point to the importance of discarding assumptions and conclusions about emotion processes
based on data using low risk samples when investigating processes among unique samples such
as children with an incarcerated parent. It also appears likely that children in this sample have a
different emotion socialization history than children in low risk environments which may result
in their developing emotion regulation skills that may not lead to adaptive outcomes.
In sum, these five studies represent the scant knowledge base available on emotion regulation
and associated processes in children of incarcerated parents. Notably, four of these studies have
focused on children with an incarcerated mother and thus, it is not clear whether differences
would exist in emotion regulation processes for children of incarcerated fathers or for children
who have both parents incarcerated. Clearly, this field is ripe for additional inquiry, as the protec-
tive effects of children’s adaptive emotion regulation appear to be a promising avenue of investiga-
tion, with important implications for intervention programs.

Limitations and future research directions


Given the sociological and public health importance of understanding the effects of parental
incarceration on children’s and adolescents’ functioning, it is surprising that the body of research
examining this topic is sparse, although there has been steady growth in published work during
the past decade (Johnson & Easterling, 2012). The field of developmental research examining
emotion regulation in children living in more typical developmental contexts is still relatively
new with a surge of interest and publication activity only witnessed since 2000 (Adrian, Zeman,
& Veits, 2011). Thus, it is not surprising that the investigation of emotion regulation processes
in a highly unique sample, such as children of incarcerated parents is lagging behind the general
trends noted in the emotion and parental incarceration literatures. Further, conducting research
with this sample has a unique set of methodological and logistical challenges that provide numer-
ous avenues for future research. Five such areas of limitation and challenge are discussed next.
First, much of the past research has used either small sample sizes or large community-╉based
databases which each incur their own set of limitations. One of the striking aspects of conducting
research using incarcerated parent samples is the wide diversity inherent in this population. For
example, incarcerated parents differ in their history of crime and incarceration with a wide range
in the number of times they have been incarcerated, the type of facility in which they are held (jail
vs. prison) which has implications for child visitation, the degree of contact with their child prior
to and during incarceration, the amount of exposure the child had to the parent’s criminal activi-
ties and so forth. Thus, the use of small samples inherently poses difficulties with generalization of
results unless the samples are selected carefully based on specific criteria that are communicated
clearly to other researchers. Further, the type of statistical analyses that can be conducted with
small samples is limited. The use of large databases avoids some of the generalizability issues but
the questions that can be answered are limited by the data that were collected, the methods used,
and the length of time since the inception of the study given that some measures and methods may
be outdated. Another issue related to generalizability is the use of datasets from different countries
in which the practices related to incarceration, criminality, and delinquency differ in important
ways. Much of the research examining the effects of paternal incarceration on sons’ function-
ing has relied on the Cambridge Study in Delinquent Development (Murray & Farrington, 2005;
Murray, Janson, & Farrington, 2007), yet this database relies on information collected during the
1960s in England and focuses solely on males. Given the significantly higher incarceration rate in
the U.S. than England, and different sociocultural mores that may affect youth functioning, the
findings from these studies may not generalize to American samples. The same argument can be
5
6
3

Limitations and future research directions 365

made for findings emerging from the use of American databases, such as the Fragile Families and
Child Wellbeing Study (Wildeman, 2010) with respect to their applicability to other nations. Thus,
future research endeavors examining the effects of parental incarceration on children’s emotional
and psychological functioning must carefully report the demographics of their samples, elucidate
the areas in which generalizability may be limited, and may be sensitive to cultural differences.
Second, one of the most difficult issues facing researchers is that of selection bias (Johnson
& Easterling, 2012). That is, children of an incarcerated parent differ from children without an
incarcerated parent on numerous factors other than the incarceration dimension. Thus, differ-
ences observed between incarcerated vs. non-​incarcerated groups of individuals may not only
be due to the incarceration itself but could also be due to a host of other factors (e.g., poverty,
poor parenting, prior psychopathology). Some research has not employed comparison groups to
determine whether maladaptive outcomes are due primarily to incarceration-​specific experiences
or whether they are the result of children living in environments characterized by high levels of
stress and disadvantage that may explain differences seen between children of incarcerated vs.
non incarcerated parents. Along these lines, it is very challenging to determine the appropriate
comparison group. For example, does parental absence due to incarceration yield more deleteri-
ous outcomes than parental absence due to other family variables, such as divorce or military
deployment? Should the comparison group focus on controlling for living environments (e.g.,
high stress, parenting characteristics, housing, access to medical care) that approximate those of
children with an incarcerated parent? Researchers have responded to these challenges in a variety
of ways. Some research has not included a comparison group but instead examined high and low
exposure to incarceration-​specific risk experiences within a sample (e.g., Zeman et  al., 2016).
Other research has used large-​scale databases and created comparisons between children with
and without an incarcerated parent (e.g., Murray et al., 2007) or with various types of parental
separation due to death, hospital, incarceration, or family discord (Murray & Farrington, 2005).
Still, others have collected data from community-​based samples to create multiple comparison
groups based on different histories of parental separation (e.g., Dallaire & Zeman, 2013). Other
research has not used a control group (e.g., Lotze et al., 2010). Although the decision about the
nature of the comparison group(s) is ideally guided by the research question being investigated,
practical considerations in this research field also play an important role.
A third limitation concerns the reliance on self-​report and on the use of paper and pencil mea-
sures. Although self-​report is often the appropriate source when obtaining information about
private, internal processes (Zeman, Klimes-​Dougan, Cassano, & Adrian, 2007), having multiple
sources of information on the children’s and the incarcerated parent’s functioning adds strength
to the validity of the findings and the conclusions that can be drawn (Adrian et al., 2011). Further,
relying only on the incarcerated parents’ report of their incarceration history and the effects of
their incarceration on family members may yield a biased perspective due to social desirability,
retrospective memory concerns, and possible malingering (Houck & Loper, 2002). Further, due
to a potentially erratic history of parental contact with the child before and during incarceration,
parents’ report of their child’s functioning may not be accurate. The challenges of incorporating
observational methods, often considered the gold standard in developmental research with young
children and those in elementary school, pose considerable logistics issues, particularly if the
observation is of an interaction between the child and the incarcerated parent. Many children
have limited or no visitation with their incarcerated parent (Poehlmann-​Tynan et al., 2015), thus,
limiting the opportunity for incorporating observational methods. Further, the type of physical
arrangement in the jail for parent-​child visitation may preclude video-​or audio-​taping (e.g., plexi-​
glass barrier), although Poehlmann-​Tynan and colleagues (2015) were able to conduct an obser-
vational study investigating children’s responses to visitation. Thus, it is evident why research to
63

366 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

date has relied primarily on self-​or parent-​report using survey methods, but researchers should
endeavor to find creative ways to obtain data that provides a more complete perspective. For
example, using momentary ecological assessment tools might provide real time indicators of chil-
dren’s and caregivers’ emotional reactions and regulatory efforts when faced with exposure to
stressors in the environment. Further, the use of psychophysiological indicators of stress and emo-
tional reactivity would also supplement and add validity to the findings derived from self-​and
other-​report concerning emotional processes.
Fourth, greater attention to the moderating effect of parent and child gender is warranted.
Regarding the importance of examining gender, much of the literature has examined parental
incarceration without taking into consideration how maternal vs. paternal incarceration may
exert unique effects on children. It is likely that sons and daughters may respond differently to
the absence of a same-​vs. opposite-​sex parent. A  body of research has indicated the negative
outcomes for sons of incarcerated fathers including antisocial behavior as well as internalizing
problems (Besemer, van der Geest, Murray, Bijleveld, & Farrington, 2011; Murray & Farrington,
2005, 2008). However, some of this research has not included girls in the samples and thus, it is
not known if the antisocial outcomes are specific to all children or just boys. Using the National
Longitudinal Survey of Youth, Huebner and Gustafson (2007) identified 31 children with incar-
cerated mothers and compared this sample to 1666 children with no incarcerated mother. They
found that 26% of the children with incarcerated mothers were convicted in adulthood vs. 10%
in the comparison group. However, research has yet to make direct comparisons within the same
sample of the effects of maternal vs. paternal imprisonment on sons vs. daughters. Further, the
research examining emotion regulation in incarcerated parent samples has primarily investigated
these processes in children with incarcerated mothers, yet the emotion development literature
indicates the importance of fathers as well as mothers in emotion socialization processes (e.g.,
Cassano & Zeman, 2010; Cassano, Perry-​Parrish, & Zeman, 2007; Lunkenheimer et al., 2007).
Relatedly, little research has considered the role of the caregiver in helping to ameliorate some
of the negative sequalae associated with parental incarceration (Cecil, McHale, Strozier, & Pietsch,
2008). For example, children may have a more consistent relationship with the caregiver such
as a grandparent when their mother is incarcerated. The nature of this relationship is key to
understanding how some children with an incarcerated parent display more resilience than other
children. For example, Mackintosh et al. (2006) found that when children perceived that their
caregivers responded to them with warmth and support, fewer behavior problems emerged.
Fifth, the negative effects of incarceration on emotional development may also differ depending
on the developmental timing of the child’s exposure to and experience of parental separation due
to incarceration. It may be that there are sensitive periods for emotional development in which
parental separation is particularly pernicious. For example, the first 12–​18 months of age are con-
sidered critical for the development of attachment (Sroufe, 2005) and thus, maternal absence, in
particular, at this stage could have detrimental effects that could have far reaching implications for
all spheres of development. Although parental socialization of emotion is important through all
stages of childhood and adolescence (Klimes-​Dougan & Zeman, 2007), the groundwork for later
emotional development occurs during the toddler and preschool years (Denham, 1998). Thus,
parental, and particularly maternal absence during these years may have a lasting negative impact
on emotional development. Further, having poor emotion socialization models during the early
childhood years also has been shown to lead to negative psychological, social, and academic out-
comes (Zeman et al., 2006). Given the difficulty with recruiting samples of youth with an incarcer-
ated parent, many studies have used wide age ranges in their samples, which are not of sufficient
size to allow for analyses by child age or developmental stage in order to answer these questions.
Even the landmark ACEs study (Anda et al., 2001) did not take into account the age of youth at
7
6
3

Summary and conclusion 367

time of parental imprisonment. Thus, future research needs to carefully consider the role of devel-
opmental status when investigating the effect of parental incarceration on children’s functioning.

Summary and conclusion


The effect of parental incarceration on children’s adjustment is a topic of societal importance given
the high rates of incarceration in the US in particular, and the proportion of inmates who are
parents. President Obama (2015) recently tweeted: “If we make investments early in our children,
we will reduce the need to incarcerate those kids. One study found that for every dollar we invest
in pre-╉K, we save at least twice that in reduced crime. We recognize that every child deserves
opportunity. Not just some. Not just our own.” A growing body of research clearly indicates that
children of incarcerated parents are at risk for negative psychological socio-╉emotional, educa-
tional, and health outcomes (Anda et al., 2001). Thus, understanding the processes that may lead
to resilience in this population is of key importance so that the potential negative effects can be
mitigated, and the intergenerational transmission of antisocial outcomes disrupted.
One potential area for preventive intervention is in the domain of emotional development. Poor
emotion management skills including emotion dysregulation have been linked to most forms of
psychopathology as well as poorer health, educational, and social outcomes (e.g., Bradley, 2003,
Cicchetti et  al., 1995; Keenan, 2000; Zeman et  al., 2013). The few studies conducted examin-
ing emotion regulation skills in children of incarcerated parents (mostly incarcerated mothers)
provides preliminary evidence that those children who evidence more adaptive emotional devel-
opment skills exhibit fewer negative psychological outcomes. Although this body of research is
small, these preliminary findings identify a pressing need to more fully investigate the types of
emotional processes that may be critical for children living in the high-╉risk context of parental
incarceration. Although it is important to document the areas in which children of an incar-
cerated parent may evidence less well developed skills in specific emotional competencies (e.g.,
emotion understanding, emotion identification), the focus of research should also include under-
standing the mechanisms that underlie children’s under-╉developed or maladaptive emotion skills,
and developing methods for bolstering these skills. Specifically, future research should examine
how key socialization figures (e.g., caregivers, teachers, peers, siblings) socialize or teach children
how to manage their negative and positive emotions in ways that are sensitive to the demands in
different social contexts. Intervention programs that help children and caregivers learn to express
and manage emotions constructively during high intensity emotional situations (e.g., visitations,
family reunification) may help children to learn adaptive coping methods that can help them
more successfully navigate the challenges present in their environments.

References
Adrian, M., Zeman, J., & Veits, G. (2011). Methodological implications of the affect revolution: A 35-╉year
review of emotion regulation assessment in children. Journal of Experimental Child Psychology, 110,
171–╉197. doi:10.1016/╉j.jecp.2011.03.009
Aldao, A. (2013). The future of emotion regulation research: Capturing context. Perspectives on
Psychological Science, 8, 155–╉172. doi:10.1177/╉1745691613504116
Anda R. F., Felitti V. J., Chapman D. P., Croft J. B., Williamson, D. F., Santelli, J., … Marks, J. S. (2001).
Abused boys, battered mothers, and male involvement in teen pregnancy. Pediatrics, 107, e19.
doi:10.1542/╉peds.107.2.e19
Arditti, J. A., Lambert-╉Shute, J., & Joest, K. (2003). Saturday morning at the jail: Implications
of incarceration for families and children. Family Relations, 52, 195–╉204. doi:10.1111/╉
j.1741-╉3729.2003.00195.x
8
6
3

368 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

Armsden, G. C. (1986). Attachment to parents and peers in late adolescence: Relationship to affective


status, self-​esteem, and coping with loss, threat and challenges (Doctoral dissertation). Retrieved from
Dissertation Abstracts International. (1987-​54327-​001).
Armsden, G. C., & Greenberg, M. T. (1987). The inventory of parent and peer attachment: Relationships to
well-​being in adolescence. Journal of Youth and Adolescence, 16, 427–​454. doi:10.1007/​BF02202939
Besemer, S., van der Geest, V., Murray, J., Bijleveld, C. C. J. H., & Farrington, D. P. (2011). The relationship
between parental imprisonment and offspring offending in England and the Netherlands. British
Journal of Criminology, 51, 413–​437. doi:10.1093/​bjc/​azq072
Blair, C. B. (2002). School readiness: Integrating cognition and emotion in a n neurobiological
conceptualization of children’s functioning at school entry. American Psychologist, 57, 111–​127.
doi:10.1080/​15377900903379125
Bradley, S. J. (2003). Affect regulation and the development of psychopathology. New York: Guilford.
doi:10.1097/​00004583-​200112000-​00022
Bretherton, I., Ridgeway, D., & Cassidy, J. (1990). Assessing internal working models of the attachment
relationship: An attachment story completion task for 3-​year-​olds. In M. T. Greenberg, D. Cicchetti, &
E. M. Cummings (Eds.), Attachment in the preschool years: Theory, research, and intervention (pp. 273–​
308). Chicago: University of Chicago Press. doi:10.1037/​0012-​1649.28.5.759
Bureau of Justice Statistics. (2015). FAQ and Detail. Retrieved Sept. 13, 2015 from http://​www.bjs.gov/​
index.cfm?ty=qa&iid=322
Campos, J. J., Frankel C. N., & Camras, L. (2004). On the nature of emotion regulation. Child Development,
75, 377–​394. doi:0009-​3920/​2004/​7402-​0010
Cassano, M., & Zeman, J. (2010). The role of parental expectations and gender in the socialization of
sadness regulation in middle childhood. Developmental Psychology, 46, 1214–​1226. doi:10.1037/​
a0019851
Cassano, M., Zeman, J., & Perry-​Parrish, C. (2007). Influence of gender on parental socialization of
children’s sadness regulation. Social Development, 16, 210–​231. doi:10.1111/​j.1467-​9507.2007.00381.x
Cassidy, J. (1994). Emotion regulation: Influences of attachment relationships. Monographs of the Society for
Research in Child Development, 59, 228–​249. doi:10.1111/​j.1540-​5834.1994.tb01287.x
Cassidy, J., Ziv, Y., Cooper, G., Hoffman, K. T., Powell, B., & Karfgin, A. (2007, April). TAMAR’s
Children: Enhancing attachment security in the infants of women in a jail-​diversion program. In
J. Poehlmann (Chair), Incarcerated mothers, their children, and families: Attachment and parenting.
Symposium at the biennial meeting of the Society for Research in Child Development, Boston, MA.
Cecil, D. K., McHale, J., Strozier, A., & Pietsch, J. (2008). Female inmates, family caregivers, and young
children’s adjustment: A research agenda and implications for corrections programming. Journal of
Criminal Justice, 36, 513–​521. doi:10.1016/​j.jcrimjus.2008.09.002
Chaplin, T. M. (2006). Anger, happiness, and sadness: Associations with depressive symptoms in late
adolescence. Journal of Youth and Adolescence, 35, 977–​986. doi:10.1007/​s10964-​006-​9033-​x
Chapman, D. P., Whitfield, C. L., Felitti, V. J., Dube, S. R., Edwards, V. J., & Anda, R. F. (2004). Adverse
childhood experiences and the risk of depressive disorders in adulthood. Journal of Affective Disorders,
82, 217–​225. doi:10.1016/​j.jad.2003.12.013
Cicchetti, D., Ackerman, B. P., & Izard, C. E. (1995). Emotions and emotion regulation in developmental
psychopathology. Development and Psychopathology, 7, 1–​10. doi:10.1017/​s0954579400006301
Clear, T. R. (2008). Communities with high incarceration rates. Crime & Justice: A Review of Research, 37,
97–​132.
Cole, P. (1986). Children’s spontaneous control of facial expression. Child Development, 57, 1309–​1321.
doi:10.2307/​1130411
Cole, P. M., Martin, S. E., & Dennis, T.A. (2004). Emotion regulation as a scientific construct: Challenges
and directions for child development research. Child Development, 75, 317–​333. doi:0009-​3920/​2004/​
7502-​0002
9
6
3

Summary and conclusion 369

Cole, P. M., Tamang, B. L., & Shrestha, S. (2006). Cultural variations in the socialization of young children’s
anger and shame. Child Development, 77, 1237–​1251. doi: 10.1111/​j.1467-​8624.2006.00931.x
Dallaire, D. H. (2007). Incarcerated mothers and fathers: A comparison of risks for children and families.
Family Relations, 56, 440–​453. doi:10.1111/​j.1741-​3729.2007.00472.x
Dallaire, D. H., Ciccone, A., & Wilson, L. (2012). The family drawings of at-​risk children: Concurrent
relations with contact with incarcerated parents, caregiver behavior and stress. Attachment & Human
Development, 14, 161–​183. doi:10.1080/​14616734.2012.661232
Dallaire, D. H., & Wilson. L. (2010). The impact of exposure to parental criminal activity, arrest, and
sentencing on children’s academic competence and externalizing behavior. Journal of Child and Family
Studies, 19, 404–​418. doi:10.1007/​s10826-​009-​9311-​9
Dallaire, D., & Zeman, J. (2013). Empathy as a protective factor for children with incarcerated parents. In
J. Poehlmann & J. M. Eddy (Eds.), Relationship processes and resilience in children with incarcerated
parents. Monographs of the Society for Research in Child Development, 78, 7–​25. doi:10.1111/​
mono.12018
Dallaire, D., Zeman, J., Borowski, S., & Poon, J. (2014, March). Emotion coping strategies and externalizing
behaviors in youth with an incarcerated mother. C. Cavanaugh (Chair). Families behind bars: Familial
justice system contact and adolescent development. Talk presented at the biennial meeting of the
Society for Research in Adolescence. Austin, TX.
Dallaire, D., Zeman, J., & Thrash, T. (2015). Children’s experiences of maternal incarceration-​specific
risks: Predictions to psychological maladaptation. Journal of Clinical Child and Adolescent Psychology,
44, 109–​122. doi:10.1080/​15374416.2014.913248
Denham, S. A. (1998). Emotional development in young children. New York, NY: Guilford Press.
Denham, S. A., Bassett, H. H., & Wyatt, T. (2007). The socialization of emotional competence. In J. Grusec
& P. Hastings (Eds.), The handbook of socialization (pp. 614–​637). New York: Guilford Press.
Dube, S. R., Anda, R. F., Felitti, V. J., Chapman, D. P., Williamson, D. F., & Giles, W. H. (2001). Childhood
abuse, household dysfunction, and the risk of attempted suicide throughout the life span: Findings
from the adverse childhood experiences study. Journal of the American Medical Association, 286, 3089–​
3096. doi:10.1001/​jama.286.24.3089
Dunn, J., & Brown, J. (1994). Affect expression in the family, children’s understanding of emotions, and
their interactions with others. Merrill-​Palmer Quarterly, 40, 120–​137.
Ellis, L. K., & Rothbart, M. K. (1999). Early Adolescent Temperament Questionnaire—​Revised. Available
from Mary Rothbart, University of Oregon, Maryroth@oregon.uoregon.edu.
Fabes, R. A., Poulin, R. E., Eisenberg, N., & Madden-​Derdich, D. A. (2002). The Coping with Children’s
Negative Emotions Scale (CCNES): Psychometric properties and relations with children’s emotional
competence. Marriage & Family Review, 34, 285–​310. doi:10.1300/​J002v34n03_​05
Felitti, M. D., Vincent, J., Anda, M. D., Robert, F., Nordenberg, M. D., Williamson, M. S., … & James,
S. (1998). Relationship of childhood abuse and household dysfunction to many of the leading causes
of death in adults: The Adverse Childhood Experiences (ACE) Study. American Journal of Preventive
Medicine, 14, 245–​258. doi:10.1016/​S0749-​3797(98)00017-​8
Fury, G., Carlson, E., & Sroufe, L.A. (1997). Children’s representations of attachment relationships in
family drawings. Child Development, 68, 1154–​1164. doi:10.1111/​j.1467-​8624.1997.tb01991.x
Gjelsvik, A., Dumont, D. M., & Nunn, A. (2013). Incarceration of a household member and Hispanic
health disparities: Childhood exposure and adult chronic disease risk behaviors. Preventing Chronic
Disease, 10, E69. doi:10.5888/​pcd10.120281
Gjelsvik A., Dumont D. M., Nunn A., & Rosen D. (2014). Adverse childhood events: Incarceration of
household members and health-​related quality of life in adulthood. Journal of Health Care for the Poor
and Underserved, 25, 1169–​1182. doi: 10.1353/​hpu.2014.0112
Glaze, L. E., & Maruschak, L. M. (2008). Bureau of Justice statistics special report: Parents in prison and their
minor children. Washington, DC: Bureau of Justice Statistics.
0
7
3

370 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

Gottman, J. M., Katz, L. F., & Hooven, C. (1997). Meta-​emotion: How families communicate emotionally.
Mahwah, NJ: Erlbaum.
Gross, J. J., & Levenson, R. W. (1997). Hiding feelings: The acute effects of inhibiting negative and positive
emotion. Journal of Abnormal Psychology, 106(1), 95. doi:10.1037/​0021-​843X.106.1.95
Gross, J. J., & Muñoz, R. F. (1995). Emotion regulation and mental health. Clinical Psychology: Science &
Practice, 2, 151–​164. doi:10.1111/​j.1468-​2850.1995.tb00036.x
Gumora, G., & Arsenio, W. F. (2002). Emotionality, emotion regulation, and school performance in middle
school children. Journal of School Psychology, 40, 395–​413. doi:10.1016/​S0022-​4405(02)00108-​5
Hanlon, T. E., Blatchley, R. J., Bennett-​Sears, T., O’Grady, K. E., Rose, M., & Callaman, J. M. (2005).
Vulnerability of children of incarcerated addict mothers: Implications for preventive intervention.
Children and Youth Services Review, 27, 67–​84.
Harcourt, B. (2011). The illusion of free markets: Punishment and the myth of natural order. Cambridge,
MA: Harvard University Press.
Haskins, A. R. (2014). Unintended consequences: Effects of paternal incarceration on child school
readiness and later special education placement. Sociological Science, 1, 141–​157. doi:10.15195/​v1.a11
Hatzenbuehler, M. L., Keyes, K., Hamilton, A., Uddin, M., & Galea, S. (2015). The collateral damage
of mass incarceration: Risk of psychiatric morbidity among nonincarcerated residents of high-​
incarceration neighborhoods. American Journal of Public Health, 105, 138–​143. doi:10.2105/​
AJPH.2014.302184
Houck, K., & Loper, A. B. (2002). The relationship of parenting stress to adjustment among mothers in
prison. American Journal of Orthopsychiatry, 72, 548–​558. doi:10.1037/​0002-​9432.72.4.548
Huebner, B. M., & Gustafson, R. (2007). The effect of maternal incarceration on adult offspring
involvement in the criminal justice system. Journal of Criminal Justice, 35, 283–​296.
Izard, C., Fine, S., Schultz, D., Mostow, A., Ackerman, B., & Youngstrom, E. (2001). Emotion knowledge
as a predictor of social behavior and academic competence in children at risk. Psychological Science, 12,
18–​23. doi:10.1111/​1467-​9280.00304
James, D. J., & Glaze, L. E. (2006). Bureau of Justice statistics special report: Mental health problems of prison
and jail inmates. Washington, DC: Bureau of Justice Statistics.
Johnson, E. I., & Esterling, B. (2012). Understanding unique effects of parental incarceration on
children: Challenges, progress, and recommendations. Journal of Marriage and Family, 74, 342–​356.
doi:10.1111/​j.1741-​3737.2012.00957.x
Johnson E. I., & Waldfogel, J. (2002). Parental incarceration: Recent trends and implications for child
welfare. Social Service Review, 76, 460–​479. doi:10.1086/​341184
Justice Policy Institute (2002). Cellblocks or classrooms? The Funding of higher education and corrections
and tis impact on African American men. Retrieved Sept. 13, 2015 from http://​www.justicepolicy.org/​
uploads/​justicepolicy/​documents/​02-​09_​rep_​cellblocksclassrooms_​bb-​ac.pdf
Keenan, K. (2000), Emotion dysregulation as a risk factor for child psychopathology. Clinical
Psychology: Science and Practice, 7, 418–​434. doi:10.1093/​clipsy.7.4.418
Kerns, K. A. (2008). Attachment in middle childhood. In J. Cassidy & P. R. Shaver (Eds), Handbook of
Attachment (2nd edition, pp. 366–​382). New York, NY: Guilford Press.
Klimes-​Dougan, B., & Zeman, J. (2007). Introduction to the Special Issue of Social Development: Emotion
socialization in childhood and adolescence. Social Development, 16, 203–​209.
Kring, A. M., & Sloan, D. S. (2010). Emotion regulation and psychopathology. New York: Guilford Press.
Legerski, J. P., Biggs, B. K., Greenhoot, A. F., & Sampilo, M. L. (2015). Emotion talk and friend responses
among early adolescent same-​sex friend dyads. Social Development, 24, 20–​38. doi:10.1111/​sode.12079
Lengua, L. J. (2002). The contribution of emotionality and self-​regulation to the understanding of children’s
response to multiple risk. Child Development, 73, 144–​161. doi:10.1111/​1467-​8624.00397
Lotze, G. M., Ravindran, N., & Myers, B. J. (2010). Moral emotions, emotion self-​regulation, callous-​
unemotional traits, and problem behavior in children of incarcerated mothers. Journal of Child and
Family Studies, 19, 703–​713. doi:10.1007/​s10826-​010-​9358-​7
1
7
3

Summary and conclusion 371

Lunkenheimer, E. S., Shields, A. M., & Cortina, K. S. (2007). Parental emotion coaching and dismissing in
family interaction. Social Development, 16, 232–​248. doi:10.1111/​j.1467-​9507.2007.00382.x
Mackintosh, V. H., Myers, B. J., & Kennon, S. S. (2006). Children of incarcerated mothers and their
caregivers: Factors affecting the quality of their relationship. Journal of Child and Family Studies, 15,
579–​594. doi:10.1007/​s10826-​006-​9030-​4
Markus, H. R., & Kitayama, S. (1994).The cultural construction of self and emotion: Implications for
social behavior. In S. Kitayama & H. R. Markus (Eds.), Emotion and culture: Empirical studies of mutual
influence (pp. 89–​130). Washington, DC: American Psychological Association. doi:10.1037/​10152-​003
McDowell, D. J., & Parke, R. D. (2000). Differential knowledge of display rules for positive and negative
emotions: Influences from parents, influences on peers. Social Development, 9, 83–​104. doi:10.1111/​
1467-​9507.00136
Miller, K. M. (2007). Risk and resilience among African American children of incarcerated parents. Journal
of Human Behavior in the Social Environment, 15(2-​3), 25–​37. doi:10.1300/​J137v15n02_​03
Minton, T. D. (2013). Jail inmates at midyear 2012—​statistical tables. Bureau of Justice Statistics 2013.
Murray J., & Farrington D. P. (2005). Parental imprisonment: Effects on boys’ antisocial behaviour and
delinquency through the life-​course. Journal of Child Psychology and Psychiatry, 46, 1269–​1278.
doi:10.1111/​j.1469-​7610.2005.01433.x
Murray, J., & Farrington, D. P. (2008). The effects of paternal imprisonment on children. Crime and Justice,
37, 133–​206. doi:10.1086/​520070
Murray, J., Farrington, D. P., & Sekol, I. (2012).Children’s antisocial behavior, mental health, drug use,
and educational performance after parental incarceration: A systematic review and meta-​analysis.
Psychological Bulletin, 138, 175–​210. doi:10.1037/​a0026407
Murray, J., Janson, C. G., & Farrington, D. P. (2007). Crime in adult offspring of prisoners: A cross-​national
comparison of two longitudinal samples. Criminal Justice and Behavior, 34, 133–​149. doi:10.1177/​
0093854806289549
Myers, B. J., Smarsh, T., Amlund-​Hagen, K., & Kennon, S. (1999). Children of incarcerated mothers.
Journal of Child and Family Studies, 8, 11–​25.doi:10.1023/​A:1022990410036
Obama, B. (2015). http://​www.bustle.com/​articles/​97376-​the-​best-​quotes-​from-​president-​obamas-​
criminal-​justice-​reform-​speech-​that-​demanded-​a-​long-​overdue-​solution
Parke, R. D., & Clarke-​Stewart, K. A. (2001). Effects of parental incarceration on young children.
Paper presented at the National Policy Conference, From Prison to Home: The Effect of Incarceration
and Reentry on Children, Families and Communities (January 30–​31, 2002). Washington, DC: U.S.
Department of Health and Human Services, The Urban Institute.
Perry-​Parrish, C., & Zeman, J. (2011). Relations among sadness regulation, peer acceptance, and social
functioning in early adolescence: The role of gender. Social Development, 20, 135–​153. doi:10.1111/​
j.1467-​9507.2009.00568.x
Phillips, S. D., Burns, B. J., Wagner, H. R., & Barth, R. P. (2004). Parental arrest and children involved
with child welfare services agencies. American Journal of Orthopsychiatry, 74, 174. doi:10.1037/​
0002-​9432.74.2.174
Poehlmann, J. (2005a). Representations of attachment relationships in children of incarcerated mothers.
Child Development, 76, 679–​696. doi:10.1111/​j.1467-​8624.2005.00871.x
Poehlmann, J. (2005b). Children’s family environments and intellectual outcomes during maternal
incarceration. Journal of Marriage and Family, 67, 1275–​1285. doi:10.1111/​j.1741-​3737.2005.00216.x
Poehlmann, J., Dallaire, D., Loper, A. B., & Shear, L. D. (2010). Children’s contact with their incarcerated
parents: research findings and recommendations. American Psychologist, 65, 575–​598. doi:10.1037/​
a0020279
Poehlmann-​Tynan, J., Runion, H., Burnson, C., Maleck, S., Weymouth, L., Pettit, K., & Huser M. (2015).
Young children’s behavioral and emotional reactions to plexiglass and video visits with jailed parents.
In J. Poehlmann-​Tynan (Ed.), Children’s contact with incarcerated parents: Implications for policy and
intervention (pp. 39–​58). New York, NY: Springer, Inc.
2
7
3

372 Children’s and Adolescents’ Emotion Regulation in the Context of Parental Incarceration

Poon, J., Zeman, J., Dallaire, D., & Sanders, W. (2012, August). Display rule usage in children of incarcerated
mothers. Talk presented at the annual meeting of the American Psychological Association, Orlando, FL.
Saarni, C. (1979). Children’s understanding of display rules for expressive behavior. Developmental
Psychology, 15, 424–​429. doi:10.1037/​0012-​1649.15.4.424
Saarni, C. (1984). An observational study of children’s attempts to monitor their expressive behavior. Child
Development, 55, 1504–​1513. doi:0009-​3920/​84/​5504-​0002S01.00
Saarni, C. (1999). The development of emotional competence. Guilford Press.
Sameroff, A. J., Bartko, W. T., Baldwin, A., Baldwin, C., & Siefer, R. (1998). Family and social influences
on the development of child competence. In M. Lewis & C. Feiring (Eds.), Families, risk, and
competence (pp. 161–​185). Mahwah, NJ: Erlbaum.
Shlafer, R. J., & Poehlmann, J. (2010). Attachment and caregiving relationships in families affected
by parental incarceration. Attachment & Human Development, 12, 395–​415. doi:10.1080/​
14616730903417052
Shields, A., & Cicchetti, D. (1997). Emotion regulation among school-​age children: The development and
validation of a new criterion Q-​sort scale. Developmental Psychology, 33, 906–​916. doi: http://​psycnet.
apa.org/​doi/​10.1037/​0012-​1649.33.6.906
Shipman, K., & Zeman, J. (1999). Emotional understanding: A comparison of physically maltreating and
nonmaltreating mother-​child dyads. Journal of Clinical Child Psychology, 28, 407–​417. doi:10.1207/​
S15374424jccp280313
Siegel, D. J. (2007). The mindful brain: Reflection and attunement in the cultivation of well-​being. New York,
NY: W. W. Norton & Company.
Simonds, J., Kieras, J. E., Rueda, M. R., & Rothbart, M. (2007). Effortful control, executive attention, and
emotional regulation in 7-​to 10-​year-​old children. Cognitive Development, 22, 474–​488. doi:10.1016/​
j.cogdev.2007.08.009
Sroufe, L. A. (2005). Attachment and development: A prospective, longitudinal study from birth to
adulthood. Attachment & Human Development, 7(4), 349–​367. doi:10.1080/​14616730500365928
Supplee, L. H., Skuban, E. M., Shaw, D. S., & Prout, J. (2009). Emotion regulation strategies and later
externalizing behavior among European American and African American children. Development and
Psychopathology, 21, 393–​415. doi:10.1017/​S0954579409000224
Thompson, R. A. (2008). Attachment-​related mental representations: Introduction to the special issue.
Attachment and Human Development, 10, 347–​358. doi:10.1080/​14616730802461334
Thompson, R. A. (1994). Emotion regulation: A theme in search of definition. In N. A. Fox (Ed.), The
development of emotion regulation and dysregulation: Biological and behavioral aspects. Monographs
of the Society for Research in Child Development, 59, 25–​52.
Travis, J., Western, B., & Redburn, S. (2014). The growth of Incarceration in the U.S.: Exploring Causes and
Consequences. Washington DC: National Research Council.
Trentacosta, C. J., & Izard, C. E. (2007). Kindergarten children’s emotion competence as a predictor of their
academic competence in first grade. Emotion, 7, 77–​88. doi:10.1037/​1528-​3542.7.1.77
Trice, A. D., & Brewster, J. (2004). The effects of maternal incarceration on adolescent children. Journal of
Police and Criminal Psychology, 19, 27–​35. doi:10.1007/​bf02802572
Turney, K., & Wildeman, C. (2015). Detrimental for some? Heterogeneous effects of maternal incarceration
on child wellbeing. Criminology & Public Policy, 14, 125–​156. doi:10.1111/​1745-​9133.12109
Wagner, P. & Sakala, L. (2014). Mass incarceration: The whole pie. Retrieved Sept. 13, 2015 from http://​www.
prisonpolicy.org/​reports/​pie.html
Walden, T. A., Harris, V. S., & Catron, T. F. (2003). How I Feel: A self-​report measure of emotional arousal
and regulation for children. Psychological Assessment, 15, 399–​412. doi:10.1037/​1040-​3590.15.3.399
Walmsley, R. (2013). World prison population list. King’s College London: International Centre for Prison
Studies. The Institute for Democracy and Conflict Resolution website.
3
7

Summary and conclusion 373

Western, B., & Wildeman, C. (2009). The black family and mass incarceration. The ANNALS of the
American Academy of Political and Social Science, 621, 221–​242. doi:10.1177/​0002716208324850
Whitson, S., & El Sheikh, M. (2003). Moderators of family conflict and children’s adjustment and health.
Journal of Emotional Abuse: Interventions, Research and Theories of Psychological Maltreatment, Trauma
and Nonphysical Aggression, 3, 47–​73. doi:10.1300/​J135v03n01_​03
Wilbur, M. B., Marani, J. E., Appugliese, D., Woods, R., Siegel, J. A., Cabral, H. J., & Frank, D. A. (2007).
Socioemotional effects of fathers’ incarceration on low-​income, urban, school-​aged children. Pediatrics,
120(3), 678–​685. doi:10.1542/​peds.2006-​2166
Wildeman, C. (2010). Paternal incarceration and children’s physically aggressive behavior: Evidence from
the Fragile Families and Child Wellbeing Study. Social Forces, 89, 285–​309. doi:10.1353/​sof.2010.0055
Zehr, A. (2011). What Will Happen To Me? Good Books: New York City, NY.
Zeman, J., Cassano, M., & Adrian, M. (2013). Socialization influences on children’s and adolescents’
emotional self-​regulation processes: A developmental psychopathology perspective. In K. Barrett, G.
Morgan, & N. Fox (Eds.), Handbook of Self-​Regulatory Processes in Development: New Directions and
International Perspectives (pp. 79–​107). Routledge: NYC.
Zeman, J., Cassano, M., Perry-​Parrish, C., & Stegall, S. (2006). Emotion regulation in children and
adolescents. Journal of Developmental and Behavioral Pediatrics, 27(2), 155–​168. doi:10.1097/​
00004703-​200604000-​00014
Zeman, J., Cassano, M., Suveg, C., & Shipman, K. (2010). Initial validation of the Children’s Worry
Management Scale. Journal of Child and Family Studies, 19, 381–​392. doi:10.1007/​s10826-​009-​9308-​4
Zeman, J., Dallaire, D., & Borowski, S. (i2016). Maternal emotion socialization in children of incarcerated
mothers. Social Development, 25(1), 66–​81.
Zeman, J., Klimes-​Dougan, B., Cassano, M., & Adrian, M. (2007). Measurement issues in emotion
research with children and adolescents. Clinical Psychology: Science and Practice, 14, 377–​401.
doi:10.1111/​j.1468-​2850.2007.00098.x
Zeman, J., Shipman, K., & Penza-​Clyve, S. (2001). Development and initial validation of the Children’s
Sadness Management Scale. Journal of Nonverbal Behavior, 25, 187–​205. doi:10.1023/​A:1010623226626
Zeman, J., Shipman, K., & Suveg, C. (2002). Anger and sadness regulation: Predictions to internalizing and
externalizing symptomatology in children. Journal of Clinical Child and Adolescent Psychology, 31, 393–​
398. doi:10.1207/​S15374424JCCP3103_​11
4
7
3

Chapter 18

Children Exposed to Traumatic Stress


Brandon G. Scott & Carl F. Weems

Traumatic stress
Children may be exposed to several types of traumatic events across development, such as endur-
ing abuse or maltreatment at the hands of relatives, authority figures, or peers (Chapter 15), living
in negative, uncontrollable family environments (e.g., divorce, parental incarceration, bereave-
ment; Chapters 16, 17, and 19), or experiencing mass trauma due to the widespread violence, war/╉
terrorism, or natural disasters within the child’s respective community (September 11th terrorist
attacks, Hurricane Katrina). This chapter primarily focuses on this latter type of global traumatic
stress, as it impacts a large number of youth at one time and across multiple ecologies (ontogenic,
microsystem, or macrosystem; Weems & Overstreet, 2008). Research suggests that mass trauma
exposure is related to several negative posttraumatic outcomes in youth (e.g., Aber, Gershoff,
Ware, & Kotler, 2004; Eisenberg & Silver, 2011; La Greca, Silverman, Vernberg, & Roberts, 2002;
Norris, Friedman, & Watson, 2002; Osofsky, Osofsky, Kronenberg, Brennan, & Hansel, 2009;
Weems & Overstreet, 2008).
In this chapter, we draw from previous child trauma models (La Greca, Silverman, Vernberg, &
Prinstein, 1996; Lanius, Frewen, Vermetten, & Yehuda, 2010; Masten, & Narayan, 2012; Pynoos,
Steinberg, & Piacentini, 1999; Weems & Overstreet, 2008), and past research findings (e.g., Jeney-╉
Gammon et al., 1993; Kithakye et al., 2010; Marsee, 2008; La Greca et al., 1996; Punamäki et al.,
2014; Russoniello et al., 2002) to present a theoretical framework illustrating how youths’ emotion
regulation in the aftermath of natural disasters, war, or terrorism may affect posttraumatic stress
outcomes (e.g., resilience or posttraumatic stress disorder symptoms, anxiety, depression, trau-
matic grief, and aggression). Specifically, we propose that emotion regulation, at multiple levels of
analysis (neurobiological, cognitive, and behavioral), is critical to understanding youth reactions
and outcomes (both negative, such as the development of posttraumatic stress disorder [PTSD],
but also resilience) following trauma exposure. We also provide a brief overview of evidence-╉
based interventions aimed at alleviating posttraumatic stress reactions and trauma-╉related symp-
tomology highlighting techniques that focus on increasing emotional self-╉efficacy and helping
youth acquire adaptive emotion regulation skills.

Mass trauma exposure and posttraumatic stress outcomes


Traumatic stress exposure
The defining features of traumatic stress exposure and subsequent posttraumatic stress outcomes
(e.g., posttraumatic stress disorder) were recently updated and revised in the American Psychiatric
Association’s Diagnostic and Statistical Manual—╉Fifth Edition (DSM-╉5, APA 2013). The DSM-╉5
(APA, 2013) defines traumatic stress exposure as experiencing actual or threatened death, serious
injury, or sexual violation, which may result from, 1) “direct exposure” or witnessing a traumatic
5
7
3

Mass trauma exposure and posttraumatic stress outcomes 375

event in person; or 2) “indirect or remote exposure,” such as learning of a traumatic event that hap-
pened to a close family member or friend, or experiences such as first-╉hand, repeated or extreme
exposure to aversive details of a traumatic event. Note that the DSM-╉5 does not consider exposure
to traumatic events via electronic media, such as broadcast over the internet or television, as a
qualifying traumatic event. An important difference between the DSM-╉5 and its predecessor (i.e.,
Diagnostic and Statistical Manual—╉Fourth Edition; DSM-╉IV-╉TR [APA, 2000]) is that the pres-
ence or absence of an emotional reaction (e.g., fear or horror; Criterion A2 in DSM-╉IV-╉TR) does
not determine whether an event is traumatic.
Epidemiological studies in the United States have shown that 25% of children and adolescents
in the general population will likely be exposed to at least one traumatic event by the age of 16
and that up to 20% of these trauma-╉exposed children also experience academic, emotional, and
physical difficulties across development (Costello, Erkanli, Fairbank, Angold, 2002; Copeland,
Keeler, Angold, & Constello, 2007). Prevalence rates of mass trauma exposure often depend on the
type of event and geographic region (e.g., hurricanes and earthquakes have differential probability
depending on where one lives). In terms of natural disaster exposure, Becker-╉Blease, Turner, &
Finkelhor (2010) found that the lifetime risk for disaster exposure was 13.9% among a representa-
tive sample of youth (ages two to 17 years) from the United States, while another study among
1140 children in South Africa and 901 youth in Kenya showed 16% were exposed to a natu-
ral disaster (e.g., earthquake, fire, flood; Seedat, Nymami, Njenga, Vythlingum, & Stein, 2004).
Research among children living in countries that are prone to political violence or war exposure
have shown higher rates of exposure with some estimates ranging from 90% (Crotia; Kuterovac,
Dyregrov, & Stuvland, 1994) to 100% (Bosnia; Goldstein, Wampler, & Wise, 1995). Ultimately,
mass trauma exposure is potentially highly prevalent among youth and is a global phenomenon.

Posttraumatic stress outcomes


Posttraumatic stress disorder (PTSD)
The most commonly studied of the posttraumatic stress outcomes are the symptoms of PTSD
(Furr, Comer, Edmunds, & Kendall, 2010). PTSD is a complex disorder that involves chronic and
persistent heightened emotional reactivity and arousal (subjectively and physiologically) to inter-
nal and external trauma cues (e.g., faster heart rate, angry outbursts, and hypervigilance), maladap-
tive ways of regulating this emotional reactivity (e.g., cognitive or behavioral avoidance), intrusive
cognitions (e.g., memories or dreams of traumatic event) and negative cognitions and mood (e.g.,
memory loss, faulty ways of thinking, persistent negative affect and inability to experience positive
emotions; DSM-╉5; APA, 2013). Although PTSD was classified as an anxiety disorder in previous
versions of the DSM, the DSM-╉5 has relocated it to a new chapter titled trauma-╉and stress-╉related
disorders. One of the main reasons for this move was that the prominent features of PTSD are not
always fear-╉or anxiety-╉based, with many individuals mainly experiencing anhedonic or dysphoric
symptoms, aggressive behavior, or even dissociative symptoms (DSM-╉5; APA, 2013).
Prevalence rates of PTSD diagnosis following mass trauma exposure (September 11th,
Hurricane Katrina) varies considerably and seems to mainly depend on the type of trauma experi-
enced, physical proximity in relation to the trauma (personally experienced the event, versus wit-
nessed a family member experiencing the trauma), dose of exposure, time elapsed since trauma
exposure, number of traumatic events (e.g., home destroyed, separated from family) that pre-
cede or follow trauma exposure, and the developmental period the child is currently progressing
through at the time of trauma exposure (Furr et al., 2010; Attanakyake, et al., 2009). For example,
Hoven et al. (2005) reported that 10.6% of children who were exposed to September 11th met
criteria for probable PTSD at six months post-╉attack, but that this rate increased or decreased
depending on the severity of exposure (i.e., 18.4% for severe, 10.0% for moderate, and 3.6% for
6
7
3

376 Children Exposed to Traumatic Stress

mild). The prevalence rate of PTSD symptom severity for children exposed to Hurricane Andrew
decreased over time with 29.8% of children reporting severe to very severe levels at three-╉months
post-╉disaster 18.1% at seven-╉month follow-╉up and 12.5% at ten-╉month follow-╉up (La Greca et al.,
1996). In an even more recent, systemic review of 7920 youth exposed to war-╉related events (e.g.,
armed conflict), probable PTSD diagnosis ranged from 4.5 to 89.3% (pooled estimate of 47%) and
variability was mainly attributable to either study location, measurement methods or duration
since initial war exposure. For example, Weems et al. (2010) found no significant decline in PTSD
symptom severity among urban minority youth exposed to hurricane Katrina from two to two
and a half years following the storm.

Other posttraumatic stress outcomes


Although PTSD symptoms are the most often studied of posttraumatic stress outcomes (Furr
et al., 2010), a growing body of research has shown that trauma-╉exposed youth experience other
outcomes such as anxiety, depression, aggression, grief, somatic complaints, poor academic
achievement, social problems, and sleep problems (e.g., Brown et al., 2011; Hensley & Varela,
2008; Marsee, 2008; La Greca et al., 2002; Osofsky et al., 2009; Roberts et al., 2010; Terranova,
Boxer, & Morris, 2009a; Weems et al., 2013). Conversely, certain youth may also exhibit resil-
ience in the face of adversity and even experience posttraumatic growth or positive adjustment
following a traumatic event (Tedeschi & Calhoun, 2004; Prati & Pietrantoni, 2009; Weems &
Graham, 2014). One possible mechanism that may explain this wide range of posttraumatic
stress outcomes is individual differences in youths’ ability to regulate their emotional experi-
ences and expressions.

Traumatic stress exposure and negative outcomes


Previous models of child traumatic stress (La Greca et al., 1996; Lanius et al., 2010; Masten, &
Narayan, 2012; Pynoos, et al., 1999; Weems & Overstreet, 2008) and published research (Jeney-╉
Gammon et al., 1993; Kithakye et al., 2010; Marsee, 2008; La Greca et al., 1996; Polusny et al., 2011;
Punamäki et al., 2014; Russoniello et al., 2002) have emphasized the central role of heightened
emotional reactivity and dysregulation in the emergence of PTSD symptoms and other negative
posttraumatic stress outcomes, such as greater anxiety and aggression. Research has provided ini-
tial support for a concurrent mediational path from traumatic stress exposure to PTSD symptoms
(Polusny et al., 2011) and aggression (Marsee, 2008) via experiential avoidance and dysregulated
emotional expression. Drawing largely from this past theoretical and empirical knowledge base,
we present a theoretical framework in Figure 18.1 of childhood traumatic stress exposure, which
posits emotion regulation as a potential mediating and causal pathway from mass traumatic stress
exposure to both PTSD symptomology and other negative outcomes (e.g., internalizing and exter-
nalizing problems, academic difficulties).
We posit that emotional dysregulation is an inherent aspect of PTSD and related negative out-
comes (e.g., PTSD symptoms, anxiety, and aggression), and involves maladaptive attempts at
regulating positive and negative emotions and emotion regulation deficits across neurobiologi-
cal, cognitive, and behavioral systems. As outlined in Figure 18.1, Trauma Exposure (direct or
indirect) may lead to Immediate Posttraumatic Stress Reactions, that include the emergence of
negative cognitions (lack of perceived control over self or environment), heightened physiologi-
cal arousal (e.g., faster heart rate), and negative affect (feelings of anxiety, anger, or shame). In
turn, these Immediate Posttraumatic Stress Reactions may become problematic as youth rely on
maladaptive automatic and effortful emotion regulation or dysfunctional emotion regulatory pro-
cesses, at multiple levels of the individual, for immediate reduction of this distress (see Emotion
73
Potential Negative Outcomes

Short-Term Outcomes
• Internalizing Problems
• Externalizing Problems
• Interpersonal Problems
Automatic Effortful • Poor Academic Achievement
• Biological Changes

Long-Term Outcomes
Emotion Regulation Mechanisms
Immediate Posttraumatic • Affective Disorder Onset
Stress Reactions • Substance Use
Behavioral
• Academic Failure
Negative Cognitions • Avoidance/Escape
• Exaggerated or Blunted
• Emotional Suppression
• Low Perceived Control HPA Axis Functioning
Mass Trauma • Negative Memories
Cognitive
Exposure • Negative Schemas
• Experiential Avoidance
• Natural Disaster Physiological Arousal • Rumination
• War Conflict • Distraction
• Increased Heart Rate
• Terrorism • Attention to Threat
• Cortisol Secretion
Neurobiological
Negative Affect Secondary Posttraumatic
• Fear/Anxiety • Prefrontal Cortex Stress Reactions
• Anger • Increased Parasympathetic
• Shame Activation
Physiological/Emotional
• Cortisol Reactivity
• Maintained or Increased Arousal
• Emotional Numbing

Cognitive Response
• Attentional Bias
• Cognitive Bias
• Memory Bias

Figure 18.1  An Emotion Regulation Model of Traumatic Stress Exposure in Youth and Potential Negative Outcomes
8
7
3

378 Children Exposed to Traumatic Stress

Regulation Mechanisms in Figure 18.1; note that mechanisms on a continuum, such that a effort-
ful process may become more automatic over time or typically automatic processes come under
effortful control). Thus, the inability to effectively regulate their Immediate Posttraumatic Stress
Reactions will give rise to Secondary Posttraumatic Stress Reactions, that include cognitive biases
(sustained attention to threat, poor self-╉efficacy, or trauma memories), increased or maintained
heightened emotional arousal, and prolonged negative affect. Finally, Secondary Posttraumatic
Reactions may lead to further reliance of maladaptive or dysfunctional mechanisms via a negative
feedback loop, thus maintaining the experience of negative emotional distress (subjectively and
physiologically) and negative cognitions over time.
A vicious cycle of heightened emotional reactivity and dysregulation, coupled with negative
thought patterns, may result in short-╉and long-╉term negative sequelae (as aforementioned in the
preceding Posttraumatic Stress Outcomes section). In the short-╉term and as shown in past research,
youth may begin to experience internalizing problems (e.g., anxiety, depression; Copeland et al.,
2007), externalizing problems (e.g., aggression; Marsee, 2008; Scott, Lapré, Marsee, & Weems,
2014), interpersonal problems (e.g., peer bullying and victimization; Terranova et  al., 2009a),
poor academic achievement (Scott et  al., 2014; Weems et  al., 2013)  or certain neurobiological
changes across development (atypical maturation in amygdala volumes; Weems, Scott, Russell,
Reiss, & Carrión, 2013). Long-╉term effects may become even more serious with affective disorder
development (e.g., PTSD, Anxiety Disorders, and Major Depressive Disorder; Hoven et al., 2005;
La Greca et al., 1996), substance use (Wagner et al., 2009), academic failure (Porche, Fortuna, Lin,
& Alegria, 2011), and changes in neurobiological functioning (De Bellis, 2001; Weems & Carrión,
2007). The following sections further develop each facet of the model.

Defining emotion and emotion regulation


Although a standard definition of emotion has yet to be developed, most emotion theorists agree
upon several key features (Campos et al., 2004; Cole, et al., 2004; Gross, 1998a; Gross & Thompson,
2007). In this chapter we adopt Gross and Thompson’s (2007) process model of emotion because
it provides a comprehensive overview of emotion based on several emotion theories (Campos
et al., 2004; Cole et al., 2004; Frijda, 1986; Gross, 1998a; James, 1884; Thompson, 1994) and thus
allows for a solid foundation in establishing a working definition of emotion regulation. Gross and
Thompson (2007) suggests that most emotion theories agree upon three core features of emotion,
which are: an emotion 1) occurs when an individual attends to an internal or external event that
pertains to achieving a personal and enduring goal (e.g., avoiding dangerous situation) or self-╉
efficacious and transient goal (e.g., passing a test), but the meaning of each goal will produce a
specific emotion; 2) arises within a flexible, multi-╉system structure and involves independent but
also interactive changes among cognitive (e.g., “I have no control over bad things happening to
me”), behavioral (e.g., not returning to the physical site of trauma or similar settings), and affec-
tive (subjective and physiological; “I feel scared” and faster heart rate, respectively); and 3) is mal-
leable and may be modulated at any time or level of the individual and may vary across a number
of situational domains.
This third feature relates to the child’s ability to regulate emotion responses and how youth try
to regulate such emotions. Again drawing from the framework of Gross and Thompson, emotion
regulation is the automatic or effortful modulation (i.e., enhance, decline, or maintain) of both
negative and positive emotions at multiple levels of the analysis (i.e., neurobiological, cognitive,
and behavioral; Cole et al., 2004; Gross, 1998a; 1999; Thompson, 1994). Emotion regulation may
occur at any given point of the emotional generative process and may involve modifications to the
latency, intensity, rise time, magnitude, and duration of an emotion (Gross & Thompson, 2007).
9
7
3

Traumatic stress exposure and negative outcomes 379

Immediate posttraumatic stress reactions


As shown in the Immediate Posttraumatic Stress Reaction box of Figure 18.1, exposure to mass
trauma that involves life threat may result in posttraumatic stress reactions that are synonymous
with symptoms outlined in the DSM-╉5 as part of the PTSD criterion (APA, 2013). First, trauma
exposure may result in negative, intrusive cognitions, such as challenging one’s sense of control
and self-╉efficacy in containing the actual or perceived threat (e.g., low perceived control) or the
activation of emotion latent memories or schemas of past negative life events. Second, direct
exposure to life threat may increase activity of the limbic-╉hypothalamic-╉pituitary-╉adrenal (LHPA)
axis and autonomic nervous system as part of a normative fight-╉flight reaction, thus inducing
hyperarousal (Heim & Nemeroff, 2001; Porges, Doussard-╉Roosevelt, & Maiti, 1994; Thayer and
Lane, 2000). Research has shown that fear reactions are associated with elevations in the secretion
of cortisol, a corticosteroid hormone produced by the adrenal cortex that can be assayed from
blood, urine, or saliva samples (see Nader & Weems, 2011 for a review) and greater sympathetic
activation can be assessed using heart rate and skin conductance measurement during stressful
conditions (e.g., increased heart rate; Culbert, Kajander, & Reaney, 1996; De Young, Kenardy, &
Spence, 2007; Weems et al., 2005). Third, trauma exposure may result in negative affect that may
be difficult to regulate (e.g., anxiety, anger, or shame; Norris et al., 2002), further increasing the
risk for the development of child psychopathology. As a whole, these posttraumatic stress reac-
tions may occur simultaneously (thus covarying with one another), have some influence on each
other, or even lead to the occurrence of the other.

Emotion Regulation Mechanisms


Immediate posttraumatic stress reactions may have a significant effect on the way that children
regulate their emotions in the aftermath of a natural disaster or during wartime. For example,
heightened arousal and hypervigilance to perceived or actual threat may promote the use of
automatic or effortful mechanisms associated with emotion regulation to be activated within
the child (e.g., avoidance, emotional suppression). Although these regulatory mechanisms
respond in an adaptive manner, as to allow for the immediate reduction of intense arousal or
to suppress negative cognitions, in the long run they may either serve as maladaptive mecha-
nisms that could lead to greater arousal, more negative cognitions, and poor outcomes across
emotional, behavioral, cognitive, and social domains. Moreover, these mechanisms will even-
tually become the symptoms of the emotional and behavioral disorders (e.g., PTSD, anxiety
disorders) that subsequently develop from their actions and also help maintain these disorders
across development.
Research has shown that mass trauma-╉exposed youth tend to use a greater number of emo-
tion regulation strategies than non-╉exposed youth (Russoniello et  al., 2002)  and high levels of
PTSD, anxiety, depression symptoms and general mental health distress are related to greater fre-
quency (or intensity) of emotion regulation use (Asarnow et al., 1999; Jeney-╉Gammon et al., 1993;
Kithakye et al., 2010; Punamäki et al., 2014; Russoniello et al., 2002). Thus, it appears that greater
engagement in emotion regulation strategies does not necessarily protect youth from developing
emotional and behavioral problems following a traumatic event, but that more likely, individual
differences in the specific types (or patterns) of emotion regulation mechanisms utilized plays a
larger role in post-╉trauma functioning.

Neurobiological mechanisms
At the neurobiological level, theory and research suggest that emotional dysregulation
may largely stem from structural changes in prefrontal cortices of trauma-╉exposed youth,
0
8
3

380 Children Exposed to Traumatic Stress

maladaptive parasympathetic-​mediated control over stress responses, and both diurnal cortisol
release over time and cortisol reactivity to stress (Carrión et al., 2001; Carrión, Weems, & Reiss,
2007; Carrión, Weems, Richert, Hoffman, & Reiss, 2010; De Bellis, 2001; Feldman, Vengrober,
Eidelman-​Rothman, & Zagoory-​Sharon, 2013; Lou et al., 2012; Richert, Carrión, Karchemskiy,
& Reiss, 2006; Scheeringa, Zeanah, Myers, Putnam, 2004; Scott & Weems, 2014; Vigil, Geary,
Granger, & Flinn, 2010; Weems & Carrión, 2007). However, it is important to note that, relative
to other types of trauma, the study of neurobiological mechanisms among mass trauma-​exposed
youth is quite limited. Nevertheless, we rely on these small, but consistent past research findings
(e.g., Feldman et al., 2013; Luo et al., 2012 Scheeringa et al., 2004; Scott & Weems, 2014; Vigil
et al., 2010; Yehuda et al., 1995; Yehuda, 2006), to illustrate how traumatic exposure may affect
neurobiological mechanisms related to emotion regulation and, in turn, their relation to mental
health problems.
Neuroimaging research comparing prefrontal cortical regions, that are theoretically linked to
cognitive emotion regulation (medial frontal cortex, orbital prefrontal cortex, dorsal prefrontal
cortex), has shown differentiated brain development and functional impairment among trauma-​
exposed and non-​exposed youth (see Carrión and Wong, 2012 for in-​depth review). For exam-
ple, Richert et al. (2006) found that 23 youth with PTSD had larger volumes of gray matter in
the middle-​inferior and ventral regions of the prefrontal cortex as compared to 24 children with
no PTSD and that decreased gray matter volume of the dorsal prefrontal cortex increased the
functional impairment of those youth with PTSD. The authors suggested that the larger volume
of the middle-​inferior and ventral regions may reflect frequent prefrontal lobe activity aimed at
inhibiting emotional posttraumatic stress responses (thus maintaining PTSD symptoms), while
decreased dorsal volume may signify a predisposition towards difficulty engaging in cognitive
emotion regulation strategies, such as reappraisal. In terms of functionality, Carrión, Garrett,
Menon, Weems, and Reiss (2008) found that youth experiencing PTSD symptoms had reduced
middle frontal cortex activity and increased medial frontal activity during memory and executive
functioning tasks, as compared to an age and gender-​matched healthy control sample of youth.
Together these findings suggest two possibilities: 1) youth may have predisposed deficits in reg-
ulating emotion using higher-​order cognitive processing (Emotion Regulation Mechanisms in
Figure 18.1), or 2) traumatic stress exposure and subsequent maladaptive reactivity and regula-
tion may alter youths’ brain development and functionality in centers of the brain largely involved
in emotion regulation (Short or Long-​Term Outcomes in Figure 8.1).
Accumulating research also has shown that maladaptive patterns of parasympathetic-​mediated
control over the heart (i.e., indexed by heart rate variability [HRV] or respiratory sinus arrhyth-
mia [RSA]) is related to PTSD symptoms, anxiety, and aggression in preschoolers and adolescents
with trauma exposure (Scheeringa et al., 2004; Scott & Weems, 2014). Parasympathetic-​mediated
control involves controlled modulation of emotions via a negative feedback loop between several
afferent and efferent neurobiological structures thought to be responsible for top-​down emotion
regulation (e.g., nucleus ambiguus, medial prefrontal cortex) and the heart’s sinoartial (SA) node
(i.e., pacemaker) along the vagus nerve (Beauchaine, 2001; Oschner & Gross, 2005; Porges et al.,
1994; Thayer & Lane, 2000). Theoretically, the PNS’s function is to help youth maintain bodily
homeostasis during times of rest with lower resting HRV indexing a deficit in PNS-​mediated
control over the heart it serves as a proxy for general emotional dysregulation (Beauchaine, 2001;
Porges, 2007; Thayer & Lane, 2000). Moreover, parasympathetic nervous system (PNS)-​mediated
control allows youth to quickly and efficiently regulate the amount of time and degree of control
the sympathetic nervous system (SNS) has over bodily sub-​systems (Porges et al., 1994; Thayer
& Lane, 2000). Porges (2007) posits that greater parasympathetic suppression is a flexible, adap-
tive response to stress, which in turn promotes functional behavior, while increased or blunted
1
8
3

Traumatic stress exposure and negative outcomes 381

parasympathetic activation is a rigid and maladaptive response that leads to poor physical and
mental health outcomes.
In a study with 80 adolescents (ages 11–╉17  years) with whom most experienced Hurricane
Katrina or the BP Oil Spill (n  =  76), we found that lower resting HRV was related to greater
anxiety problems (Scott & Weems, 2014). We also found evidence to support Porges’s (2007) the-
ory in that greater parasympathetic activation (i.e., increase in HRV from baseline to cognitive
stress task) was observed among the adolescents with higher levels of anxiety and aggression.
Schreeringa et al. (2004) found that among 144 preschool children (62 trauma-╉exposed and 62
non-╉exposed) those with high levels of PTSD had greater RSA withdrawal to a trauma reminder,
but only for those youth whose parents exhibited low positive discipline during a clean-╉up task.
Though a number of factors may explain these contradictory findings, developmentally younger
children with PTSD may display a more normative strong sympathetic-╉mediated response to
stressful situations (HRV or RSA withdrawal), but as youth continue to encounter repeated stress
responses for longer periods of time they may develop a vulnerability for emotion dysregulation
(low resting HRV) or engage in greater overcontrolling responses, such as the HRV attenuation
stress response found in our study (Scott & Weems, 2014).
Changes in neural mechanisms following stress (Carrión, Weems, & Reiss, 2007; Carrión et al.,
2010)  and susceptibility to dysregulation in the normative stress response may also character-
ize individual risk for mental health problems among mass trauma victims. The persistent and
intense taxing of the stress response system over time may lead to a physiological dysregulation
of the system. Research suggests that after a period of relative cortisol hypersecretion elevated
levels may reverse in trauma exposed youth (De Bellis, 2001; Weems & Carrión, 2007) to relatively
low levels of cortisol in diurnal patterns and/╉or blunted cortisol reactivity in response to stress
(Feldman, Vengrober, Eidelman-╉Rothman, & Zagoory-╉Sharon, 2013; Yehuda, 2006). This low or
blunted cortisol response may result from an enhanced negative feedback loop at the pituitary-╉
adrenal level of the axis (Yehuda et al., 1995). A proposed mechanism for this sensitization is an
increased number of glucocorticoid receptors in the LHPA axis and, hence, facilitation of this
negative feedback loop (Yehuda, 2006). Dysregulation in the stress response system has been
associated with several mental health disorders among trauma-╉exposed youth, including PTSD,
anxiety, and depression (Feldman et al., 2013; Gunnar, 2001; Vigil et al., 2010; Weems & Carrión,
2009; Yehuda, 2006).

Cognitive mechanisms
At the cognitive level, youth seem to have a relatively large number of emotion regulation strat-
egies to use, though these higher-╉order, top-╉down regulatory processes are taxing to the indi-
vidual and more accessible to older children and adolescents (Gross, 1998b; Oschner & Gross,
2005; Richards & Gross, 2000). For example, some trauma-╉exposed youth may rely on cogni-
tive strategies that reduce emotional distress quickly (thus negatively reinforcing use), but at the
expense of chronic efforts to control internal reactions (e.g., emotional suppression, experiential
avoidance) and thus maintaining or increasing negative affect and physiological arousal over the
course of development (Polusny et al., 2011; Richards & Gross, 2000). Additionally, the use or
capability to use certain emotion regulation strategies may vary across development. Following
the September 11th attacks, Wadsworth and colleagues (2004) found an increase in the use of
emotion-╉focused coping and decreases in rumination and disengagement across adolescence to
adulthood. Cardeña and colleagues (Cardeña, Dennis, Winkel, & Skitka, 2005) reported that ado-
lescents were more likely to use distraction and disengagement than adults, who tended to use
strategies such as planning and acceptance. Furthermore, younger children exposed to trauma
may still be developing the cognitive skills to fully process the event or engage in higher-╉order
2
8
3

382 Children Exposed to Traumatic Stress

cognitive emotion regulation (Shiner, 1998), such as reappraisal, and may be more susceptible to
engaging in temperamentally-╉driven emotional coping, such as using avoidance or withdrawal
(e.g., becoming sick and staying home from school).
A few studies have begun to examine the relation between specific cognitive emotion regu-
lation stratigies and negative outcomes in mass trauma-╉exposed youth. In one study, Polusny
et al. (2011) found that experiential avoidance (i.e., attempts at controlling, escaping, or avoid-
ing negative internal experiences) concurrently mediated the relation between disaster exposure
and PTSD symptoms among 288 adolescents exposed to severe tornados. In two other studies,
Prinstein, La Greca, Vernberg and Silverman (1996) and Noppe, Noppe and Bartell (2006) found
that distraction was associated with more severe symptoms of PTSD among hurricane exposed
youth and greater anxiety among youth remotely-╉exposed to September 11th, respectively.
Wadsworth et al. (2004) also found among 168 sixth to eight graders exposed to September 11th,
greater use of secondary control (cognitive restructuring, positive thinking, or acceptance) and
less use of involuntary engagement strategies (e.g., rumination) was related to fewer anxiety prob-
lems among all youth.

Behavioral mechanisms
At the behavioral level, youth have access to strategies that are accessible from early childhood to
regulate various emotions (e.g., fear or aggression; Buss & Goldsmith, 1998; Calkins & Johnson,
1998; Diener & Mangelsdord, 1999)  and involve such behaviors as avoidance or withdrawal,
approach, suppression of emotional expression, problem solving and comfort or help seeking.
Research has shown that some of these behavioral emotion regulation strategies may shift fol-
lowing exposure to a traumatic event from typically adaptive to maladaptive strategies (Kennedy,
Charlesworth, & Chen, 2004). For instance, Kennedy et al. (2004) found in a prospective study 1-╉
2 months before and after the September 11th attacks that children remotely affected (i.e., watched
media coverage and lived in San Francisco, California) reported using more active coping strate-
gies (e.g., “think about it” and “talk to someone”) before September 11th, but after September 11th
used more avoidant coping strategies (“draw, read, or write” or “play a game”), which they viewed
as effective.
Research has shown that avoidant coping is consistently associated with greater PTSD symp-
toms (e.g., La Greca et al., 1996; Vernberg, Silverman, La Greca, & Prinstein, 1996), while active
coping (e.g., problem-╉focused) is related to lower depression symptoms in youth hurricane sur-
vivors (e.g., Jeney-╉Gammon et al., 1993). Pina et al. (2008) found that avoidant coping behaviors
(i.e., repression, avoidant actions) predicted post-╉Katrina PTSD and anxiety symptoms, which is
consistent with other research (Norris et al., 2002). Terranova, Boxer and Morris (2009b) exam-
ined predictors of PTSD symptoms in a sample of 152 sixth grade school children from southeast
Louisiana (neighboring Orleans parish) assessed at one-╉and-╉a-╉half months and eight months after
Katrina and found that negative coping (a combination of internalizing, externalizing and avoid-
ant coping) was associated with PTSD at one-╉and-╉a-╉half months. Peer victimization (i.e., being
bullied) was predictive of change in PTSD (PTSD symptoms at Time 2 controlling for symptoms
at Time 1) and results further indicated that negative coping interacted with level of hurricane
exposure to predict change in PTSD, such that high negative coping and high exposure was asso-
ciated with the highest PTSD symptoms at Time 2.
Though avoidance or withdrawal strategies have a strong and consistent relation with youth
problems post-╉trauma, other strategies have shown specific associations. For example, Wadsworth
et  al., (2004) found that greater use of involuntary disengagement (e.g., emotional numbing,
escape) strategies and less use of primary control strategies (e.g., problem solving) were associated
with increased anxiety in girls exposed to September 11th. Furthermore, research also has shown
3
8

Traumatic stress exposure and negative outcomes 383

that comfort and help seeking is associated with overall better mental health in youth exposed
to a natural disaster (i.e., 2008 Chinese Earthquake; Zhang et al., 2010) and the absence of PTSD
diagnosis among preschool children (ages one-╉and-╉a-╉half to five years) continuously exposed to
war conflicts.

Secondary traumatic stress reactions


The engagement in maladaptive emotion regulation mechanisms or the inability of such mech-
anisms to modulate emotion effectively, will likely result in further increased or maintained
emotional arousal, negative affect, and the activation of more negative cognitive responses (i.e.,
attention bias towards threat, general cognitive biases, poor self-╉efficacy, memory bias for nega-
tive and trauma-╉related information; Moradi, Taghavi, Neshat-╉Doost, Yule, & Dalgleish, 1999;
Scott & Weems, 2014; Weems, Russell, Graham, Neill, & Banks, 2015). For example, unsuccess-
ful attempts to reduce their negative emotions (or produce positive emotions) may diminish
youths’ self-╉efficacy to handle frightening or stressful situations (external events) or to effec-
tively manage their emotions (internal events). In turn, this low perceived control may lead to
further use of maladaptive emotion regulation mechanisms via a negative feedback loop (as
illustrated with arrow in Figure 18.1 from the Secondary Posttraumatic Reactions to Emotion
Regulation Mechanisms) or to Potential Negative Outcomes (Chorpita & Barlow, 1998). Scott
et al., (2014) recently found that poor anxiety control beliefs were associated with lower rest-
ing heart rate variability (an objective measure related to emotion regulation) among disaster-╉
exposed youth. Weems et  al. (accepted) also found that poor anxiety control beliefs were
associated with more PTSD and Generalizing Anxiety Disorder symptoms above and beyond
trauma exposure among 1048 Hurricane Katrina-╉exposed third to 12th graders. However, this
relation was moderated by youths’ age, such that it only remained significant for youth over
12 years of age. Weems et al. (2015) suggests findings are consistent with cognitive develop-
mental theory in that younger children may tend to attribute outcomes out of their control to
their own behavioral responses and as they mature control beliefs become more realistic and
predictive of youths’ emotional problems.

Vulnerability and moderating factors


A number of individual and environmental factors may place children at risk or protect them
from negative outcomes. For example, at the individual level, research has suggested that pre-╉
existing characteristics of the child can affect the impact of disaster exposure on mental health
outcomes. In particular, previous research has documented that pre-╉hurricane trait anxiety and
negative affect levels (which may represent the emotional dysregulation these youth are expe-
riencing) predict PTSD symptoms above and beyond trauma exposure (La Greca et  al., 1998;
Weems, Pina, et al., 2007), while emotion regulation capabilities may protect youth from the ill
effects of trauma exposure (e.g., Kithakye et al., 2010; Punamäki et al., 2014). In addition, youth
with certain temperamental qualities related to emotion regulatory processes may be at greater
risk for having emotional and behavioral problems following a traumatic event. For example,
Lengua, Long, Smith, and Meltzoff (2005) found in a prospective study before and after the
September 11th attacks that greater inhibitory control was marginally related to fewer post-╉attack
PTSD symptoms (after controlling for pre-╉attack PTSD symptoms) in 142 children (ages nine to
13 years). In another longitudinal study with Hurricane Katrina youth victims, Terranova et al.
(2009b) found that Time 1 effortful control moderated the relation between Time 1 hurricane
exposure and Time 2 PTSD symptoms with a positive relation exhibited for those youth with low
effortful control.
4
8
3

384 Children Exposed to Traumatic Stress

Halberstadt, Thompson, Parker, and Dunsmore (2008) have demonstrated that specific parent-
ing beliefs of youths’ emotions increases the odds of the child using a specific type of emotion-╉
related coping strategy among 51 youth remotely exposed to the September 11th terrorist attacks.
That is, youth of parents who believed youth emotion was valuable was related to the child’s use of
typically adaptive coping strategies (i.e., problem-╉solving, support-╉seeking, and emotion-╉oriented
coping), while their belief that their child’s emotions were dangerous was related to typically mal-
adaptive coping strategies (i.e., avoidance and distraction). In a second study, Polusny et al. (2011)
found that the concurrent mediated relation between adolescents’ disaster exposure and PTSD
symptoms via experiential avoidance was stronger among those youth with parents experienc-
ing greater PTSD-╉related distress. In a third study, Hendricks and Borstein (2007) found, that
relations between PTSD arousal symptoms and both attentional control and cognitive avoidance
among 97 adolescents directly exposed to the September 11th attacks, disappeared once distal
contextual factors were statistically controlled for in the model (i.e., maternal personality charac-
teristics and adolescent’s perceptions of parenting), suggesting that family factors may also play a
large role in the development of PTSD symptoms beyond individual factors.

Evidence-╉based interventions for trauma-╉exposed children


Helping children and their families effectively process traumatic experiences and cope with the
deleterious impact of mass trauma exposure is a difficult, but much needed undertaking. Robust
empirical evidence supporting the efficacy and/╉or effectiveness of interventions for mass trauma-╉
exposed youth is scant. As La Greca and Silverman (2009) note, there are several ethical and
methodological challenges that have thwarted stringent tests of manualized interventions in the
immediate aftermath or recovery phases (i.e., 1st year post-╉trauma) of mass trauma. Examples of
such roadblocks include the inability to ethically withhold treatment from exposed youth (i.e.,
lack of no-╉treatment control), wide spread displacement of youth and their families from the
community after exposure, breakdown of the community’s infrastructure, and limited immediate
funds or mental health resources in affected areas. Despite these limitations, a few interventions
have been developed and empirically tested during the recovery (within one year post-╉trauma)
and long-╉term recovery phases (one year or more following initial trauma exposure), such as cog-
nitive behavioral therapy (CBT), school-╉based interventions, and Eye Movement Desensitization
and Reprocessing (EMDR). In this section, we focus on those interventions that meet criteria as a
probably efficacious or well-╉established for reducing negative posttraumatic stress reactions and
which target emotion regulation skill deficits in trauma-╉exposed youth.

Cognitive behavioral therapy (CBT)


Cognitive behavioral therapy (CBT) has long been a first-╉line intervention for anxiety and stress-╉
related disorders, such as PTSD, and is one of a few treatments with robust empirical support of
efficacy and effectiveness in reducing negative post-╉trauma reactions. Several variants of indi-
vidualized, exposure-╉based CBT interventions, that include common core features, have been
examined among youth exposed to natural disasters, terrorism, and war (March et  al., 1998;
Taylor & Weems, 2011). One of the more widely known and supported interventions for trauma-╉
exposed youth is trauma-╉focused cognitive behavioral therapy (TF-╉CBT). Although TF-╉CBT was
initially designed for youth exposed to child sexual abuse, it has since been modified and widely
adapted to help children exposed to other types of trauma (Cohen & Mannarino, 1996; 2008).
TF-╉CBT incorporates affective modulation training, trauma narratives to assist in cognitive pro-
cessing of the events and emotional reactions, and guided cognitive and affective modification
during in vivo exposure to traumatic cues into the traditional CBT format and relies heavily on
5
8
3

Evidence-based interventions for trauma-exposed children 385

parental involvement. As a whole, TF-╉CBT does meet criteria for a well-╉established treatment (see
Silverman et al., 2008 for review) for reducing PTSD symptoms, depressive symptoms, shame, and
other emotional and behavioral problems (e.g., Cohen & Mannarino, 1996; Cohen, Deblinger,
Mannarino, & Steer, 2004; Deblinger & Heflin, 1996; Deblinger, Stauffer & Steer, 2001; Deblinger,
Steer, & Lippman, 1999; March et al., 1998).

School-╉based interventions
Increased risk created by disasters and terrorism within the family environment can be offset by
the presence of protective factors within other microsystems surrounding the child. For example,
school-╉based mental health interventions represent a protective factor within the school micro-
system that can counterbalance the negative developmental outcomes associated with mass
trauma exposure (Pynoos, Goenijian, & Steinberg, 1998). School-╉based interventions also allow
for the delivery of mental health services to a large group of youth and may serve youth whose
parents are not able to or are reluctant to seek out services in the community, or live in an area
that is struggling to meet the demand of services (i.e., reduction in mental health providers, but a
greater number of youth in need of mental health services).
One of the more promising school-╉based interventions for traumatized youth is the Cognitive
Behavioral Intervention for Trauma in Schools, which meets criteria for a “probably efficacious”
treatment of reducing PTSD symptoms (CBITS; Stein et  al., 2003; Jaycox et  al., 2010; Kataoka
et al., 2003). CBITS is composed of ten group sessions and one to three individual sessions and
is intended for school-╉aged youth. The therapeutic goals and components of CBITS closely align
with those of TF-╉CBT, as they focus on psychoeducation, affective modulation skills training, cog-
nitive processing, trauma narrative, in-╉vivo exposure, safety development, and include optional
sessions with parents and teachers. In a randomized field trial, Jaycox et al. (2010) compared the
efficacy of CBITS against TF-╉CBT among 195 hurricane-╉exposed youth at 15  months follow-
ing Hurricane Katrina. They found that the CBITS program was just as effective in significantly
reducing PTSD symptoms following treatment. Although the CBITS providers were limited in
their ability to tailor the intervention for each child to their specific trauma histories, as typically is
done in TF-╉CBT, CBITS was more accessible to youth, as 98% (91% completed treatment) began
the CBITS program as opposed to 23% in a community mental health clinic (15% completed).
Thus, implementing a trauma-╉focused CBT program in a naturalistic setting may bring services to
a large group of youth in a community whose citizens and infrastructure is struggling to recover
from a devastating natural disaster.
Another promising school-╉based treatment targeting mass trauma-╉exposed youth is the Grief
and Trauma Intervention (GTI; Salloum, Garfield, Irwin, Anderson, & Francois, 2009, Salloum
& Overstreet 2008; 2012). GTI is a 10-╉week school-╉based trauma and grief focused interven-
tion that was specifically designed for elementary-╉aged children (seven to 12 years of age) who
have experienced either trauma or grief associated with death, disaster, or violence. The primary
treatment components of GTI directly parallel those emphasized in TF-╉CBT (e.g., psychoeduca-
tion, affective modulation training, trauma narrative), though the active role of parents in GTI is
limited to one parent session and targets other post-╉trauma reactions, such as traumatic grief and
depression.
Salloum and Overstreet (2008) first tested the effectiveness of delivering GTI among 56 youth
(second to sixth graders) exposed to Hurricane Katina via after-╉school programs and in-╉school
mental health services (The LAST Project) just four months post-╉Katrina. Youth were randomly
assigned to an individual or group modality. Findings showed that both treatment modalities
led to fewer PTSD, depression, and traumatic grief symptoms at post-╉treatment and at a 20-╉day
6
8
3

386 Children Exposed to Traumatic Stress

follow-╉up. In another study (Salloum & Overstreet, 2012), they randomly assigned 72 youth (sec-
ond to sixth graders) exposed to Hurricane Katrina to receive either GTI with coping skills plus
a trauma or loss narrative or GTI with coping skills only. Both treatment groups showed signifi-
cant reductions of PTSD, depression, and traumatic grief symptoms, though the GTI group who
received the trauma or loss narrative reported expressing their thoughts and feelings more, while
the GTI group who received coping skills-╉only training reported more ways of coping.
Evidence for the efficacy of similar treatment programs among youth exposed to hurricanes
and war/╉terrorism also has been demonstrated (earthquakes, Goenjian et al., 2005: war, Layne
et al., 2001; 2008). For example, Layne and colleagues (Layne et al., 2001; 2008) found sup-
port for implementing a trauma and grief group intervention (17 sessions vs. 10 sessions for
GTI) in war torn communities. More specifically, Layne et al. (2008) found in a randomized
control trial that reductions in PTSD, depressive, and maladaptive grief only occurred for the
treatment group and not for an active control group (classroom-╉based psychoeducational and
skills training) among 127 war-╉exposed youth in central Bosnia. Altogether, findings suggest
that GTI and similar treatment protocols are probably efficacious and it will be important for
future trials to include a control group (e.g., wait-╉list, well-╉established treatment for PTSD,
such as TF-╉CBT).
Weems et  al. (2009; 2014)  also demonstrated the efficacy of a school-╉based intervention on
reducing posttraumatic stress symptoms in the post-╉disaster environment, even when the inter-
vention was not specifically focused on treating post-╉traumatic reactions. In one study (Weems
et al., 2009), a prospective intervention design was utilized with a sample of 94 ninth graders from
New Orleans exposed to Hurricane Katrina and its aftermath. Thirty youth with elevated test
anxiety completed a primarily behavioral and exposure-╉based (e.g., relaxation training combined
with gradual exposure to anxiety-╉provoking test-╉related stimuli) group administered, test anxiety
reduction intervention. Findings suggested a statistically significant effect of the intervention on
test anxiety levels and academic performance with evidence of positive secondary effects on post-
traumatic stress symptoms (PTS). Moreover, change in test anxiety predicted change in PTS and
there appeared to be no negative effects on natural PTS symptom decline.

Eye movement desensitization and reprocessing


Eye movement desensitization and reprocessing (EMDR) is an upcoming and promising treat-
ment for PTSD symptoms among traumatized youth (Fernandez, 2007). EMDR is centered upon
the adaptive information processing (AIP) model, which asserts that traumatic memories are
stored without fully processing the information, which interferes with adaptive responses to
abate related cognitive or emotional reactions, and thus, one continues to experience chronic
and intense reactivity (e.g., fear or anxiety) to trauma cues (e.g., memories; Shapiro, 2007).
Moreover, neural networks storing unprocessed trauma memories do not assimilate positive
experiences and may even block positive information from being accessed or processed in other
networks. EMDR involves eight phases aimed at providing children with self-╉regulation skills to
manage post-╉trauma reactions, identify or build adaptive positive networks to link with negative
networks; process negative memories through desensitization; reduce distress to negative mem-
ories; integrate and strengthen positive cognitive connections; improve self-╉efficacy; and elimi-
nate negative bodily sensations to the negative memories (Shapiro, 2007). Educational meetings
are scheduled with the parents to help them understand the posttraumatic stress reactions their
child is experiencing and how their child can adaptively manage their posttraumatic stress reac-
tions. Parents are also encouraged to attend each child session to serve as a family support system
and to help the child through the processing experience.
7
8
3

Trauma-focused cognitive behavioral therapy (TF-CBT) 387

Only a limited number of studies have examined the efficacy of EMDR in youth exposed to
mass trauma. In one study, Fernandez (2007) showed significant reductions in PTSD symptoms
and PTSD diagnosis (61% vs. 9% from pre-╉to post-╉treatment) among 27 children (ages seven
to 11 years) who survived a primary school building collapse (27 of 59 children died) during the
2002 Molise earthquake in Italy. However, the lack of a control group limits one’s ability to dis-
cern whether reductions were due to the passage of time. In a more recent randomized trial, De
Roos et al. (2011) compared the efficacy of EMDR to CBT in 52 youth (ages four to 18 years) who
were exposed to a large fireworks factory explosion in the Netherlands. Youth in both interven-
tions showed significant post-╉treatment reductions in PTSD, anxiety, depression, and behavioral
problems and effects were maintained at three-╉month follow up, though EMDR produced these
results using fewer sessions. As with the Grief and Trauma Intervention, EMDR is a promising
treatment option for mass trauma-╉exposed youth, but further research with a control group is
needed to elucidate the efficacy and effectiveness of EMDR to ameliorate negative posttraumatic
outcomes.

Trauma-╉focused cognitive behavioral therapy (TF-╉CBT)


TF-╉CBT is an individualized, component-╉based intervention (i.e., length of treatment may vary
and movement forward is not dictated by completing a session) that helps youth and their
parents learn to process and cope with various posttraumatic stress reactions, such as PTSD
symptoms, anxiety, depression, and grief (Cohen & Mannarino, 2008). TF-╉CBT is appropriate
to treat trauma-╉exposed youth ranging in age from three to 18 years. The primary therapeutic
goal of TF-╉CBT is to assist trauma-╉exposed children and adolescents in progressively master-
ing the skills to understand and manage the cognitive, affective, and behavioral reactions that
occur post-╉trauma. Moreover, TF-╉CBT is aimed to guide children and their parents through
the cognitive processing of the initial trauma exposure and secondary adversities (e.g., depleted
finances, lost home) that often happen in the aftermath of such mass trauma events. As such,
parents serve as an integral part of the TF-╉CBT program, as they often co-╉experience the trau-
matic event (e.g., natural disaster) directly or indirectly via their child and may also experience
negative reactions. Thus, helping parents learn to cope with their own post-╉trauma reactions
will not only lessen parental and familial distress, but also provides parents with an opportunity
to model adaptive coping skills for their children and encourage them to practice between treat-
ment sessions. Nevertheless, Cohen and Mannarino (2008) are quick to note that TF-╉CBT is
still a child-╉focused intervention. Parents’ with severe PTSD symptoms or other psychopathol-
ogy should be referred for individual treatment as their symptoms may interfere with adaptive
parenting practices.

TF-╉CBT treatment components


The TF-╉CBT program consists of multiple components that envelop more specific therapeutic
goals and is aimed at helping youth to systematically develop skills and self-╉efficacy for dealing
with the various types of post-╉traumatic reactions they may experience. The component-╉based
method allows for individualization of treatment needs and progression. That is, some youth may
need as few as 12 sessions to process the trauma or learn coping skills, while other youth may need
as many as 20 sessions to accomplish the same goals. Several components have strong connection
to emotion regulation. For example, relaxation skills training, where youth and their parents are
then taught methods to reduce physiological stress reactions (e.g., focused breathing or mindful-
ness) and encouraged to apply these skills when posttraumatic stress reactions are experienced
between sessions.
83

388 Children Exposed to Traumatic Stress

Another component “Affective Expression and Modulation” occurs when the therapist asks the
child-╉parent dyad to share and talk about their feelings and to engage in written or verbal exer-
cises (e.g., produce a list of feelings in three minutes) and games (e.g., Emotional Bingo, Mitlin,
1998) aimed at helping the child identify their feelings. The dyad also learns a number of affective
modulation strategies (i.e., thought interruption, positive imagery, positive self-╉talk, and problem
solving through social skill building) to combat negative and intrusive cognitive-╉affective states
via guided instruction and in-╉session practice. In Cognitive Coping and Processing the child-╉
parent dyad is taught how to generate alternative thoughts that may aid them in changing the way
they feel.
Three studies have provided some support for the efficacy of TF-╉CBT among youth exposed
to mass trauma. In one study, Jaycox et al. (2010) showed that TF-╉CBT was just as efficacious
in reducing PTSD symptoms as the school-╉based CBITS (Jaycox, 2003), though the effective-
ness of such treatment in disaster settings is still debatable given the low frequency of youth
beginning or completing TF-╉CBT (23% and 15%, respectively). In a second study among 306
youth exposed to the September 11th World Trade Center terrorist attacks, researchers for
the CATS Consortium (2010) compared the effects of TF-╉CBT (12–╉20 sessions; adolescents
received the trauma and grief component therapy for adolescents [Layne et al., 2002], which
has comparable components) to a brief CBT program that included a parent component (4
sessions) at six-╉months post-╉treatment. Though youth were not randomly assigned to the
conditions given ethical and methodological issues cited by La Greca and Silverman (2009),
the researchers did employ a regression discontinuity (needs-╉based assignment), which allows
for comparison of regression slopes and intercepts across treatment groups (instead of mean
level comparisons). The results showed reductions in PTSD symptoms from pre-╉intervention
to six-╉month follow-╉up across both groups. However, youth assigned to the TF-╉CBT group had
greater trauma exposure and more environmental adversity (e.g., victimization) at baseline
and showed more clinical improvement on average as they moved from probable PTSD criteria
(as based upon the PTSD Reactions Index; Mpre-╉treatment = 36.61 vs. M6-╉month follow-╉up = 17.03;
scale ranges from zero to 80) to mild PTSD symptom criteria following treatment. In a third
study, Schreeringa et al. (2011) showed that TF-╉CBT may even be a viable treatment option
for younger children who experience traumatic events. Specifically, they found that PTSD
symptoms significantly decreased from pre-╉to post-╉treatment for those traumatized pre-╉
school children (three to six years of age) who received and completed treatment immediately
upon entering the study, as compared to wait-╉list control children assessed during the same
time frame.

Conclusions
In summary, mass trauma-╉exposed youth utilize a number of mechanisms or strategies that span
across neurobiogical, cognitive, and behavioral domains in an effort to regulate their negative
emotional responses following trauma exposure (Kennedy et al., 2011). Review of the findings
presented in this chapter further suggests that negative outcomes of PTSD symptoms, anxiety,
and depression are associated specifically with certain neurobiogical markers of poor emotion
regulation, such as structural and functional changes in prefrontal cortical regions, low HRV at
rest and blunted or increased HRV to stress, and lower cortisol responses over time. In addi-
tion, PTSD has been associated with the frequent use of maladaptive cognitive strategies such as,
experiential avoidance, distraction, and rumination; and behavioral emotion regulation strategies
such as avoidance, escape, and emotional suppression (Feldman et al., 2013; Jeney-╉Gammon et al.,
9
8
3

Conclusions 389

1993; Kithakye et al., 2010; Noppe et al., 2006; Pina et al., 2008; Polusny et al., 2011; Prinstein
et al. 1996; Punamäki et al., 2014; Russoniello et al., 2002; Scott & Weems, 2014; Terranova et al.,
2009b;Vigil et al., 2010; Wadsworth et al., 2004). However, and as with many other psychological
processes, emotion regulation likely does not occur in isolation and youth may attempt to use
many strategies (either consciously or non-​consciously) during a stressful situation. Conceptually,
youth may automatically respond to a particular trauma reminder or stressful event at first, but
if distress continues, may progressively move through more effortful emotion regulation strat-
egies until emotional distress subsides. Thus, over time youth may develop certain sequential
patterns of responding to both positive and negative emotional events. It would be beneficial for
researchers to use mixed modeling approaches (e.g., latent growth models; see Gullone, Hughes,
King, & Tonge, 2010 for example of using such methods to measure change in emotion regulation
over time) to examine patterns of emotion regulation mechanisms and whether they are uniquely
related to posttraumatic stress outcomes among trauma-​exposed youth.
Additionally, emerging research has already begun to test the mediating processes of emo-
tion regulation as proposed in our conceptual framework and has shown evidence for a link
between mass trauma exposure and negative outcomes via emotion regulation (Polusny et  al.,
2011; Marsee, 2008). However, it is still relatively early to draw definitive conclusions on causal
inferences, as these two studies were cross-​sectional and focused on a single cognitive emotion
regulation strategy and general emotional dysregulation. Furthermore, it seems that contextual
factors, such as temperament and parenting behaviors, influence whether youth develop or con-
tinue to have problems following trauma (Halberstadt et al., 2008; Hendricks and Borstein, 2007;
Kithakye et al., 2010; Lengua et al., 2005; Punamäki et al., 2014; Polusny et al., 2011; Terranova
et al., 2009b). Future research will need to address the mediating and moderating role of emotion
regulation in the development and maintenance of youth outcomes using both prospective (i.e.,
pre-​and post-​trauma) and longitudinal designs that can capture time dependent changes in emo-
tion regulation mechanisms and posttraumatic stress outcomes independently and in relation to
one another. Researchers will also need to carefully consider developmental factors, as younger
children may be more prone to using maladaptive strategies (e.g., avoidance) as their access to
high-​order cognitive abilities is limited.
Most of the aforementioned studies also relied on broad coping measures to capture emotion
regulation strategies, though the strategies measured may go beyond specifically changing or
modulating an emotional response and tell us little about youths’ ability to modulate emotions
in non-​stressful situations (Compas et al., 2014). For example, seeking social support or com-
fort from parents, peers, or teachers during a stressful time may serve to ease worries about
losing a house or a loved one following a natural disaster (regulation of negative cognitions)
or blaming others may reduce stress (not necessarily related to an emotion) through migrat-
ing responsibility on environmental factors. Researchers also need to take caution when using
coping measures to predict child psychopathology symptoms or other negative posttraumatic
stress reactions as there is considerable overlap in measure constructs and items (e.g., behav-
ioral avoidance is a DSM-​5 defined symptom of PTSD) and thus increases the odds of relations
being driven by shared method variance (Pfefferbaum, Noffsinger, Wind, & Allen, 2014).
In closing, it is important to note that effective interventions that target emotion regulation
skills are available to youth experiencing mental health difficulties following mass trauma expo-
sure. However, little is known about whether change in emotion regulation (increase in efficacy
or greater or lesser use of specific strategies) is the mechanism that leads to reductions in negative
outcomes following treatment or whether these skills require booster sessions to solidify their
flexible and adaptive usage.
0
9
3

390 Children Exposed to Traumatic Stress

References
Aber, J. L., Gershoff, E. T., Ware, A., & Kotler, J. A. (2004). Estimating the effects of September 11th and
other forms of violence on the mental health and social development of New York City’s youth: A matter
of context. Applied Developmental Science, 8, 111–╉129. doi:10.1207/╉s1532480xads0803_╉2
American Psychiatric Association. (2000). Diagnostic and statistical manual of mental disorders. (4th ed.,
text rev.). Washington, DC: Author.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders. (5th ed.).
Washington, DC: Author.
Asarnow, J. R., Glynn, S. M., Pynoos, R. S., Nahum, J., Guthrie, D., Cantwell, D. P., & Franklin, B.A.
(1999). When the earth stops shaking: earthquake sequelaee among children diagnosed for pre-╉
earthquake psychopathology. Journal of the American Academy of Child and Adolescent Psychiatry, 38,
1016–╉1023. doi:10.1097/╉00004583-╉199908000-╉00018
Attanakyake, V., McKay, R., Joffres, M., Singh, S., Burkle, F. Jr & Mills, E. (2009). Prevelance of mental
disorders among children exposed to war: A systematic review of 7,920 children. Medicine, Conflict,
and Survival, 25, 4–╉19. doi:10.1080/╉13623690802568913
Beauchaine, T. P. (2001). Vagal tone, development, and Gray’s motivational theory: Toward an
integrated model of autonomic nervous system functioning in psychopathology. Development and
Psychopathology, 13, 183–╉214. doi:10.1017/╉S0954579401002012
Becker-╉Blease, K. A., Turner, H. A., & Finkelhor, D. (2010). Disasters, victimization, and children’s mental
health. Child Development, 81, 1040–╉1052. doi:10.1111/╉j.1467-╉8624.2010.01453.x
Buss, K. A., & Goldsmith, H. H. (1998). Fear and anger regulation in infancy: Effects on the temporal
dynamics of affective expression. Child Development, 69, 359–╉374. doi:10.1111/╉j.1467-╉8624.1998.
tb06195.x
Brown, T. H., Mellman, T. A., Alfano, C. A., & Weems, C. F. (2011). Sleep fears, sleep disturbance, and
PTSD symptoms in minority youth exposed to Hurricane Katrina. Journal of Traumatic Stress, 24, 1–╉6.
doi:10.1002/╉jts.20680
Calkins, S. D., & Johnson, M. C. (1998). Toddler regulation of distress to frustrating events: Temperamental
and maternal correlates. Infant Behavior & Development, 21, 379–╉395. doi:10.1016/╉
S0163-╉6383(98)90015-╉7
Campos, J. J., Frankel, C. B., & Camras, L. (2004). On the nature of emotion regulation. Child
Development, 75, 377–╉394. doi:10.1111/╉j.1467-╉8624.2004.00681.x
Cardeña, E., Dennis, J. M., Winkel, M., & Skitka, L. J. (2005). A snapshot of terror: Acute posttraumatic
responses to the September 11 attack. Journal of Trauma and Dissociation, 6, 69–╉84. doi:10.1300/╉
J229v06n02_╉07
Carrión, V. G., Garrett, A., Menon, V., Weems, C. F. & Reiss, A. L. (2008). Posttraumatic stress symptoms
and brain function during a response-╉inhibition task: An fMRI study in youth. Depression and Anxiety,
25, 514–╉526. doi:10.1002/╉da.20346
Carrión, V. G., Weems, C. F., Eliez, S., Patwardhan, A., Brown, W., Ray R., & Reiss, A. L. (2001).
Attenuation of frontal lobe asymmetry in pediatric PTSD. Biological Psychiatry, 50, 943–╉951.
doi:10.1016/╉S0006-╉3223(01)01218-╉5
Carrión, V. G., Weems, C. F., & Reiss, A. L. (2007). Stress predicts brain changes in children: A pilot
longitudinal study on youth stress, PTSD, and the hippocampus. Pediatrics, 119, 509–╉516. doi:10.1542/╉
peds.2006-╉202
Carrión, V. G., Weems, C. F., & Reiss, A. L. (2007). Stressing about posttraumatic stress disorder: In reply.
Pediatrics, 120, 234–╉235. doi:10.1542/╉peds.2007-╉1024
Carrión, V. G., Weems, C. F., Richert, K., Hoffman, B., & Reiss, A. L. (2010). Decreased prefrontal cortical
volume associated with increased bedtime cortisol in traumatized youth. Biological Psychiatry, 68, 491–╉
493. doi:10.1016/╉j.biopsych.2010.05.010
1
9
3

Conclusions 391

Carrión, V. G., & Wong, S. S. (2012). Can traumatic stress alter the brain? Understanding the implications
of early trauma on brain development and learning. Journal of Adolescent Health, 51, S23–​S28.
doi:10.1016/​j.jadohealth.2012.04.010
CATS Consortium (2010). Implementation of CBT for youth affected by the World Trade Center
disaster: Matching need to treatment intensity and reducing trauma symptoms. Journal of Traumatic
Stress, 23, 699–​707. doi: 10.1002/​jts.20594
Chorpita, B. F., & Barlow, D. H. (1998). The development of anxiety: The role of control in the early
environment. Psychological Bulletin, 124, 3–​21. doi:10.1037//​0033-​2909.124.1.3
Cohen, J. A., Deblinger, E., Mannarino, A. P., & Steer, R. A. (2004). A multisite, randomized controlled
trial for children with sexual abuse-​related PTSD symptoms. Journal of the American Academy of Child
and Adolescent Psychiatry, 43, 393–​402. doi:10.1097/​00004583-​200404000-​00005
Cohen, J. A., & Mannarino, A. P. (1996). A treatment outcome study for sexually abused preschool
children: Initial findings. Journal of the American Academy of Child and Adolescent Psychiatry, 35,
42–​50. doi:10.1097/​00004583-​199601000-​00011
Cohen, J. A., & Mannarino, A. P. (2008). Trauma-​focused cognitive behavioral therapy for children and
parents. Child and Adolescent Mental Health, 13, 158–​162. doi:10.1111/​j.1475-​3588.2008.00502.x
Cole, P. M., Martin, S. E., & Dennis, T. A. (2004). Emotion regulation as a scientific
construct: Methodological challenges and directions for child development research. Child
Development, 75, 317–​333. doi:10.1111/​j.1467-​8624.2004.00673.x
Compas, B. E., Jaser, S. S., Dunbar, J. P., Watson, K. H., Bettis, A. H., Gruhn, M. A., & Williams, E. K.
(2014). Coping and emotion regulation from childhood to early adulthood: Points of convergence and
divergence. Australian Journal of Psychology, 66, 1–​11. doi:10.1111/​ajpy.12043
Copeland, W. E., Keeler, G., Angold, A., & Constello, E. J. (2007). Traumatic events and posttraumatic
stress in childhood. Archives of General Psychiatry, 64, 577, 584. doi:10.1001/​archpsyc.64.5.577
Costello, E. J., Erkanli, A., Fairbank, J. A., & Angold, A. (2002). The prevalence of potentially traumatic
events in childhood and adolescence. Journal of Traumatic Stress, 15, 99–​112. doi:10.1023/​
A:1014851823163
Culbert, T. P., Kajander, R. L., & Reaney, J. B. (1996). Biofeedback with children and adolescents: Clinical
observations and patient perspectives. Developmental and Behavioral Pediatrics, 17, 342–​350.
doi:10.1097/​00004703-​199610000-​00009
De Bellis, M. D. (2001). Developmental traumatology: The psychobiological development of maltreated
children and its implications for research, treatment, and policy. Development and Psychopathology, 13,
539–​564.
Deblinger, E. H., & Heflin, A. H. (1996). Treating sexually abused children and their nonoffending
parents: A cognitive behavioral approach. Thousand Oaks: Sage.
Deblinger, E., Stauffer, L. B., & and Steer, R. A. (2001). Comparative efficacies of supportive and cognitive
behavioral group therapies for young children who have been sexually abused and their nonoffending
mothers. Child Maltreatment, 6, 332–​343. doi:10.1177/​1077559501006004006
Deblinger, E., Steer, R. A., & Lippmann, J. (1999). Two-​year follow-​up study of cognitive-​behavioral
therapy for sexually abused children suffering post-​traumatic stress symptoms. Child Abuse & Neglect,
23, 1371–​1378. doi:10.1016/​S0145-​2134(99)00091-​5
de Roos, C., Greenwald, R., den Hollander-​Gijsman, M., Noorthoorn, E., van Buuren, S., & de Jongh,
A. (2011). A randomized comparison of cognitive behavioural therapy (CBT) and eye movement
desensitisation and reprocessing (EMDR) in disaster-​exposed children. European Journal of
Psychotraumatology, 2, 5694. doi:10.3402/​ejpt.v2i0.5694
De Young, A. C., Kenardy, J. A., & Spence, S. H. (2007). Elevated heart rate as a predictor of PTSD six
month following accidental pediatric injury. Journal of Traumatic Stress, 20, 751–​756. doi:10.1002/​
jts.20235
2
9
3

392 Children Exposed to Traumatic Stress

Diener, M. L., & Mangelsdorf, S. C. (1999). Behavioral strategies for emotion regulation in
toddlers: Associations with maternal involvement and emotional expressions. Infant Behavior and
Development, 22, 569–​583. doi:10.1016/​S0163-​6383(00)00012-​6
Eisenberg, N., & Silver, R. C. (2011). Growing up in the shadow of terrorism: Youth in America after
September 11th. American Psychologist, 66, 468. doi:10.1037/​a0024619
Feldman, R., Vengrober, A., Eidelman-​Rothman, M., & Zagoory-​Sharon, O. (2013). Stress reactivity in
war-​exposed young children with and without posttraumatic stress disorder: Relations to maternal
stress hormones, parenting, and child emotionality and regulation. Development and Psychopathology,
25, 943–​955. doi:10.1017/​S0954579413000291
Fernandez, I. (2007). EMDR as treatment of post-​traumatic reactions: A field study on child victims of an
earthquake. Educational & Child Psychology, 24, 65–​72.
Frijda, N. H. (1986). The emotions: Studies in emotion and social interaction. Cambridge, England:
Cambridge University Press.
Furr, J. M, Comer, J. S., Edmunds, J. M., & Kendall, P. C. (2010). Disasters and youth: A meta-​analytic
examination of posttraumatic stress. Journal of Consulting and Clinical Psychology, 78, 765–​780.
doi:10.1037/​a0021482
Goenjian, A. K., Walling, D., Steinberg, A. M., Karayan, I., Najarian, L. M., & Pynoos, R. (2005). A
prospective study of posttraumatic stress and depressive reactions among treated and untreated
adolescents 5 years after a catastrophic disaster. American Journal of Psychiatry, 162, 2302–​2308.
doi:10.1176/​appi.ajp.162.12.2302
Goldstein, R. D., Wampler, N. S., & Wise, P. H. (1995). War experiences and distress symptoms of Bosnian
children. Pediatrics, 100, 873–​878. doi:10.1542/​peds.100.5.873
Gross, J. J. (1998a). The emerging field of emotion regulation: An integrative review. Review of General
Psychology, 2, 271–​299. doi:10.1037/​1089-​2680.2.3.271
Gross, J. J. (1998b). Antecedent-​and response-​focused emotion regulation: Divergent consequences for
experience, expression, and physiology. Journal of Personality and Social Psychology, 74, 224–​237.
doi:10.1037/​0022-​3514.74.1.224
Gross, J. J. (1999). Emotion regulation: Past, present, future. Cognition & Emotion, 13(5), 551–​573.
Gross, J. J., & Thompson, R. A., (2007). Emotion regulation: Conceptual Foundations. In J. J. Gross (Ed),
Handbook of Emotion Regulation (pp. 3–​24). New York, NY US: Guilford Press.
Gullone, E., Hughes, E. K., King, N. J., & Tonge, (2010). The normative development of emotion regulation
strategy use in children and adolescents: a 2-​year follow-​up study. The Journal of Child Psychology and
Psychiatry, 51, 567–​574. doi:10.1111/​j.1469-​7610.2009.02183.x
Gunnar, M. R. (2001). The role of glucocorticoids in anxiety disorders: A critical analysis. In: Vasey M. W.,
Dadds M. R. (Eds). The developmental psychopathology of anxiety. (pp. 143–​159) New York, NY: Oxford
University Press.
Halberstadt, A. G., Thompson, J. A., Parker, A. E., & Dunsmore, J. C. (2008). Parents’ emotion‐related
beliefs and behaviours in relation to children’s coping with the 11 September 2001 terrorist attacks.
Infant and Child Development, 17, 557–​580. doi:10.1002/​icd.569
Heim, C., & Nemeroff, C. B. (2001). The role of childhood trauma in the neurobiology of mood and anxiety
disorders: preclinical and clinical studies. Biological Psychiatry, 49, 1023–​1039. doi:S0006-​3223(01)01157-​X
Hendricks, C., & Bornstein, M. H. (2007). An ecological analysis of early adolescents’ trauma reactions
to 9-​11 in the Washington, DC, area. Applied Developmental Science, 11, 71–​88. doi: 10.1080/​
10888690701384905
Hensley, L., & Varela, R. E. (2008). PTSD symptoms and somatic complaints following Hurricane
Katrina: The role of trait anxiety and anxiety sensitivity. Journal of Clinical Child and Adolescent
Psychology, 37, 542–​552. doi:10.1080/​15374410802148186
Hoven, C. W., Duarte, C. S., Lucas, C. P., Wu, P., Mandell, D. J., Goodwin, R. D., et al. (2005).
Psychopathology among New York City public school children 6 months after September 11. Archives
of General Psychiatry, 62, 545–​552. doi:10.1001/​archpsyc.62.5.545
3
9

Conclusions 393

James, W. (1884). II.—​What is an emotion?. Mind, 34, 188–​205.


Jaycox LH. Cognitive-​behavioral intervention for trauma in schools. Longmont, CO: Sopris West
Educational Services; 2003.
Jaycox, L. H., Cohen, J. A., Mannarino, A. P., Walker, D. W., Langly, A. K., Gegenheimer, K. L., Scott,
M., & Schonla, M. (2010). Children’s mental health care following Hurricane Katrina: A field trial of
trauma-​focused psychotherapies. doi:10.1002/​jts.20518
Jeney-​Gammon P., Daugherty T. K., Finch A. J., Belter R. W., Foster K. Y. (1993). Children’s coping styles
and report of depressive symptoms following a natural disaster. Journal of Genetic Psychology, 154, 259–​
267. doi:10.1080/​00221325.1993.9914739
Kataoka, S. H., Stein, B. D., Jaycox, L. H., Wong, M., Escudero, P., Tu, W., Zaragoza, C., & Fink,
A. (2003). A school-​based mental health program for traumatized latino immigrant children.
Journal of the American Academy of Child and Adolescent Psychiatry, 42, 311–​318. doi:10.1097/​
00004583-​200303000-​00011
Kennedy, C., Charlesworth, A., & Chen, J. (2004). Disaster at a distance: Impact of 9.11.01 televised new coverage
on mothers’ and children’s health. Journal of Pediatric Nursing, 19, 329–​339. doi: 10.1016/​j.pedn.2004.09.003
Kithakye M., Morris A. S., Terranova A. M., Myers S. S. (2010) The Kenyan political conflict and children’s
adjustment. Child Development, 81, 1114–​1128. doi:10.1111/​j.1467-​8624.2010.01457.x
Kuterovac, G., Dyregrov, A., & Stuvland, R. (1994). Children in war: a silent majority under stress. British
Journal of Medical Psychology, 67, 363–​375. doi:10.1111/​j.2044-​8341.1994.tb01804.x
La Greca, A. M., Silverman, W. K. (2009). Treatment and prevention of posttraumatic stress reactions in
children and adolescents exposed to disasters and terrorism: What is the evidence? Child Development
Perspectives, 3, 4–​10. doi:10.1111/​j.1750-​8606.2008.00069.x
La Greca, A. M., Silverman, W. K., Vernberg, E. M., & Prinstein, M.J. (1996). Symptoms of posttraumatic
stress in children after Hurricane Andrew: A prospective study. Journal of Consulting and Clinical
Psychology, 64, 712–​723. doi:10.1037/​0022-​006X.64.4.712
La Greca, A. M., Silverman, W. K., Vernberg, E. M., & Roberts, M. C. (Eds.). (2002). Helping Children Cope
with Disasters and Terrorism. Washington, DC: American Psychological Association.
La Greca, A. M., Silverman, W. K., & Wasserstein, S. B. (1998). Children’s predisaster functioning as
a predictor of posttraumatic stress following Hurricane Andrew. Journal of Consulting and Clinical
Psychology, 66, 883–​892. doi:10.1037/​0022-​006X.66.6.883
Lanius, R. A., Frewen, P. A., Vermetten, E., & Yehuda, R. (2010). Fear conditioning and early life
vulnerabilities: two distinct pathways of emotional dysregulation and brain dysfunction in PTSD.
European Journal of Psychotraumatology, 1, 5467. doi:10.3402/​ejpt.v1i0.5467
Layne C. M., Saltzman W. R., Pynoos R. S., Steinberg A. M. Trauma and Grief Component Therapy.
New York State Office of Mental Health; New York, NY: 2002.
Layne, C. M., Pynoos, R. S., Saltzman, W. R., Arslanagić, B., Black, M., Savjak, N., … & Houston, R.
(2001). Trauma/​grief-​focused group psychotherapy: school-​based postwar intervention with
traumatized Bosnian adolescents. Group Dynamics: Theory, Research, and Practice, 5, 277. doi:10.1037/​
1089-​2699.5.4.277
Lengua, L. J., Long, A. C., Smith, K. I., & Meltzoff, A. N. (2005). Pre-​attack symptomatology and
temperament as predictors of children’s responses to the September 11 terrorist attacks. Journal of Child
Psychology and Psychiatry, 46, 631–​645. doi:10.1111/​j.1469-​7610.2004.00378.x
Luo, H., Hu, X., Liu, X., Ma, X., Guo, W., Qiu, C., … & Li, T. (2012). Hair cortisol level as a biomarker
for altered hypothalamic-​pituitary-​adrenal activity in female adolescents with posttraumatic
stress disorder after the 2008 Wenchuan earthquake. Biological Psychiatry, 72, 65–​69. doi:10.1016/​
j.biopsych.2011.12.020
March, J. S., Amaya-​Jackson, L., Murray, M. C., & Schulte, A. (1998). Cognitive‐behavioral psychotherapy
for children and adolescents with Posttraumatic Stress Disorder after a single‐incident stressor.
Journal of the American Academy of Child & Adolescent Psychiatry, 37, 585–​593. doi:10.1097/​
00004583-​199806000-​00008
4
9
3

394 Children Exposed to Traumatic Stress

Marsee, M. A. (2008). Reactive aggression and posttraumatic stress in adolescents affected by


Hurricane Katrina. Journal of Clinical Child and Adolescent Psychology, 39, 519–​529. doi:10.1080/​
15374410802148152
Masten, A. S., & Narayan, A. J. (2012). Child development in the context of disaster, war, and
terrorism: Pathways of risk and resilience. Annual Review of Psychology, 63, 227–​257. doi:10.1146/​
annurev-​psych-​120710-​100356
Mitlin, M. (1998). Emotional bingo. Los Angeles: Western Psychological Services.
Moradi, A. R., Taghavi, M. R., Neshat Doost, H. T., Yule, W. and Dalgleish, T. (1999). The performance of
children and adolescents with PTSD on the Stroop colour-​naming task. The Journal of Child Psychology
and Psychiatry, 29, 415–​419. doi:10.1017/​S0033291798008009
Nader, K. O. & Weems, C. F. (2011). Understanding and assessing cortisol levels in children and
adolescents. Journal of Child and Adolescent Trauma, 4, 318–​338. doi:10.1080/​19361521.2011.624059
Noppe, I. C., Noppe, L. D., & Bartell, D. (2006). Terrorism and resilience: Adolescents’ and teachers’
responses to September 11, 2001. Death Studies, 30, 41–​60. doi:10.1080/​07481180500348761
Norris, F. H., Friedman, M. J., & Watson, P. J. (2002). 60,000 disaster victims speak: Part II. Summary
and implications of the disaster mental health research. Psychiatry, 65, 240–​260. doi:101521/​
psyc.65.3.240.20169
Oschner, K.N., & Gross, J.J. (2005). The cognitive control of emotion. TRENDS in Cognitive Sciences, 9,
242–​249. doi:10.1016/​j.tics.2005.03.010
Osofsky, H. J., Osofsky, J. D., Kronenberg, M., Brennan, A., & Hansel, T. C. (2009). Posttraumatic
stress symptoms in children after Hurricane Katrina: Predicting the need for mental health services.
American Journal of Orthopsychiatry, 79, 212–​220. doi: 10.1037/​a0016179
Pfefferbaum, B., Sweeton, J. L., Newman, E., Varma, V., Nitiéma, P., Shaw, J. A., … & Noffsinger, M. A.
(2014). Child disaster mental health interventions, part I: Techniques, outcomes, and methodological
considerations. Disaster Health, 2, 46–​57. doi:10.4161/​dish.27534
Pina, A. A., Villalta, I. K., Ortiz, C. D., Gottschall, A. C., Costa, N. M., & Weems, C. F. (2008). Social
support, discrimination, and coping as predictors of posttraumatic stress reactions in youth survivors
of Hurricane Katrina. Journal of Clinical Child and Adolescent Psychology, 37, 564–​574. doi:10.1080/​
15374410802148228
Polusny, M. A., Ries, B. J., Meis, L. A., DeGarmo, D., McCormick-​Deaton, C. M., Thuras, P., & Erbes,
C. R. (2011). Effects of parents’ experiential avoidance and PTSD on adolescent disaster-​related
posttraumatic stress symptomatology. Journal of Family Psychology, 25, 220–​229. doi:10.1037/​a0022945
Porche, M. V., Fortuna, L. R., Lin, J., & Alegria, M. (2011). Childhood Trauma and Psychiatric Disorders as
Correlates of School Dropout in a National Sample of Young Adults. Child Development, 82, 982–​998.
doi:10.1111/​j.1467-​8624.2010.01534.x
Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74, 116–​143. doi:10.1016/​
j.biopsycho.2006.06.009
Porges, S. W., Doussard-​Roosevelt, J. A., & Maiti, A. K. (1994). Vagal tone and the physiological regulation
of emotion. Monographs of the society for research in child development, 59, 167–​186. doi:10.1111/​
j.1540-​5834.1994.tb01283.x
Prati, G., & Pietrantoni, L. (2009). Optimism, social support, and coping strategies as factors contributing
to posttraumatic growth: A meta-​analysis. Journal of Loss and Trauma, 14, 364–​388. doi:10.1080/​
15325020902724271
Prinstein, M. J., La Greca, A. M., Vernberg, E. M., & Silverman, W. K. (1996). Children’s coping
assistance: How parents, teachers, and friends help children cope after a natural disaster. Journal of
Clinical Child Psychology, 25, 463–​475. doi:10.1207/​s15374424jccp2504_​11
Punamäki, R. L., Peltonen, K., Diab, M., & Qouta, S. R. (2014). Psychosocial interventions and emotion
regulation among war-​affected children: Randomized control trial effects. Traumatology, 20, 241–​252.
doi:10.1037/​h0099856
5
9
3

Conclusions 395

Pynoos, R. S., Goenjian, A. K., Steinberg. A. M. (1998). A public mental health approach to the post-​
disaster treatment of children and adolescents. Psychiatric Clinics of North America, 7, 195–​210.
Pynoos, R. S., Steinberg, A. M., & Piacentini, J. C. (1999). A developmental psychopathology model of
childhood traumatic stress and intersection with anxiety disorders. Biological Psychiatry, 46, 1542–​
1554. doi:10.1016/​S0006-​3223(99)00262-​0
Richards, J. M., & Gross, J. J. (2000). Emotion regulation and memory: The cognitive costs of keeping one’s
cool. Personality Processes and Individual Differences, 79, 410–​424. doi:1O.1037/​70O22-​3514.79.3.410
Richert, K. A., Carrión, V. G., Karchemskiy, A., & Reiss, A. L. (2006). Regional differences of the prefrontal
cortex in pediatric PTSD: an MRI study. Depression and Anxiety, 23, 17–​25. doi:10.1002/​da.20131
Roberts, Y. H., Mitchell, M. J., Witman, M. W., & Taffaro, C. (2010). Mental health symptoms in youth
affected by Hurricane Katrina. Professional Psychology: Research and Practice, 41, 10–​18. doi:10.1037/​
a0018339
Russoniello, C. V., Skalko, T. K., O’Brien, K., McGhee, S. A., Bingham-​Alexander, D., & Beatley, J. (2002).
Childhood posttraumatic stress disorder and efforts to cope after Hurricane Floyd. Behavioral Medicine,
28, 61–​71. doi:10.1080/​08964280209596399
Salloum, A., Garfield, L., Irwin, A., Anderson, A., & Francois A. (2009). Grief and trauma group
therapy with children after Hurricane Katrina. Social Work with Groups, 32, 67–​79. doi:10.1080/​
01609510802290958
Salloum, A., & Overstreet, S. (2008) Evaluation of individual and group grief and trauma interventions
for children post disaster. Journal of Clinical Child & Adolescent Psychology, 37, 495–​507. doi:10.1080/​
15374410802148194
Salloum, A., & Overstreet, S. (2012). Grief and trauma intervention for children after disaster: Exploring
coping skills versus trauma narration. Behaviour Research and Therapy, 50, 169–​179. doi:0.1016/​
j.brat.2012.01.001
Scheeringa, M. S., Zeanah, C. H., Myers, L., & Putnam, F. (2004). Heart period and variability findings in
preschool children with posttraumatic stress symptoms. Biological Psychiatry, 55, 685–​691. doi:10.1016/​
j.biopsych.2004.01.006
Scheeringa, M. S., Weems, C. F., Cohen, J. A., Amaya-​Jackson, L., & Guthrie, D. (2011). Trauma-​
focused cognitive-​behavioral therapy for posttraumatic stress disorder in three through six year-​old
children: A randomized clinical trial. Journal of Child Psychology and Psychiatry, and Allied Disciplines,
52(8), 853–​860. http://​doi.org/​10.1111/​j.1469-​7610.2010.02354.x
Scott, B. G., & Weems, C. F. (2014). Resting vagal tone and vagal response to stress: Associations with
anxiety, aggression, and perceived anxiety control among youths. Psychophysiology, 51, 718–​727.
doi:10.1111/​psyp.12218
Scott, B. G., Larpré, G. E., Marsee, M. A., & Weems, C. F. (2014). Aggressive behavior and its associations
with posttraumatic stress and academic achievement following a natural disaster. Journal of Clinical
Child and Adolescent Psychology, 43, 43–​50. doi:10.1080/​15374416.2013.807733
Seedat, S., Nyamai, C., Njenga, F., Vythilingum, B., & Stein, D. J. (2004). Trauma exposure and post-​
traumatic stress symptoms in urban African schools. The British Journal of Psychiatry, 184, 169–​175.
doi:10.1192/​bjp.184.2.169
Shapiro, F. (2007). EMDR and case conceptualization from an adaptive information processing perspective.
In Shapiro, F., Kaslow, F. W., & Maxfield, L. (Eds.), Handbook of EMDR and Family Therapy Processes
(pp. 3–​34). Guilford Press: New York.
Shiner, R. L. (1998). How shall we speak of children’s personalities in middle childhood? A preliminary
taxonomy. Psychological Bulletin, 124, 308–​332. doi:10.1037/​0033-​2909.124.3.308
Silverman, W. K., Ortiz, C. D., Viswesvaaran, C., Burns, B. J., Kolko, D. J., Putnam, F. W., & Amaya-​
Jackson, L. (2008). Evidence-​based psychosocial treatments for children and adolesscents exposed
to traumatic events. Journal of Clinical Child and Adolescent Psychology, 37, 156–​183. doi:10.1080/​
15374410701818293
6
9
3

396 Children Exposed to Traumatic Stress

Silverman, W. K., Pina, A. A., Viswesvaran, C. (2008). Evidence-​based psychosocial treatments for phobic
and anxiety disorders in children and adolescents. Journal of Clinical Child and Adolescent Psychology,
37, 105–​130. doi:10.1080/​15374410701817907
Stein, B. D., Jaycox, L. H., Kataoka, S. H., Wong, M., Tu, W., Elliot, M. N., & Fink, A. (2003). A mental
health intervention for school children exposed to violence. The Journal of the American Medical
Association, 290, 603–​611. doi:10.1001/​jama.290.5.603
Taylor, L. K., & Weems, C. F. (2011). Cognitive-​behavior therapy for disaster-​exposed youth with
posttraumatic stress: Results from a multiple-​baseline examination. Behavior Therapy, 42, 349–​363.
doi:10.1016/​j.beth.2010.09.001
Tedeschi, R. G., & Calhoun R. G. (2004). Posttraumatic growth: Conceptual foundations and empirical
evidence. Psychological Inquiry, 14, 1–​18. doi:10.1207/​s15327965pli1501_​01
Terranova, A. M., Boxer, P., & Morris, A. S. (2009a). Changes in children’s peer interactions following a
natural disaster: How predisaster bullying and victimization rates changed following Hurricane Katrina.
Psychology in the Schools, 46, 333–​347. doi:10.1002/​pits.20379
Terranova, A. M., Boxer, P., & Morris, A. S. (2009b). Factors influencing the course of posttraumatic
stress following a natural disaster: Children’s reactions to Hurricane Katrina. Journal of Applied
Developmental Psychology, 30, 344–​355. doi:10.1016/​j.appdev.2008.12.017
Thayer, J. F., & Lane, R. D. (2000). A model of neurovisceral integration in emotion regulation and
dysregulation. Journal of Affective Disorders, 61, 201–​216. doi:10.1016/​S0165-​0327(00)00338-​4
Thompson, R. A. (1994). Emotion regulation: A theme in search of definition. Monographs of the Society for
Research in Child Development, 59, 25–​52. doi:10.1111/​j.1540-​5834.1994.tb01276.x
Vernberg, E. M., Silverman, W. K., La Greca, A. M., & Prinstein, M. J. (1996). Prediction of posttraumatic
stress symptoms in children after Hurricane Andrew. Journal of Abnormal Psychology, 105, 237–​248.
doi:10.1037/​0021-​843X.105.2.237
Vigil, J. M., Geary, D. C., Granger, D. A., & Flinn, M. V. (2010). Sex differences in salivary cortisol,
alpha‐amylase, and psychological functioning following Hurricane Katrina. Child Development, 81,
1228–​1240. doi:10.1111/​j.1467-​8624.2010.01464.x
Wadsworth, M. E., Gudmundsen, G. R., Raviv, T., Ahlkvist, J. A., McIntosh, D. N., Kline, G.H.,
… & Burwell, R. A. (2004). Coping with terrorism: Age and gender differences in effortful and
involuntary responses to September 11th. Applied Developmental Science, 8, 143–​157. doi:10.1207/​
s1532480xads0803_​4
Wagner, K. D., Brief, D. J., Vielhauer, M. J., Sussman, S., Keane, T. M., & Malow, R. (2009). The Potential
for PTSD, Substance Use, and HIV Risk Behavior among Adolescents Exposed to Hurricane Katrina.
Substance Use & Misuse, 44, 1749–​1767. doi:10.3109/​10826080902963472
Weems, C. F., & Carrión, V. G. (2009). Diurnal salivary cortisol in youth: Clarifying the nature of post
traumatic stress dysregulation. Journal of Pediatric Psychology, 34, 389–​395. doi:10.1093/​jpepsy/​jsn087
Weems, C. F., & Graham, R. A. (2014). Resilience and trajectories of posttraumatic stress among youth
exposed to disaster. Journal of Child and Adolescent Psychopharmacology, 24(1), 2–​8.
Weems, C. F. & Overstreet, S. (2008). Child and adolescent mental health research in the context of
Hurricane Katrina: An ecological-​needs-​based perspective and introduction to the special section.
Journal of Clinical Child and Adolescent Psychology, 37, 487–​494. doi:10.1080/​15374410802148251
Weems, C. F., Russell, J.D., Graham, R. A., Neill, E. L. & Banks, D. M. (2015). Developmental differences
in the linkages between anxiety control beliefs and post-​traumatic stress in youth. Depression &
Anxiety, 32, 356–​363. doi:10.1002/​da.22319
Weems, C. F., Scott, B. G., Graham, R. A., Banks, D. M. Russell, J.D. Taylor, L. K., Cannon, M. Varela,
R. E., Scheeringa, M. S. Perry A. M. & Marino, R., C. (2014). Fitting anxious emotion focused
intervention into the ecology of schools: Results from a test anxiety program evaluation. Prevention
Science. doi:10.1007/​s11121-​014-​0491-​1
7
9
3

Conclusions 397

Weems, C. F., Scott, B. G., Russell, J. D., Reiss, A. L., & Carrión, V. G. (2013). Developmental variation
in amygdala volumes among children with a history of exposure to traumatic stress. Developmental
Neuropsychology, 38, 481–​495. doi:10.1080/​87565641.2013.820307
Weems, C. F., Taylor, L. K., Costa, N. M., Marks, A. B., Romano, D. M., Verrett, S. L., & Brown, D. M.
(2009). Effect of a school-​based test anxiety intervention in ethnic minority youth exposed to Hurricane
Katrina. Journal of Applied Developmental Psychology, 30, 218–​226. doi:10.1016/​j.appdev.2008.11.005
Weems, C. F., Piña, A. A., Costa, N. M., Watts, S. E., Taylor, L. K., & Cannon, M. F. (2007). Pre-​disaster
trait anxiety and negative affect predict posttraumatic stress in youth after hurricane Katrina. Journal of
Consulting and Clinical Psychology, 75, 154–​159. doi:10.1037/​0022-​006X.75.1.154
Weems, C. F., & Carrión, V. G. (2007). The association between PTSD symptoms and salivary cortisol in
youth: The role of the time since the trauma. Journal of Traumatic Stress, 20, 903–​907. doi:10.1002/​
jts.20251
Weems, C. F., Scott, B. G., Taylor, L. K., Cannon, M. F., Romano, D. M., & Perry, A. M. (2013). A
theoretical model of continuity in anxiety and links to academic achievement in disaster exposed
school children. Development and Psychopathology, 25, 729–​738. doi:10.1017/​S0954579413000138
Weems, C. F., Taylor, L. K., Cannon, M. F., Marino, R. C., Romano, D. M., Scott, B. G., … & Triplett, V.
(2010). Post traumatic stress, context, and the lingering effects of the Hurricane Katrina disaster among
ethnic minority youth. Journal of Abnormal Child Psychology, 38(1), 49–​56.
Weems, C. F., Zakem, A. H., Costa, N. M., Cannon, M. F., & Watts, S. E. (2005). Physiological response
and childhood anxiety: Association with symptoms of anxiety disorders and cognitive bias. Journal of
Clinical Child and Adolescent Psychology, 34, 712–​723. doi:10.1207/​s15374424jccp3404_​13
Yehuda, R., Kahana, B., Binder-​Brynes, K., Southwick, S. M., Mason, J. W., & Giller, E. L. (1995). Low
urinary cortisol excretion in Holocaust survivors with posttraumatic stress disorder. American Journal
of Psychiatry, 152, 982–​986. doi:10.1016/​0006-​3223(94)91000-​6
Yehuda, R. (2006). Advances in understanding neuroendocrine alterations in PTSD and their therapeutic
implications. Annals of the New York Academy of Sciences, 1071, 137–​166. doi:10.1196/​annals.1364.012
Zhang, Y., Kong, F., Wang, L., Chen, H., Gao, X., Tan, X., … & Liu, Y. (2010). Mental health and coping
styles of children and adolescent survivors one year after the 2008 Chinese earthquake. Children and
Youth Services Review, 32, 1403–​1409. doi:10.1016/​j.childyouth.2010.06.009
8
9
3

Chapter 19

Adolescents who Engage in Nonsuicidal


Self-╉Injury (NSSI)
David Voon & Penelope Hasking

Nonsuicidal self-injury
Nonsuicidal self-╉injury (NSSI) is the deliberate damage to the body in the absence of fatal intent
(Nock, 2009). As well as being a symptom criterion for a diagnosis of Borderline Personality Disorder
(BPD; American Psychiatric Association, APA, 2013), NSSI is uniquely associated with symptoms
of depression, anxiety, substance abuse, and reduced well-╉being (Dilberto & Nock, 2008; Giletta,
Scholte, Engels, Ciairano, & Prinstein, 2012; Hankin & Abela, 2011; Hilt, Nock, Llloyd-╉Richardson,
& Prinstein, 2008). NSSI is distinguishable from other self-╉harm behaviors, such as substance use,
where the harmful consequences of the behavior are usually unintended, and from suicidal behavior
where the consequences are intended to be fatal (Nock, 2012). Yet, although distinct from suicidal
behavior, NSSI is a risk factor for later suicide (Klonsky, May, & Glenn, 2013; Whitlock et al., 2013).
Highlighting the transdiagnostic nature of NSSI (for discussion see Bentley, Cassiello-╉Robbins,
Vittorio, Sauer-╉Zavala, & Barlow, 2015), and the significant impact of the behavior on psychological
health and well-╉being, the APA recently included NSSI as a condition requiring further research in
the Diagnostic and Statistical Manual of Mental Disorders, Fifth Edition (DSM-╉5; APA, 2013).
Theoretical perspectives on NSSI suggest poor emotion regulation is critical to the aetiology
and maintenance of the behavior. Regulation of negative emotional states is frequently cited as the
predominant motivation for engaging in the behavior (Klonsky, 2009; Martin et al., 2010; Nock
& Prinstein, 2004; Nock, Prinstein, & Sterba, 2009), and evidence-╉based interventions focus-
ing on improving individuals’ skills and capacity for emotion regulation hold promise for the
treatment of NSSI (e.g., Gratz & Tull, 2011). “Emotion regulation” in this context refers to a set
of responses that contribute to initiating, maintaining and modifying the occurrence, intensity,
duration and expression of emotion (Gross, 1998a, 1998b). These responses are both conscious
and unconscious, and may be both effortful and automatic (Gross, 1998a, 1998b; Koole, 2009).
Specific emotion regulation processes that have been implicated in NSSI, which will be focused
on in this chapter, include rumination, cognitive reappraisal, and expressive suppression (Armey
& Crowther, 2008; Hasking, Momeni, Swannell, & Chia, 2008; Hasking, Coric, Swannell, Martin,
Thompson, & Frost, 2010; Hilt, Cha, & Nolen-╉Hoeksema, 2008; Martin et al., 2010).
In the next sections, we describe the nature and extent of NSSI, differentiating it from suicidal
behavior. We broadly discuss the role of emotion regulation in the development, expression, and
maintenance of the behavior, and discuss specific emotion regulation processes and how they
might be implicated in NSSI. Finally, we discuss intervention approaches for NSSI and the evidence
related to their effectiveness. We conclude with the observation that although there are currently
no treatments developed specifically for NSSI among adolescents, and that treatment approaches
for self-╉harm (broadly defined) have inconclusive evidence regarding their effectiveness for this
population, findings from both empirical and intervention research, provide promising leads.
93

THEORETICAL PERSPECTIVES ON NSSI 399

The nature and extent of NSSI


Although NSSI can occur throughout the lifespan, typically it first occurs during adoles-
cence, making this a prime developmental period for prevention and early intervention efforts
(Hankin & Abela, 2011; Jacobson & Gould, 2007; Nock, 2009). International prevalence rates
of NSSI among adolescents have been estimated at 12.5% to 23.6%, with rates decreasing to
approximately 13% in young adults (Muehlenkamp, Claes & Plener, 2012; Swannell, Martin,
Page, Hasking & St. John, 2014). Common forms of NSSI include skin cutting, scratching, self-╉
battery, biting, and burning; with boys/╉men more likely to engage in self-╉battery than girls.
Much of the work describing NSSI observes no gender differences in the prevalence of the
behavior; however, gender differences are reported in the preferred means of NSSI (Bjärehed,
Wångby-╉Lundh, & Lundh, 2012; Sornberger, Heath, Toste, & McLouth, 2012; You, Leung, Fu &
Lai, 2011). A recent meta-╉analysis reported a slightly higher prevalence among women (Plener,
Schumacher, Munz & Groschwitz, 2015), with this difference particularly apparent in clinical
samples (Bresin & Schoenleber, 2015).
In early work, researchers studied NSSI alongside suicidal behavior, arguing both are
characterized by intentional self-╉directed harm. Terms such as “parasuicide” and “deliberate
self-╉harm” have been used to describe NSSI, without differentiating behaviors which may be
performed with or without conscious suicidal intent. Conflating the two behaviors however
obscures their differences. Apart from the absence of fatal intent, several authors have noted
that suicidal and non-╉suicidal self-╉injurious behaviors differ in prevalence, frequency and
methodologies (Hamza, Stewart & Willoughby, 2012; Klonsky et  al., 2013). Highlighting the
contrast between NSSI and suicidal behavior, rates of suicidal behavior among adolescents are
estimated at between 1.3% to 10.1% (Bridge, Goldstein & Brent, 2006). NSSI typically involves
methods that are non-╉lethal, while suicide attempts typically involve more lethal methods such
as hanging, poisoning, and drug overdose (Bridge et al., 2006; McNamara, 2013). While both
behaviors do occur in the context of intense negative emotional states, Walsh (2005) observed
that “the intent of the self-╉injuring person is not to terminate consciousness (as in suicide) but
instead, to modify it” (p. 7).
Although frequent NSSI increases risk of suicidal thoughts and behavior, it ought to be noted
that individuals who attempt suicide and/╉or experience suicidal ideation are equally likely to
report concurrent or future NSSI (Whitlock et al., 2013). In other words, examination of the tem-
poral relationship between NSSI and suicidal thoughts and behaviors does not support a direct
causal link. While NSSI may serve as a “gateway” to concurrent and/╉or future suicidal thoughts
and behaviors, several psychological and social indicators including lack of social connectedness,
less meaning in life, and poorer access to mental health services, can play a critical role in increas-
ing suicide risk. Nonetheless, reducing the frequency of NSSI among individuals who engage in
the behavior is likely to be protective.

Theoretical perspectives on NSSI


There are several perspectives on why individuals might engage in NSSI to regulate their emo-
tional states. Nock (2009) postulates that there are distal, proximal and NSSI-╉specific vulnera-
bilities which increase the likelihood of NSSI. Briefly, he identifies distal vulnerabilities such as a
genetic predisposition to emotional reactivity and sensitivity, and invalidating childhood environ-
ments; proximal vulnerabilities such as maladaptive coping and poor communication; and NSSI-╉
specific vulnerabilities, including social learning and the release of endogenous opioids associated
with physical injury, which might explain why some individuals engage in the behavior while
others do not. It is beyond the scope of this chapter to discuss these various vulnerabilities. In
04

400 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

this section we present theories and perspectives which highlight emotion regulation as a critical
factor in understanding NSSI.

Invalidating childhood experiences and NSSI


Linehan (1993) attempted to explain self-╉harm behaviors such as NSSI among individuals with
BPD and theorized that the propensity to engage in NSSI to regulate negative affective states
may be due to invalidating childhood experiences. According to Linehan, individual vulnerabili-
ties such as sensitivity to emotional stimuli and emotional reactivity interact with invalidating
childhood environments to contribute to emotion regulation deficits in later life. The child learns
to control his/╉her emotional expressiveness and, over time, becomes more restricted in his/╉her
capacity to appropriately identify and regulate emotional arousal, tolerate distress, and trust his/╉
her emotional responses. As a consequence, emotional expression becomes increasingly charac-
terized by extreme inhibition or extreme disinhibition. Emotions are experienced as intense with
limited ability to effectively regulate them except through self-╉harm behavior.
Extending Linehan’s theory, Lang and Sharma-╉Patel (2011) suggest that in invalidating child-
hood environments, individuals’ emotions are rarely reflected, accepted or clarified; furthermore,
they have limited models of effective emotion regulation and limited opportunities to practice
these during childhood. These experiences also contribute to representations of the world as
threatening, others as unreliable and the self as inept. Consequently, individuals develop a height-
ened sense of danger and experience high levels of arousal when confronted with stressful situa-
tions and life events. They may perceive that they have limited internal and external resources to
cope with stressors and use NSSI to help alleviate the arousal associated with stressful situations.
Empirical support for the contribution of invalidating childhood experiences to engagement in
NSSI exists predominantly in studies on childhood abuse. There is evidence from cross-╉sectional
studies that childhood physical and sexual abuse (Fliege, Lee, Grimm & Klapp, 2009; Gratz, 2003;
Gratz & Chapman, 2007) and emotional neglect (Fliege et al., 2009; Gratz, 2003) are associated
with NSSI. More pertinent to this discussion are research findings that highlight emotion dysreg-
ulation as a distinguishing characteristic among individuals with abuse histories who engage in
NSSI and individuals with similar histories who do not. Several studies report emotion dysregu-
lation as a significant predictor of NSSI over and above the presence of childhood maltreatment
(Gratz, 2006; Gratz & Chapman, 2007; Gratz & Roemer, 2008; Muehlenkamp, Kerr, Bradley &
Larson, 2010). In this regard, emotion dysregulation refers to a constellation of factors including
non-╉acceptance of emotional experience, difficulties with goal-╉directed behaviors and impulse
control, lack of emotional awareness, limited number, and restricted access to emotion regula-
tion strategies, and lack of emotional clarity. Sim, Adrian, Zeman, Cassano and Friedrich (2009)
found that poor awareness of emotion and reluctance to express emotion were partial media-
tors of the abuse-╉NSSI relationship among female adolescents in a clinical setting. Swannell and
colleagues (2012) reported difficulties with identifying and articulating emotion (alexithymia)
partially mediated the relationships between physical abuse and neglect, and NSSI among females
aged 18 years and over. Taken together, the empirical evidence suggests while invalidating child-
hood experiences may contribute to the aetiology of NSSI, poor emotion regulation is the likely
mechanism explaining this increased risk. As such, poor emotion regulation is likely to maintain
NSSI for people raised in invalidating childhood environments.

Maladaptive coping, experiential avoidance and NSSI


Lang and Sharma-╉Patel’s (2011) observation that individuals who self-╉injure perceive that they
have limited resources to cope with stressors suggests that lack of effective coping strategies or
1
0
4

Theoretical perspectives on NSSI 401

maladaptive coping styles play a role in NSSI. Indeed, Haines and Williams (1997) reported that
people who self-╉injure have difficulty coping with problems. Others have found that reliance on
avoidant coping strategies is associated with the presence of NSSI among adolescents (Evans,
Hawton, & Rodham, 2005), young adults (Andover, Pepper, & Gibb, 2007; Borrill, Fox, Flynn, &
Roger, 2009; Brown, Williams, & Collins, 2007), and adult prisoners (Kirchner, Forns, & Mohino,
2008). These studies suggest individuals who engage in NSSI are less likely to use coping strategies
aimed at resolving problems and are more likely to engage in avoidant behaviors which potentially
maintain high levels of emotional arousal.
It is unsurprising then that self-╉injurers are also more likely to engage in emotion-╉focused cop-
ing strategies (Borrill et al., 2009; Mikolajczak, Petrides & Hurry, 2009); perhaps in an attempt to
cope with the resultant negative emotions arising from problem avoidance. Interestingly, Williams
and Hasking (2010) reported that, when experiencing psychological distress, individuals who
relied on avoidant coping strategies were more likely to have higher scores on a measure of NSSI
which took into account the frequency, recency, severity and range of methods of self-╉injury used.
Further, the relationship between psychological distress and NSSI was positive among study par-
ticipants who did not rely on emotion-╉focused coping strategies, suggesting that inability to cope
effectively with distressing emotions is implicated in NSSI.
The cross-╉sectional nature of the above studies preclude conclusions regarding causal relation-
ships; however, they indicate that NSSI is associated with a constellation of behaviors aimed at
avoiding problems which, in turn, maintain emotional turmoil occasioned by these problems.
Lack of effective emotion-╉focused coping may leave individuals with no other recourse than
to engage in NSSI to alleviate their negative emotional states. Such a dynamic is proposed in
the Experiential Avoidance Model of NSSI (Chapman, Gratz, & Brown, 2006) which conceptual-
izes self-╉injury as a means of avoiding unwanted internal experiences such as bodily sensations,
thoughts, memories and emotions, and the events and contexts that occasion them.
The Experiential Avoidance Model of NSSI provides a theoretical framework that makes sense
of the empirical findings in the coping literature and accounts for the initial motivation for prob-
lem avoidance as a means to avoid or escape from unwanted thoughts, memories and emotions
associated with stressful situations and life events. According to the model, limited access to effec-
tive strategies to regulate emotional arousal is a key factor, as their absence leaves individuals to
contend with their unwanted emotions which are often experienced as intense. Without effective
strategies to modulate their emotional states, individuals use NSSI to further escape from them.
Chapman and colleagues (2006) contend that the behavior is maintained through behavioral rein-
forcement and may become rule governed through verbal rules which specify that NSSI is related
to feeling better.

Functional accounts of NSSI


The observations by Chapman and colleagues echo Nock and Prinstein’s (2004) Functional Model
of NSSI, which draws on commonly reported motivations for engaging in NSSI. According to
this perspective, NSSI serves two general functions in the intra-╉and interpersonal domains
respectively. In the intrapersonal domain, NSSI assists individuals to feel better. NSSI is negatively
reinforced when it reduces or brings to an end a negative emotional state (automatic-╉negative
reinforcement); and is positively reinforced when it induces a desired emotional state (automatic-╉
positive reinforcement) such as feeling something rather than feeling numbed. In the interper-
sonal domain, NSSI is behaviorally reinforced as it assists individuals to escape from others’
demands (social-╉negative reinforcement) as well as to gain attention from others or to have access
to resources as a consequence of their behavior (social-╉positive reinforcement). It ought to be noted
2
0
4

402 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

that while Nock and Prinstein identified two functional domains for NSSI, they acknowledged
that regulating emotional states (i.e., the intrapersonal domain) was more commonly reported by
people who self-╉injure. Accordingly, the authors observed that the social/╉interpersonal function
of NSSI may be secondary to the intrapersonal function of feeling better.

Rumination, emotional cascades and NSSI


The theoretical perspectives and empirical findings discussed above suggest high emotional
arousal, low distress tolerance and lack of effective emotion regulation strategies are implicated in
NSSI. However, they provide little explanation for the intensity of negative emotional experience
that is associated with the behavior. Distal factors such as a genetic predisposition for emotional
reactivity and sensitivity might contribute somewhat to the experience of emotional turmoil
(Nock, 2009). Proximal factors, such as the types of responses individuals engage to initiate, main-
tain and modify the occurrence, intensity, duration and expression of emotion might also play a
role in the underlying processes of NSSI.
Selby and Joiner (2009) proposed one such response type in the Emotional Cascade Model for
Dysregulated Behaviours. They suggest that the intense negative emotional states that precede or
co-╉occur with NSSI are “emotional cascades” which begin with minute negative emotional stimuli
which become amplified and intensified by a vicious cycle of rumination. That is, by “a tendency
to repetitively think about the causes, situational factors, and consequences of one’s negative emo-
tional experience … continuously thinking about and focusing attention on emotionally relevant
stimuli” (Selby & Joiner, 2009; p.  220). The contribution of rumination in amplifying negative
emotional states has previously been reported for depression, anxiety, worry and anger (Calmes
& Roberts, 2007; Harrington & Blakenship, 2002; Muris, Roelofs, Meesters & Boomsma, 2004;
Nolen-╉Hoeksema, 2000; Peled & Moretti, 2007). Moreover, other studies show that ruminating
on one’s sad mood increased distress regarding current concerns (Conway, Csank, Holm, & Blake,
2000), and that the tendency to ruminate about negative inferences following stressful events
had a larger effect on the number, rate and duration of depressive episodes than ruminating on
depressed mood alone (Robinson & Alloy, 2003).
There is empirical support for the contribution of rumination to NSSI among adults and ado-
lescents (Armey & Crowther, 2008; Bjärehed & Lundh, 2008; Borrill et al., 2009; Hilt, Cha, et al.,
2008). In an early test of the Emotional Cascades Model, Selby, Connell, and Joiner (2010) found
significant direct effects of rumination and painful and provocative life events on NSSI. Further,
they reported that individuals who had experienced more painful and provocative life events and
who had greater ruminative tendencies were more likely to engage in frequent NSSI than indi-
viduals who had lower ruminative tendencies. In a more recent test of the model Selby, Franklin,
Carson-╉Wong, and Rizvi (2013) found that individuals who engaged in more frequent rumina-
tion, and individuals who experienced greater changes in levels of negative emotion, reported
more NSSI episodes. As would be expected from the model, the researchers also found that indi-
viduals with greater fluctuation in both rumination and negative affect had more episodes of
NSSI. However, fluctuations in rumination also predicted NSSI, even when levels of daily nega-
tive emotion were stable. This contradicted the Emotional Cascade Model, as the former should
only predict NSSI in the context of more frequent and greater fluctuations of negative emotion.
Nonetheless, when taken together, there is some preliminary empirical support for the model.

Gross’ process model and NSSI


The process model of emotion regulation developed by John Gross (1998a; 1998b) is one of the
most widely used frameworks in the emotion regulation field (Gullone, Hughes, King & Tonge,
3
0
4

Theoretical perspectives on NSSI 403

2010; Webb, Miles, & Sheeran, 2012). It describes five emotion regulation processes that contrib-
ute to how emotions might be experienced and expressed. The first of these emotion regulation
responses is situation selection which describes the process whereby an individual chooses to enter
into and engage in different types of situations which might evoke different emotions. Individuals
might choose to engage in NSSI as an alternative to engaging in situations which they anticipate
will cause distress (McKenzie & Gross, 2014). They may also engage in NSSI as a means to enter
new and more desired situations (e.g,. going to hospital and receiving medical attention).
The second is situation modification—╉where individuals might choose to alter the situation in
which they find themselves so as to increase or decrease the likelihood of experiencing specific
emotions. Consistent with Nock and Prinstein’s (2004) Functional Model of NSSI, individuals
engage in NSSI to elicit alternative responses from others which in turn modifies the situations
in which they find themselves (e.g., increased caregiving or reductions in external demands from
others; McKenzie & Gross, 2014).
Attentional deployment refers to the selective attention to different aspects of a situation.
Therefore, an individual might choose to attend to aspects of a situation that are likely to evoke
specific types of emotional responses or may choose to disregard or distract from these emo-
tional cues. Importantly, Webb and colleagues (2012) observe that attentional deployment can
be applied to internal experiences such as memories where the individual is re-╉immersed in the
initial situation that gave rise to the emotion that is then re-╉experienced. Although not conceptu-
alized within Gross’ model of emotion regulation, Webb and colleagues (2012) classified rumina-
tion as an example of attentional deployment which is consistent with its definition in Selby and
Joiner’s (2009) Emotional Cascade Model for Dysregulated Behaviours as rumination continually
focuses attention on emotionally relevant stimuli. Several of the theories described above suggest
that NSSI is a form of attentional deployment, with the Experiential Avoidance Model of NSSI
positing that it distracts from aversive internal experiences.
A fourth emotional regulation response is cognitive change which refers to the interpretations
and appraisals placed on emotionally relevant stimuli. McKenzie and Gross (2014) suggest NSSI
might be a mechanism by which individuals change their self-╉views: Transforming higher-╉order
self-╉construals that may be overwhelmed by responsibilities and external demands and, therefore,
lead to negative emotional states, to lower-╉order self-╉construals focusing on bodily experiences
which may not evoke the same intensity of emotion. Alternatively, individuals may engage in
cognitive change to reinterpret and ascribe different meanings to specific emotional stimuli so as
to change their emotional experience.
Finally, individuals can choose to limit the expression of the emotional response through
response modulation and in doing so regulate the behavioral, experiential, and/╉or physiological
expression of the emotion. Engaging in NSSI might be a form of response modulation as the
subsequent release of endogenous opioids soothes the physiological arousal that accompanies
negative emotional states (McKenzie & Gross, 2014).

Cognitive reappraisal and expressive suppression


Two emotional regulation strategies from Gross’ process model—╉cognitive reappraisal and expres-
sive suppression—╉have been explicitly applied to NSSI. Cognitive reappraisal is a process whereby
the emotional salience of a situation is reduced through cognitive change. Expressive suppression,
on the other hand, is a response modulation process which specifically aims to reduce emotion-
ally expressive behavior. Among adults, cognitive reappraisal is related to a greater experience and
expression of positive emotion but reduced experience and expression of negative emotion (Gross
& John, 2003). Conversely, expressive suppression leads to reduced experience and expression of
4
0

404 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

positive emotion; and increased experience of negative emotion and lower expression of negative
emotion, which has been associated with negative health outcomes (Gross & John, 2003). Among
adolescents, the reduced use of cognitive reappraisal in conjunction with a greater tendency to
engage in expressive suppression is related to depressive symptomatology, school refusal and anx-
iety (Betts, Gullone, & Allen, 2009; Hughes, Gullone & Watson, 2011; Hughes, Gullone, Dudley,
& Tonge, 2010). These studies hint at the protective effect of reappraisal and suggest expressive
suppression is associated with maladaptive outcomes in psychological health and reduced well-​
being among adolescents.
Given the contribution of cognitive reappraisal and expressive suppression in the regulation
of negative states, it may be speculated that the ability to effectively engage in reappraisal might
reduce the likelihood of engaging in NSSI, as it is likely to reduce the intensity and duration of
negative emotion. Expressive suppression, on the other hand, would therefore increase the likeli-
hood of engaging in NSSI as it has the tendency to increase the experience of negative emotion.
In support, among adolescents aged 13–​18 years (Hasking et al., 2010), and young adults aged
18–​30  years (Williams & Hasking, 2010), cognitive reappraisal was negatively correlated with
NSSI; whereas, expressive suppression had a positive correlation. Comparisons between groups
of 18–​30 year olds who did not engage in NSSI, those who engaged in infrequent and low sever-
ity NSSI, and those who engaged in frequent (at least once per month) NSSI which resulted in
wounds requiring first aid, revealed significant group differences in mean scores for expressive
suppression, but not for cognitive reappraisal (Hasking et al., 2008). The moderate/​severe group
had the highest mean score for expressive suppression, while those who did not engage in NSSI
had the lowest. In other work, while there were differences in the use of cognitive reappraisal
between self-​injurers and non-​self-​injurers aged ten years and above, no differences were found
for the use of expressive suppression (Martin et  al., 2010). In the latter study, individuals who
self-​injured were 3.3 times more likely to report difficulty with using reappraisal to regulate their
emotional states compared with those who did not self-​injure.
The discrepant findings in the above studies may be due to different criterion variables under
investigation. Hasking and colleagues (Hasking et al., 2008, 2010; Williams & Hasking, 2010) were
interested in frequency, recency and severity of NSSI; whereas, Martin and colleagues (2010)
focused on NSSI history. It might be that the two emotion regulation processes have differential
contributions to the presence of NSSI (i.e,. whether and when individuals engage in the behavior),
and to the severity of the behavior (i.e,. the extent to which they engaged in the behavior in regard
to frequency, recency, severity etc.). Cognitive reappraisal may be more pertinent in the former
case; while, expressive suppression may be related to the latter (see Voon, Hasking & Martin,
2014a).
Such an observation is consistent with findings from earlier studies on a related construct of
emotional inexpressivity. A personality trait which confers a tendency to restrict displays of emo-
tions regardless of the valence of the emotion or the manner of expression, it is similar to expres-
sive suppression. Research findings show that although emotional inexpressivity did not reliably
distinguish undergraduates with and without a history of NSSI (Gratz, 2006; Gratz & Chapman,
2007), it was significantly associated with frequency of NSSI among women who engaged in NSSI
(Gratz, 2006; Gratz & Roemer, 2008).
A set of studies from a longitudinal dataset produced mixed findings. In two, neither cognitive
reappraisal nor expressive suppression predicted first episode NSSI among adolescents (Andrews,
Martin, Hasking & Page, 2014; Tatnell, Kelada, Hasking & Martin, 2014); although, findings by
Voon, Hasking, and Martin (2014b) suggest reappraisal may protect against NSSI onset in younger
cohorts. Further, Voon, Hasking, and Martin (2014c) suggest persistent and increasing use of cog-
nitive reappraisal may have a slight protective effect in reducing medical severity of NSSI although
5
0
4

Intervention approaches for NSSI 405

it did not contribute to changes in frequency and duration of the behavior over a two-╉year period.
Use of expressive suppression, while differentiating youth who self-╉injured from youth who did
not, had no bearing on NSSI over the same two-╉year period. Finally, Andrews, Hasking, Martin,
and Page (2013) reported adolescents who continued to engage in NSSI 12-╉months from baseline
reported less tendency to engage in both cognitive reappraisal and expressive suppression com-
pared with adolescents who stopped self-╉injuring, which is consistent with the general consensus
that NSSI is associated with deficits in emotion regulation.
The above studies suggest cognitive reappraisal and expressive suppression are pertinent con-
structs in the underlying processes of NSSI, although further research is required to clarify their
roles. Given NSSI is used as a means of emotion regulation, with the assumption being that this
strategy is used when other emotion regulation techniques are lacking, increased understanding
of what kinds of emotion regulation processes among adolescents are most effective in modulat-
ing the negative emotional states that precede or co-╉occur with NSSI can be beneficial in refining
interventions.

Intervention approaches for NSSI


Most interventions for NSSI have been developed within the context of addressing self-╉harm
behaviors broadly which include self-╉injury with fatal intent and non-╉direct methods of harm.
Accordingly, in this section, the term “self-╉harm” is used to denote all self-╉injurious behaviors
regardless of intent and includes overdose, while NSSI is used to specifically refer to self-╉directed
physical violence without fatal intent. There are few interventions that have been developed spe-
cifically for NSSI, and of the interventions that have been evaluated, most have been with adult
participants rather than adolescents. With these caveats in mind, current treatment interventions
for self-╉harm draw on a range of psychotherapeutic approaches including Cognitive-╉Behavioral
Therapy, Dialectical Behavior Therapy, and Mentalisation-╉Based Therapy (Brausch & Girresch,
2012; Kerr, Muehlenkamp & Turner, 2010; Ougrin et al., 2012; Stoffers, et al., 2012; Washburn
et al., 2012). These interventions have, for the most part, favored a comprehensive approach which
addresses cognitive and emotional triggers for NSSI. Evaluations indicate such an approach is
likely to be promising, particularly where it also includes skills building in managing emotions.
However, more rigorous evaluations of existing interventions are required before firm conclusions
regarding efficacious treatments for adolescent NSSI can be made.

Cognitive-╉behavioral therapy (CBT)
CBT is a treatment approach that comprises both cognitive and behavioral components (Stoffers
et al., 2012). One of the earliest interventions applied to self-╉harm behaviors is a form of CBT
known as Problem Solving Therapy (Brausch & Girresch, 2012; Washburn et al., 2012). The inter-
vention assists individuals to cope with and resolve problems and includes cognitive restructuring
to engender a more positive orientation to problems, as well as skills training in coping and ratio-
nal problem-╉solving. Use of this intervention to address self-╉injurious behaviors draws on early
conceptualisations of such behaviors as a general deficit in coping skills. However, the strength of
evidence for the intervention was weak (Brausch & Girresch, 2012; Washburn et al., 2012). While
initial evaluations were promising (showing a trend towards reductions in self-╉harm behaviors),
the intervention did not produce statistically significant differences compared to controls.
The need for more comprehensive intervention approaches to self-╉harm behaviors led to the
development of Manual Assisted Cognitive-╉Behavioral Therapy (MACT; Washburn et al., 2012).
This intervention integrates CBT with solution-╉focused therapy and includes a bibliotherapy
component aimed at improving emotion regulation, and coping with negative cognitions (Kerr et
6
0
4

406 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

al., 2010). Conducted over six sessions, MACT includes a functional analysis of self-╉harm behav-
iors, education on emotion regulation and problem-╉solving strategies, management of negative
thinking, management of substance use, and relapse prevention (Weinberg, Gunderson, Hennen,
& Cutter, 2006). Two early evaluations of MACT showed reductions in frequency of self-╉harm
and duration between self-╉harm episodes in the intervention group, but these outcomes were not
significantly different from similar reductions in the control group (see Evans et al., 1999; Tyrer
et al., 2003). Kerr and colleagues (2010) noted that the non-╉significant results in these studies
could be due to the heterogeneity in how the intervention was delivered. Following these trials,
Weinberg and colleagues (2006) evaluated the efficacy of MACT in reducing NSSI and suicide
attempts among women with BPD (aged 18–╉40 years). Participants were randomly assigned to a
MACT intervention or treatment-╉as-╉usual (TAU). The authors reported significant reductions in
frequency of NSSI post-╉treatment as well as at six-╉month follow-╉up. Moreover, NSSI severity was
significantly lower compared with TAU at follow-╉up.
Further emphasising the utility of focussing on factors other than coping with and resolving
problems, Slee, Garnefski, van der Leeden, Arensman, and Spinhoven (2008) developed and eval-
uated a CBT intervention to address deliberate self-╉harm among 15-╉35 year olds and reported
significant reductions over nine months in the number of self-╉harm episodes among the interven-
tion group compared with TAU. The intervention comprised 12 individual sessions focussing on
identifying cognitive and emotional factors maintaining self-╉harm behaviors, and included the
use of cognitive and behavioral strategies to address these maintaining factors. Strategies include
addressing cognitive distortions, emotion regulation, and problem-╉solving. A  follow-╉up study
(Slee, Spinhoven, Garnefski & Arensman, 2008), showed that improved emotion regulation par-
tially mediated reductions in self-╉harm following the intervention.
More recently, Taylor and colleagues (2011) evaluated the efficacy of a similar intervention
(Manualised Cognitive-╉ Behavioral Therapy) developed specifically for adolescents aged 12–╉
18 years. This intervention comprised eight to 12 individual therapy sessions utilising a standard
manual which included modules on identifying cognitive and emotional triggers for self-╉harm
behavior, as well as modules teaching skills in managing cognitive distortions, mindfulness,
problem-╉solving, and assertiveness. Preliminary findings were promising and showed reductions
in frequency of deliberate self-╉harm post-╉treatment and at three-╉month follow-╉up. However, the
study did not include a control group and, therefore, inferences regarding its efficacy cannot be
made conclusively.

Dialectical behavior therapy (DBT)


Of all interventions addressing self-╉harm, DBT has received the most attention with treatment
efficacy assessed via numerous evaluations including randomized controlled trials (Stoffers
et al., 2012; Washburn et al., 2012). Developed by Marcia Linehan (1993) to treat BPD, it targets
the main areas of dysregulation in BPD through validating individuals’ experiences and assist-
ing them to increase their coping skills across a number of domains (interpersonal function-
ing, cognitive functioning, behavioral functioning and sense of self). It comprises a combination
of individual therapy, skills training, and brief telephone counselling/╉coaching components.
Together these aim to improve individuals’ capacity to accept the negative emotions that motivate
them: to engage in self-╉harm; to tolerate aversive situations, thoughts and emotions; to identify,
appraise, and modulate their emotional experiences; and to improve interpersonal relationships.
Importantly, therapy progresses through a number of stages with the initial stage focussing pri-
marily on reducing self-╉harm. In this regard, DBT has been described as a specific treatment for
self-╉harm rather than treating it as a peripheral consequence of psychopathology (Feigenbaum,
2010; Lynch & Cozza, 2009).
7
0
4

Intervention approaches for NSSI 407

Stage 1 DBT as developed by Linehan is designed to be completed over a 12-╉month period with
weekly one-╉hour individual therapy sessions, and two-╉and-╉a-╉half-╉hour group skills training ses-
sions covering the following topic areas over two rotations:
1. Mindfulness
2. Distress tolerance
3. Emotion regulation
4. Interpersonal effectiveness
The explicit focus on emotion regulation not only underscores the importance of emotion regu-
lation in dysregulated behaviors, but differentiates DBT from similar therapies. Similar to CBT-╉
based interventions described above, individual therapy sessions focus on the identification of
cognitive, emotional and situational triggers for target self-╉harm behaviors, and counselling/╉
coaching on the use of appropriate cognitive and behavioral skills to cope with these triggers
(Koerner & Dimeff, 2007). This is achieved through the use of chain analysis which proceeds with
identifying antecedent factors leading up to the focal behavior (typically self-╉harming behaviors
such as NSSI). These antecedent factors include situational, social, cognitive and emotional fac-
tors. The aim of chain analysis is to identify points at which individuals may interrupt the chain of
events leading up to the behavior. As such, unlike CBT-╉based approaches, chain analysis does not
just focus on temporally proximate antecedent triggers. For example, an individual may identify
feeling angry as a precursor to NSSI. The negative emotional state may be preceded by a comment
from a family member or friend. Negative cognitions arising from the comment may also be iden-
tified, as well as general factors such as feeling physically unwell or stressed which can contribute
to the individual’s overall vulnerability or sensitivity. Using chain analysis, the therapist will work
with the client to recognize the vulnerabilities associated with feeling unwell/╉stressed and nega-
tive cognitions and to identify strategies to prevent future recurrence of the behavioral chain.
Evaluations of DBT among adults with BPD have demonstrated reductions in self-╉harm among
participants. Stoffers and colleagues (2012) reported that the pooled effect from three trials
undertaken between 2001 and 2005 showed significant reductions compared with TAU. However,
a more recent Australian trial (Carter, Wilcox, Lewin, Conrad & Bendit, 2010) did not find sig-
nificantly different results between a modified DBT program and TAU.
DBT has been adapted for adolescents (DBT-╉A) by decreasing duration of treatment to a 16-╉
week program, using age-╉appropriate terminology and inclusion of family members in the skills
training groups (Groves, Backer, van den Bosch, & Miller, 2012). However, these programs have
not been subjected to randomized control trials and results are mixed (Brausch & Girresch, 2012;
Kerr et al., 2010; Washburn et al., 2012). Non-╉significant group differences were reported when
comparing DBT-╉A with TAU on suicide attempts (Rathus & Miller, 2002) and self-╉harm (Katz,
Cox, Gunasekara & Miller, 2004). Two other studies reported significant post-╉treatment reduc-
tions in self-╉harm (James, Taylor, Winmill & Alfoadari, 2008) and NSSI (Fleischhaker et al., 2011);
although the absence of a control group limits conclusions regarding the efficacy of these inter-
ventions among adolescents.

Mentalization-╉based therapy (MBT)
MBT draws on psychodynamic theories (Kerr et  al., 2010; Stoffers et  al., 2012), and aims to
“strengthen patients’ capacity to understand their own and others’ mental states in attachment
contexts in order to address their difficulties with affect, impulse regulation, and interpersonal
functioning which act as triggers for acts of suicide and self-╉harm” (Bateman & Fonagy, 2009,
p. 1355). Thus, MBT assists with improved interpersonal function by building individuals’ capac-
ity to mentalize and be aware of how thoughts and emotions influence their own and others’
8
0
4

408 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

behaviors (Kerr et al., 2010). Stoffers and colleagues (2012) noted that, comparing the interven-
tion with TAU, MBT achieved significant reductions in self-╉harm among adults in two trials
undertaken in 1999 and 2009.
More recently, Rossouw and Fonagy (2012) reported findings from a randomized control trial
of MBT for adolescents. Significant group differences among adolescents randomly assigned to
the MBT treatment group versus TAU controls were found. Those in the treatment group had
lower scores on self-╉harm at the end of the 12-╉month period, and showed greater reductions
in self-╉harm over the course of treatment. Although results are promising, further replication is
required.

Emotion regulation group therapy


Finally, following research on the impact of emotion dysregulation on NSSI, Gratz, and Gunderson
(2006) developed a 14-╉week emotion regulation group intervention specifically for NSSI. Drawing
on a range of extant psychotherapeutic approaches including Acceptance and Commitment
Therapy and DBT, the intervention focused on building emotional awareness and acceptance of
emotions (versus emotional avoidance). Initial results among a group of women with BPD were
positive and showed significant between-╉group differences (i.e., intervention vs TAU) in reduc-
tions in NSSI. Findings were replicated in a subsequent study extending the intervention to more
diverse settings and groups of women with BPD (Gratz & Tull, 2011). Importantly, in a follow-╉up
study, which analysed data collated from the above trials (Gratz, Levy & Tull, 2012), decreases in
NSSI and emotion dysregulation were reported among the intervention groups (RCT and open
trial completers)1 but not among controls. Moreover, a mediation analysis showed that emotion
dysregulation mediated the relationship between intervention and outcome. Thus, the available
evidence indicates that achieving reductions in emotion dysregulation is a worthwhile goal and
that the intervention is promising (Gratz, Tull & Levy, 2014). However, it has not been applied to
adolescents.

Clinical guidance on the management and treatment


The foregoing discussion shows that emotion regulation is a critical factor in understanding the
aetiology, and more importantly, the maintenance of NSSI. Specific emotion regulation pro-
cesses such as rumination and expressive suppression have emerged in the research as possible
processes which can play a role in the escalation of negative emotional states that precede or
co-╉occur with the behavior, while engaging in cognitive reappraisal might be useful as a de-╉
escalation strategy. Assisting individuals in managing their propensity to engage in ruminative
thinking and focusing attention on minute emotional stimuli may be useful; incorporating skills
in distracting oneself from one’s emotional distress in intervention programs for NSSI may be
warranted, as distraction has been shown to be a more adaptive response to distress than rumi-
nation (see Nolen-╉Hoeksema, 1991). Mindfulness skills may also have some utility in this regard
as both distraction and mindfulness have been reported to be effective (compared with problem-╉
solving) for reducing rumination among adolescents (Hilt & Pollak, 2012). Indeed, these specific
skills are incorporated in DBT in the distress tolerance and mindfulness skills training modules
(Linehan, 1993). Additionally, improving individuals’ effectiveness in the use of cognitive reap-
praisal to reduce the emotional salience of stressful situations and life events might also be a

1 The intervention groups comprised of the “RCT” group in Gratz & Gunderson (2006) and the “open trial
completers” in Gratz & Tull (2011). The control group were the TAUs in Gratz & Gunderson (2006).
9
0
4

Conclusion 409

useful component in interventions for NSSI. As described above, reappraisal features in several
existing interventions, in the form of cognitive restructuring to address cognitive distortions
and beliefs.
Although specific interventions for NSSI in adolescence have yet to be developed, the above
observations dovetail with available intervention approaches for self-╉harm behaviors generally.
Emotion regulation skills that feature in these intervention approaches include: Cognitive restruc-
turing to address cognitive distortions and negative thinking styles such as in CBT-╉based inter-
ventions; acceptance of emotions through mindfulness and use of distraction as featured in DBT;
and awareness of emotion and their contribution to actions and behaviors as in DBT (see emotion
regulation module) and MBT. The current state of intervention research for self-╉harm behaviors
does not at present point to any single component which influences treatment outcome. It is likely
that a combination of these skills is warranted as the research suggests comprehensive approaches
which address cognitive and emotional triggers for NSSI have greater utility. Further examination
of how specific emotion regulation processes contribute to or modulate the underlying emotional
distress which precedes or accompanies NSSI is warranted to assist with identifying further areas
for inclusion in treatment for the behavior.
Glenn, Franklin, and Nock (2014) highlight that in addition to emotion regulation skills train-
ing, effective interventions for self-╉harm also include:  1)  A  focus on improving interpersonal
functioning, 2)  high intensity and frequent sessions, and 3)  incorporation of interventions to
address other co-╉occurring maladaptive behaviors or risk factors such as substance use. These
may be applicable to treatment interventions for NSSI among adolescents.
In the absence of empirically tested interventions for NSSI among adolescents, Washburn and
colleagues (2012) distilled from the available literature additional elements to guide the man-
agement and treatment of NSSI among this population group. They suggest that assessment for
NSSI is important and should, at a minimum, aim to improve understanding of current and past
behaviors including methods, locations, frequency, severity, urges and age of onset. Motivational
enhancement (e.g., through motivational interviewing) may be necessary both prior to and dur-
ing treatment. Additionally, cognitive and behavioral interventions can be useful to address self-╉
derogatory and distorted beliefs about NSSI (e.g., through Socratic questioning and thought
monitoring), and include contingency management and behavioral activation. Dialectical strate-
gies such as acceptance and tolerance of distress may also be useful to address the urge to engage
in NSSI.
More broadly, Washburn and colleagues suggest individuals may benefit from interpersonal
approaches to understand and modify maladaptive interpersonal styles which may coincide with
the negative emotional states that precede or accompany NSSI. Skills training is likely to be central
and should focus on improving emotion regulation, problem-╉solving, interpersonal and commu-
nication skills. In addition, treatment may need to focus on physical factors such as body image
and physical self-╉care.
Of particular importance is the need to address “social contagion” when working with groups
(Washburn et al., 2012). It ought to be noted that some DBT skills training groups address this by
expressly discouraging discussion of the types of self-╉harm behaviors participants engage in. Finally,
“contracts for safety” or “no-╉harm agreements” can be either ineffective or harmful, and treatment
should focus on contingency management and relapse prevention instead (Washburn et al., 2012).

Conclusion
NSSI is a behavior that typically begins during adolescence and adversely impacts on psychologi-
cal health and well-╉being. Frequency of NSSI during adolescence predicts its maintenance into
0
1
4

410 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

adulthood, as well as increasing risk for suicide behavior. Consequently, prevention and early
intervention addressing the behavior among adolescents is of critical importance.
The general consensus is that NSSI is a behavior that functions to alleviate negative emo-
tional states. While distal factors such as a genetic predisposition to emotional sensitivity and
reactivity may interact with invalidating childhood environments to predispose individuals
toward a vulnerability to engaging in NSSI, poor emotion regulation is likely to maintain the
behavior. It is, therefore, unsurprising that interventions to address behaviors such as NSSI
have focused on building individuals’ capacity for emotion regulation. While there are no spe-
cific interventions that have been developed solely for NSSI among adolescents, the existing
interventions for self-╉harm behaviors provide promising leads. Research findings on specific
emotion regulation processes (e.g., cognitive reappraisal, expressive suppression, and rumi-
nation) and how they may be implicated in increasing or decreasing the emotional arousal
associated with NSSI provide further clues that can assist with developing better targeted
interventions among adolescents. This is particularly important as the evidence for effective
interventions for adolescents addressing self-╉harm behaviors such as NSSI remain inconclu-
sive. Thus, replication and adaptations of existing intervention approaches accompanied by
rigorous evaluation is therefore, warranted.

References
American Psychiatric Association (2013). Diagnostic and Statistical Manual of Mental Disorders, Fifth
Edition. Arlington, VA: American Psychiatric Association.
Andover, M. S., Pepper, C. M., & Gibb, B. E. (2007). Self-╉mutilation and coping strategies in a college
sample. Suicide & Life-╉Threatening Behaviour, 37, 238–╉243.
Andrews, T., Martin, G., Hasking, P., & Page, A. (2013). Predictors of continuation and cessation of non-╉
suicidal self-╉injury. Journal of Adolescent Health, 53, 40–╉46. doi:10.1016/╉j.jadohealth.2013.01.009
Andrews, T., Martin, G., Hasking, P., Page, A. (2014). Predictors of onset for non-╉suicidal self-╉injury
within a community-╉based sample of adolescents. Prevention Science, 15, 850–╉859. doi:10.1007/╉
s11121-╉013-╉0412-╉8
Armey, M. F., & Crowther, J. H. (2008). A comparison of linear versus non-╉linear models of aversive self-╉
awareness, dissociation, and non-╉suicidal self-╉injury among young adults. Journal of Consulting and
Clinical Psychology, 76, 9–╉14. doi:10.1037/╉0022-╉006X.76.1.9
Bateman, A., & Fonagy, P. (2009). Randomized controlled trial of outpatient mentalization-╉based treatment
versus structured clinical management for Borderline Personality Disorder. American Journal of
Psychiatry, 166, 1355–╉1364.
Bentley, K. H., Cassiello-╉Robbins, C. F., Vittorio, L., Sauer-╉Zavala, S., & Barlow, D. H. (2015). The
association between nonsuicidal self-╉injury and the emotional disorders: A meta-╉analytic review.
Clinical Psychology Review, 37, 72–╉88. doi:10.1016/╉j.cpr.2015.02.006
Betts, J., Gullone, E., & Allen, J. S. (2009). An examination of emotion regulation, temperament, and
parenting style as potential predictors of adolescent depression risk status: A correlational study. British
Journal of Developmental Psychology, 27, 473–╉485. doi:10.1348/╉026151008X314900
Bjärehed, J., & Lundh, L. (2008). Deliberate self-╉harm in 14-╉year-╉old adolescents: How frequent is it, and
how is it associated with psychopathology, relationship variables, and styles of emotional regulation?
Cognitive Behaviour Therapy, 37, 26–╉37.
Bjärehed, J., Wångby-╉Lundh, M., & Lundh, L. (2012). Nonsuicidal self-╉injury in a community sample of
adolescents: Subgroups, stability, and associations with psychological difficulties. Journal of Research on
Adolescence, 22, 678–╉693. doi:10.1111/╉j.1532-╉7795.2012.00817.x
Borrill, J., Fox, P., Flynn, M., & Roger, D. (2009). Students who self-╉harm: Coping style, rumination and
alexithymia. Counselling Psychology Quarterly, 22, 361–╉372.
14

Conclusion 411

Brausch, A. M., & Girresch, S. K. (2012). A review of empirical treatment studies for adolescent
nonsuicidal self-​injury. Journal of Cognitive Psychotherapy: An International Quarterly, 26, 3–​18.
doi:10.1891/​0889-​8391.26.1.3
Bresin, K., & Schoenleber, M. (2015). Gender differences in the prevalence of nonsuicidal self-​
injury: A meta-​analysis. Clinical Psychology Review, 38, 55–​64. doi:10.1016/​j.cpr.2015.02.009
Bridge, J. A., Goldstein, T. R., & Brent, D. A. (2006). Adolescent suicide and suicidal behaviour. Journal of
Child Psychology and Psychiatry, 47, 372–​394. doi:10.1111/​j.1469-​7610.2006.01615.x
Brown, S. A., Williams, K., & Collins, A. (2007). Past and recent deliberate self-​harm: Emotion and coping
strategy differences. Journal of Clinical Psychology, 63, 791–​803. doi:10.1002/​jclp.20380
Calmes, C. A., & Roberts, J. E. (2007). Repetitive thought and emotional distress: Rumination and worry as
prospective predictors of depressive and anxious symptomatology. Cognitive Therapy and Research, 30,
343–​356. doi:10.1007/​s10608-​006-​9026-​9
Carter, G. L., Willcox, C. H., Lewin, T. J., Conrad, A. M., & Bendit, N. (2010). Hunter DBT
project: Randomised controlled trial of dialectical behaviour therapy in women with borderline
personality disorder. Australian & New Zealand Journal of Psychiatry, 44, 162–​173.
Chapman, A. L., Gratz, K. L., & Brown, M. Z. (2006). Solving the puzzle of deliberate self-​harm: The
experiential avoidance model. Behaviour Research and Therapy, 44, 371–​394. doi:10.1016/​
j.brat.2005.03.005
Conway, M., Csank, P. A. R., Holm, S. L., & Blake, C. K. (2000). On assessing individual differences in
rumination on sadness. Journal of Personality Assessment, 75, 404–​425.
Dilberto, T. L., & Nock, M. K. (2008). An exploratory study of correlates, onset, and offset of non-​suicidal
self-​injury. Archives of Suicide Research, 12, 219–​231. doi:10.1080/​13811110802101096
Evans, E., Hawton, K., & Rodham, K. (2005). In what ways are adolescents who engage in self-​harm
or experience thoughts of self-​harm different in terms of help-​seeking, communication and coping
strategies? Journal of Adolescence, 28, 573–​587. doi:10.1016/​j.adolescence.2004.11.001
Evans, K., Tyrer, P., Catalan, J., Schmidt, U., Davidson, K., Dent, J., Tata, P., Thornton, S., Barber, J., &
Thompson, S. (1999). Manual-​assisted cognitive-​behaviour therapy (MACT): A randomized controlled
trial of a brief intervention with bibliotherapy in the treatment of recurrent deliberate self-​harm.
Psychological Medicine, 29, 19–​25.
Feigenbaum, J. (2010). Self-​harm—​The solution not the problem: The Dialectical Behaviour Therapy
Model. Psychoanalytic Psychotherapy, 24, 115–​134. doi:10.1080/​02668731003707873
Fleischhaker, C., Bohme, R., Sixt, B., Bruck, C., Schneider, C., & Schulz, E. (2011). Dialectical Behavioral
Therapy for Adolescents (DBT-​A): A clinical trial for patients with suicidal and self-​injurious behavior
and borderline symptoms with a one-​year follow-​up. Child and Adolescent Psychiatry and Mental
Health, 5:3. http://​www.capmh.com/​content/​5/​1/​3
Fliege, H., Lee, J., Grimm, A., & Klapp, B. F. (2009). Risk factors and correlates of deliberate self-​
harm behaviour: A systematic review. Journal of Psychosomatic Research, 66, 477–​493. doi:10.1016/​
j.jpsychores.2008.10.013
Giletta, M., Scholte, R. H. J., Engels, R., Ciairano, S., & Prinstein, M. J. (2012). Adolescent non-​suicidal
self-​injury: A cross-​national study of community samples from Italy, the Netherlands and the United
States. Psychiatry Research, 197, 66–​72. doi:10.1016/​j.psychres.2012.02.009
Glenn, C. R., Franklin, J. C., & Nock, M. K. (2014). Evidence-​based psychosocial treatments for self-​
injurious thoughts and behaviours in youth. Journal of Clinical Child and Adolescent Psychology, 44(1)
doi:10.1080/​15374416.2014.945211
Gratz, K. L. (2003). Risk factors for and functions of deliberate self-​harm: An empirical and conceptual
review. Clinical Psychology: Science and Practice, 10, 192–​205. doi:10.1093/​clipsy/​bpg022
Gratz, K. L. (2006). Risk factors for deliberate self-​harm among female college students: The role and
interaction of childhood maltreatment, emotional inexpressivity, and affect intensity/​reactivity.
American Journal of Orthopsychiatry, 76, 238–​250. doi:10.1037/​002-​9432.76.2.238
2
1
4

412 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

Gratz, K. L., & Chapman, A. L. (2007). The role of emotional responding and childhood maltreatment in
the development and maintenance of deliberate self-​harm among male undergraduates. Pyschology of
Men & Masculinitiy, 8, 1–​14. doi:10.1037/​1524-​9220.8.1.1
Gratz, K. L., & Gunderson, J. G. (2006). Preliminary data on an acceptance-​based emotion regulation
group intervention for deliberate self-​harm among women with Borderline Personality Disorder.
Behavior Therapy, 37, 25–​35.
Gratz, K. L., & Roemer, L. (2008). The relationship between emotion dysregulation and deliberate self-​
harm among female undergraduate students at an urban commuter university. Cognitive Behaviour
Therapy, 37, 14–​25. doi: 10.1080/​16506070701819524
Gratz, K. L., & Tull, M. T. (2011). Brief report: Extending research on the utility of an adjunctive emotion
regulation group therapy for deliberate self-​harm among women with Borderline Personality pathology.
Personality Disorders: Theory, Research, and Treatment, 2, 316–​326. doi:10.1037/​a0022144
Gratz, K. L., Levy, R., & Tull, M. T. (2012). Emotion regulation as a mechanism of change in Acceptance-​
based Emotion Regulation Group Therapy for deliberate self-​harm among women with Borderline
Personality pathology. Journal of Cognitive Psychotherapy: An International Quarterly, 26, 365–​380.
doi:10.1891/​0889-​8391.26.4.365
Gratz, K. L., Tull, M. T., & Levy, R. (2014). Randomised controlled trial and uncontrolled 9-​month
follow-​up of an adjunctive emotion regulation group therapy for deliberate self-​harm among
women with borderline personality disorder. Psychological Medicine, 44, 2099–​2112. doi:10.1017/​
S0033291713002134
Gross, J. J. (1998a). Antecedent-​and response-​focused emotion regulation: Divergent consequences for
experience, expression, and physiology. Journal of Personality and Social Psychology, 74, 224–​237.
Gross, J. J. (1998b). The emerging field of emotion regulation: An integrative review. Review of General
Psychology, 2, 271–​299.
Gross, J. J., & John, O. P. (2003). Individual differences in two emotion regulation processes: Implications
for affect, relationships, and well-​being. Journal of Personality and Social Psychology, 85, 348–​362.
doi:10.1037/​0022-​3514.85.2.348
Groves, S., Backer, H. S., van den Bosch, W., & Miller, A. (2012). Dialectical Behaviour Therapy
with Adolescents. A Review. Child and Adolescent Mental Health, 17, 65–​75. doi:10.1111/​
j.1475-​3588.2011.00611.x
Gullone, E., Hughes, E. K., King, N. J., & Tonge, B. (2010). The normative development of emotion
regulation strategy use in children and adolescents: A 2-​year follow-​up study. Journal of Child
Psychology and Psychiatry, 51, 567–​574. doi:10.1111/​j.1469-​7610.2009.02183.x
Haines, J., & Williams, C. L. (1997). Coping and problem solving of self-​mutilators. Journal of Clinical
Psychology, 53, 177–​186.
Hamza, C. A., Stewart, S. L., & Willoughby, T. (2012). Examining the link between nonsuicidal self-​injury
and suicidal behaviour: A review of the literature and an integrated model. Clinical Psychology Review,
32, 482–​495. doi:10.1016/​j.cpr.2012.05.003
Hankin, B. L., & Abela, J. R. Z. (2011). Nonsuicidal self-​injury in adolescence: Prospective rates
and risk factors in a 2½ year longitudinal study. Psychiatry Research, 186, 65–​70. doi:10.1016/​
j.psychres.2010.07.056
Harrington, J. A., & Blankenship, V. (2002). Ruminative thoughts and their relation to depression and
anxiety. Journal of Applied Social Psychology, 32, 465–​485.
Hasking, P. A., Coric, S. J., Swannell, S., Martin, G., Thompson, H. K., & Frost, A. D. J. (2010). Emotion
regulation and coping as moderators in the relationship between personality and self-​injury. Journal of
Adolescence, 33, 767–​773. doi:10.1016/​j.adolescence.2009.12.006
Hasking, P., Momeni, R., Swannell, S., & Chia, S. (2008). The nature and extent of non-​suicidal self-​
injury in a non-​clinical sample of young adults. Archives of Suicide Research, 12, 208–​218. doi:10.1080/​
13811110802100957
3
1
4

Conclusion 413

Hilt, L. M., Cha, C. B., & Nolen-​Hoeksema, S. (2008). Nonsuicidal self-​injury in young adolescent
girls: Moderators of the Distress-​Function relationship. Journal of Consulting and Clinical Psychology,
76, 63–​71. doi:10.1037/​0022-​006X.76.1.63
Hilt, L. M., Nock, M. K., Lloyd-​Richardson, E. E., & Prinstein, M. J. (2008). Longitudinal study of
nonsuicidal self-​injury among young adolescents: Rates, correlates, and preliminary test of an
interpersonal model. Journal of Early Adolescence, 28, 455–​469. doi:10.1177/​0272431608316604
Hilt, L. M., & Pollak, S. D. (2012). Getting out of rumination: Comparison of three brief interventions in a
sample of youth. Journal of Abnormal Child Psychology, 40, 1157–​1164. doi:10.1007/​s10802-​012-​9638-​3
Hughes, E. K., Gullone, E., & Watson, S. D. (2011). Emotional functioning in children and adolescents
with elevated depressive symptoms. Journal of Psychopathology and Behavioural Assessment, 33, 335–​
345. doi: 10.1007/​s10862-​011-​9220-​2
Hughes, E. K., Gullone, E., Dudley, A., & Tonge, B. (2010). A case-​control study of emotion regulation
and school refusal in children and adolescents. Journal of Early Adolescence, 30, 691–​706. doi:10.1177/​
0272431609341049
Jacobson, C. M., & Gould, M. (2007). The epidemiology and phenomenology of non-​suicidal self-​injurious
behaviour among adolescents: A critical review of the literature. Archives of Suicide Research, 11, 129–​
147. doi:10.1080/​13811110701247602
James, A. C., Taylor, A., Winmill, L., & Alfoadari, K. (2008). A preliminary community study of dialectical
behaviour therapy with adolescent females demonstrating persistent, deliberate self-​harm. Child and
Adolescent Mental Health, 13, 148–​152.
Katz, L. Y., Cox, B. J., Gunasekara, S., & Miller, A. L. (2004). Feasibility of dialectical behavior therapy for
suicidal adolescent inpatients. Journal of the American Academy of Child and Adolescent Psychiatry, 43,
276–​282.
Kerr, P. L., Muehlenkamp, J. J., & Turner, J. M. (2010). Nonsuicidal self-​injury: A review of current
research for family medicine and primary care physicians. Journal of the American Board of Family
Medicine, 23, 240–​259.
Kirchner, T., Forns, M., & Mohino, S. (2008). Identifying the risk of deliberate self-​harm among young
prisoners by means of coping typologies. Suicide & Life-​Threatening Behaviour, 38, 442–​448.
Klonsky, E. D. (2009). The functions of self-​injury in young adults who cut themselves: Clarifying the
evidence for affect-​regulation. Psychiatry Research, 166, 260–​268. doi:10.1016/​j.psychres.2008.02.008
Klonsky, E. D., May, A. M., & Glenn, C. R. (2013). The relationship between nonsuicidal self-​injury and
attempted suicide: Converging evidence from four samples. Journal of Abnormal Psychology, 122, 231–​
237. doi: 10.1037/​a0030278
Koerner, K., & Dimeff, L. A. (2007). Overview of Dialectical Behaviour Therapy. In Dimeff, L. A. &
Koerner, K. (eds.), Dialectical Behaviour Therapy in Clinical Practice: Applications across disorders and
settings (pp. 1–​18). New York: The Guilford Press.
Koole, S. L. (2009). The psychology of emotion regulation: An integrative review. Cognition & Emotion, 23,
4–​41. doi:10.1080/​02699930802619031
Lang, C. M., & Sharma-​Patel, K. (2011). The relation between childhood maltreatment and self-​
injury: A review of the literature on conceptualisation and intervention. Trauma, violence, and abuse,
12(1), 23–​37. doi: 10.1177/​1524838010386975
Linehan, M. (1993). Skills Training Manual for Treating Borderline Personality Disorder. (New York: Guilford
Publication).
Lynch, T. R., & Cozza, C. (2009). Behaviour therapy for nonsuicidal self-​injury. In Understanding
Nonsuidicidal Self-​Injury: Origins, Assessment, and Treatment. Nock, M.K. (ed.). pp. 221–​250.
Washington: American Psychological Association.
Martin, G., Swannell, S., Harrison, J., Hazell, P., & Taylor, A. (2010). The Australian National
Epidemiological Study of Self-​Injury (ANESSI). Brisbane, Australia: Centre for Suicide Prevention
Studies.
4
1

414 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

McNamara, P. M. (2013). Adolescent suicide in Australia: Rates, risk and resilience. Clinical and Child
Psychology and Psychiatry, 18, 351–​369. doi: 10.1177/​1359104512455812
McKenzie, K.C., & Gross, J.J. (2014). Nonsuicidal self-​injury: An emotion regulation perspective.
Psychopathology, 47, 207–​219. doi:10.1159/​000358097
Mikolajczak, M., Petrides, K. V., & Hurry, J. (2009). Adolescents choosing self-​harm as an emotion
regulation strategy: The protective role of trait emotional intelligence. British Journal of Clinical
Psychology, 48, 181–​193. doi:10.1348/​014466508X386027
Muehlenkamp, J. J., Claes, L., & Plener, P. L. (2012). International prevalence of adolescent non-​suicidal
self-​injury and deliberate self-​harm. Child and Adolescent Psychiatry and Mental Health, 6(10), 1–​9.
Muehlenkamp, J. J., Kerr, P. L., Bradley, A. R., & Larsen, M. A. (2010). Abuse subtypes and nonsuicidal
self-​injury: Preliminary evidence of complex emotion regulation patterns. Journal of Nervous and
Mental Disease, 198, 258–​263. doi:10.1097/​NMD.0b013e3181d612ab
Muris, P., Roelofs, J., Meesters, C., & Boomsma, P. (2004). Rumination and worry in nonclinical
adolescents. Cognitive Therapy and Research, 28, 539–​554.
Nock, M. K. (2009). Why do people hurt themselves? New insights into the nature and functions of self-​
injury. Current Directions in Psychological Science, 18(2), 78–​83.
Nock, M. K. (2012). Future directions for the study of suicide and self-​injury. Journal of Clinical Child &
Adolescent Psychology, 41, 255–​259. doi:10.1080/​15374416.2012.652001
Nock, M. K., & Prinstein, M. J. (2004). A functional approach to the assessment of self-​mutilative
behaviour. Journal of Consulting and Clinical Psychology, 72, 885–​890. doi:10.1037/​0022-​006X.72.5.885
Nock, M. K., Prinstein, M. J., & Sterba, S. K. (2009). Revealing the form and function of self-​injurious
thoughts and behaviours: A real-​time ecological assessment study among adolescents and young adults.
Journal of Abnormal Psychology, 118, 816–​827. doi:10.1037/​a0016948
Nolen-​Hoeksema, S. (1991). Responses to depression and their effects on the duration of depressive
episodes. Journal of Abnormal Psychology, 100, 569–​582.
Nolen-​Hoeksema, S. (2000). The role of rumination in depressive disorders and mixed anxiety/​depressive
symptoms. Journal of Abnormal Psychology, 109, 504–​511. doi:10 I037//​0021-​843X.1093.504
Ougrin, D., Tranah, T., Leigh, E., Taylor, L., & Asarnow, J. R. (2012). Practitioner review: Self-​
harm in adolescents. Journal of Child Psychology and Psychiatry, 53, 337–​350. doi:10.1111/​
j.1469-​7610.2012.02525.x
Peled, M., & Moretti, M. M. (2007). Rumination on anger and sadness in adolescence: Fueling of fury and
deepening of despair. Journal of Clinical Child & Adolescent Psychology, 36, 66–​75.
Plener, P.L., Schumacher, T.S., Munz, L.M., & Groschwitz, R.C. (2015). The longitudinal course of non-​
suicidal self-​injury and deliberate self-​harm: A systematic review of the literature. Borderline Personality
Disorder and Emotion Dysregulation, 2, 2. doi:10.1186/​s40479-​014-​0024-​3.
Rathus, J. H., & Miller, A. L. (2002). Dialectical behaviour therapy adapted for suicidal adolescents. Suicide
& Life-​Threatening Behaviour, 32, 146–​157.
Robinson, M. S., & Alloy, L. B. (2003). Negative cognitive styles and stress-​reactive rumination interact to
predict depression: A prospective study. Cognitive Therapy and Research, 27, 275–​292.
Rossouw, T. I., & Fonagy, P. (2012). Mentalization-​based treatment for self-​harm in
adolescents: A randomized controlled trial. Journal of the American Academy of Child and Adolescent
Psychiatry, 51, 1304–​1313.
Selby, E. A., & Joiner, T. E. (2009). Cascades of emotion: The emergence of Borderline Personality
Disorder from emotional and behavioural dysregulation. Review of General Psychology, 13, 219–​229.
doi: 10.1037/​a0015687
Selby, E. A., Connell, L. D., & Joiner, T. E. (2010). The pernicious blend of rumination and fearlessness in
non-​suicidal self-​injury. Cognitive Therapy and Research, 34, 421–​428. doi:10.1007/​s10608-​009-​9260-​z
Selby, E. A., Franklin, J., Carson-​Wong, A., & Rizvi, S. L. (2013). Emotional cascades and self-​
injury: Investigating instability of rumination and negative emotion. Journal of Clinical Psychology, 69,
1213–​1227. doi:10.1002/​jclp.21966
5
1
4

Conclusion 415

Sim, L., Adrian, M., Zeman, J., Cassano, M., & Friedrich, W. N. (2009). Adolescent deliberate self-​
harm: Linkages to emotion regulation and family emotional climate. Journal of Research on Adolescence,
19, 75–​91.
Slee, N., Garnefski, N., van der Leeden, R., Arensman, E., & Spinhoven, P. (2008). Cognitive-​behavioural
intervention for self-​harm: randomised controlled trial. British Journal of Psychiatry, 192, 202–​211.
doi:10.1192/​bjp.bp.107.037564
Slee, N., Spinhoven, P., Garnefski, N., & Arensman, E. (2008). Emotion regulation as mediator of
treatment outcome in therapy for deliberate self-​harm. Clinical Psychology and Psychotherapy, 15, 205–​
216. doi: 10.1002/​cpp.577
Sornberger, M. J., Heath, N. L., Toste, J. R., & McLouth, R. (2012). Nonsuicidal self-​injury and
gender: Patterns of prevalence, methods, and locations among adolescents. Suicide and Life-​threatening
Behavior, 42(3), 266–​278. doi:10.1111/​j.1943-​278X.2012.00088.x
Stoffers, J. M, Völlm, B. A., Rücker, G., Timmer. A., Huband, N., & Lieb, K. (2012). Psychological therapies
for people with borderline personality disorder. Cochrane Database of Systematic Reviews 2012, 8(16)
Art.No.:CD005652. doi:10.1002/​14651858.CD005652.pub2.
Swannell, S., Martin, G., Page, A., Hasking, P., & St John, N. (2014). Prevalence of non-​suicidal self-​injury
(NSSI) in non-​clinical samples; Systematic review, meta-​analysis and meta-​regression. Suicide and Life-​
Threatening Behavior, 44, 273–​303.
Swannell, S., Martin, G., Page, A., Hasking, P., Hazell, P., Taylor, A., & Protani, M. (2012). Child
maltreatment, subsequent non-​suicidal self-​injury and the mediating roles of dissociation, alexithymia
and self-​blame. Child Abuse & Neglect, 36, 572–​584. doi:10.1016/​j.chiabu.2012.05.005
Tatnell, R., Kelada, L., Hasking, P., & Martin, G. (2014). Longitudinal analysis of adolescent NSSI: The
role of intrapersonal and interpersonal factors. Journal of Abnormal Child Psychology, 42, 885–​896.
doi:10.1007/​s10802-​013-​9837-​6
Taylor, L. M. W., Oldershaw, A., Richards, C., Davidson K., Schmidt, U., Simic, M. (2011). Development
and pilot evaluation of a manualized cognitive-​behavioural treatment package for adolescent self-​harm.
Behavioural and Cognitive Psychotherapy, 39, 619–​625. doi: 10.1017/​S1352465811000075
Tyrer, P., Thompson, S., Schmidt, U., Jones, V., Knapp, M., Davidson, K., Catalan, J., Airlie, J., Baxter, S.,
Byford S., et al. (2003). Randomized controlled trial of brief cognitive behaviour therapy versus treatment
as usual in recurrent deliberate self-​harm: The POPMACT study. Psychological Medicine, 33, 969–​976.
Voon, D., Hasking, P., & Martin, G. (2014a). The roles of emotion regulation and ruminative thoughts in
non-​suicidal self-​injury. British Journal of Clinical Psychology, 53, 95–​113.
Voon, D., Hasking, P., & Martin, G. (2014b). Emotion regulation processes in adolescent non-​suicidal self-​
injury: What difference does a year make? Journal of Adolescence, 37, 1077–​1087.
Voon, D., Hasking, P., & Martin, G. (2014c). Change in emotion regulation strategy use and its impact on
adolescent non-​suicidal self-​injury: A three-​year longitudinal analysis using latent growth modelling.
Journal of Abnormal Psychology, 123, 487–​498.
Walsh, B. W. (2005). Treating self-​injury: A practical guide. New York, NY: Guilford Press.
Washburn, J. J., Richardt, S. L., Styer, D. M., Gebhardt, M., Juzwin, K. R., Yourek, A., & Aldridge, D.
(2012). Psychotherapeutic approaches to non-​suicidal self-​injury in adolescents. Child and Adolescent
Psychiatry and Mental Health, 6:14, 1–​8.
Webb, T. L., Miles, E., & Sheeran, P. (2012). Dealing with feeling: A meta-​analysis of the effectiveness of
strategies derived from the process model of emotion regulation. Psychological Bulletin, 138, 775–​808.
doi:10.1037/​a0027600
Weinberg, I., Gunderson, J. G., Hennen, J., & Cutter, C. J. (2006). Manual assisted cognitive treatment for
deliberate self-​harm in Borderline Personality Disorder patients. Journal of Personality Disorders, 20,
482–​492.
Whitlock, J., Muehlenkamp, J., Eckenrode, J., Purington, A., Abrams, G. B., Barriera, P., & Kress, V.
(2013). Nonsuicidal self-​injury as a gateway to suicide in young adults. Journal of Adoelscent Health, 52,
486–​492.
6
1
4

416 Adolescents who Engage in Nonsuicidal Self-injury (NSSI)

Williams, F., & Hasking, P. (2010). Emotion regulation, coping and alcohol use as moderators in the
relationship between non-​suicidal self-​injury and psychological distress. Prevention Science, 11, 33–​41.
doi:10.1007/​s11121-​009-​0147-​8
You, J., Leung, F., Fu, K., & Lai, C. M. (2011). The prevalence of nonsuicidal self-​injury and different
subgroups of self-​injurers in Chinese adolescents. Archives of Suicide Research, 15, 75–​86. doi: 10.1080/​
13811118.2011.540211
Yurkowski, K., Martin, J., Levesque, C., Bureau, J., Lafontaine, M., & Cloutier, P. (2015). Emotion
dysregulation mediates the influence of relationship difficulties on non-​suicidal self-​injury behaviour in
young adults. Psychiatry Research. http://​dx.doi.org/​10.1016/​j.psychres.2015.05.006
7
1
4

Part IV

Epilogue
8
1
4
9
1
4

Chapter 20

Transdiagnostic Approaches to Emotion


Regulation: Basic Mechanisms
and Treatment Research
Brian C. Chu, Junwen Chen, Christina Mele,
Andrea Temkin, & Justine Xue

Transdiagnostic approaches
The broad construct of emotion regulation refers to multiple and diverse processes that individu-
als use to modulate emotional experience and involves neuro-╉anatomical, neuro-╉chemical, physi-
ological, behavioral, cognitive, and interpersonal systems, among others (Gross, 1998). Systems
are divided into multiple domains for greater specificity (e.g., the cognitive system can be under-
stood across adaptive and maladaptive cognitive strategies; automatic and conscious processes),
and domains can be separated further into sub-╉domains (e.g., automatic cognitive processes can
be broken down into particular strategies, such as suppression, distraction, and cognitive avoid-
ance). Each of these regulatory processes can be defined, assessed, and conceptualized along con-
tinuous or (likely arbitrarily defined) discrete dimensions.
These multiple regulatory systems and domains operate in close coordination, as activation of
one system typically calls for response (either activation or inhibition) of other systems (Werner
& Gross, 2010). As an example, behavioral problem solving (generating and weighing options;
choosing and acting on decisions) calls for the activation of multiple cognitive, physiological, and
behavioral systems. It is no surprise that such coordinated emotion regulation processes have
been implicated in the onset, maintenance, and amelioration of nearly every psychological disor-
der or problem set (Kring & Sloan, 2010).
Such a complex and influential system deserves an epistemological approach that honors
the inter-╉dependent nature of emotion regulatory systems. An explicit transdiagnostic research
agenda recognizes this complexity and inter-╉relatedness amongst systems and has the potential
to optimize research efforts across pathology groups and regulatory systems. The modern-╉day
transdiagnostic research agenda emerged from two parallel efforts (Ehrenreich-╉May & Chu, 2013;
Mansell et al., 2009). First, pathology researchers were interested in explaining the tremendous
co-╉occurrence (comorbidity) amongst related disorders and increasing understanding of what
mechanisms accounted for commonalities in presentation, impairment, and development across
disorders. Until the turn of the millennium, clinical research was frequently conducted in the
context of single-╉disorder research agendas. Efforts by experts across diverse fields of clinical psy-
chology (Barlow, Allen, & Choate, 2004; Fairburn, Cooper, & Shafran, 2003; Harvey, Watkins,
Mansell, & Shafran, 2004)  attempted to summarize converging lines of evidence generated in
isolation of each other. Transdiagnostic work signaled a desire to identify and integrate these con-
verging lines of research that helped explain the commonalities and distinctions across disorders
and problem sets.
0
2
4

420 Transdiagnostic Approaches to Emotion Regulation

A second aim of transdiagnostic research sought to economize and enhance the effects of
evidence-​ based treatments for psychological disorders. Most evidence-​ based intervention
research has produced treatment protocols designed to treat a single disorder, resulting in literally
hundreds of distinct treatment protocols to treat individual problem areas (Chorpita & Daleidan,
2009), despite the fact that the majority of protocols were primarily comprised of a highly over-
lapping set of clinical practices (e.g., in vivo exposure, problem solving). Thus, a second goal of
transdiagnostic research has been to consolidate evidence-​based interventions into a more effi-
cient set of empirically-​supported practices. It has also been hypothesized that treatments might
gain robustness to the degree that common practices targeted core underlying mechanisms that
maintained the diverse classes of pathology.
A transdiagnostic framework might be particularly relevant for children and adolescents. The
high rates of comorbidity (the co-​occurrence of two or more disorders) seen in adult popula-
tions are even higher in children and adolescents, where both within-​class (e.g., multiple anxi-
ety diagnoses), and across-​class comorbidity (e.g., diagnosis of anxiety and conduct disorder)
make comorbidity the rule rather than the exception (Angold, Costello, & Erkanli, 1999; Garber
& Weersing, 2010). The field has become increasingly aware of the importance of dimensional
conceptualizations of distress and multiple-​domain outcomes (functional impairment, symptom;
Achenbach, 2005). Dimensional models are particularly relevant where great symptom overlap
exists across disorders, as they do in youth, and where rapid development leads to transitory
symptoms across developmental stages. In addition, multiple informants (e.g., youth, parent,
teacher, doctor, coach) add complexity to any diagnostic picture that may be best accommodated
by dimensional and multi-​domain models.
A transdiagnostic approach may also help explain divergent trajectories and multifinality (the
case where a single risk factor leads to the subsequent expression of different disorders), key con-
cepts in developmental pathology (Nolen-​Hoeksema & Watkins, 2011). As one example, longitu-
dinal evidence suggests that many, but not all, teens and young adults who develop depression first
display evidence of anxiety earlier in life. Which teens and adults ultimately develop depression,
which retain their anxiety disorders, and which show remission from anxiety? Transdiagnostic
research encourages simultaneous evaluations of multiple processes (risk factors, mediators,
moderators) across disorders. Such an approach permits unique understanding of the relative
impact of multiple processes that lead to unifying and distinctive outcomes.
As implied in its name, transdiagnostic research aims to identify or change mechanistic pro-
cesses that unify related disorders and problem sets. However, it is critical to note that the bound-
aries of each disorder class (e.g., anxiety disorders) and specific disorder (e.g., generalized anxiety
disorder) are respected even as researchers aim to identify commonalities (Ehrenreich-​May &
Chu, 2013). It would be an error to mistake transdiagnostic research with prior attempts to “lump”
all of diagnostic categories into a single or limited set of general personality traits or distress fac-
tors (Taylor & Clark, 2009). Transdiagnostic research acknowledges that traditional diagnostic
categories, represented by classification systems, such as the Diagnostic and Statistical Manual of
Mental Disorders (DSM-​5; American Psychiatric Association, 2013), still retain helpful organiz-
ing heuristics. After all, individual anxiety disorders still share many more commonalities with
each other than they do with conduct or schizophrenic disorders. Thus, transdiagnostic research
encourages unifying frameworks while retaining the knowledge we have gained from the science
of diagnosis.
This framework is consistent with initiatives to identify the dimensional neurobiological under-
pinnings of psychological disorders. The recently initiated National Institute for Mental Health’s
(NIMH) Research Domain Criteria (RDoC) project (NIMH, 2011) aims to bring mental health
research in line with other areas of medicine that base diagnostic systems on underlying biology
1
2
4

CROSS-​S ECTIONAL RESEARCH WITH ADULTS 421

and not just clinical presentations of symptoms (Insel, 2013). RDoC emphasizes underlying bio-
logical and behavioral mechanisms as the root units of understanding pathology. In this way,
transdiagnostic research can provide a bridge between traditional classification-╉oriented research
and tomorrow’s dimensional, mechanism-╉oriented conceptualization of emotional distress.
The current chapter reviews the state of emotion regulation research from a transdiagnostic
framework. The first half of the review focuses on experimental, survey, and neurobiological
research that explores etiological and maintenance roles of emotion regulation across disorders.
The second half of our review evaluates transdiagnostic psychological interventions that incor-
porated emotion regulation techniques or assessed emotion regulation processes as outcomes.
By taking a transdiagnostic approach to our review, we hoped to demonstrate how investigations
that take a multi-╉disorder approach can help the field understand more about both the underlying
process and the organizing disorders of multiple problems in a unifying framework.

Experimental, survey, and neurobiological designs


To complete a review of basic research on transdiagnostic emotion regulation processes, includ-
ing cross-╉sectional and longitudinal designs, we identified empirical studies that (1) included par-
ticipant samples identified with two or more diagnostic or problem-╉areas, and (2) assessed any
emotional regulation process and evaluated that process across disorders. The goal of this review
was to evaluate the strengths and weaknesses of the extant literature, to reflect on the current
knowledge base of the emotion regulation processes that unify or distinguish diverse disorders,
and to offer recommendations for future directions in transdiagnostic research.
Studies that recruited children, adolescent, or adult samples were included in this review
because of limited available research focused on emotion regulation across multiple disorders.
Types of methodology included cross-╉sectional, longitudinal, clinical trial, and neuro-╉biological
(e.g., imaging) studies. We focus our review on evaluating the approach used to define and com-
pare disparate disorder groups, the assessment tools implemented, and the clarity of results in
establishing common and unique emotion processes that affect pathology.

Cross-╉sectional research with adults


Most adult research involving multiple disorder classes or symptom profiles has taken advan-
tage of cross-╉sectional research designs. Table 20.1 details study design, participants, measures,
and results for cross-╉sectional research. Most of this research has shown heterogeneity across the
types of emotion regulation strategies examined. Cognitive emotion regulation strategies (e.g.,
rumination, reappraisal, worry) have been defined as cognitive responses to emotionally provok-
ing events in which an individual attempts to modulate his/╉her emotional reaction (e.g., Aldao
& Nolen-╉Hoeksema, 2010, Campbell-╉Sills & Barlow, 2007). The literature has also labeled cer-
tain emotion regulation strategies as “adaptive” or “maladaptive” based on their function across
contexts. Adaptive emotion regulation strategies (e.g., acceptance, reappraisal, problem-╉solving)
are defined as strategies that are negatively correlated with psychopathology severity. Conversely,
maladaptive strategies (e.g., rumination, suppression) have been associated with the contribution
to and maintenance of disorders (e.g., Aldao & Nolen-╉Hoeksema, 2010; Conklin et al., 2015).
Cognitive emotion regulation strategies have been examined as a key mechanistic process across
internalizing disorders. Aldao and Nolen-╉Hoeksema (2010) examined four cognitive emotion
regulation strategies (rumination, thought suppression, reappraisal, and problem-╉solving) across
two internalizing disorders (anxiety and depression) and a non-╉internalizing disorder (eating dis-
orders) among 252 undergraduate students. The strategies were also categorized as either adaptive
(reappraisal, problem-╉solving) or maladaptive (suppression, rumination [including brooding and
24
Table 20.1  Adult cross-​sectional

Citation Sample Characteristics and ER Measures ER components (all that Processes found to be Processes found to be
Design were assessed) universal disorder specific
Aldao & Nolen-​ N = 252; Undergraduate COPE (Problem-​Solving Rumination, thought Rumination, suppression, and None
Hoeksema (2010) students with symptoms of subscale), ERQ, RRS, suppression, reappraisal, and reappraisal were significantly
depression, anxiety, and eating WBSI problem-​solving related to anhedonic
disorders depression (.44), anxious
arousal (.24), and eating
disorders symptoms (.15)
Conklin et al., N = 81; Clinical trial patients Brief COPE Maladaptive strategies and Not analyzed Not analyzed
(2015) with comorbid alcohol use adaptive strategies
and anxiety disorders: GAD
(67.9%), social phobia
(50.6%), panic disorder (9.9%)
Desrosiers et al. N = 187; Mood and Anxiety DERS, ERQ, FFMQ, Rumination, reappraisal, Simple mediation Reappraisal linked to
(2013) Disorder Clinic patients, ages PSWQ, RRS worry, and nonacceptance model: Rumination linked depression; Worry linked to
17–​81, with GAD (42.9%), to anxiety and depression; anxiety
MDD (20.1%), or social phobia Multiple medation
(12.2%) model: Rumination linked to
depression
Vine & Aldao N = 211; Undergraduate ATTC, DERS Emotional clarity Emotional clarity deficits Specific indirect
(2014) students, ages 18–​32, linked to anhedonic pathways: anhedonic
with seven symptom depression, social anxiety, depression and shifting
types: anhedonic depression, borderline personality, binge attention; social anxiety and
anxious arousal, social anxiety, eating, and substance abuse acceptance and strategies;
borderline personality, binge symptoms borderline personality
eating, restrictive eating, and symptoms and shifting and
alcohol use. strategies; alcohol use and
impulse
Gruber et al. N = 60; Adults with euthymic CCL, GRS, PSWQ Rumination, worry, and BP and INS significantly linked None
(2008) bipolar disorder (n = 21), negative automatic thoughts to rumination and worry;
insomnia (n = 19), and non-​ finding not significant when
clinical control (n = 20) analysis controlled for anxiety
and depressive symptoms
3
2
4
Brockmeyer et al. N = 140; Women, ages 18-​ ER self-​report: DERS Experience and differentiation MDD and AN linked to MDD sample had increased
(2012) 65, with MDD (41), anorexia of emotions and attenuation elevated difficulties in difficulties in attenuating and
nervosa (39), healthy controls and modulation of emotions experience and differentiation modulating emotions
(60) of emotions
Queen & N = 76; Youth, ages 12–​18, CBCL, CEMS, EESC, Cognitive reappraisal, Higher comorbidity linked Possibility that expressive
Ehreneich-​May with comorbid anxiety and ERQ-​CA, PANAS emotion suppression, to higher youth reported reluctance, emotional
(2014) depression (57.9%) or a emotional awareness, negative affect, expressive expression, emotional
primary anxiety disorder expressive reluctance, reluctance, emotional awareness are linked to
(42.1%) emotional inhibition, positive expression, emotional depression, but lack of
affect, negative affect awareness; Parent depression-​only group makes
report: inhibition of sadness this difficult to determine
Garnefski et al. N = 271; Youth, ages 12–​18, CERQ Self and other blame, Catastrophizing Internalizing: self-​
(2005) with Internalizing symptoms rumination, catastrophizing, blame and rumination;
(8.9%), Externalizing putting into perspective, Externalizing: positive
symptoms (8.9%), Comorbid positive refocusing, positive refocusing; Cognitive ER
internalizing and externalizing reappraisal, acceptance, and strategies explained more of
symptoms (4.8%), “No planning the variance for internalizing
problems” (38%) than externalizing symptoms

Table Abbreviations
ER self-​report: Attentional Control Scale = ATTC, Brief COPE, Child Behavior Checklist = CBCL, Children’s Anger Management Scale = CAMS, Children’s Emotion Management Scale = CEMS,
Children’s Response Styles Questionnaire = CRSQ-​Rumination, Children’s Sadness Management Scale = CSMS, Cognitions Checklist = CCL, Cognitive Emotion Regulation Questionnaire = CERQ,
COPE, Difficulties in Emotion Regulation Scale = DERS, Emotion Expression Scale for Children = EESC, Emotion Regulation Questionnaire = ERQ, Emotion Regulation Questionnaire for Children
and Adolescents = ERQ-​CA, The Five Facet Mindfulness Questionnaire = FFMQ, Global Rumination Scale = GRS, modified Differential Emotions Scale = mDES, Positive and Negative Affect
Schedule = PANAS, Penn State Worry Questionnaire = PSWQ, Ruminative Response Scale = RRS, White Bear Suppression Inventory = WBSI
Symptom self-​report: Anxiety Sensitivity Index = ASI, Bech-​Rafaelsen Mania Scale = BRMS, Beck Anxiety Inventory = BAI, Beck Depression Inventory = BDI, The Binge-​Eating Scale = BES, The Brief
Fear of Negative Evaluation = BFNE, Children’s Depression Inventory = CDI, Children’s Eating Attitudes Test = chEAT, Clinical Global Impression-​Severity = CGI-​S, Child Symptom Inventory = CSI,
Depression Anxiety and Stress Scale = DASS, The Eating Disorders Attitude Test = EAT-​26, Eating Disorders Examination-​Questionnaire (EDE-​Q), Global Assessment of Functioning = GAF; DSM-​IV Axis
V, Hamilton Anxiety Rating Scale = HAM-​A, Hamilton Rating Scale for Depression = HRSD, Inventory of Depressive Symptomatology, IDS-​C, Liebowitz Social Anxiety Scale = LSAS, Lifetime Interference
Measure = LIM, McLean Screening Instrument for Borderline Personality Disorder = MSI-​BPD, Mood and Anxiety Symptom Questionnaire = MASQ, Mood and Anxiety Symptom Questionnaire-​Short
form = MASQ-​SF, Obsessive Compulsive Drinking Scale = OCDS, Overall Anxiety Severity and Impairment Scale = OASIS, Quick Inventory of Depressive Symptomology = QIDS, Revised Children’s
Anxiety and Depression Scales = RCADS/​RCADS-​Parent version, Revised Peer Experiences Questionnaire = RPEQ-​aggressor version, The Short Michigan Alcohol Screening Test = SMAST-​G, The Social
Interaction Anxiety Scale = SIAS, State-​Trait Anxiety Inventory = STAI, Youth Self-​Report = YSR
Semi-​structured interviews: Anxiety Disorders Interview Schedule = ADIS-​IV, Anxiety Disorders Interview Schedule–​Child Version, Child and Parent report forms = ADIS-​IV-​C/​P, Diagnostic Interview
Schedule for Children = DISC, Duke Structural Interview for Sleep Disorders = DSISD, Insomnia Diagnostic Interview = IDI, Structured Clinical Interview for DSM Disorders = SCID-​I
FMRI and Physiological Measures: High Frequency heart rate variability = HRV-​HF, Functional Magnetic Resonance Imaging = fMRI, Respiratory Sinus Arrhythmia = RSA, The Zephyr[TM]
BioHarness[TM] device = wireless physiological monitoring system
4
2

424 Transdiagnostic Approaches to Emotion Regulation

pondering]). Participants completed a battery of emotion regulation questionnaires and symptom


measures in one sitting. As expected, adaptive strategies were negatively correlated with anxiety/​
depression and eating disorder symptoms, and maladaptive strategies were positively correlated
with symptoms. A structural equation model also revealed that rumination (brooding, ponder-
ing) and suppression had positive loadings, and reappraisal had a negative loading, on the single
latent factor of cognitive emotion regulation. Problem solving did not load onto the latent fac-
tor of cognitive emotion regulation. Model loadings suggested that adaptive strategies evidenced
weaker relationships with distress than maladaptive strategies. One key finding revealed that
rumination, suppression, and reappraisal were significantly related to all three types of psychopa-
thology: anhedonic depression, anxious arousal, and eating disorder symptoms. Overall, a latent
factor of cognitive emotion regulation was significantly associated with symptoms of all three
disorders, providing evidence for a role for cognitive emotion regulation strategies across anxiety,
depression, and eating pathology.
Conklin et  al. (2015) examined the relationship between adaptive and maladaptive emo-
tion regulation strategies in 81 adults with comorbid anxiety and alcohol use disorders. Data
were used from a clinical trial in which participants were randomized to one of four cognitive
behavioral conditions: Unified Protocol for Transdiagnostic Treatment of Emotional Disorders
or progressive muscle relaxation, and both conditions combined with either venlafaxine, or pill
placebo. A different set of adaptive strategies were assessed using a revised BRIEF COPE measure
(Carver, 1997) to capture active coping, planning, use of emotional support, use of instrumental
support, positive reframing, acceptance, religion, sharing emotions (e.g., venting), and humor.
Maladaptive strategies included denial, self-​blame, behavioral disinhibition, and substance use.
Change in adaptive and maladaptive strategies before and after treatment was the primary inter-
est. The key results showed that the change in use of maladaptive strategies was positively related
to symptom severity at pre-​and post-​treatment. Decreases in the use of maladaptive strategies
were significantly related to a decrease in symptom severity across treatment. Use of adaptive
strategies was associated with lower symptom severity only for individuals with a higher use of
maladaptive strategies at pre-​treatment. These findings helped to demonstrate an association
between emotion regulation strategies and the severity of psychopathology at repeated points in
treatment and provided supporting data to establish emotion regulation as potential mediators
of treatment outcomes. However, they did not examine differential relations between emotion
regulation and either disorder class or differential change in disorder status. Thus, we have limited
data to draw conclusions about the relative importance (commonality, distinctiveness) for these
emotion regulation strategies across anxiety and alcohol use.
Desrosiers, Vine, Klemanski, and Nolen-​Hoeksema (2013) sought to provide further specific-
ity on which cognitive emotion regulation strategies mediate associations between mindfulness
and distress. Reappraisal, nonacceptance, rumination, and worry were examined in relation to
depression and anxiety. Participants included a clinical sample of 187 adults, ages 18–​71, seek-
ing treatment at a mood and anxiety disorders clinic. Simple mediation model analyses revealed
that rumination significantly mediated associations between mindfulness and both anxiety and
depression, whereas multiple mediation analyses reveal that rumination showed specificity to
depression symptoms. That is, lesser use of mindfulness was associated with greater depression,
but this relation was mediated to the extent that individuals ruminated. While this finding sug-
gests that rumination operates as a transdiagnostic mechanism across depression and anxiety,
reappraisal was found to mediate only depression symptoms and worry was found to mediate only
anxiety symptoms. The authors provide several plausible explanations for these differences. While
rumination and worry both involve repetitive thinking, worry typically involves over-​engagement
about future negative outcomes, which is more closely associated with anxiety symptoms, and
5
2
4

Cross-sectional research with adults 425

rumination typically involves over-​engagement with both present and past outcomes, which is
associated with both anxiety and depressive symptoms. Reappraisal’s specificity to depressive
symptoms might be related to the benefits of mindfulness, which can help individuals to adopt a
nonjudgmental stance and potentially disengage from their repetitive thinking.
Rumination and worry were examined with negative automatic thoughts in a study assessing
differences between euthymic bipolar I disorder (BP; n = 21), insomnia (INS; n = 19), and non-​
clinical comparisons (NC; n = 20) (Gruber, Eidelman, & Harvey, 2008). While BP and INS par-
ticipants reported significantly more rumination and worry than NC participants, BP and INS
participants did not differ from one another in rumination and worry. The study also found that
the BP group reported significantly more negative automatic thoughts than the NC group; how-
ever, BP and INS groups again did not differ from one another in negative automatic thoughts.
In follow-​up analysis, rumination and worry were no longer elevated in BP and INS compared
to NC after controlling for anxiety and depressive symptoms (Gruber et al., 2008). These results
highlight important methodological and conceptual issues when studying transdiagnostic pro-
cesses. Controlling for anxiety and depression may be important to adjust for any spurious effects
of general distress or symptom severity when comparing a mechanism across disorders. At the
same time, removing the effects of general distress may also be removing important explanatory
processes that may explain commonalities across disorders. We simply may not have the measures
that can adequately parse out the variance that is associated with unique transdiagnostic processes
from general distress.
Based on these studies (Aldao & Nolen-​Hoeksema, 2010; Desrosiers et al., 2013), there is evi-
dence that some cognitive emotion regulation strategies might operate transdiagnostically, while
others might operate more uniquely to certain disorders. In particular, rumination might be a
transdiagnostic mechanism across depression and anxiety, whereas worry might relate more
closely to anxiety. The relationship between reappraisal and pathology may be more mixed. While
Aldao and Nolen-​Hoeksema (2010) found that reappraisal was significantly related to anxiety
and depression, Desrosiers et al. (2013) found that reappraisal was significantly related to depres-
sion only. Another interesting finding discovered through latent class analysis was that the con-
struct of cognitive emotion regulation may be more strongly influenced by maladaptive cognitive
strategies than adaptive cognitive strategies. Adaptive strategies also appear to be helpful because
they decrease the use of maladaptive strategies more than because they are positive in-​and-​of
themselves.
Another way to examine emotion regulation difficulties across disorders is to investigate the
central facets of emotion regulation rather than certain constructs (e.g., rumination, worry).
Using the Difficulties in Emotion Regulation Scale (DERS), Brockmeyer et  al. (2012) exam-
ined difficulties 1) experiencing and differentiating emotions (i.e., non-​acceptance of emotional
responses, lack of emotional awareness, and lack of emotional clarity) and 2)  attenuating and
modulating emotions (i.e., difficulties in engaging in goal-​directed behavior, impulse control dif-
ficulties, and limited access to effective emotion regulation strategies) in women diagnosed with
major depression (MDD) or anorexia nervosa (AN), or considered healthy controls (N = 140).
Overall, individuals with both MDD and AN reported greater difficulty regarding the experience
and differentiation of emotions compared to healthy controls. However, difficulties in attenuating
and modulating emotions was significantly higher only in individuals with MDD, and there was
no significant difference between the AN or NC groups (Brockmeyer et al., 2012). These results
provide examples of both disorder-​specific and transdiagnostic emotion regulation processes.
Research has also focused on how specific facets of emotion regulation mediate the relation-
ship between broader emotion regulation processes and psychopathology. Vine and Aldao (2014)
examined whether deficits in the broad construct of emotional clarity would correlate with
6
2
4

426 Transdiagnostic Approaches to Emotion Regulation

psychopathology (i.e., symptoms of anhedonic depression, social anxiety, borderline personality,


binge eating, restrictive eating, anxious arousal, and substance abuse) by way of impaired emotion
regulation. Participants were 211 undergraduate students who completed the DERS and multiple
self-╉report symptom measures. Multiple regression analyses revealed that deficits in emotional
clarity were significantly related to five symptom types:  anhedonic depression (β  =  .56), social
anxiety (β = .49), borderline personality (β = .47), binge eating (β = .52), and alcohol use (β = .21);
however, deficits were not significantly related to restrictive eating and anxious arousal. These
results show that deficits in emotional clarity were more closely related to anhedonic depres-
sion and borderline personality disorder symptoms. Multiple mediation model analyses revealed
that there were several emotion regulation pathways that significantly related to the aforemen-
tioned symptoms. Individuals with anhedonic depression experienced a significantly lower abil-
ity to shift attention (Attentional Control Scale Shifting), which was associated with depression
severity. Individuals with social anxiety showed significant difficulties with acceptance (DERS
Acceptance) and access to strategies (DERS Strategies). Individuals with borderline personality
symptoms showed that the specific indirect pathways through shifting and strategies were sig-
nificant, with the effect of the “strategies” pathway being significantly larger than the “shifting”
pathway (e.g., deficits in emotional clarity  shift attention  borderline personality symptoms).
That is, individuals who demonstrated deficits in emotional clarity also reported lower ability
to shift attention, which in turn was associated with higher levels of borderline symptoms. No
specific indirect pathways were significant for individuals with binge eating symptoms. There
was one significant specific indirect pathway through impulse (DERS Impulse) for participants
with problematic alcohol use, which showed that individuals with deficits in emotional clarity
tended to report difficulty regulating impulsive behavior. These findings show that a single emo-
tion regulation process, such as emotional clarity, may predict multiple types of pathology in a
transdiagnostic fashion. However, these relationships appear to be mediated by different emotion
regulation mechanisms, depending on the type of pathology.

Cross-╉sectional research with youth


Evidence of disorder-╉specific and transdiagnostic emotion regulation processes has also been
found in cross-╉sectional research with youth (see Table 20.1). Queen and Ehrenreich-╉May (2014)
examined affect and emotion regulation strategies (cognitive reappraisal, emotional suppression,
emotional awareness, expressive reluctance, emotional inhibition) as a function of comorbidity in
adolescents (12–╉18-╉years-╉old). The study compared 44 youth with diagnosed comorbid anxiety
and depressive disorders with 32 youth with diagnosed anxiety and no comorbid depressive disor-
der. Parents and youth were assessed using a combination of semi-╉structured interviews and self-╉
report questionnaires. Results found that youth with comorbidity were rated as having a higher
Clinical Severity Rating for emotional disorders and higher Lifetime interference measures based
on a semi-╉structured interview. They did not, however, have higher Clinical Global Impression
(CGI) ratings overall. Youth self-╉report found individuals with comorbidity reported higher levels
of negative affect, expressive reluctance, and emotional suppression, in addition to poorer emo-
tional awareness and lower positive affect. Parents reported that youth had higher inhibition of
sadness, but did not report differences between inhibition of worry or anger between the two
groups. Youth with comorbid anxiety and depression appear to demonstrate greater impairment
and greater emotion regulation difficulties. While these results do not directly implicate common
or unique processes, they do suggest that depression may convey greater risk for unique emotion
regulation difficulties, specifically in awareness and experience of emotions. However, without
a depression only group, it is unclear if the presence of depression alone that is linked to higher
7
2
4

Cross-sectional research with youth 427

dysregulation, or if it is something unique about the combination of anxiety and depression that
is of key importance. Further examination of expressive reluctance, emotional expression, and
emotional awareness may play an important role in the development or maintenance of comorbid
anxiety and depression.
In addition to emotion regulation strategies involving awareness and recognition of different
emotions, cognitive emotion regulation strategies have also been thought to influence disor-
der trajectories, and may be particularly important in youth given the vast changes to cognitive
capacity that occur between childhood and adolescence. Garnefski, Vivian Kraaij, and Marije van
Etten (2005), for example, compared a number of cognitive emotion regulation components (i.e.,
self and other blame, rumination, catastrophizing, putting into perspective, positive refocusing,
positive reappraisal, acceptance, and planning) among 271 12-​to-​18-​year-​old teens who demon-
strated internalizing symptoms, externalizing symptoms, comorbid internalizing, and external-
izing symptoms, or no significant symptoms. It was hypothesized that self-​blame, rumination,
catastrophizing, and lack of positive reappraisal would be most highly associated with internal-
izing and externalizing symptoms. Regression analyses were completed to determine the relation-
ship between these cognitive strategies and internalizing problems (while controlling for gender,
age and number of externalizing symptoms), and between cognitive strategies and externaliz-
ing problems (while controlling for gender, age and number of internalizing symptoms). Results
showed that adolescents with internalizing problems (either alone or appearing with externaliz-
ing symptoms) reported significantly higher levels of self-​blame and rumination than those with
externalizing problems alone or the healthy controls. No differences were found between inter-
nalizing and externalizing youth with regard to catastrophizing. Adolescents with externalizing
problems were more likely to report positive refocusing. With the exception of catastrophizing,
there was no overlap between the specific cognitive emotion regulation strategies that predicted
internalizing and externalizing problems, and cognitive emotion regulation strategies were able to
explain more of the variance of internalizing problems than of externalizing problems. This sug-
gests that cognitive emotion regulation strategies in general may play a larger role in contributing
to internalizing problems than to externalizing problems, and that these mechanisms may play a
key role in distinguishing between the two symptom profiles.
The results of these adult and youth cross-​sectional studies highlight some of the ways in which
emotion regulation strategies can be conceptualized and how they might operate uniquely or
transdiagnostically across psychopathology. Emotion regulation mechanisms can be further
classified based on their hypothesized function, such as cognitive strategies (e.g., rumination,
worry, reappraisal), and adaptive or maladaptive strategies (e.g., Aldao & Nolen-​Hoeksema,
2010; Desrosiers et al., 2013). Certain emotion regulation strategies have also been examined as
mediating variables (Vine & Aldao, 2014). One interesting takeaway from the adult studies seems
to be the greater link of maladaptive emotion regulation strategies to pathology than adaptive
emotion regulation strategies. These initial findings suggest that future research, and potentially
clinical work, should focus more on how the reduction of maladaptive strategies, rather than the
increase of adaptive strategies, relates to the development and maintenance of pathology over
time. Additionally, as it stands, the above studies suggest that cognitive emotion regulation strate-
gies may play an influential role in distinguishing between disorders within a youth population.
Overall, the strengths of these studies include examining an array of emotion regulation strategies
in both clinical and nonclinical populations, the use of multi-​method assessments (diagnostic and
semi-​structured interviews, such as the ADIS, SCID), and mixed methods (e.g., multiple media-
tion models, measurement, and structural models).
Unfortunately, there is a paucity of cross-​sectional studies looking at emotion regulation
across disorders in youth, and additional research could help to further clarify specific emotion
8
2
4

428 Transdiagnostic Approaches to Emotion Regulation

regulation strategies that may differentiate between or underlie multiple different disorders in
children and adolescents. With further assessment, this information may be beneficial in inform-
ing which mechanisms can be targeted to produce the most effective change in youth with dif-
ferent disorder profiles. It is important to note that while cross-╉sectional research can help detect
possible links between strategies and symptomatology, the design limits examination of how the
relationship between emotion regulation strategies and psychopathology might evolve over time.
Additionally, despite some use of multimodal assessment, there is heavy reliance on self-╉report
measures in the majority of cross-╉sectional studies, which lends itself to shared variance concerns
and does not provide any directions for specific behavioral or physiological indicators of emotion
regulation strategies or symptomology. Within both the adult and youth studies, recruitment of
nonclinical (e.g., undergraduate students, community youth) samples, which restricts the range
of psychopathology and might weaken the relationship to emotion regulation strategies. Overall,
however, these cross-╉sectional studies provide a foundation for future studies, especially prospec-
tive and experimental studies, which are needed to assess if these processes contribute to the
etiology or maintenance of disorders.

Biomarker and fMRI Studies


Psychophysiogical and advanced imaging research provide exciting directions for understanding
emotion regulation across disorders and diverse problem sets (Table 20.2). Gruber, Mennin, Fields,
Purcell, and Murray (2015) looked at mean levels of intra-╉individual changes in high-╉frequency
heart rate variability (HRV-╉HF), which has been associated with dysfunctional experiencing of
positive emotions. Participants included 18–╉60-╉years-╉old adults, 21 of whom met criteria for
bipolar disorder, 17 of whom met criteria for MDD, and 28 healthy controls. Additional mea-
sures assessed mood symptoms, global functioning, and state and trait positive affectivity. Results
found no group differences in mean HRV-╉HF between bipolar disorders I, MDD, and controls.
However, there was greater HRV-╉HF instability in individuals with bipolar but this instability was
not associated with dimensional measures of positive affect. The results support previous find-
ings that bipolar disorders are linked to higher perceived variability in positive emotions, which
is subsequently linked to increased depressive symptoms. The authors concluded that the study
highlights the potential for research to move beyond laboratory and questionnaire based studies,
and emphasized the importance of assessment over time.
More recently, Burklund, Craske, Taylor, and Lieberman (2015) conducted a functional mag-
netic resonance imaging (fMRI) study, which included participants with Social Phobia (SP) with-
out comorbidity (n = 30), with comorbid depression (n = 18), with comorbid anxiety (n = 19),
and healthy controls (n = 15). Participants were scanned while completing an emotion regula-
tion task that involved affect labeling (e.g., labeling photographs of emotional facial expressions).
fMRI data analyses revealed that individuals with pure SP and all comorbidity types showed an
upregulation of amygdala activity compared to healthy controls during affect labeling, suggesting
altered emotion regulation capacity in SP. Further, individuals with comorbid depression showed
significant upregulation, or increased amygdala activity, compared to the other groups. This high-
lights the significant impact of comorbidity within disorders and the importance of considering
the role of comorbidity when examining emotion regulation processes across disorders. Burklund
et al. (2015) also found that SP individuals with comorbid anxiety and depression had greater
right ventral lateral prefrontal cortex (RVLPFC) activity, or amygdala reactivity, compared to SP
individuals without comorbidity. These results highlight the role of comorbid symptoms when
examining emotion regulation processes even when controlling for a single primary disorder (i.e.,
SP as in this study). These findings suggest that the individual’s difficulty with downregulating
9
2
4

Biomarker and fMRI Studies 429

Table 20.2  Adult fMRI and biomarker studies

Citation Sample ER and ER Processes found Processes


Characteristics Physiological components to be universal found to
Measures (all that were be disorder
assessed) specific
Gruber et al. N = 66; Ages ER physiological High Mean HRV-​HF HRV-​HF
(2015) 18–​60, with measure: The Frequency consistent across instability
BD (n = 21), Zephyr[TM] heart rate groups seen in
MDD (n = 17), BioHarness[TM] variability individuals
healthy controls device (measures (HRB-​HF) with BD
(n = 28) of parasympathetic
nervous system
activity) ER self-​
report: mDES (trait
and state PA)
Burklund N = 82; Ages ER physiological Upregulation Greater right Individuals
et al. (2015) 18–​45, with measure: fMRI of amygdala ventral lateral with
SP (n = 30), SP labeling and activity prefrontal cortex comorbid
and comorbid reactivity task (RVLPFC) activity, depression
depression or amygdala showed
(n = 18), SP reactivity in SP greater
with comorbid individuals with upregulation
anxiety (n = 19), comorbid anxiety compared to
and healthy and depression other groups
controls (n = 15)
Manber Ball n = 64: Adults ER physiological Reappraisal Less activation None
et al. (2013) with GAD (n measure: fMRI of emotional in PFC in both
= 23), panic (activation in responses, GAD and PD
disorder (n = prefrontal cortex maintenance participants
18), and healthy (PFC) during of emotional
controls (n = emotion regulation) responses
23)

neural emotion responses to negative stimuli likely relates to deficits in emotion regulation capac-
ity. It is difficult to isolate the unique contributions of depression and anxiety disorders other than
SP here, but findings suggest that greater pathology contributes to greater amygdala upregulation,
which reflects greater emotion regulation dysregulation. It would take extra steps to determine the
magnitude and type of effects each disorder contributes to amygdala dysfunction.
Ball, Ramsawh, Campbell-​Sills, Paulus, and Stein (2013) conducted an fMRI study including
participants with primary diagnoses of generalized anxiety disorder (GAD; n = 23), panic disor-
der (PD; n = 18), and healthy controls (HCs; n = 23). Participants completed self-​report measures
and were scanned while completing a task that required them to reappraise or maintain their
emotional responses to negative images. The study was designed to test the hypothesis that GAD
and PD would evidence hypo-​activation in the prefrontal cortex (PFC) during emotion regu-
lation attempts. Analyses of self-​report measures revealed individuals with GAD reported the
least reappraisal use in daily life. Reappraisal use was inversely associated with anxiety severity
and functional impairment. fMRI data analyses showed that HCs had greater activation during
both reappraisal and maintenance in brain areas important for emotion regulation (i.e., dorsolat-
eral and dorsomedial PFC), whereas GAD and PD participants showed less activation in these
0
3
4

430 Transdiagnostic Approaches to Emotion Regulation

areas, even as activation levels did not differ in the two clinical groups. Furthermore, those with
the least PFC activation reported the greatest anxiety severity and impairment. These results pro-
vide cross-╉method evidence that cognitive reappraisal and maintenance of emotional response
might have transdiagnostic properties across GAD and PD.
Although in nascent stages, heart rate and fMRI studies provide emerging evidence for physio-
logical and neural functions of emotion regulation processes. Gruber et al. (2015) demonstrated
that inconsistency in heart rate variability may be specifically linked to bipolar disorder, compared
to MDD or healthy controls, and additional studies suggest that some forms of neural reactiv-
ity is consistent across more severe symptoms of pathology. Burklund et  al. (2015) found that
participants with comorbid anxiety and depression exhibited increased amygdala reactivity com-
pared to SP individuals without comorbidity (indicative of greater fear reactivity), and Ball et al.
(2013) found that individuals with GAD and PD exhibited less reactivity in prefrontal cortex areas
important for emotional regulation. These results provide evidence of several neural transdiag-
nostic mechanisms across anxiety and depression. Burklund et al. (2015) also found that there
are unique characteristics for individuals with comorbid depression, as they exhibited increased
amygdala activity compared to individuals with comorbid anxiety. Unfortunately, we could not
identify any youth-╉based studies that investigated biomarkers of emotion regulation processes
across multiple disorders. The more invasive assessment procedures used in this research may
make investigators cautious about using child and adolescent samples; investigators and human
subjects review boards may also expect a greater foundational evidence base before including a
youth population. As it stands, limited research exists to inform the field on developmental dif-
ferences of biological markers of emotion regulation. In summary, these studies underscore the
importance of expanding research to include physiological and neurological measures of emotion
regulation in order to gain greater specificity regarding potential transdiagnostic or disorder spe-
cific strategies across pathology.

Longitudinal designs
Cross-╉sectional studies can highlight important links between emotion regulation strategies and
disorders, but they do little to explain the development of regulation processes over time or their
reciprocal relations with socio-╉emotional distress. Longitudinal studies help explain how particu-
lar processes contribute to or maintain particular symptoms over time and clarify convergent and
divergent developmental trajectories (see Table 20.3).
McLaughlin, Hatzenbuehler, Mennin, and Nolen-╉Hoeksema (2011) conducted a longitudinal
assessment of the reciprocal relationship between psychopathology and emotion regulation skills.
Participants included 1065 sixth to eighth graders who completed self-╉report measures in school
at baseline and after seven months. Emotional understanding, adaptive expression of negative, and
cognitive emotion management strategies were targeted as emotion regulation processes. Results
demonstrated that all types of emotion regulation were inter-╉related and were positively associ-
ated with all four types of symptomology (anxiety, anger, eating pathology, depression), which
were also inter-╉related. Interestingly, analyses found that a one factor model best fit the data,
such that emotional understanding, dysregulated expression of sadness and anger, and ruminative
responses were better combined into one factor of emotion dysregulation rather than examined as
separate constructs. Using the single factor model, results showed that emotion dysregulation at
Time 1 predicted Time 2 anxiety, anger, and eating pathology, but not depression. However, Time
1 anxiety, anger, eating pathology, and depression did not predict emotion regulation at Time 2.
A number of interesting implications result. Unlike other studies that appear to find distinctions
between different elements of emotion regulation as they relate to different disorders, this study
1
3
4
Table 20.3  Child longitudinal

Citation Sample Characteristics ER Measures ER components (all that Processes found to be Processes found to be
were assessed) universal disorder specific
McLaughlin et al. N = 1065; Ages 11–​14 ER self-​report: CAMS, CRSQ-​ Emotional understanding, Four ER components best T1 ER did not predict
(2011) with anxiety, depression, Rumination, CSMS, EESC expression of anger and fit into a single emotion depression at T2
agression, and eating -​Assessment at T1 and T2 sadness, and cognitive dysregulation factor. T1 ER
pathology (7 months later) emotion management ratings predicted T2 anxiety,
strategies (rumination) aggression, and eating
pathology
Vasilev et al. (2009) N = 212; Ages 8–​12 with ER physiological RSA, nonacceptance Higher DERS linked to greater Access to ER strategies,
conduct disorder (n = 30), measure: RSA ER self-​ of emotional response, emotional withdrawal (as impulse control, and
depression (n = 28), report: DERS -​Assessment at difficulteis engaging in goal indicated by RSA) in response acceptance of emotional
comorbid conduct and 3 time points (each one year directed behaviors, impulse to sadness induction at Time response were most linked
depression (n = 80), control apart) control difficulties, lack of 1, but increased response at to reactivity
group (n = 69) emotional awarenes, limited Time 3. Lower DERS scores
access to ER strategies, lack of linked to stable reactivity
emotional clarity across all three time points
Pang & Beauchaine N = 159; Ages 8–​12 with ER physiological RSA TI higher levels of depression T1 comorbid depression
(2013) conduct disorder (n = 30), measures: RSA -​Assessment and conduct disorder and conduct disorder
depression (n = 28), at 3 time points (each one linked to lower baseline predicted lower baseline
comorbid conduct and year apart) RSA. Depression and RSA, over and above the
depression (n = 80), control conduct disorder samples main effects. Comorbid
group (n = 69) demonstrated elevated depression and conduct
emotional reactivity in disorder samples had
response to sadness induction highest RSA reactivity
at each of the time points
2
3
4

432 Transdiagnostic Approaches to Emotion Regulation

suggests that one broad construct actually fit the data more accurately. This may be due, in part,
to the four particular inter-​related elements of emotion regulation examined in this study; how-
ever, other ER constructs may not fit as well into a single factor. Dysregulation in emotion man-
agement also predicted subsequent symptomology at a later time point, but the reverse relation
was not found. Although a formal cross-​panel analysis was not performed to rule out additional
confounding covariates, these results provide initial support for a prospective role of emotion
regulation strategies in promoting anxiety, anger, and eating pathology. Further research in why
depression was not predicted by earlier emotion regulation dysregulation is warranted.
Hoping to examine the relationship between self-​reported emotion regulation and physiologi-
cal indices over time, Vasilev, Crowell, Beauchaine, Mead, and Gatzke-​Kopp (2009) conducted
a three-​year study with 212 eight to 12 year olds who were categorized in one of four groups:
non-​clinical controls, conduct problems, depressive problems, or comorbid conduct and depres-
sive problems. Physiological assessment of emotion regulation consisted of measuring respira-
tory sinus arrhythmia (RSA), which captures changes in heart rate, and has been shown to be
related to symptoms of depression, anxiety, self-​injury, and disruptive behavior (Beauchaine et
al., 2007; Crowell et al., 2005; Shannon, Beauchaine, Brenner, & Neuhaus, 2007; Silk, Steinberg, &
Morris, 2003). At each of the three time points, examiners assessed baseline and change in RSA
during a sadness induction task. The study aimed to link RSA at each assessment point to year
three self-​reported difficulties in emotion regulation, assessed by the self-​reported DERS (i.e.,
non-​acceptance of emotional response, difficulties engaging in goal directed behaviors, impulse
control difficulties, lack of emotional awareness, limited access to emotion regulation strategies,
and lack of emotional clarity). It is normative for individuals to demonstrate increasing baseline
levels of RSA during adolescence, and results from this study found that increasing baseline levels
of RSA at each of the time points was associated with greater emotional awareness at year three.
In comparison, individuals with relatively stable baseline RSA at each time demonstrated greater
emotion regulation difficulties at year three. Regarding change in RSA rate following the sadness
induction task, those who scored higher on the DERS tended to have greater emotional with-
drawal (as indicated by RSA) in response to sadness induction at Time 1, but increased response
at Time 3. Those who scored low on the DERS demonstrated relatively stable reactivity across
all three time points. Authors suggested that low baseline RSA may actually be associated in
increased instability in emotional reaction in response to triggers.
Additional analysis found that access to emotion regulation strategies, impulse control, and
acceptance of emotional response were most linked to reactivity. In related work, Pang and
Beauchaine (2013) looked at baseline RSA and changes in RSA across the different disorder
groups. At Time 1, higher levels of depression and conduct disorder were linked to lower RSA at
baseline, indicating increased emotional withdrawal. The interaction of depression and conduct
disorder also predicted lower baseline RSA, over and above the main effects. Individuals with
depression and conduct disorder also demonstrated greater emotional reactivity in response to
sadness induction at each of the time points, though individuals with comorbid depression and
conducted disorder displayed the highest levels of reactivity. None of the disorder profiles were
linked to trajectory changes in baseline RSA or RSA reactivity over time.
Collectively, these studies demonstrate the ability of researchers to begin exploring how emo-
tion regulation can impact change over time. The first study provides support for the notion that
poor ER plays a temporal role in the development of symptomology, such that emotion regula-
tion dysregulation precedes distress. Longitudinal studies using physiological markers of distress
have found a link between RSA and specific self-​reported emotion regulation strategies, such as
emotional awareness and acceptance, impulse control, and access to emotion regulation strate-
gies. The studies demonstrated that while both depression and conduct disorder were linked to
34

Summary of cross-sectional and longitudinal research 433

greater difficulties, comorbidity was associated with even higher levels of deficit. It appears that
individuals who demonstrated developmentally normative rates of RSA over adolescence may
actually respond most appropriately to emotional cues. In contrast, individuals with blunted base-
line levels of RSA demonstrate difficulty with emotion regulation when presented with triggers.
Interestingly, all of these studies have been conducted with youth.

Summary of cross-╉sectional and longitudinal research


This review of cross-╉sectional, biomarker, advanced imaging, and longitudinal studies across
youth and adult populations has highlighted several key themes in investigating emotion regula-
tion as a transdiagnostic process. Currently, as demonstrated by the studies reviewed in this chap-
ter, emotion regulation can be characterized by an array of different constructs (e.g., cognitive,
adaptive, maladaptive) and specific emotion regulation strategies (i.e., rumination, emotional
understanding, emotional expression). Further, each of these different constructs and strategies
can be defined and measured in different ways depending on the particular self-╉report assess-
ments used by each research team. While this allows for increased breadth in studying emotion
regulation, it also poses a challenge for the field in that it becomes difficult to compare and syn-
thesize results across studies and to conclusively highlight the most influential strategies. Despite
these limitations, within the research that currently exists, it appears that cognitive emotion regu-
lation strategies show promise as a transdiagnostic mechanism. Approximately 35% (n = 5) of
reviewed studies found evidence for some cognitive emotion regulation strategies (e.g., worry,
rumination) to operate transdiagnostically across disorders, 35% (n = 5) found evidence for emo-
tional response strategies (e.g., emotional clarity, emotional expression), and 21% (n = 4) found
evidence for transdiagnostic physiological reactivity mechanisms.
The methodological and design features reviewed within these studies highlight a number of
important considerations for the field to take into account in future research. Cross-╉sectional
research has been an efficient method to provide insights into which mechanisms merit additional
examination. As noted, the importance of these findings will likely increase as the field further
clarifies which strategies to examine and how to examine them. One positive step demonstrated in
this review is the use of biological and physiological markers of emotion regulation and their link
to self-╉report measures. Not only might this help demonstrate validity of self-╉report measures,
but adds an important objective component to examining emotion regulation strategies across
disorders. As a number of these studies show, research utilizing physiological and neurological
assessment offers a unique and cutting-╉edge approach to this area of research. This, in addition to
longitudinal design, can help explain the nuances of how particular emotion regulation strategies
and emotion regulation as a whole can influence the trajectory of different disorders over time.
Thus far, there is support for the impact of early emotion regulation difficulties in contributing to
the development of pathology rather than the other way around. These findings have potentially
important implications for the most effective and efficient ways to both prevent and treat different
disorders. Self-╉report was a primary assessment used in all studies (e.g., Ruminative Response
Scale). Approximately 64% used multimodal assessments, with 35% (n = 5) using clinical inter-
views (e.g., SCID-╉IV) and 28% (n = 4) using physiological assessments combined with self-╉report.

Clinical trials research


To evaluate the evidence for transdiagnostic mechanisms represented in treatment research, a
literature search was conducted using PsycINFO, PubMed and Google Scholar with the key-
words “transdiagnostic AND (treatment OR intervention) AND (emotion regulation OR emo-
tional regulation)”. Studies that involved interventions with a component of emotion regulation
4
3

434 Transdiagnostic Approaches to Emotion Regulation

and used samples with comorbidities or multiple disorders were included. The literature search
resulted in 14 articles, ten of which reported results from open trials or randomized controlled
trials (RCTs) and four reported small n case studies. Effect sizes were (i.e., Cohen’s d) calculated at
post-╉treatment and follow-╉up to evaluate the effects of transdiagnostic treatments on diagnostic
status, emotion-╉related measures such as positive and negative affect, and anxiety and depressive
symptoms (Table 4) A Cohen’s d value of 0.20 indicates a small effect size, 0.50 a medium effect
size and 0.80 or greater a large effect size.

Unified protocol
Twelve of the studies focused on the Unified Protocol (UP) for treating emotional disorders.
Originally developed by Barlow, Allen, and Choate (2004), the UP protocol integrates cognitive
behavior therapy (CBT) and emotion regulation principles (Wilamowska et al., 2010). It con-
sists of five core principles that include: (1) Increasing present-╉focused awareness of emotions,
(2) increasing cognitive flexibility, (3) identifying and preventing emotional avoidance and mal-
adaptive emotion-╉driven behaviors, (4) increasing awareness and tolerance of emotion-╉related
physiological sensations, (5) exposing to bodily and environmental triggers of emotional experi-
ences (Farchione et al., 2012). The aim of UP is to treat the symptoms of anxiety and mood dis-
orders through tolerance of emotions and modifying maladaptive emotion regulation strategies
(Wilamowska et al., 2010). A summary of these studies is presented in Table 20.4. The effects of
UP have been studied mostly as an individual treatment for adults with a principal anxiety or uni-
polar depressive disorder (above 18 years old; Bullis, Fortune, Farchione, & Barlow, 2014; Ellard,
Fairholme, Boisseau, Farchione, & Barlow, 2010; Farchione et al., 2012). There are also studies
using participants with chronic pain (Allen, Tsao, Seidman, Ehrenreich-╉May, & Zeltzer, 2012),
Bipolar Disorder, and Borderline Personality Disorder (Ellard, Deckersbach, Sylvia, Nierenberg,
& Barlow, 2012; Sauer-╉Zavala, Bentley, & Wilner, 2015). The UP has been applied to children
(seven to 12 years old; Bilek & Ehrenreich-╉May, 2012), adolescents (12–╉17 years old; Trosper,
Buzzella, Bennett, & Ehrenreich, 2009; Ehrenreich, Goldstein, Wright, & Barlow, 2009; Queen,
Barlow, & Ehrenreich-╉May, 2014) and adults (Bullis et al., 2015; Ornelas Maia, Braga, Nunes,
Nardi, & Silva, 2013) in group format.

Unified protocol as an individual intervention for adults


The UP was first pilot tested by Ellard, Fairholme, Boisseau, Farchione, and Barlow (2010).
This study examined the extent to which participants learned and used emotion regulation
skills during treatment. The protocol included 15 treatment sessions. The efficacy of the treat-
ment was tested through independent evaluator-╉rated and self-╉reported improvements in
participants’ anxiety, and depressive symptoms, and functional impairment. Although emo-
tion regulation was not directly measured, levels of participants’ positive and negative affect
were assessed, which were seen as a consequence of emotion regulation (Barlow et al., 2004).
Eighteen participants with multiple disorders completed an average of 13 treatment sessions.
Results showed significant improvement in clinical severity ratings (CSR) for their princi-
pal diagnosis, self-╉reported anxiety and depressive symptoms, negative affect and functional
impairment. Fifty-╉six percent of participants were classified as in the normal range of negative
affect at post-╉treatment.
The study defined a responder as ≥ 30% change on at least two outcomes of principal diagnosis
(i.e., CSR, functional impairment, diagnosis specific measure) and high end-╉state functioning as
a loss of the principal diagnosis (i.e., CSR ≤ 3) and within the normal range of at least one of the
functional impairment or diagnosis measures. Based on these criteria, 56% of the participants
5
3
4
Table 20.4  Clinical Trials Research

Citation Sample Characteristics and Assessments ER components (all Treatment Effect size (Cohen’s d)
Design that were assessed)
Pre-​posttreatment Pre-​follow-​up
Allen et al Case study: N = 2, 14–​17 years Pain interview, Children’s Somatization Lack of emotion UP for the Treatment N/​A N/​A
(2012) old, chronic pain with comorbid Inventory (CSI), Emotion Expression awareness, reluctance of Emotional
anxiety and depression. Scale for Children (EESC), Faces Pain to express negative Disorders in Youth
Assessment at posttreatment Scale-​Revised (FPS-​R), Functional emotion with Pain (UP-​YP)
and 3 month follow-​up Disability Inventory (FDI), RCADS

Bilek & Open trial, intent-​to-​treat ADIS-​IV-​C/​P, Screen for Child Anxiety N/​A Emotion Detectives Within-​group ESs: N/​A
Ehrenreich-​May (ITT): N = 22, 7–​12 years old Related Emotional Disorders-​Child and Treatment Protocol Principal diagnosis
(2012) (M = 9.79 years), anxiety Parent Reports (SCARED), Children’s severity = 1.38
disorder and comorbid Depression Inventory-​Child and Total severity of depression and
depressive disorder. Within Parent Reports (CDI), parent and child anxiety = 1.07
group comparison at satisfaction with treatment Parent-​report anxiety = 0.49
posttreatment and 3 month Self-​report anxiety = 0.47
follow-​up Parent-​report depression = 0.54
Self-​report depression
(ITT) = 0.34
Self-​report depression
(completers) = 0.65

Bullis et al. Open trial: N = 15, 20–​52 years ADIS-​IV-​L, BDI-​II, BAI, PANAS, PSWQ, Positive and negative Individual UP Within-​group ESs: Within-​group ESs:
(2014) old (M = 32.27 years), anxiety SIAS, WSAS, OCI-​R, Albany Panic and affect Principal diagnosis Principal diagnosis
disorders (M = 2.47 diagnoses; Phobia Questionnaire (APPQ) severity = 1.74 severity = 1.98
7 with comorbid depressive Self-​report anxiety = 1.72 Self-​report anxiety = 0.96
disorder). Within group Self-​report depression = 1.19 Self-​report
comparison of treatment Positive affect = 0.39 depression = 0.38
completers at posttreatment and Negative affect = 1.00 Positive affect = 0.41
18 month follow-​up Negative affect = 0.31

Bullis et al. Open trial: N = 11, 20–​69 years ADIS-​IV-​L, Overall Anxiety Severity Total experiential Group UP Within-​group ESs: N/​A
(2015) old (M = 44.55 years), anxiety and Impairment Scale (OASIS), Overall avoidance, Behavioral Self-​report anxiety = 1.36
disorders (M = 1.27 diagnoses, Depression Severity and Impairment avoidance, Self-​report depression = 0.66
n = 8 with comorbidity). Scale (ODSIS), Multidimensional distress aversion, Total experiential
Within group comparison at Experiential Avoidance Questionnaire procrastination, avoidance = 1.12
posttreatment (MEAQ), WSAS, Quality of Life distraction and
Enjoyment and Satisfaction suppression,
Questionnaire (Q-​LES-​Q) repression and denial,
distress endurance

(continued)
6
3
4
Table 20.4  Continued

Citation Sample Characteristics and Assessments ER components (all Treatment Effect size (Cohen’s d)
Design that were assessed)
Pre-​posttreatment Pre-​follow-​up

Ehrenreich, Open trial: N = 12, 12–​17 years ADIS-​IV-​C/​P, Revised Child Anxiety and Emotion regulation of UP-​Y None reported None reported
Buzzella et al. old, anxiety (42%) or comorbid Depression Scale (RCADS), Children’s sadness, anger and
(2009) anxiety and depressive Emotion Management Scales (CEMS) worry
disorder (58%). Assessment at
posttreatment, 3 and 6 month
follow-​up

Ehrenreich et al. Multiple baseline design: N = 3, Anxiety Disorders Interview Schedule N/​A UP-​Y N/​A N/​A
(2009) 12–​16 years old, principal for DSM-​IV Child and Parent Versions
diagnosis of anxiety or (ADIS-​IV-​C/​P)
mood disorder. 2–​8 week
baseline phase. Assessment at
posttreatment and 6 month
follow-​up

Ellard et al. Clinical replication series: N = 3, MINI, Hamilton Depression Rating N/​A Individual UP N/​A N/​A
(2012) 23–​62 years old, bipolar disorder Scale, Montgomery Asberg Depression
with comorbid anxiety disorder. Rating Scale (MADRS), Young Mania
Assessment at posttreatment Rating Scale (YMRS), Clinical Global
Impression Severity and Improvement
(CGI-​S, CGI-​I), BDI-​II, BAI

Ellard et al. Open trial #1: N = 18, 18–​ Clinician assessed Anxiety Disorders Positive and negative Individual UP Within-​group ESs: N/​A
(2010) 54 years old (M = 30 years), Interview Schedule for DSM-​IV Lifetime affect Principal diagnosis
anxiety or major depressive version (ADIS-​IV-​L), Beck Depression severity = 1.20
disorder (M = 1.94 diagnoses). Inventory-​II (BDI-​II), Beck Anxiety Self-​report anxiety = 0.60
Within group comparison at Inventory (BAI), Positive and Negative Self-​report depression = 0.50
posttreatment Affect Scale (PANAS), Obsessive-​ Positive affect = 0.30
Compulsive Inventory-​Revised (OCI-​R), Negative affect = 0.53
Panic Disorder Severity Scale (PDSS-​
SR), Penn State Worry Questionnaire
(PSWQ), Social Interaction Anxiety
Inventory (SIAS), Work and Social
Adjustment Scale (WSAS)
7
3
4
Open trial #2: N = 15, 18–​ ADIS-​IV-​L, clinician rated structured Individual UP Winthin-​group ESs: Within-​group ESs:
44 years old (M = 29.73 years), interview guides for the Hamilton Principal diagnosis Principal diagnosis
principal anxiety disorder Anxiety and Depression Rating Scales severity = 1.84 severity = 2.13
(M = 2.2 diagnoses). Within (SIGH-​A, SIGH-​D), BDI-​II, BAI, PANAS, Self-​report anxiety = 0.62 Self-​report anxiety = 0.64
group comparison at Yale-​Brown Obsessive Compulsive Self-​report depression = 0.43 Self-​report
posttreatment and 6 month Scale (Y-​BOCS), PDSS-​SR, PSWQ, SIAS, Positive affect = 0.53 depression = 0.65
follow-​up WSAS Negative affect = 0.75 Positive affect = 0.27
Negative affect = 0.78

Farchione et al. Randomized controlled ADIS-​IV-​L, SIGH-​A, SIGH-​D, BDI-​II, BAI, Positive and negative Individual UP Between-​group ESs: N/​A
(2012) trial: N = 37, anxiety disorders PANAS, Y-​BOCS, PDSS-​SR, PSWQ, affect Principal diagnosis
(M = 2.16 diagnoses; 12 with SIAS, WSAS severity = 2.27
comorbid depressive disorder) Self-​report anxiety = 0.43
Between group comparison of Self-​report depression = 0.87
treatment group (n = 26, 19–​ Positive affect = 0.87
52 years old, M = 29.38 years) Negative affect = 0.42
with 16 week waitlist control
group (n = 11, 19-​43 years
old, M = 30.64 years) at
posttreatment

N = 35, treatment initiators. Within-​group ESs: Within-​group ESs: Principal


Within group comparison at Principal diagnosis diagnosis severity = 2.12
posttreatment and 6 month severity = 1.55 Self-​report anxiety = 1.04
follow-​up Self-​report anxiety = 1.18 Self-​report
Self-​report depression = 1.05 depression = 0.94
Positive affect = 0.53 Positive affect = 0.57
Negative affect = 0.85 Negative affect = 0.84

(continued)
8
3
4
Table 20.4  Continued

Citation Sample Characteristics and Assessments ER components (all Treatment Effect size (Cohen’s d)
Design that were assessed)
Pre-​posttreatment Pre-​follow-​up

Mennin et al. Open trial, ITT: N = 21 ADIS-​IV-​L, modified CGI, PSWQ, Emotional intensity, Emotion Regulation Within-​group ESs: Within-​group ESs:
(2015) (M = 35.25 years), generalized BDI-​II, Mood and Anxiety Symptom decentering, Therapy (ERT) GAD severity = 3.60 GAD severity_​3m = 3.55
anxiety disorder (GAD) with Questionnaire-​Short Form (MASQ), nonacceptance of MDD severity = 0.54 GAD severity_​9m = 3.60
comorbid major depressive State-​Trait Anxiety Inventory (STAI), negative emotions, Self-​report worry = MDD severity_​3m = 0.65
disorder (MDD; n = 11). Sheehan Disability Scale (SDS), Quality difficulty in emotional Self-​report depression = 1.28 MDD severity_​9m = 0.80
Within group comparison at of Life Inventory (QOLI), Negative situations with Emotional intensity = 0.55 Self-​report worry_​3m = 1.48
posttreatment, 3 and 9 month Intensity scale from Affect Intensity pursuing goal-​ Decentering = 1.18 Self-​report worry_​9m = 1.77
follow-​up Measure (AIM), Decentering subscale directed behaviors, Emotion regulation = 0.71 Self-​report depression_​
from Experiences Questionnaire, DERS, controlling impulses, Reappraisal = 1.26 3m = 1.22
Emotion Regulation Questionnaire lack of regulation Trait mindfulness = 0.62 Self-​report depression_​
(ERQ), Five Facet Mindfulness strategies, problems 9m = 1.41
Questionnaire (FFMQ) with emotional Emotional intensity_​
awareness, emotional 3m = 0.25
clarity, reappraisal, Emotional intensity_​
trait mindfulness 9m = 0.64
Decentering_​3m = 1.01
Decentering_​9m = 1.35
Emotion regulation_​
3m = 0.72
Emotion regulation_​
9m = 0.99
Reappraisal_​3m = 1.35
Reappraisal_​9m = 1.06
Trait mindfulness_​3m = 0.32
Trait mindfulness_​9m = 0.76

Neacsiu et al. Randomized controlled DERS, DBT Skills subscale of DBT Ways Nonacceptance of Dialectical Behavior Between-​group Ess: Between-​group ESs:
(2014) trial, ITT: N = 44, high on of Coping Checklist (DBT-​WCCL), negative emotions, Therapy-​ Skills Self-​report anxiety = 0.86 Self-​report anxiety = 0.70
emotion disorder, at least Patient Health Questionnaire-​9 difficulty in emotional Training group Self-​report depression = 0.39 Self-​report
one anxiety or depressive (PHQ-​9), OASIS, Brief Treatment situations with (DBT-​ST) Emotion regulation = 0.99 depression = 0.39
disorder (DBT-​ST: M = 2.68; History Interview (B-​THI), Addiction pursuing goal-​ DBT skills = 0.75 Emotion regulation = 0.63
ASG: M = 2.59). Between Severity Index Self-​Report Form (ASI-​ directed behaviors, DBT skills = 0.27
group comparison of treatment SR), Credibility and Expectancy of controlling impulses,
group (M = 32.27 years old) Improvement Scales (CEIS) lack of regulation
with 16 sessions, 120 min, strategies, problems
Activities-​Based Support Group with emotional
(ASG, M = 38.82 years old) at awareness, emotional
posttreatment and 2 month clarity
follow-​up
9
3
4
Ornelas et al. Open trial: N = 16, 18-​58 years Mini International Psychiatric Nil Group UP Within-​group ESs: N/​A
(2013) old (M = 35.63 years), unipolar Interview (MINI 5.0), BDI, BAI, World Self-​report anxiety = 0.95
mood disorder comorbid Health Organisation Quality of Life Self-​report depression = 1.34
with anxiety. Within group (WHOQOL-​BREF), ARIZONA scale of
comparison at posttreatment sexual function

Queen et al. Piecewise latent growth curve ADIS-​IV-​C/​P, RCADS, Revised Child Nil Individual UP-​Y Within-​group ESs: Within-​group ESs:
(2014) modelling: N = 59, 12-​17 years Anxiety and Depression Scale—​Parent Self-​report anxiety = 0.81 Self-​report anxiety_​
old (M = 15.42 years), ≥ 8 Version Self-​report depression = 0.65 3m = 1.29
sessions in previous open trial or Parent-​report anxiety = 0.48 Self-​report anxiety_​
RCT, anxiety (79.9%), unipolar Parent-​report depression = 0.63 6m = 1.34
depressive (15.3%) disorder or Self-​report depression_​
both (5.1%). Test trajectories 3m = 0.97
of anxiety and depressive Self-​report depression_​
symptoms over the course of 6m = 0.80
the UP-​Y (data from an open Parent-​report anxiety_​
trial or an RCT were analysed). 3m = 1.11
Within group comparison at Parent-​report anxiety_​
posttreatment, 3 and 6 month 6m = 1.07
follow-​up Parent-​report depression_​
3m = 0.87
Parent-​report depression_​
6m = 0.60

Sauer-​Zavala Clinical replication series: N = 5, Diagnostic Interview for Personality Nonacceptance of Individual UP N/​A N/​A
et al. (2015) borderline personality disorder Disorders-​4th Edition (DIPD-​IV), negative emotions,
with comorbid anxiety and ADIS-​IV, The Zanarini Rating Scale difficulty in emotional
mood disorders. Assessment at for Borderline Personality Disorder situations with
posttreatment (ZAN-​BPD), Depression, Anxiety and pursuing goal-​
Stress Scales, Difficulties in Emotion directed behaviors,
Regulation Scale (DERS) controlling impulses,
lack of regulation
strategies, problems
with emotional
awareness, emotional
clarity
0
4

440 Transdiagnostic Approaches to Emotion Regulation

were identified as treatment responders based on their principal diagnosis and 71% based on their
comorbid diagnoses. Thirty-​three percent of participants gained high end-​state functioning based
on their principal diagnosis and 50% based on their comorbid diagnosis.
However, the authors considered the effects as “modest,” given that 67% of the sample remained
at the clinical level at post-​treatment. To enhance the UP, the protocol was extended to 18 ses-
sions and this was further tested in 15 participants with a principal diagnosis of anxiety disor-
der in a second pilot trial (Ellard et al., 2010). Participants had at least two comorbid anxiety or
depressive disorders and they attended an average of 17 sessions of the treatment. There were
significant improvements in clinical severity ratings of principal diagnosis, self-​reported anxiety
symptoms, negative affect and functional impairment. However, self-​reported depressive symp-
toms and positive affect did not show significant changes. The percentage of responders increased
to 71% compared to the first pilot trial (56%) based on their principal diagnosis and 64% based on
their comorbid diagnoses. Similarly, based on both principal and comorbid diagnoses, high end-​
state functioning increased from 32% and 50% in the first pilot trial to 60% and 64%, respectively.
A higher proportion of participants were classified as within the normal range of negative affect
(67%). At six-​month follow-​up, only clinical severity ratings of principal diagnosis, negative affect
and functional impairment showed significant improvements. Eleven participants (73%) showed
further gains in responder status and high end-​state functioning based on their principal diagnosis
while only 50% showed improvements based on their comorbid diagnosis. However, the results
from these two trials were preliminary, given the small sample size and absence of a control group.
Subsequently, Farchione et al. (2012) conducted an RCT to investigate the efficacy of the UP and
its effects at six-​month follow-​up. The study involved 37 participants with a principal diagnosis of
anxiety disorders and at least two comorbid diagnoses. Twelve of these participants had comor-
bid depressive disorders. The UP used in this study was different from Ellard et al. (2010), which
included motivational techniques to assess for readiness for change and engagement in treatment,
additional optional emotion-​focused exposure exercises, and treatment review and relapse pre-
vention (see details in Barlow et al., 2011). Participants received a maximum of 18 sessions of
treatment. The study used the same assessment measures as Ellard et al. (2010). Compared to the
waitlist control, participants in the UP group demonstrated significantly greater improvement in
self-​reported and independent evaluator-​rated anxiety and depressive symptoms, positive and
negative affect, and functional impairment. Between groups, effect sizes ranged from 0.42 to 2.27
(Table 20.4). Responders were defined as either ≥ 30% change or ≤ 3 in clinical severity rating
of principal diagnosis plus ≥ 30% change in at least diagnosis-​specific or functional impairment
measures. The definition of high end state functioning was the same as Ellard et al. (2010). None
of the participants in the waitlist control were identified as responders or at high end state func-
tioning. In contrast, 59% participants in the UP group were identified as responders and 50%
of them at high end-​state functioning. The study also analyzed results of the treatment initiator
sample which included waitlist control participants who completed treatment after the waitlist
period (n = 35; average of 15.26 sessions completed). Within group effect sizes of clinical severity
ratings of principal diagnosis, anxiety and depressive symptoms, positive and negative affect and
functional impairment ranged from moderate to large (Table 20.4). Forty five percent of the treat-
ment initiators no longer met criteria of any clinical diagnoses at post-​treatment and 64% at six
month follow-​up. Similarly, the proportion of the participants who were identified as responders
(59%) and at high end state functioning (52%) at post-​treatment based on principal diagnosis
increased to 71% and 64% at follow-​up. Regarding comorbid diagnoses, there was an increase in
the percentage of participants who were identified as responders and at high end state functioning
at follow-​up compared to post-​treatment, from 38% to 62%, and 41% to 72%, respectively. It was
highlighted that 67% of participants with comorbid depressive disorders no longer met criteria
1
4

Summary of cross-sectional and longitudinal research 441

for any clinical diagnoses, and were identified as responders and at high end state functioning at
post-╉treatment and this increased to 89% at follow-╉up. These results demonstrated the efficacy of
UP on anxiety and depressive disorders, with further symptom improvements at six month fol-
low-╉up. The authors suggested that the effects of the UP on positive and negative affect were likely
achieved through improving individuals’ reactions toward negative emotions and their engage-
ments in positive emotional experiences. However, there were limitations including small sample
size and the lack of reliability and fidelity checks. Furthermore, given that the comparison group
was not an active treatment, conclusions on the processes of the UP in relation to therapeutic
effects cannot be drawn.
A follow-╉up study by Bullis, Fortune, Farchione, and Barlow (2014) aimed to explore the out-
comes of participants in the Farchione et al. (2012) study at 18 months after treatment (i.e., long-╉
term follow-╉up; n = 15). The results showed a significant improvement in clinical severity ratings
of principal diagnosis, number of clinical diagnoses, and clinician-╉rated and self-╉report functional
impairment at long-╉term follow-╉up in 15 treatment completers with anxiety disorders. Although
53% of participants did not meet criteria for any clinical diagnosis at long term follow-╉up, there
was no evidence of further improvements from six months to 18-╉month follow-╉up. Participants
who were identified as responders and at high end state functioning at six months maintained
their functioning at 18-╉month follow-╉up. While there was an increase in participants’ depressive
symptoms, negative affect and clinician-╉rated functional impairment from 6 months to long-╉term
follow-╉up, the average scores in these domains remained at the normal to mild range. The results
suggest that gains in treatment were maintained up to 18 month follow-╉up. However, the small
sample made it impossible to generalize the results to other diagnoses and populations.
To sum up, the three studies reviewed above revealed that UP focuses primarily on increasing
emotional awareness and changing maladaptive emotion regulation strategies. Although emotion
regulation was not directly measured in these studies, there is evidence that targeting emotion
regulation in adults could lead to improvements in anxiety and depressive symptoms, and nega-
tive or positive affect concurrently.

Unified protocol as a group intervention for adults


While the UP has been delivered as an individual therapy, Ornelas Maia, Braga, Nunes, Nardi, and
Silva (2013) aimed to evaluate a group treatment based on UP for 16 participants with moderate
to severe unipolar mood disorder with comorbid anxiety. The treatment covered the five core
principles of UP over 12 two-╉hour sessions. The MINI International Neuropsychiatric Interview
was used to identify the diagnoses and self-╉reported anxiety and depressive symptoms were mea-
sured. There were significant improvements in self-╉reported anxiety and depressive symptoms at
post-╉treatment with large within-╉group effect sizes. This study provided evidence for the feasi-
bility of UP as a group intervention for adults, which allows the opportunity for social learning
among participants. However, similar to previous studies, this study was an open trial with a small
sample size and did not measure treatment adherence and emotion regulation.
Similarly, the researchers who developed UP also pilot tested the UP as a group format in an
open trial with 11 participants to examine whether results from studies on UP as an individual
intervention could be replicated (Bullis et al., 2015). The majority of the participants had a prin-
cipal anxiety diagnosis with only one with a dysthymia diagnosis. However, eight of the 11 par-
ticipants had a comorbid disorder including other anxiety and mood disorders, attention deficit
hyperactivity disorder, and alcohol abuse. The UP was delivered in a group of five to six partici-
pants over 12 two-╉hour sessions. Participants in the group treatment completed an average of ten
treatment sessions and were assessed at pre-╉, mid-╉and post-╉treatment. Although there was no
clear description about measurements for emotion regulation, a multi-╉dimensional experiential
2
4

442 Transdiagnostic Approaches to Emotion Regulation

avoidance questionnaire (MEAQ:  Gamez, Chmielewski, Kotov, Ruggero, & Watson, 2011)  was
administered to assess the tendency to avoid negative internal experiences. This measure consisted
of six subscales including behavioral avoidance, distress aversion, procrastination, distraction and
suppression, repression and denial, and distress endurance. In addition, anxiety, depression, func-
tional impairment, and satisfaction and enjoyment in daily living were also measured.
Results showed that the UP demonstrated a strong effect on anxiety (d  =  1.36) and experi-
ential avoidance (d = 1.12) and a moderate effect on depressive symptoms (d = 0.66). UP was
mostly rated as “very” or “extremely” acceptable and participants were satisfied. The group format
provided opportunities for participants to increase their confidence in practicing skills, through
involvement in other participants’ exposure and small group exercises. Although the efficacy of
UP was demonstrated in a group intervention, the authors noted difficulties in training and moni-
toring the understanding of treatment concepts in the group setting, especially with participants
who require extensive direction. Moreover, the small sample size, reliance on self-╉report meas-
ures, and lack of control group and follow-╉up failed to ascertain the causal inferences or modera-
tors of treatment efficacy. Therefore, it was suggested to improve the treatment by devoting more
time to homework review, limiting to one objective per session, including brief individual meet-
ings with participants, and using the group intervention as a step towards intensive treatment.
In sum, the group format of the UP protocol shows promising results. Specifically, Bullis et al.
(2015) measured multiple dimensions of experiential avoidance in a study of adults with multiple
disorders. Group UP showed large effects on six domains of emotion regulation as well as anxi-
ety and depressive symptoms. Future research should include control conditions and long-╉term
follow-╉up to consolidate the efficacy of UP on emotion regulation as a group intervention for
adults with emotional disorders.

Unified protocol for adolescents


Concurrent to the development and testing of the UP with adults, the UP was also adapted for
adolescents aged 12–╉17 years old with anxiety or unipolar depressive disorders. In addition to
the modifications of developmentally appropriate language and examples in the 13 session treat-
ment, reviews with the parent at the end of every session and dedicated parent sessions in session
one and six were included. Ehrenreich, Goldstein, Wright and Barlow (2009) used a multiple
baseline design of two to eight weeks to pilot test the effects of the UP for adolescents on clinical
severity ratings of anxiety and mood symptoms in multiple diagnoses for three individual cases.
There were sustained improvements across disorders from pre-╉treatment to six month follow-╉up.
Participants found the emotion exposures to be the most helpful component of the treatment
and parents were receptive to being included in treatment. However, due to the small sample size
(n = 3) and the primary focus on anxiety disorders, the effect of UP on emotional disorder symp-
toms in a heterogeneous sample of adolescents needs to be further investigated.
To enhance participants’ motivation, Ehrenreich and colleagues further modified the treatment
protocol by including goal-╉oriented discussions and decisional balance, and extended the num-
ber of sessions to 16 with three additional parent sessions that focused greatly on individual-╉
specific emotion exposure and regulation skills (Ehrenreich, Goldstein, Wright, & Barlow, 2009).
The preliminary feasibility and effects of this UP-╉Youth (UP-╉Y) protocol were investigated in an
open trial reported in Trosper et al.(2009). Participants included 12 adolescents with a principal
anxiety disorder or a comorbid anxiety and depressive disorder. Both the adolescent and their
parent reported on changes in anxiety and depressive symptoms and emotion regulation skills for
anger, sadness and worry at post-╉treatment, three and six month follow-╉up. There was a signifi-
cant improvement in adolescent clinical severity ratings and self-╉reported anxiety and depressive
symptoms at post treatment and both follow-╉up time points. Adolescents reported a significant
3
4

Summary of cross-sectional and longitudinal research 443

improvement in worry, sadness, and overall emotion dysregulation, as well as coping with anger at
post-╉treatment. At three month follow-╉up, there was significant improvements in overall emotion
coping and dysregulation of sadness. However, there was generally no significant difference in
emotion dysregulation and coping between three and six months. These results showed that UP-╉
Y is effective in improving emotion regulation and coping with emotions in adolescents. Similar
to Ehrenreich et al. (2009), the sample was small (n = 12) and did not include participants with a
principal depressive disorder. Moreover, emotion regulation was not specifically measured.
Although anxiety and depression share similar vulnerabilities, studies have demonstrated that
anxiety and depression are also distinct from each other (Anderson & Hope, 2008). As such, to
compare the trajectory of the changes in self and parent reported anxiety and depressive symp-
toms in adolescents up to six months following UP-╉Y treatment, Queen, Barlow, and Ehrenreich-╉
May (2014) analyzed the results of 59 adolescents who completed at least eight sessions of UP-╉Y
from a total sample of 67 participants in either an open trial or an RCT. Emotion regulation was
not examined in the analysis. Participants were 12–╉17 year old adolescents with an anxiety disor-
der (79.9%), unipolar depressive disorder (15.3%) or co-╉principal anxiety and depressive disor-
ders (5.1%). Twenty-╉three participants (38.98%) were assigned a secondary comorbid depressive
disorder. The UP was flexibly conducted between eight to 21 sessions with optional parenting
skills training, motivational interviewing and safety planning depending on the needs of the indi-
vidual participant. By the end of treatment, mean scores of self and parent-╉reported anxiety and
depressive symptoms fell into the normal range. Symptom trajectories showed that self-╉reported
anxiety symptoms decreased significantly by 4.76 units every eight weeks during treatment and
1.48 units every eight weeks during three to six month follow-╉up period. The rate of change of
self-╉reported anxiety symptoms during treatment was significantly associated with the rate of
change during the follow-╉up period. Although self-╉reported depressive symptoms significantly
decreased during treatment at a similar rate of change as self-╉reported anxiety symptoms, there
were no significant reductions at follow-╉up. The rate of change of self-╉reported depressive symp-
toms during treatment was not significantly related with the rate of change during the follow-╉up
period. Parent-╉reported anxiety and depressive symptoms also demonstrated significant improve-
ment during treatment. However, similar to the adolescents’ self-╉reported depressive symptoms,
both parent-╉reported anxiety and depressive symptoms revealed no significant reduction from
post treatment to three to six month follow-╉up. Participants and parents who reported greater
severity in both anxiety and depressive symptoms at baseline demonstrated a greater rate of
improvement of symptoms during treatment but only self-╉reported anxiety symptoms showed a
reduction in rate of improvement at follow-╉up. The preliminary evidence from this study needs to
be further replicated with a larger sample with a diverse population in order to obtain trajectories
of improvement in symptoms.

Unified protocol for children


Subsequently, UP was also developed into a group-╉based intervention for younger children
aged seven to 12 years old. Bilek, and Ehrenreich-╉May (2012) aimed to examine the effect of
Emotion Detectives Treatment Protocol on self and parent-╉reported symptoms of the principal
anxiety disorder and the anxiety and depressive symptom severity ratings, given that research
had shown that co-╉morbid depressive symptoms was associated with poorer treatment out-
comes in youths with anxiety disorders. Twenty-╉two children with a principal anxiety disorder,
of which seven had comorbid depressive disorder, participated in an open trial of the 15 ses-
sion Emotion Detectives Treatment Protocol. Although half of the participants were identi-
fied to have “elevated depressive symptoms” based on self or parent report, participants with a
principal depressive disorder were excluded. At each session, parents underwent group-╉based
4

444 Transdiagnostic Approaches to Emotion Regulation

parent training after individual reviews with their child at the beginning of the session. It was
hypothesized that participants would show improvement in severity of their principal diagnosis
regardless of the severity of their depressive symptoms at pre-╉treatment. Seventy-╉three percent
of the sample was identified as treatment completers who completed 11 or more sessions. At
post-╉treatment, 77.8% and 80% of the participants no longer met criteria for an anxiety and a
depressive disorder, respectively. There was a significant improvement in clinical severity rat-
ings of both principal diagnosis and the total scores of the clinical severity ratings of anxi-
ety and depressive disorders, self-╉reported anxiety symptoms and parent-╉reported depressive
symptoms. Effect sizes ranged from 0.47 to 1.38. A  significant improvement in self-╉reported
depressive symptoms was only seen in the treatment completers. Both self-╉and parent-╉reported
depressive symptoms at baseline did not significantly predict change in clinical severity ratings
of principal diagnosis at post-╉treatment. Satisfaction of the child (M = 5.5 on an eight point
Likert Scale) and the parent with the treatment were high (M = 7.7). The Emotion Detectives
Treatment Protocol was shown to be efficacious and feasible in children with a diagnosis of a
principal anxiety disorder who also had comorbid depressive symptoms. However, analysis was
not conducted to determine effects on those with or without comorbid depression. The sample
was ethnically diverse, but the study was limited by the small sample size, lack of control group,
and lack of control for Type I errors. The small sample also did not allow for group and therapist
effects to be analyzed. More studies are required to test the efficacy of the Emotion Detectives
Treatment Protocol with other emotional disorders.
In sum, there is only limited research of UP with child and adolescent samples. In the three stud-
ies reviewed above, emotion regulation was not directly assessed despite the fact that UP retained
its focus on emotions and dysregulation of emotions and coping. Nevertheless, the improvements
in both anxiety and depressive symptoms based on self-╉and parent-╉reports provided beneficial,
albeit only preliminary, evidence for the UP as a transdiagnostic approach for emotion regulation
across emotion disorders.

Unified protocol for other principal disorders


Ellard, Deckersbach, Sylvia, Nierenberg, and Barlow (2012) explored the effects of the UP for the
commonly comorbid bipolar disorders with anxiety disorders by evaluating a series of 3 clinical
cases. Two of the cases who received 15 sessions of UP were rated as mildly ill and very much
improved by clinician-╉rated severity and improvement measures at post-╉treatment. Self-╉reported
symptoms of depression, mania and anxiety were in the normal to borderline range. Emotion
regulation was not measured. Participants identified that the helpful components of the treatment
were present-╉focused awareness skills and learning how to alter maladaptive responses. However,
as the authors pointed out, given the nature of the study design, it was impossible to determine
the long-╉term effects of the treatment. In addition, given that all three participants requested con-
tinued care to consolidate their learned skills, it was difficult to determine the optimum number
of treatment sessions for this population. Therefore, more longitudinal and RCTs are needed to
further investigate the short-╉and long-╉term efficacy, potential moderators, and mediators of UP
for comorbid bipolar and anxiety disorders.
Similar to bipolar disorders, anxiety and depressive disorders are also highly comorbid with
chronic pain in children. Taking into account the role of emotion dysregulation in exacerbating
responses to and severity of chronic pain, Allen, Tsao, Seidman, Ehrenreich-╉May, and Zeltzer
(2012) developed the UP in Youth with Pain (UP-╉YP) to address the lack of research that targets
comorbid emotional disorders with pain in adolescents with chronic pain. The UP-╉YP is an adap-
tation of the United Protocol for the Treatment of Emotional Disorders in Youth (Ehrenreich et al.,
2008), with inclusion of psychoeducation for pain and discomfort, and awareness of emotions and
5
4

Summary of cross-sectional and longitudinal research 445

exposure exercises. Although similar in length (i.e., eight to 21 sessions) as the UP-​Y, the 50-​
minute sessions were conducted within six months and were flexible to be fortnightly delivered
sessions at the end of the treatment. Results of the treatment protocol were presented in two cases
with emotion awareness and expression of negative emotions measured. In one case, there was a
decline in functional impairment, anxiety symptoms, somatization, and an increase in emotional
awareness and expression while the level of pain remained stable. There were further gains at three
month follow-​up and pain was at its lowest level. However, in the other case, no change was seen
in depressive symptoms and there were increases in somatization, pain and functional impair-
ment despite an increase in emotion regulation scores. Although results suggest that the emphasis
of increasing awareness of both physical and emotional symptoms in the UP-​YP may contribute
to its effectiveness, future study is warranted to consolidate the findings.
Given that neuroticism (i.e., the tendency to experience negative emotions uncontrollably)
has been proposed as one common underlying trait in borderline personality disorder (BPD),
anxiety and depressive disorders, Sauer-​Zavala, Bentley, and Wilner (2015) used the UP as a
relatively brief treatment to target neuroticism in participants with less severe symptoms of BPD.
Using the same UP protocol as Farchione et al. (2012), this study evaluated symptom levels and
emotion regulation skills in five BPD patients with comorbid anxiety and depressive disorder.
Effect sizes were reported as standardized mean gain (ESsg). Results showed the improvements in
BPD symptoms (ESsg = 1.06) and emotion regulation skills (ESsg = 1.29) were large in magnitude
while moderate in anxiety (ESsg =.51) and depressive symptoms (ESsg =.70). Although the UP
showed promise in treating BPD with comorbid emotional disorders, the heterogeneity in BPD
made it difficult for the protocol to be appropriate for everyone. Therefore, the authors suggested
that more studies with a larger sample size and including other personality traits are required to
further investigate the moderators and efficacy of the UP for managing BPD. In sum, given that
these studies have included measures related to emotion regulation, they provided further sup-
port for the potential of UP to improve symptoms and skills in emotion regulation in multiple
disorders.
Taken together, our review reveals that the UP has been used to treat a variety of emotion-​
related disorders, with a majority focusing on principal anxiety and depressive disorders and their
comorbidities. Overall, the UP protocol demonstrated promising effects on anxiety and depres-
sive symptoms in adults, children and adolescents, either in an individual or a group treatment
format. These results suggest the potential of the UP as a transdiagnostic approach that targets
emotion regulation across emotion disorders, which in turn leads to improvements in anxiety
and depressive symptoms, and negative or positive affect in these disorders. However, the reli-
ability of the results of these studies on the UP is limited by small sample size and few comparison
conditions. It should also be noted that in some studies, especially those conducted in children
and adolescents, the majority of the participants had a principle diagnosis of an anxiety disorder
but not a depressive disorder. The exclusion of individuals with Major Depressive Disorder (e.g.,
Bilek & Ehrenreich-​May, 2012) and a relatively small number of participants with comorbid anxi-
ety and depression symptoms require future research to include participants with a diversity of
diagnoses. This will allow a more solid conclusion for the effects of the UP as a transdiagnostic
approach. In addition, only three of the studies directly measured changes in emotion regula-
tion in their samples (Allen et al., 2012; Ehrenreich et al., 2009; Sauer-​Zavala et al., 2015). Future
research including a controlled design with larger samples, assessments of emotion regulation,
and participants with different types of emotional disorders are warranted. Furthermore, longer
duration of follow-​up, as well as more detailed analysis of outcomes, and investigating potential
moderators and mediators would increase the validity of UP as a transdiagnostic treatment tar-
geting emotion regulation.
6
4

446 Transdiagnostic Approaches to Emotion Regulation

Using dialectical behavior therapy skills training


Another intervention that was developed specifically to improve emotional regulation across
multiple disorders is Dialectical Behavior Therapy Skills Training (DBT-╉ ST). Dialectical
Behavior Therapy was originally developed by Linehan (1993) for suicide and borderline per-
sonality disorder (BPD). It proposed that dysfunctional behaviors are due to poorly regulated
emotions or are maladaptive approaches to emotion regulation. Neacsiu, Eberle, Kramer,
Wiesmann, and Linehan (2014) modified this approach to teach skills that help individuals
to be less susceptible to emotions, manage situations that trigger emotions, control attention
toward or away from emotional stimuli, interpret emotional cues, manage biological, experien-
tial, and action changes, and process emotions. The four skill modules included in the DBT-╉ST
are mindfulness (e.g., non-╉judgmental observation of the present moment), emotion regula-
tion (e.g., strategies to change emotions or the tendency to respond emotionally), interpersonal
effectiveness (e.g., assertiveness and self-╉respect), and distress tolerance (e.g., control impulses
and accept difficulties).
Neacsiu et al. (2014) examined the effects of the DBT-╉ST in reducing anxiety, depression and
emotion dysregulation in a sample of non-╉BPD adults with difficulties with emotion regulation.
Forty-╉four adults with anxiety or depressive disorders who reported high emotion dysregula-
tion participated in this study. This RCT compared the 16-╉session DBT-╉ST in a two-╉hour weekly
group therapy to a 16-╉session activities-╉based support group (ASG) of the same duration. All
participants attended an average of 13.66 sessions and were assessed at pre-╉treatment, two months
into treatment, post-╉treatment, and two-╉month follow-╉up on self-╉reported anxiety and depressive
symptoms, use of coping skills, and specific emotion regulation skills.
Results showed that both groups experienced significant decreases in emotion dysregulation over
time but the DBT-╉ST group demonstrated significantly greater and faster decrease compared to the
ASG at post-╉treatment (d = 1.86). However, at two month follow-╉up, there was a trend towards loss
of gains in DBT-╉ST group while the ASG group showed continued improvement. A similar pattern
at follow-╉up was observed for the change in anxiety and depressive symptoms. There was a signifi-
cantly faster improvement in anxiety symptoms in participants in the DBT-╉ST group compared to
the ASG group at the end of treatment (d = 1.37). On the other hand, both groups showed significant
and equal magnitude of improvement in depressive symptoms at the end of treatment (d = 0.73).
The use of DBT skills mediated the relationship between groups and improvement in 62.31% of
variance in emotion regulation, 47.63% of variance in anxiety and 42.5% of variance in depressive
symptoms. Participants in the DBT-╉ST group significantly increased their use of regulation strate-
gies and goal-╉directed behaviors (16%) compared to those in the ASG (3.5%), which was maintained
at follow-╉up. Moreover, they were more likely to attribute symptom improvement to the interven-
tion (M = 53.33%) than the ASG group (M = 26.88%). Participants in the DBT-╉ST group were also
more confident to recommend it to a friend (M = 6.58) than the ASG group (M = 4.06).
Overall, DBT-╉ST was superior to ASG in improving anxiety, emotion dysregulation and use of
DBT skills. However, with no difference in the improvement in depressive symptoms, DBT-╉ST as
a transdiagnostic intervention needs further investigation. The study was limited by the use of dif-
ferent therapists, small sample and more participants with dysthymia in the ASG group. Further
research is required to improve dropout rates, noncompliance to protocol, to explore the loss of
gains in the DBT-╉ST at follow-╉up, and to replicate these initial findings.

Emotion Regulation Therapy


Mennin, Fresco, Ritter, and Heimberg (2015) argued that dysfunction in distress disorders such
as generalized anxiety disorder (GAD) can be understood through three mechanisms. These are
7
4

Summary and conclusion 447

motivational mechanisms (i.e., the functional and directional properties of an emotional response
tendency), regulatory mechanisms (i.e., controlling emotional responses using a variety of elabora-
tive systems); and contextual learning consequences (i.e., promoting broad and flexible behavioral
strategies). Emotion Regulation Therapy, which integrates principles of CBT with experiential
therapy, targets deficits in these mechanisms that maintain GAD. Given that GAD with comorbid
major depressive disorder (MDD) is associated with severe functional impairments compared to
GAD or MDD alone, Mennin et al. (2015) conducted the first study to test the efficacy of Emotion
Regulation Therapy for GAD with comorbid depressive symptoms. Emotion Regulation Therapy
focuses on increasing awareness of properties of emotion response tendencies, developing less
and more elaborate emotion regulation strategies and exposure to different contexts, which allow
participants to develop a variety of flexible behaviors. Participants were 21 adults with principal
diagnosis of GAD, among which 11 had comorbid MDD. The treatment protocol consisted of 20
weekly sessions with each session lasting for 60 minutes, except for Sessions 11–╉16, which lasted
90 minutes for participants to practice exposure exercises. In the first half of the treatment pro-
gram, participants were taught emotion regulation strategies to respond “counteractively” rather
than “reactively”. In the second half of the treatment, they were encouraged to become more “pro-
active” in using regulation skills through in-╉session and out-╉session exposure exercises.
Results revealed significant improvements in clinical severity ratings of GAD and MDD, self-╉
reported worry and depressive symptoms, and quality of life at post-╉treatment, three and nine
months follow-╉up. This study also assessed for emotional intensity, decentering, emotion reg-
ulation skills, reappraisal and trait mindfulness associated with the model of emotion regula-
tion therapy, all of which improved significantly at post-╉treatment and nine months follow-╉up. At
three months follow-╉up, only changes in emotion regulation skills, decentering and reappraisal
were significant. High end-╉state functioning was defined as falling in the normal range on four
of six measures of GAD and three of four MDD measures. At post-╉treatment, 66.7% and 45.5%
achieved high end-╉state functioning on GAD and MDD, respectively with the proportion increas-
ing to 75% in GAD and 70% in MDD at three months follow-╉up. At nine months follow-╉up, 85%
of the sample were at high end-╉state functioning on GAD and 80% on MDD. This study provided
preliminary evidence for the efficacy of Emotion Regulation Therapy for GAD participants with
significant depressive symptoms. In addition to improvement in severity of anxiety and depres-
sive symptoms, Emotion Regulation Therapy showed large effects on different aspects of emotion
regulation. Despite these findings, ERT requires refinement to differentiate the intervention from
other interventions with similar components. Future research should include larger RCTs and
objective assessments of symptoms to determine the mechanisms targeted. Notably, this interven-
tion has not yet been used with children and adolescents.

Summary and conclusion


Evidence that establishes emotion regulatory processes as important transdiagnostic etiological,
maintenance, and treatment mechanisms is promising. Evidence from cross-╉sectional and longi-
tudinal designs provide the most support for cognitive emotion regulation strategies (e.g., rumi-
nation, reappraisal, worry, catastrophizing) as critical mechanisms across disorders, with more
recent support for emotional response strategies (e.g., emotional clarity, emotional expression).
Neuro-╉biological research has also identified important roles for amygdala functioning across
disorders. In the treatment literature, several interventions that target emotional functioning as
part of their treatment strategies demonstrate benefits for participants in primary and secondary
presenting problems. However, most trials do not specifically report outcomes for specific emo-
tion regulation processes, other than positive and negative affect. Emerging research on DBT and
8
4

448 Transdiagnostic Approaches to Emotion Regulation

ERT appear to provide good examples of treatment evaluations that actively assess and report
outcomes in emotion regulation—╉each has shown positive short-╉and long-╉term benefits of treat-
ment on improving individuals’ ability to regulate distress. This appears consistent across multiple
presenting problems.
Our review highlights the nascent stage of investigating emotion regulation processes across
disorders and problem ╉sets and with children and adolescents. Research designs, assessment
tools, and analytic approaches differ considerably across studies, making comparisons difficult. It
is recommended that future emotion regulation research make concerted efforts to recruit multi-╉
problem samples, as comparisons can aid a more refined understanding of specific and common
regulation deficits and strengths. This review has identified several model approaches, but it has
also revealed multiple areas for future growth. Greater use of prospective designs, utilizing multi-╉
domain and multi-╉source assessment, would improve experimental, longitudinal, and treatment
research.

References
Achenbach, T. (2005). Advancing assessment of children and adolescents: Commentary on evidence-╉based
assessment of child and adolescent disorders. Journal of Clinical Child and Adolescent Psychology, 34,
541–╉547. doi: 10.1207/╉s15374424jccp3403_╉9
Aldao, A., & Nolen-╉Hoeksema, S. (2010). Specificity of cognitive emotion regulation
strategies: A transdiagnostic examination. Behaviour Research and Therapy, 48(10), 974–╉983.
doi:10.1016/╉j.brat.2010.06.0 02
Aldao, A., & Nolen-╉Hoeksema, S. (2012). When are adaptive strategies most predictive of
psychopathology? Journal of Abnormal Psychology, 121(1), 276–╉281. http://╉dx.doi.org/╉10.1037/╉
a0023598
Allen, L. B., Tsao, J. C. I., Seidman, L. C., Ehrenreich-╉May, J., & Zeltzer, L. K. (2012). A unified,
transdiagnostic treatment for adolescents with chronic pain and comorbid anxiety and depression.
Cognitive and Behavioral Practice, 19, 56–╉67.
American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders, 5th ed.
(DSM-╉5). Arlington, VA: American Psychiatric Association. doi: 10.1176/╉appi.books.9780890425596
Anderson, E.R., & Hope, D.A. (2008). A review of the tripartite model for understanding the link
between anxiety and depression in youth. Clinical Psychology Review, 28, 275–╉287. doi: 10.1016/╉
j.cpr.2007.05.004
Angold, A., Costello, E. J., & Erkanli, A. (1999). Comorbidity. Journal of Child Psychology and Psychiatry,
40, 57–╉87. Retrieved from http://╉eds.b.ebscohost.com/╉
Ball, T. M., Ramsawh, H. J., Campbell-╉Sills, L., Paulus, M. P., & Stein, M. B. (2013). Prefrontal dysfunction
during emotion regulation in generalized anxiety and panic disorders. Psychological Medicine, 43(07),
1475–╉1486. doi:10.1017/╉S0033291712002383
Barlow, D., Allen, L., & Choate, M. (2004). Toward a unified treatment for emotional disorders. Behavior
Therapy, 35, 205–╉230. doi: 10.1016/╉S0005-╉7894(04)80036-╉4
Barlow, D. H., Ellard, K. K., Fairholme, C. P., Farchione, T. J., Boisseau, C. L., Allen, L. B., & Ehrenreich-╉
May, J. (2011). The unified protocol for transdiagnostic treatment of emotional disorders: Client workbook.
New York: Oxford University Press.
Beauchaine, T. P., Gatzke-╉Kopp, L. M., & Mead, H. K. (2007). Polyvagal theory and developmental
psychopathology: Emotion dysregulation and conduct problems from preschool to adolescence.
Biological Psychology, 74, 174–╉184. doi: 10.1016/╉j.biopsycho.2005.08.008
Bilek, E. L., & Ehrenreich-╉May, J. (2012). An open trial investigation of a transdiagnostic group treatment
for children with anxiety and depressive symptoms. Behavior Therapy, 43, 887–╉897.
Brockmeyer, T., Bents, H., Holtforth, M. G., Pfeiffer, N., Herzog, W., & Friederich, H. C. (2012). Specific
emotion regulation impairments in major depression and anorexia nervosa. Psychiatry Research,
200(2), 550–╉553. doi: 10.1016/╉j.psychres.2012.07.009
9
4

Summary and conclusion 449

Bullis, J. R., Fortune, M. R., Farchione, T. J., & Barlow, D. H. (2014). A preliminary investigation of the
long-​term outcome of the unified protocol for transdiagnostic treatment of emotional disorders.
Comprehensive Psychiatry, 55, 1920–​1927. http://​dx.doi.org/​10.1016/​j.comppsych.2014.07.016
Bullis, J. R., Sauer-​Zavala, S., Bentley, K. H., Thompson-​Hollands, J., Carl, J. R., & Barlow, D. H. (2015).
The unified protocol for transdiagnostic treatment of emotional disorders: Preliminary exploration of
effectiveness for group delivery. Behavior Modification, 39(2), 295–​321. doi: 10.1177/​0145445514553094
Burklund, L. J., Craske, M. G., Taylor, S. E., & Lieberman, M. D. (2015). Altered emotion regulation
capacity in social phobia as a function of comorbidity. Social Cognitive and Affective Neuroscience,
10(2), 199–​208. doi:10.1093/​scan/​nsu058
Campbell-​Sills, L., & Barlow, D. H. (2007). Incorporating emotion regulation into conceptualizations and
treatments of anxiety and mood disorders. In J. J. Gross (Ed.), Handbook of emotion regulation (pp.
542–​559). New York: Guilford Press.
Carver, C. S. (1997). You want to measure coping but your protocol’s too long: Consider the Brief COPE.
International Journal of Behavioral Medicine, 4, 92–​100.
Chorpita, B. F., & Daleiden, E. L. (2009). Mapping evidence-​based treatments for children and adolescents:
application of the distillation and matching model to 615 treatments from 322 randomized trials.
Journal of Consulting and Clinical Psychology, 77(3), 566–​579. doi: 10.1037/​a0014565
Conklin, L. R., Cassiello-​Robbins, C., Brake, C. A., Sauer-​Zavala, S., Farchione, T. J., Ciraulo, D. A., &
Barlow, D. H. (2015). Relationships among adaptive and maladaptive emotion regulation strategies
and psychopathology during the treatment of comorbid anxiety and alcohol use disorders. Behaviour
Research and Therapy, 73, 124–​130. doi: 10.1016/​j.brat.2015.08.001
Crowell, S. E., Beauchaine, T. P., McCauley, E., Smith, C. J., Stevens, A. L., & Sylvers, P. (2005).
Psychological, autonomic, and serotonergic correlates of parasuicide among adolescent girls.
Development and Psychopathology, 17(4), 1105–​1127. doi: 10.1017/​S0954579405050522
Desrosiers, A., Vine, V., Klemanski, D. H., & Nolen‐Hoeksema, S. (2013). Mindfulness and emotion
regulation in depression and anxiety: common and distinct mechanisms of action. Depression and
Anxiety, 30(7), 654–​661. doi: 10.1002/​da.22124
Ehrenreich, J. T., Buzzella, B. A., Trosper, S. E., Bennett, S. M., Wright, L. R., & Barlow, D. H. (2008). The
Unified Protocol for treatment of emotional disorders in adolescents. Unpublished manuscript, Boston
University.
Ehrenreich, J. T., Goldstein, C. R., Wright, L. R., & Barlow, D. H. (2009). Development of a unified
protocol for the treatment of emotional disorders in youth. Child & Family Behavior Therapy, 31(1),
20–​37. doi: 10.1080/​07317100802701228
Ehrenreich-​May, J. & Chu, B. C. (2013). Overview of transdiagnostic mechanisms and treatments for youth
psychopathology. In J. Ehrenreich-​May & B. C. Chu (Eds.), Transdiagnostic treatments for children and
adolescents: Principles and practice (pp. 3–​14). New York: Guilford Press.
Ellard, K.K., Deckersbach, T., Sylvia, L.G., Nierenberg, A.A., & Barlow, D.H. (2012). Transdiagnostic
treatment of bipolar disorder and comorbid anxiety with the unified protocol: A clinical replication
series. Behavior Modification, 36(4), 482–​508. doi: 10.1177/​0145445512451272
Ellard, K. K., Fairholme, C. P., Boisseau, C. L., Farchione, T. J., & Barlow, D. H. (2010). Unified protocol
for the transdiagnostic treatment of emotional disorders: Protocol development and initial outcome
data. Cognitive and Behavioral Practice, 17, 88–​101.
Fairburn, C. G., Cooper, Z., & Shafran, R. (2003). Cognitive behaviour therapy for eating
disorders: A “transdiagnostic” theory and treatment. Behaviour Research and Therapy, 41, 509–​528.
doi: 10.1016/​S0005-​7967(02)00088-​8
Farchione, T. J., Fairholme, C. P., Ellard, K. K., Bossieau, C. L., Thompson-​Hollands, J., Carl,
J. R., … Barlow, D. H. (2012). Unified protocol for transdiagnostic treatment of emotional
disorders: A randomized controlled trial. Behavior Therapy, 43, 666–​678.
Gamez, W., Chmielewski, M., Kotov, R., Ruggero, C., & Watson, D. (2011). Development of a measure
of experiential avoidance: The Multidimensional Experiential Avoidance Questionnaire. Psychological
Assessment, 23, 692–​713. doi: 10.1037/​a0023242
0
5
4

450 Transdiagnostic Approaches to Emotion Regulation

Garber, J., & Weersing, V. R. (2010). Comorbidity of anxiety and depression in youth: Implications
for treatment and prevention. Clinical Psychology: Science and Practice, 17, 293–​306. doi: 10.1111/​
j.1468-​2850.2010.01221.x
Garnefski, N., Kraaij, V., & van Etten, M. (2005). Specificity of relations between adolescents’ cognitive
emotion regulation strategies and Internalizing and Externalizing psychopathology. Journal of
Adolescence, 28(5), 619–​631. doi: 10.1016/​j.adolescence.2004.12.009
Gross, J. J. (1998). The emerging field of emotion regulation: an integrative review. Review of General
Psychology, 2(3), 271.
Gruber, J., Eidelman, P., & Harvey, A. G. (2008). Transdiagnostic emotion regulation processes in
bipolar disorder and insomnia. Behaviour Research and Therapy, 46(9), 1096–​1100. doi:10.1016/​
j.brat.2008.05.004
Gruber, J., Mennin, D. S., Fields, A., Purcell, A., Murray, G. (2015). Heart rate variability as a potential
indicator of positive valence system disturbance: A proof of concept investigation. International Journal
of Psychophysiology, 98(2), 240–​248. doi: 10.1016/​j.ijpsycho.2015.08.005
Harvey, A. G., Watkins, E., Mansell, W., & Shafran, R. (2004). Cognitive behavioural processes across
psychological disorders: A transdiagnostic approach to research and treatment. Oxford, UK: Oxford
University Press.
Insel, T. (2013, April 29). Transforming diagnosis. National Institute of Mental Health Director’s Blog.
Retrieved from http://​www.nimh.nih.gov/​about/​director/​2013/​transforming-​diagnosis.shtml
Kring, A. M., & Sloan, D. M. (Eds.). (2010). Emotion regulation and psychopathology: A transdiagnostic
approach to etiology and treatment. Guilford Press.
Linehan, M. M. (1993). Cognitive behavioral treatment of borderline personality disorder. New York,
NY: Guilford Press.
Mansell, W., Harvey, A., Watkins, E., & Shafran, R. (2009). Conceptual foundations of the transdiagnostic
approach to CBT. Journal of Cognitive Psychotherapy: An International Quarterly, 23, 6–​19. doi: 10.1891/​
0889-​8391.23.1.6
McLaughlin, K. A., Hatzenbueler, M. L., Mennin, D. S., & Nolen-​Hoeksema, S. (2011). Emotion
dysregulation and adolescent psychopathology: A prospective study. Behaviour Research and Therapy,
49(9), 544–​554. doi: 10.1016/​j.brat.2011.06.003
Mennin, D. S., Fresco, D. M., Ritter, M., & Heimberg, R. G. (2015). An open trial of emotion regulation
therapy for generalized anxiety disorder and cooccurring depression. Depression and Anxiety, 32, 614–​
623. doi: 10.1002/​da.22377
National Institute for Mental Health’s (2011, June). NIMH Research Domain Criteria (RDoC). Retrieved
from http://​www.nimh.nih.gov/​research-​priorities/​rdoc/​nimh-​research-​domain-​criteria-​rdoc.shtml
Neacsiu, A. D., Eberle, J. W., Kramer, R., Wiesmann, T., Linehan, M. M. (2014). Dialectical behavior
therapy skills for transdiagnostic emotion dysregulation: A pilot randomized controlled trial. Behaviour
Research and Therapy, 59, 40–​51. http:/​dx.doi.org/​10.1016/​j.brat.2014.05.005
Nolen-​Hoeksema, S., & Watkins, E. R. (2011). A heuristic for developing transdiagnostic models of
psychopathology explaining multifinality and divergent trajectories. Perspectives on Psychological
Science, 6, 589–​609. doi: 10.1177/​1745691611419672
Ornelas Maia, A. C., Braga, A. A., Nunes, C. A., Nardi, A. E., & Silva, A. C. (2013). Transdiagnostic
treatment using a unified protocol: Application for patients with a range of comorbid mood and anxiety
disorders. Trends in Psychiatry and Psychotherapy, 35(2), 134–​140.
Pang, K. C., Beauchaine, T. P. (2013). Longitudinal patterns of autonomic nervous system responding
to emotion evocation among children with conduct problems and/​or depression. Developmental
Psychobiology, 55(7), 698–​706. doi: 10.1002/​dev.21065
Queen, A. H., Barlow, D. H., & Ehrenreich-​May, J. (2014). The trajectories of adolescent anxiety and
depressive symptoms over the course of a transdiagnostic treatment. Journal of Anxiety Disorders, 28,
511–​521. http://​dx.doi.org/​10.1016/​j.janxdis.2014.05.007
1
5
4

Summary and conclusion 451

Queen, A. H., & Ehrenreich-​May, J. (2014). Anxiety-​disordered adolescents with and without a comorbid
depressive disorder: Variations in clinical presentation and emotion regulation. Journal of Emotional
and Behavioral Disorders, 22(3), 160–​170. doi: 10.1177/​1063426613478175
Sauer-​Zavala, S., Bentley, K. H., & Wilner, J. G. (2015). Transdiagnostic treatment of borderline personality
disorder and comorbid disorders: A clinical replication series. Journal of Personality Disorders, 29,
179–​195.
Shannon, K. E., Beauchaine, T. P., Brenner, S. L., Neuhaus, E., & Gatzke-​Kopp, L. (2007). Familial
and temperamental predictors of resilience in children at risk for conduct disorder and depression.
Development and Psychopathology, 19(3), 701–​727. doi: 10.1017/​S0954579407000351
Silk, J. S., Steinberg, L., & Morris, A. S. (2003). Adolescents’ emotion regulation in daily life: Links to
depressive symptoms and problem behavior. Child Development, 74(6), 1869–​1880. doi: 10.1046/​
j.1467-​8624.2003.00643
Taylor, S., & Clark, D. A. (2009). Transdiagnostic cognitive-​behavioral treatments for mood and anxiety
disorders: Introduction to the special issue. Journal of Cognitive Psychotherapy: An International
Quarterly, 23, 3–​5. doi: 10.1891/​0889-​8391.23.1.3
Trosper, S. E., Buzzella, B. A., Bennett, S. M., & Ehrenreich, J. T. (2009). Emotion regulation in youth
with emotional disorders: Implications for a unified treatment approach. Clinical Child and Family
Psychology Review, 12, 234–​254. doi: 10.1007/​s10567-​009-​0043-​6
Werner, K., & Gross, J. J. (2010). Emotion regulation and psychopathology: A conceptual framework.
In Kring, A. M., & Sloan, D. M. (Eds.), Emotion regulation and psychopathology: A transdiagnostic
approach to etiology and treatment (pp. 13–​37). New York, NY, US: Guilford Press.
Wilamowska, Z. A., Thompson-​Hollands, J., Fairholme, C. P., Ellard, K. K., Farchione, T. J., & Barlow, D.
H. (2010). Conceptual background, development and preliminary data from the unified protocol for
transdiagnostic treatment of emotional disorders. Depression and Anxiety, 27, 882–​890. doi: 10.1002/​
da.20735
Vasilev, C. A., Crowell, S. E., Beauchaine, T. P., Mead, H. K., & Gatzke-​Kopp, L. M.
(2009). Correspondence between physiological and self-​report measures of emotion
dysregulation: A longitudinal investigation of youth with and without psychopathology. Journal of Child
Psychology and Psychiatry, 50(11), 1357–​1364. doi:10.1111/​j.1469-​7610.2009.02172.x
Vine, V., & Aldao, A. (2014). Impaired emotional clarity and psychopathology: A transdiagnostic deficit
with symptom-​specific pathways through emotion regulation. Journal of Social and Clinical Psychology,
33(4), 319–​342. doi:10.1521/​jscp.2014.33.4.319
2
5
4
Index

A incarceration, parental  351


abuse and neglect  305–​20 pharmacotherapy 293
assessment 307 proactive 
definition  305–​6 conduct disorder  135–​6
ecological model  307 development  137–​8
emotion regulation  311–​12 reactive 
deficits, adaptive outcomes of  317–​18 abused/​neglected children and adolescents  314
deficits, expression of  314–​15 conduct disorder  135–​6
intervention, implications for  318–​19 development  136–​7, 138
problems, development of  312–​14 reactive vs. proactive  24
propagation and maintenance of outcomes  315–​17 self-​regulation deficits  21–​2, 24
nonsuicidal self-​injury  400 Fast Track PATHS  27
outcomes  310–​11 socialization 63
emotion regulation deficits  317–​18 traumatic stress exposure  376, 381
propagation and maintenance of  315–​17 agoraphobia (AG)  156
prevalence 306 alcohol myopia therapy  217
risk factors  307–​10 alcohol use/​misuse 
academic achievement  abused/​neglected children and adolescents  316
divorce, children of  333–​4, 335, 336, 337, 340 abusive parents  309
incarceration, parental  351, 355, 360 affective lability  220
optimism 341 comorbidity 
traumatic stress exposure  376, 378 anxiety disorders  424
Acceptance and Commitment Therapy  posttraumatic stress disorder  219–​20
nonsuicidal self-​injury  408 diagnosis 213
Research Domain Criteria  98 distress tolerance  219
adaptive emotion responses  43 dual-​process theory  221
adaptive information processing (AIP) model  386 emotion dysregulation  217–​19
adulthood, transition from adolescence to  emotion regulation strategies  424
abuse and neglect  311, 316 emotional clarity  426
Attention Deficit Hyperactivity Disorder  120 emotional inertia  217
depression 174 impulsivity 25
divorce, children of  334, 335 intervention example  223
substance use disorders  213 mindfulness 220, 221
Adverse Child Experiences (ACEs) study  353, 354, 366–​7 prevalence 211, 212
affective disorders  see anxiety disorders; depression alexithymia 197
affective lability  215 autism spectrum disorder  240
abused/​neglected children and adolescents  314, 316 nonsuicidal self-​injury  400
borderline personality disorder  267 alpha-​adrenergic receptor-​mediated effects  47
substance use disorders  216–​17, 218, 220, 224 alternative reinforcers  220, 224
Affective Lability Scales (ALS)  220 American Psychiatric Association (APA) 
affective Posner paradigm  286, 287 history of mental health/​disorder  79
age factors  see also Diagnostic and Statistical Manual of Mental
abuse and neglect  307–​8 Disorders
conduct disorder  amygdala 
course 131 abused/​neglected children and adolescents  313
onset 130, 131 anxiety disorders  428–​9, 430
prevalence 131 Attention Deficit Hyperactivity Disorder  118, 288
divorce, children of  338 autism spectrum disorder  238, 239
eating disorders  196 borderline personality disorder  266
incarceration, parental  360, 362, 366–​7 depression  428–​9, 430
substance use disorders  211, 212, 213 labeling irritable emotion  295
aggression  major depressive disorder  289
abused/​neglected children and adolescents  314, 315 measures of emotional reactivity  48, 50, 51, 52, 53
Attention Deficit Hyperactivity Disorder  119, narrow phenotype bipolar disorder  288, 289
120, 121 reappraisal 
autism spectrum disorder  243 detached vs. positive  53
borderline personality disorder  268 vs. distraction  52
conduct disorder  135–​8 vs. suppression  51
co-​occurring externalizing and internalizing problems  33 severe mood dysregulation  287, 288–​9
cultural factors  63 social phobia  428–​9
impulsivity 26 traumatic stress exposure  378
4
5

454 Index

anger  for adults  434, 440–​2


abused/​neglected children and adolescents  314, 315 for children  443–​4
autism spectrum disorder  242, 244 neuroticism 445
borderline personality disorder  267 aripiprazole 242, 293
conduct disorder  134–​5 arousal levels 
cultural factors  67, 68–​9, 70 Attention Deficit Hyperactivity Disorder  116
detection by abused/​neglected children and cultural factors  68
adolescents 313, 318 atomoxetine 122
disruptive mood dysregulation disorder  283, 294 Attachment Family Drawing task  357
incarceration, parental  360, 361–​2, 364 attachment styles  4
inhibition/​maladaptive expression  10 borderline personality disorder  271
longitudinal research  430, 432 emotion regulation  312
rumination 402 incarceration, parental  356–​8, 362, 366
severe mood dysregulation  289, 290, 294 insecure 4
Stop Signal Reaction Time Task (SSRT) outcomes  116 borderline personality disorder  268
Anger Control Training  292 incarceration, parental  351
anorexia nervosa (AN)  196–​207 psychopathology 10
case example  204–​6 mentalization-​based treatment for adolescents  271
central facets of emotion regulation  425 attention 
emotion regulation  196–​8, 200 control of 133
treatments  198–​203 co-​occurring externalizing and internalizing
Children’s Hospital at Westmead program  201–​6 problems 32
evidence-​based  198–​200 internalizing problems  28, 31
antisocial behavior  reappraisal vs. distraction  52
abused/​neglected children and adolescents  316 depression  175–​6, 426
incarceration, parental  351, 354, 355, 366 nonsuicidal self-​injury  403
anxiety disorders  154–​65 processes 117
abused/​neglected children and adolescents  310, 311, 315 reactive aggression  136–​7
behavioral factors  159 severe mood dysregulation  285–​7
cognitive behavioral therapy  162–​4 traumatic stress exposure  384
cognitive factors  158 Attention Deficit Hyperactivity Disorder (ADHD)  113–​23
comorbidity  clinical implications  121
alcohol use disorder  424 comorbidity  115–​16
anorexia nervosa  197, 200 severe mood dysregulation  294
Attention Deficit Hyperactivity Disorder  115 development  119–​20
autism spectrum disorder  241, 242, 243, 244 diagnostic criteria  113
bipolar disorder  444 differential diagnosis  282
borderline personality disorder  445 emotion dysregulation  118–​19
depression  426–​7, 430, 440–​4 executive dysfunctions  116–​18
nonsuicidal self-​injury  398 gender differences  114
oppositional defiant disorder  282 impulsivity 25
pain, chronic  444–​5 interventions 293
social phobia  428 neural mechanisms  117, 118, 286, 288
substance use disorders  213–​14 pharmacotherapy 293
cultural factors  70 prevalence 114
diagnoses  154–​6 psychosocial impairment  114–​15
Dialectical Behavior Therapy Skills Training  446 temperamental pathways to  21
divorce, children of  335, 336, 340 transition from adolescence to adulthood  120
effects  10–​11 treatment  121–​2
emotion regulation strategies  421–​5 autism spectrum disorder (ASD)  235–​53
Emotion Regulation Therapy  446–​7 cognitive and behavioral influences  240–​1
emotional clarity  426 emotion regulation  237–​44
etiology  156–​7 interventions  241–​53
exposure therapy  97 irritability 293
functional magnetic resonance imaging  428–​30 parent management training  292
longitudinal research  430, 432 pharmacotherapy 293
parental reactions  160–​1 prevalence 236
physiological factors  159–​60 Secret Agency Society-​Operation Regulation
prevalence 9, 10 program  244–​53
rumination 402 autobiographical memory (AM)  176–​7, 181
self-​regulation deficits  21, 22 autonomic nervous system (ANS) 
internalizing problems  27, 28, 30, 31 Attention Deficit Hyperactivity Disorder  117–​18
severe mood dysregulation as predictor  282 sympathetic branch (SNS) 
transdiagnostic mechanisms  420, 425 activation measures  46–​7
traumatic stress exposure  378, 379, 381, 383, 387 anxiety disorders  159, 160
Unified Protocol  434–​45 traumatic stress exposure  380
for adolescents  442–​3 traumatic stress exposure  379
54

Index 455

avoidance strategies  Research Domain Criteria  98


abused/​neglected children and adolescents  317 Unified Protocol  434, 445
anxiety disorders  158, 159, 160–​1 Bradley-​Lang view of emotion  81–​3, 87–​8
borderline personality disorder  269 Research Domain Criteria  90, 102
nonsuicidal self-​injury  401 brain development  7–​8, 259
substance use disorders  214 brain regions 
traumatic stress exposure  379, 381, 382, 384 effortful control processes  20
Avoidant/​Restrictive Food Intake Disorder extended process model  87
(ARFID) 196, 197 fear conditioning  90–​1
emotional clarity  426 process model  85, 90–​1
treatments 199 reactive control processes  20
reappraisal 91
B see also neural mechanisms
battered child syndrome  305 bulimia nervosa  196, 197, 206
behavioral activation (BA)  182–​3, 184–​5 treatments  198, 199, 206–​7
behavioral inhibition 
anxiety disorders  157, 161 C
Attention Deficit Hyperactivity Disorder  115, callous unemotional (CU) traits 
116, 117 comorbid Attention Deficit Hyperactivity
internalizing problems  29–​30, 31 Disorder  115–​16
behavioral therapy  121–​2 conduct disorder  131–​2, 134, 135–​6, 140
beta-​adrenergic receptor-​mediated effects  47 externalizing 25
between-​subjects research  45 incarceration, parental  360
binge eating disorder  196, 197, 206 Cambridge Study in Delinquent Development  355, 364
emotional clarity  426 catastrophizing 427
treatments  199, 206–​7 chain analysis  407
bioinformational model of emotion  82–​3 challenge appraisal  47
biological and physiological aspects of emotion Child Behaviour Checklist  119–​20
regulation  7–​8, 43–​54 Children of Divorce Coping with Divorce  344
adaptive emotion responses  43 Children of Divorce Intervention Program (CODIP)  342
definition of emotion regulation  43 Children’s Emotion Management Scales  361, 362
neural activation associations  48, 50 cigarette use 
outcome measures  45–​8 affective lability  220
process measures  48–​51 hedonic capacity  220
reappraisal  impulsivity 25
detached vs. positive  53–​4 mindfulness 220
vs. distraction  52–​3 prevalence 211
vs. suppression  51–​2 clinical trials research, transdiagnostic mechanisms
research methods  44–​5 in  433–​4, 435–​9
biomarker studies  428–​30 coaching, emotion  359
biosocial theory  261–​2 cognitive analytic therapy (CAT)  270–​1
biphasic model of emotion  82 cognitive behavior therapy (CBT) 
bipolar disorder (BD)  abused/​neglected children and adolescents  316, 318
abused/​neglected children and adolescents  310 alcohol dependence  316
comorbid anxiety disorders  444 anorexia nervosa  199
heart rate variability  428, 430 anxiety disorders  162–​4
narrow phenotype  see narrow phenotype bipolar Attention Deficit Hyperactivity Disorder  122, 294
disorder autism spectrum disorder  242, 243, 244
rumination, worry, and negative automatic Secret Agent Society-​Operation Regulation
thoughts 425 program 245, 250
Unified Protocol  434, 444 borderline personality disorder  270
body posture  46 depression  174, 181–​5
borderline personality disorder (BPD)  259–​73 major depressive disorder  98
attachment-​based theory  264 disruptive mood dysregulation disorder  291, 292
comorbid nonsuicidal self-​injury  398, 400, 406, Feeling Thermometer  294
407, 408 nonsuicidal self-​injury  405–​6, 409
developmental theories  261–​2 severe mood dysregulation  294
diagnosis  259–​61 substance use disorders  223
dialectical behavior therapy  406, 407, 446 traumatic stress exposure  384–​5, 387–​8
emotion dysregulation  261–​4 Unified Protocol  434
emotional clarity  426 Cognitive Behavioral Intervention for Trauma in Schools
empirical evidence  264–​70 (CBITS) 385
interventions  270–​3 Cognitive Emotion Regulation Questionnaire
neuroticism 445 (CERQ) 268
parasympathetic nervous system influence on the cognitive functioning  355
heart 49 cognitive reappraisal  see reappraisal
prevalence 261 collectivism  64, 65, 66, 68, 71
6
5
4

456 Index

community-​based programs for children of divorce  343 delayed gratification  6–​7


community risk factors, abuse and neglect  310 Attention Deficit Hyperactivity Disorder  120
comorbidity, and transdiagnostic approaches  420 development 8
conduct disorder (CD)  129–​42 externalizing problems  23, 24
aggression  135–​6 deliberate self-​harm  see nonsuicidal self-​injury
proactive  137–​8 delinquency 26
reactive  136–​7 depression  171–​85
antisocial behavior trajectories  130 abused/​neglected children and adolescents  310, 311,
with callous unemotional traits  131–​2, 134, 135–​6, 140 315, 316
classification 129 adverse child experiences  354
comorbidity  comorbidity 
Attention Deficit Hyperactivity Disorder  115 anorexia nervosa  197
depression  432–​3 anxiety disorders  426–​7, 430, 440–​4
co-​occurring externalizing and internalizing Attention Deficit Hyperactivity Disorder  115, 116
problems 33 autism spectrum disorder  241, 243, 244
development of emotion regulation  132–​3 borderline personality disorder  445
diagnosis 129, 130 depression  432–​3
impulsivity 25 disruptive mood dysregulation disorder  282–​3
longitudinal research  432–​3 generalized anxiety disorder  447
parasympathetic nervous system effects on the nonsuicidal self-​injury  398
heart 49 oppositional defiant disorder  282
parent management training  291 pain, chronic  444–​5
prevalence and course  130–​2 social phobia  428
with severe anger dysregulation  134–​5 substance use disorders  213–​14
temperament and emotionality  133 cultural factors  70–​1
temperamental pathways to  21 Dialectical Behavior Therapy Skills Training  446
treatment  138–​9 divorce, children of  335, 336, 340, 341
functional family therapy  140–​1 emotion regulation 
Incredible Years  141–​2 central facets  425
Multisystemic Therapy  139–​40 difficulties  174–​81
conduct problems (CP)  130 strategies  421–​5
with callous unemotional traits  132, 134 Emotion Regulation Therapy  447
conscience, development of  138 emotional clarity  426
context-​sensitive regulation  284–​5 genetic factors  10
Cool Kids program  162 Research Domain Criteria  92, 97
Cool Little Kids program  162, 163 heart rate variability  428, 430
Coping Cat program  162, 164 incarceration, parental  351
Coping Power Program  138 interventions 
Coping with Adolescent Stress program  182 cognitive behavioral therapy  181–​5
Coping with Children’s Negative Emotions Scales  363 evidence-​based  181–​3
cortisol response to stress  379, 380, 381 longitudinal research  430, 432–​3
measures  47–​8 nature of  171–​2
court-​connected programs for children of divorce  342 prevalence  9, 172–​3
covert externalizing  24 psychological theories of  174
cross-​cultural psychology  61, 62, 64, 65–​7, 71 and reappraisal  6
cross-​sectional research  433–​47 relapse rates  10
with adults  421–​6, 427 Research Domain Criteria  92, 97, 98
biomarker and fMRI studies  428–​30 risk factors  173–​4
with youth  426–​8 rumination 402
cultural factors  60–​71 self-​regulation deficits  21, 22
abuse and neglect  310 co-​occurring externalizing and internalizing
Attention Deficit Hyperactivity Disorder  114 problems 32, 33
cross-​cultural research  61, 62, 64, 65–​7, 71 internalizing problems  27, 28, 29, 30, 31
in emotion regulation  8 severe mood dysregulation as predictor  282
externalizing problems  26 and suppression  6
incarceration, parental  363, 364–​5 transdiagnostic mechanisms  420, 425
internalizing problems  30–​1 traumatic stress exposure  378, 379, 381, 382
models of self and emotion regulation  64–​5 interventions  385–​6, 387
Research Domain Criteria  91 Unified Protocol  434–​45
scripts  67–​9, 70 for adolescents  442–​3
self-​control  33 for adults  434, 440–​1, 442
socialization  61–​4, 67, 71 for children  443–​4
Cyberball task  265–​6 neuroticism 445
developmental psychopathology model  262
D Diagnostic and Statistical Manual of Mental
decision making  177–​9 Disorders (DSM) 
default mode network  289 anxiety disorders  154–​5, 156
7
5
4

Index 457

Attention Deficit Hyperactivity Disorder  113, 114 protective factors for adjustment to parental
autism spectrum disorder  235 separation 341
borderline personality disorder  259–​61 risk factors for adjustment to parental separation  341
conduct disorder  129, 130, 131 dopamine receptors  97
depression 171 drug abuse  see substance use disorders
disruptive mood dysregulation disorder  281, 282 dual-​process theory  221
history of mental health/​disorder  80
nonsuicidal self-​injury  398 E
posttraumatic stress disorder  379, 389 Early Adolescent Temperament Scale-​Revised  360
Research Domain Criteria  90, 91, 102 eating disorders  196–​207
substance use disorders  211, 212 abused/​neglected children and adolescents  311
transdiagnostic approaches  420 case example  204–​6
traumatic stress exposure  374–​5, 379, 389 central facets of emotion regulation  425
Dialectical Behavior Therapy (DBT)  446 emotion regulation  196–​8, 200, 421–​4
abused/​neglected children and adolescents  318 emotional clarity  426
borderline personality disorder  270, 406, 407 longitudinal research  430, 432
Research Domain Criteria  98 treatments  198–​203
eating disorders  200, 206–​7 Children’s Hospital at Westmead program  201–​3
nonsuicidal self-​injury  406–​7, 408, 409 evidence-​based  198–​200
Dialectical Behavior Therapy Skills Training Ecological Momentary Assessment (EMA)
(DBT-​ST)  446 methodology 267
Difficulties in Emotion Regulation Scale (DERS)  educational attainment  see academic achievement
borderline personality disorder  267–​8, 269 effortful control  19–​20, 34
substance use disorders  217, 219 biological measures  48–​51
transdiagnostic approaches  425, 426, 432 co-​occurring externalizing and internalizing
dimensional models  420–​1 problems 32
disabilities 307, 308 externalizing problems  23–​4, 26
disappointment display rule paradigm  362–​3 internalizing problems  27, 28, 29, 30, 31
disruptive mood dysregulation disorder (DMDD)  281–​96 ego depletion effects  48
diagnosis 281 ego-​resiliency  31
differential diagnosis  282–​3 electromyography 46
emotion (dys)regulation  283–​91 emotion 3
interventions  291–​5 as action preparation  81–​3
labeling irritable emotion  295 bioinformational model  82–​3
parental contingencies for behavior, consistency in  295 biphasic model  82
prevalence 282 in childhood and adolescence  4
severe mood dysregulation  281–​2 definitions  3, 263, 378
distraction  functions  3–​4, 132, 133
nonsuicidal self-​injury  408, 409 perspectives  81–​7
vs. reappraisal  52–​3 three-​systems view  83, 84
substance use disorders  217 Emotion Acceptance Behavior Therapy (EABT)  200
traumatic stress exposure  381 emotion awareness  238, 249
distress management  emotion coaching  359
anorexia nervosa  201–​2, 203 Emotion Detectives Treatment Protocol 
case example  204 anxiety disorders  163
Dialectical Behavior Therapy Skills Training  446 Unified Protocol  443–​4
emotion regulation strategies  424 emotion dysregulation  see emotion regulation
nonsuicidal self-​injury  401 Emotion-​Focused Cognitive-​Behavioral Therapy
distress thermometers  202 (ECBT)  162, 164, 165
distress tolerance  215 emotion regulation (ER)  5–​6
borderline personality disorder  267, 268 as action change  81–​3
substance use disorders  216, 218, 219, 224 benefits 79
Distress Tolerance Scale (DTS)  219 in children and adolescents  9–​11
divalproex  definitions  5, 43, 62, 237, 263–​4, 359, 378
Attention Deficit Hyperactivity Disorder  293 development  8–​9, 132–​3
disruptive mood dysregulation disorder  293 and executive function  7
severe mood dysregulation  293 extended process model (EPM)  85–​8
divorce, children of  331–​46 Research Domain Criteria  98, 99–​101
effects 331 function  6–​7
emotion regulation  337 group therapy  408, 409
for at-​risk children  337–​8 influencing factors  7–​8
as protective adjustment  340 process model  see process model of emotion regulation
impact of parental separation  332–​6 research methods  43–​5
influential factors  338–​40 strategies 237
interventions  341–​4 cross-​sectional research  421, 427
mediating factors  336–​7 as transdiagnostic mechanism  433
prevalence  331–​2 Emotion Regulation Checklist  360, 363
8
5
4

458 Index

Emotion Regulation Therapy  446–​7 emotion regulation strategies  427


Emotion Regulation Training (ERT)  270 incarceration, parental  360, 361–​2
emotional abuse  parasympathetic nervous system influence on the
definition 306 heart 49
outcomes 311 traumatic stress exposure  376, 378, 382
prevalence 306 extinction learning  97–​8
risk factors  308 exuberance 
see also abuse and neglect externalizing problems  26
emotional arousal  internalizing problems  29
abused/​neglected children and adolescents  312, 314 eye movement desensitization and reprocessing
incarceration, parental  358, 359 (EMDR)  386–​7
nonsuicidal self-​injury  400, 401 eyetracking 46
Emotional Cascade Model 
borderline personality disorder  262, 269 F
nonsuicidal self-​injury  402, 403 face emotion labeling 
emotional clarity  425–​6 disruptive mood dysregulation disorder  290, 293
emotional competence  137 narrow phenotype bipolar disorder  288, 289
development 132, 133 severe mood dysregulation  288–​91
emotional inertia  217 facial action coding system (FACS)  46
emotional inexpressivity  404 facial expressions, measures  46
Emotional Processing Theory  98 family factors 
emotional reactivity  see emotionality abuse and neglect  308–​10
emotional sensitivity  265–​7, 270 divorce, children of  336, 339–​40, 345
emotionality 215 family interventions 
abused/​neglected children and adolescents  314 anorexia nervosa  198–​200, 201, 202, 206
conduct disorder  135, 432 case example  204, 205–​6
conscience development  138 Attention Deficit Hyperactivity Disorder  121
depression 432 borderline personality disorder  271–​2, 273
nonsuicidal self-​injury  402, 410 divorce, children of  334
reactive aggression  136 functional family therapy  140–​1
substance use disorders  215–​16, 219, 224 mentalization-​based treatment for
and temperament  133 adolescents  271–​2, 273
endophenotypes 97 self-​regulation deficits  27
environmental factors  family relationships 
emotion regulation  7, 8 anorexia nervosa  199, 204
psychopathology 9 Attention Deficit Hyperactivity Disorder  114
escape strategies  158, 159, 160–​1 borderline personality disorder  262
ethnicity factors  conduct disorder  131
incarceration, parental  362–​3 substance use disorders  224
substance use disorders  212 Fast Track Promoting Alternative Thinking Strategies
event-​related potentials (ERPs)  48 (PATHS) 27
executive function (EF)  7 fear behaviors 
Attention Deficit Hyperactivity Disorder  115, callous unemotional externalizing  25
116–​18, 119, 120 internalizing problems  30
depression  180–​1 Research Domain Criteria  97–​8
development 8 fear conditioning  90
effortful control  19 Feeling Thermometer  294, 295
externalizing problems  23–​4 fight-​flight response  46
internalizing problems  29 focused breathing  387
posttraumatic stress disorder  380 Fragile Families and Child Well Being (FFCWB)
self-​regulation  26, 27 study  354, 355, 365
substance use disorders  213 friendships  see peer relationships
experiential avoidance  see avoidance strategies functional family therapy (FFT)  140–​1
Experiential Avoidance Model of NSSI  401 functional magnetic resonance imaging (fMRI)
experimental research  44–​5 studies  428–​30
exposure therapy  97–​8 Functional Model of NSSI  401–​2, 403
expressive behavior, measuring  46
expressive suppression  see suppression G
extended process model (EPM)  85–​8 gender factors 
Research Domain Criteria  98, 99–​101 Attention Deficit Hyperactivity Disorder  114, 115
externalizing problems  21, 22–​7, 33–​4 autism spectrum disorder  236
abused/​neglected children and adolescents  311, 315, conduct disorder  131
316, 317 depression 
comorbid Attention Deficit Hyperactivity Disorder  115 prevalence 172
co-​occurring internalizing problems  28, 32–​3 risk factors  173
divorce, children of  332 development of emotional regulation  9
interventions 342 divorce, children of  338–​9
9
5
4

Index 459

eating disorders  196 incidence 352


incarceration, parental  360, 362, 366 maternal vs. paternal  352–​4
nonsuicidal self-​injury  399 types of 352
proactive aggression  138 visitation  358–​9, 365–​6
substance use disorders  212, 213, 220 incarceration-​specific risk experiences (ISRE)  361,
generalized anxiety disorder (GAD)  155–​6 363, 365
abused/​neglected children and adolescents  310 Incredible Years (IY)  141
comorbidity  conduct disorder  141–​2
major depressive disorder  447 self-​regulation  26
oppositional defiant disorder  282 independent model of the self  64–​5, 66–​7,
Emotion Regulation Therapy  446–​7 68–​9, 70
functional magnetic resonance imaging  429–​30 individualism  64, 65, 66, 68, 71
traumatic stress exposure  383 inhibitory control 
genetic factors  Attention Deficit Hyperactivity Disorder  116, 122
conduct disorder  132, 142 co-​occurring externalizing and internalizing
depression  10, 173–​4 problems 32
Research Domain Criteria  92, 97 internalizing problems  28–​9, 31
eating disorders  196 insecure attachment  4
emotion regulation  7, 8 borderline personality disorder  268
nonsuicidal self-​injury  399, 402, 410 incarceration, parental  351
psychopathology 10 psychopathology 10
Research Domain Criteria  92, 97 insomnia (INS)  425
substance use disorders  224 insula 48
gratification, delayed  see delayed gratification intellectual development  4
Grief and Trauma Intervention (GTI)  385–​6 intellectual disability (ID), and autism spectrum
Gross-​Ochsner view of emotion  81, 83–​8 disorder  235–​7
Research Domain Criteria  90, 98, 99–​101, 102 interventions 244
guilt 360 intelligence, and reactive aggression  137
interdependent model of the self  64–​5, 66–​7, 68–​9, 70
H internalizing problems  21, 27–​32, 33–​4
happiness, cultural factors  66–​7, 68 abused/​neglected children and adolescents  311, 315,
heart, parasympathetic nervous system influence on 316, 317
the  49–​51 comorbid Attention Deficit Hyperactivity Disorder  115
heart rate variability (HRV)  conduct disorder  130
high-​frequency (HF-​HRV)  co-​occurring externalizing problems  28, 32–​3
parasympathetic nervous system influence on the divorce, children of  332
heart 49, 50 interventions 342
transdiagnostic approaches  428 emotion regulation strategies  421–​5, 427
parasympathetic nervous system influence on the incarceration, parental  360, 361–​2, 366
heart 49, 50 parasympathetic nervous system influence on the
transdiagnostic approaches  428, 430 heart 49
traumatic stress exposure  380–​1, 383 traumatic stress exposure  376, 378, 382
hedonic capacity  218, 220 International Classification of Diseases (ICD) 
Helping Young People Early (HYPE) program  270 depression  171–​2
How I Feel questionnaire  361 history of mental health/​disorder  80
Internet-​based programs for children of divorce  344
I interpersonal therapy (IPT)  293–​4
illicit drug use  INTOVIAN Tool  307
emotion dysregulation  219 invalidating childhood experiences  400
intervention example  223 Inventory of Parent and Peer Attachment Inventory
prevalence 211, 212 (IPPA) 357
see also substance use disorders irritability 
impulsivity  abused/​neglected children and adolescents  314
abused/​neglected children and adolescents  315 abusive parents  309
alcohol misuse  426 autism spectrum disorder  242
Attention Deficit Hyperactivity Disorder  115, 117, 118 borderline personality disorder  267
self-​regulation deficits  20, 21, 33 disruptive mood dysregulation disorder  281
co-​occurring externalizing and internalizing differential diagnosis  282
problems 32 emotion (dys)regulation  283
externalizing problems  22, 24, 25–​6 divorce, children of  335
internalizing problems  29, 30, 31 etiopathological model  283–​4
incarceration, parental  351–​67 generalized anxiety disorder  282
attachment  356–​8 labeling  294–​5
context  352–​6 major depressive disorder  282–​3
effects on children  354–​6 oppositional defiant disorder  282
emotion processes  356–​64 pharmacotherapy 293
emotion regulation  359–​64 severe mood dysregulation  282, 283, 294
0
6
4

460 Index

J depression  176–​7
James, William  81, 82 posttraumatic stress disorder  380
working 116
K mental health and disorder, history of  79–​81
Kids First  344 mentalization based therapy (MBT) 
Kids First Center  344 for adolescents (MBT-​A)  271–​3
Kids’ Turn program  343 borderline personality disorder  271–​3
for families (MBT-​F)  271
L nonsuicidal self-​injury  407–​8, 409
language development  132, 133 methylphenidate 122
LAST Project  385 mindfulness 
late positive potentials (LPPs)  48 anorexia nervosa  201, 205
limbic-​hypothalamic-​pituitary-​adrenal (LHPA) axis  Attention Deficit Hyperactivity Disorder  122
autism spectrum disorder  240 autism spectrum disorder  243
traumatic stress exposure  379, 381 Secret Agent Society-​Operation Regulation
Limited Prosocial Emotions (LPE)  see callous program  245, 248, 249, 250
unemotional (CU) traits depression  424–​5
lithium 293 Dialectical Behavior Therapy Skills Training  446
longitudinal research designs  430–​47 emotion regulation strategies  424
nonsuicidal self-​injury  408, 409
M substance use disorders  214, 218, 220–​1
major depressive disorder (MDD)  training 122
central facets of emotion regulation  425 traumatic stress exposure  387
comorbidity  Mindfulness Attention Scale (MAAS)  220
generalized anxiety disorder  447 substance use disorders  221
oppositional defiant disorder  282 MINI International Neuropsychiatric Interview  441
decision making  178 Monitoring the Future (MTF) survey  210–​11
diagnosis 171, 172 motivational interviewing  409
Emotion Regulation Therapy  447 Movie Task for the Assessment of Social Cognition  272
executive functioning  180 multi-​dimensional experiential avoidance questionnaire
genetic factors  92, 97 (MEAQ)  441–​2
heart rate variability  428, 430 Multisystemic Therapy (MST)  139
memory 177 conduct disorder  139–​40
neural mechanisms  289
prevalence 172 N
Research Domain Criteria  92, 97, 98 narrow phenotype bipolar disorder (NP-​BD)  282,
risk factors  173, 174 283, 190
traumatic stress exposure  378 attention-​emotion interactions, dysregulated  286
Unified Protocol  445 context-​sensitive regulation, decreased  284–​5
Manual Assisted Cognitive-​Behavioral Therapy interventions  292–​3
(MACT)  405–​6 social-​emotional stimuli, misinterpretation of  288
Manualized Cognitive-​Behavioral Therapy  406 National Institute for Mental Health (NIMG), Research
marijuana use  Domain Criteria  80, 88, 90, 420–​1
distress tolerance  219 National Longitudinal Study of Youth  366
emotion dysregulation  219 National Survey on Drug Use and Health
intervention example  223 (NSDUH) 211, 212
prevalence 211, 212 natural disaster exposure 
mass trauma exposure  see traumatic stress exposure emotion regulation mechanisms  383
maternal factors  lifetime risk  375
abuse and neglect  308, 309, 313 posttraumatic stress disorder  376
anxiety disorders  161 see also traumatic stress exposure
Attention Deficit Hyperactivity Disorder  114 negative automatic thoughts  425
cultural scripts  69 negative divorce effect  334
divorce, children of  336, 339, 340, 341 neglect 
incarceration  352–​4, 366 definition 306
attachment  356, 357, 358 emotion regulation problems, development of  313,
emotion regulation  360, 361, 363 314, 315
psychopathology 10 nonsuicidal self-​injury  400
socialization  62, 63, 132–​3 outcomes 311
see also parental factors prevalence 306
Maudsley Model of Anorexia Nervosa Treatment for risk factors  308, 309–​10
Adults (MANTRA)  200 see also abuse and neglect
medication  see pharmacotherapy neural measures of emotional reactivity  48, 50
meditation 243 reappraisal 48, 50
memory  detached vs. positive  53–​4
autobiographical (AM)  176–​7, 181 vs. suppression  51–​2
1
6
4

Index 461

neural mechanisms  severe mood dysregulation  294


abuse and neglect  313–​14 traumatic stress exposure  386, 387, 388
anxiety disorders  159–​60 parental factors 
Attention Deficit Hyperactivity Disorder  117, 118, abuse and neglect  308–​10
286, 288 emotion regulation problems, development of  312
autism spectrum disorder  238–​9 anxiety disorders  157, 160–​1
borderline personality disorder  266 Attention Deficit Hyperactivity Disorder  114, 120
conduct disorder  134 borderline personality disorder  262
disruptive mood dysregulation disorder  290 conduct disorder  131
labeling irritable emotion  294–​5 cultural scripts  69
major depressive disorder  289 depression 173, 174
research methods  45 development of emotional regulation  9
severe mood dysregulation  283, 285, 286–​7, 288–​90, 294 disruptive mood dysregulation disorder  295
transdiagnostic approaches  428–​30 divorce  see divorce, children of
traumatic stress exposure  379–​81 eating disorders  196, 200
neuroticism  incarceration  see incarceration, parental
parasympathetic nervous system effects on the heart  49 posttraumatic stress disorder  384
Unified Protocol  445 poverty stressors  133
New Beginnings Program (NBP)  343–​4 psychopathology 9, 10
nonsuicidal self-​injury (NSSI)  398–​410 self-​regulation  26–​7
clinical guidance on management and treatment  408–​9 severe mood dysregulation  295
comorbid borderline personality disorder  268, 269, socialization  62, 63, 132–​3
270, 271 substance use disorders  224
Experiential Avoidance Model  401 see also maternal factors; paternal factors
extent of 399 Parenting through Change  344
Functional Model  401–​2, 403 Parents and Children Making Connections—​Highlighting
interventions  405–​8 Attention (PCMC-​A) program  26–​7
nature of 399 parietal cortex  287
theoretical perspectives  399–​405 paternal factors 
nucleus accumbens  48 Attention Deficit Hyperactivity Disorder  114
divorce, children of  334, 336–​7, 345
O incarceration  352–​4, 355, 366
observational research  44 attachment 357, 358
one-​session treatment (OST)  162 psychopathology 10
online programs for children of divorce  344 see also parental factors
oppositional defiant disorder (ODD)  peer relationships 
comorbidity 282 abused/​neglected children and adolescents  317, 318
Attention Deficit Hyperactivity Disorder  115, 121 Attention Deficit Hyperactivity Disorder  114
and conduct disorder, relationship between  130, 134 divorce, children of  333–​4, 335, 341
differential diagnosis  282 reactive aggression  136
parent management training  291 traumatic stress exposure  378, 382
optimism 341 perseveration 241
orbitofrontal cortex  118 personality development  4
overanxious disorder  see generalized anxiety disorder personality disorders  311
over-​general memory (OGM)  176–​7, 181 personality traits  341
pharmacotherapy 
P Attention Deficit Hyperactivity Disorder  121–​2
pain, chronic  434, 444–​5 autism spectrum disorder  242
panic disorder (PD)  156 disruptive mood dysregulation disorder  292–​3
abused/​neglected children and adolescents  310 phobias 156, 159
functional magnetic resonance imaging  429–​30 abused/​neglected children and adolescents  310
parasuicide  see nonsuicidal self-​injury cognitive behavioral therapy  162
parasympathetic nervous system (PNS)  social  see social anxiety
influence on the heart  49–​51 physical abuse 
measures 45 definition 305
traumatic stress exposure  380–​1 emotion regulation problems, development of  314, 315
parent education/​parent management training (PMT)  nonsuicidal self-​injury  400
abuse and neglect  319 outcomes 311
anxiety disorders  162, 163–​4 prevalence 306
Attention Deficit Hyperactivity Disorder  121, 294 risk factors  307–​8, 310
autism spectrum disorder  243–​4, 292 see also abuse and neglect
conduct disorder  141–​2, 291 physiological aspects  see biological and physiological
disruptive mood dysregulation disorder  291–​2, 295 aspects of emotion regulation
divorce  343–​4, 345 posterior cingulate 
Incredible Years  141–​2 narrow phenotype bipolar disorder  289
oppositional defiant disorder  291 severe mood dysregulation  287, 289
2
6
4

462 Index

posttraumatic stress disorder (PTSD)  375–​6, 378, 388–​9 cultural factors  66, 67, 70


abused/​neglected children and adolescents  311, 315–​16 depression 180
behavioral mechanisms  382 detached  53–​4
cognitive mechanisms  382 vs. distraction  52–​3
emotion regulation mechanisms  379 divorce, children of  338, 340, 341
immediate posttraumatic stress reactions  379 executive function  7
interventions  384, 385–​6, 387, 388 generalized anxiety disorder  429
neurobiological mechanisms  380, 381 internalizing disorders  29, 421–​5
substance use disorders  218, 219–​20, 224 neural measures  48, 50, 51–​4
vulnerability and moderating factors  383–​4 nonsuicidal self-​injury  398, 403–​5, 408–​9, 410
posttraumatic stress symptoms (PTS)  parasympathetic nervous system influence on the
abused/​neglected children and adolescents  316 heart 49
immediate  376–​8, 379 and pathology, relationship between  425
interventions 386 positive  53–​4
secondary  377, 378, 383 process model  85
poverty  Research Domain Criteria  90, 91
abuse and neglect  309–​10 strategies, variety among  53–​4
conduct disorder  133 substance use disorders  214, 217
incarceration, parental  354 vs. suppression  51–​2
prefrontal cortex (PFC)  sympathetic nervous system measures  46–​7
autism spectrum disorder  238 traumatic stress exposure  382
generalized anxiety disorder  429–​30 Reflective Functioning Questionnaire for Youth  272
panic disorder  429–​30 relaxation strategies 
social phobia  428–​9 autism spectrum disorder  244
traumatic stress exposure  379, 380 Secret Agent Society-​Operation Regulation
pregenual anterior cingulate cortex (pgACC)  313 program 248
proactive externalizing  24–​5 traumatic stress exposure  387
problem solving  Research Domain Criteria (RDoC)  80–​1, 88–​102, 420–​1
internalizing disorders  421–​4 conceptual scientific advancement  98–​102
substance use disorders  214, 225 empirical scientific advancement  92–​8
Problem-​Solving Skills Training (PSST)  292, 405 perspectives of emotion  83
process model of emotion regulation  44, 84–​5, 90–​1 respiratory sinus arrhythmia (RSA) 
divorce, children of  340 borderline personality disorder  269
nonsuicidal self-​injury  402–​3 conduct disorder  135, 432–​3
traumatic stress exposure  378 depression  432–​3
progressive muscle relaxation  parasympathetic nervous system effects on the
autism spectrum disorder  244 heart  49–​50
comorbid anxiety and alcohol use disorders  424 traumatic stress exposure  380–​1
Promoting Alternative Thinking Strategies (PATHS) restrictive eating 
curriculum 319 emotional clarity  426
psychoeducation  see also Avoidant/​Restrictive Food Intake Disorder
anorexia nervosa  202, 203, 206 risperidone 242, 293
borderline personality disorder  272 rumination 215
traumatic stress exposure  385, 386 autism spectrum disorder  243
psychopathology in children and adolescents  9–​11 bipolar disorder and insomnia compared  425
psychopathy 315, 316 borderline personality disorder  262, 269
psychosomatic complaints  336, 337 depression 179, 80
psychotherapy  externalizing problems  33, 427
anorexia nervosa  198, 199 internalizing problems  29, 33, 421–​5, 427
Attention Deficit Hyperactivity Disorder  121 nonsuicidal self-​injury  398, 402, 403, 408, 410
autism spectrum disorder  243 substance use disorders  217
disruptive mood dysregulation disorder  291–​2, 294 as transdiagnostic mechanism  425
traumatic stress exposure  381
Q
Q methodology  346 S
sadness 67
R sandcastles program  344
Radically-​Open Dialectical Behavior Therapy schizophrenia 311
(RO-​DBT)  200 school-​based interventions 
reactive attachment disorder  311 divorce, children of  342
reactive control  20 teacher training 
externalizing problems  22–​3, 24 Attention Deficit Hyperactivity Disorder  121
internalizing problems  27, 29–​30, 31 Incredible Years  141
reappraisal  5–​6, 44 traumatic stress exposure  385–​6
anxiety disorders  158 Secret Agent Society-​Operation Regulation (SAS-​OR)
autism spectrum disorder  239 program  244–​53
biological measures  45 Seeking Safety  318
3
6
4

Index 463

selective mutism (SM)  155 strange situation experiment  4


self-​blame  33, 427 stress 
self-​injurious behavior  see nonsuicidal self-​injury abuse and neglect  309, 310, 317
Self Referent Encoding Task (SRET)  176 autism spectrum disorder  239
self-​regulation  18, 33–​4 in caregivers  354
conceptual issues  18–​20 cortisol response to  379, 380, 381
co-​occurring problems  32–​3 measures  47–​8
definition 18 depression 173, 174
effortful control  19–​20 divorce, children of  339, 344
externalizing problems  22–​7 incarceration, parental  354, 358–​9, 361
internalizing problems  27–​32 nonsuicidal self-​injury  400
maladjustment  21–​2 traumatic  see traumatic stress exposure
reactive control  20 striatum 287
separation, parental  see divorce, children of Stroop task 
separation anxiety disorder (SAD)  155, 159, 160 Attention Deficit Hyperactivity Disorder  118
separation distress  28, 30 emotion regulation  116
severe mood dysregulation (SMD)  281–​2, 296 substance use disorders (SUDs)  210–​25
emotion (dys)regulation  283–​91 abused/​neglected children and adolescents  311, 316
interventions  291–​4 abusive parents  309
labeling irritable emotion  295 comorbidity 
parental contingencies for behavior, consistency in  295 borderline personality disorder  268
sexual abuse  nonsuicidal self-​injury  398
definition  305–​6 composite constructs and dual-​process theory  221–​2
emotion regulation problems, development of  314, 315 diagnosis  212–​13
nonsuicidal self-​injury  400 divorce, children of  343
prevalence 306 emotion regulation  213–​21
risk factors  308 attributes  215–​16
see also abuse and neglect intervention example  222–​3
shame  strategies  213–​14
borderline personality disorder  267 emotional clarity  426
incarceration, parental  360 impulsivity 25
traumatic stress exposure  385 incarceration, parental  354, 355
shyness 30 prevalence  210–​12
single parents  308 relapse rates  10
Skills Training in Affect and Interpersonal traumatic stress exposure  378
Regulation (STAIR)/​Prolonged Exposure  318–​19 suicidal behavior  399
smoking behavior  see cigarette use dialectical behavior therapy  407, 446
Snaith-​Hamilton Pleasure Scale (SHAPS)  220 Supportive Parenting for Anxious Childhood Emotions
social anxiety 155, 160 (SPACE) program  163–​4, 165
divorce, children of  340 suppression 5, 6
emotional clarity  426 abused/​neglected children and adolescents  317
functional magnetic resonance imaging  428–​9 Attention Deficit Hyperactivity Disorder  116
self-​regulation deficits  22 cultural factors  8, 66, 67, 70
internalizing problems  29 divorce, children of  335, 338, 340
social cognitive neuroscience  83 executive function  7
social development  4 internalizing disorders  29, 421–​4
social-​emotional stimuli, misinterpretation of  287–​90 nonsuicidal self-​injury  398, 403–​5, 408, 410
social factors  60–​71 parasympathetic nervous system effects on the heart  49
abuse and neglect  310 process model  85
Attention Deficit Hyperactivity Disorder  114 vs. reappraisal  51–​2
divorce, children of  333, 336 substance use disorders  214
social learning theory  309 traumatic stress exposure  379, 381
social phobia  see social anxiety sympathetic nervous system (SNS) 
social relationships  see peer relationships activation measures  46–​7
social withdrawal  22 anxiety disorders  159, 160
socialization  61–​4, 67, 71, 132–​3 traumatic stress exposure  380
abused/​neglected children and adolescents  312 systemic family therapy  198
anxiety disorders  161 Systems Training for Emotional Predictability and Problem
cultural scripts  69 Solving (STEPPS)  270
divorce, children of  339
incarceration, parental  359, 363–​4, 366, 367 T
soothability 215 teacher training 
substance use disorders  216, 224 Attention Deficit Hyperactivity Disorder  121
specific phobias (SP)  156, 159 Incredible Years  141
abused/​neglected children and adolescents  310 temperament 
cognitive behavioral therapy  162 and emotionality  133
Stengel, Erwin  80 substance use disorders  215–​16
4
6

464 Index

theory of mind  240–​1 for adults 


thinking about reward in young people (TRY) group intervention  441–​2
program  182–​3, 184–​5 individual intervention  434–​41
thought suppression  see suppression bipolar disorder  444
threat appraisal/​system  for children  443–​4
abused/​neglected children and adolescents  318 neuroticism 445
biological aspects  47 pain, chronic  444–​5
disruptive mood dysregulation disorder  290 Unified Protocol for the Treatment of Emotional Disorders
divorce, children of  340 in Youth (UP-​Y)  444
severe mood dysregulation  290 anxiety disorders  163, 165
three-​systems view of emotion  83, 84 Unified Protocol for Transdiagnostic Treatment of
Tools of the Mind intervention  26 Emotional Disorders  424
transdiagnostic approaches  419–​48 Unified Protocol in Youth with Pain (UP-​YP)  444–​5
biomarker and fMRI studies  428–​30
clinical trials  433–​4, 435–​9 V
cross-​sectional research  433–​47 vagal tone  49
with adults  421–​6, 427 venlafaxine 424
with youth  426–​8 ventral striatum  118
dialectical behavior therapy skills training  446 verbal intelligence  137
emotion regulation therapy  446–​7 visual cortex  48
experimental, survey, and neurobiological designs  421
longitudinal designs  430–​47 W
Unified Protocol  434–​45 website-​based programs for children of divorce  344
transference-​focused psychotherapy (TFP)  271 within-​subjects research  45
transition from adolescence to adulthood  see adulthood, working memory  116
transition from adolescence to World Health Organization 
Trauma Adaptive Recovery Group Education and child abuse, definition  305
Therapy 318 history of mental health/​disorder  79–​80
trauma-​focused cognitive behavioral therapy International Classification of Diseases (ICD) 
(TF-​CBT)  384–​5, 387–​8 depression  171–​2
traumatic stress exposure  374–​89 history of mental health/​disorder  80
interventions  384–​8 worry 
outcomes  375–​84 anxiety disorders  425
prevalence 375 bipolar disorder and insomnia compared  425
internalizing disorders  424–​5
U
Unified Protocol (UP)  434–​45 Y
for adolescents  442–​3 Youth Risk Behavior Surveillance Study  211

S-ar putea să vă placă și