(Updesh Kumar) Suicidal Behaviour Assessment of P

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 405

Suicidal Behaviour

ii
Suicidal Behaviour
Assessment of People-at-Risk

Edited by
Updesh Kumar
Manas K. Mandal
Copyright © Updesh Kumar and Manas K. Mandal, 2010

All rights reserved. No part of this book may be reproduced or utilized in any form
or by any means, electronic or mechanical, including photocopying, recording or
by any information storage or retrieval system, without permission in writing from
the publisher.

First published in 2010 by

Sage Publications India Pvt Ltd


B 1/I-1 Mohan Cooperative Industrial Area
Mathura Road, New Delhi 110 044, India
www.sagepub.in
Sage Publications Inc
2455 Teller Road
Thousand Oaks, California 91320, USA
Sage Publications Ltd
1 Oliver’s Yard, 55 City Road
London EC1Y 1SP, United Kingdom
Sage Publications Asia-Pacific Pte Ltd
33 Pekin Street
#02-01 Far East Square
Singapore 048763

Published by Vivek Mehra for Sage Publications India Pvt Ltd, typeset in
10/12pt Minion by Star Compugraphics Private Limited, Delhi and printed at
Chaman Enterprises, New Delhi.

Library of Congress Cataloging-in-Publication Data

Suicidal behaviour: assessment of people-at-risk/edited by Updesh Kumar,


Manas K. Mandal.
╅╅╇ p. cm.
â•… Includes bibliographical references and index.
1. Suicidal behavior—Risk factors—Testing. I. Kumar, Updesh. II. Mandal,
Manas K.

RC569.S865 616.85′8445—dc22 2010 2009049002

ISBN:╇ 978-81-321-0299-1 (HB)

The Sage Team:╇ Rekha Natarajan, Pranab Jyoti Sarma, Mathew P J and Trinankur
Banerjee
Dedicated to

Dr (Mrs) Sanjukta Mandal


and
Mrs Anju Walia
vi
Contents

List of Tables ix
List of Figures xi
List of Abbreviations xiii
Foreword xvii
Preface xix

Section I—Risk Assessment: Theoretical Issues

1. Psychological Perspectives on Suicidal Behaviour 3


Rory C. O’Connor
2. Empirically Based Assessment of Suicide Risk 20
Chad E. Morrow, Craig J. Bryan and Kathryn Kanzler Appolonio
3. Neurobiological Basis of Suicidal Ideation 42
Jitendra Kumar Trivedi and Sannidhya Varma
4. Problem-solving Ability and Repeated Deliberate Self-harm 65
Carmel McAuliffe
5. Suicide and Homicide: Theoretical Issues 91
Swati Mukherjee, Updesh Kumar and Manas K. Mandal
6. Cultural Issues in Suicide Risk Assessment 107
Erminia Colucci
7. Gender Issues in Suicide Risk Factor Assessment 136
Peter Osvath, Viktor Voros and Sandor Fekete
8. Developmental Issues in Risk Factor Assessment 152
Kimberly A. Van Orden and Alec L. Miller

vii
Suicidal Behaviour

╇ 9. Reporting Suicide: Impact on Suicidal Behaviour 173


Farah Kidwai

Section II—Assessment: People-at-Risk

10. Suicide: Its Assessment and Prediction 193


Pritha Mukhopadhyay
11. Substance Use and Suicidal Behaviour 230
Nishi Misra, Amri Sabharwal and Updesh Kumar
12. Suicide Risk in Bipolar Disorder 256
Maurizio Pompili, Marco Innamorati, Enrica De Simoni,
Ilaria Falcone, Gaspare Palmieri, Laura Sapienza
and Roberto Tatarelli
13. Depression and Suicide 278
Eva Schaller and Manfred Wolfersdorf
14. The Suicidal Soldier 297
Lars Mehlum and Latha Nrugham
15. Suicidal Ideation and Behaviour among Asian Adolescents 324
Angel Nga-man Leung, Cathy Yui-chi Fong
and Catherine Alexandra McBride-Chang

About the Editors and Contributors 343


Author Index 355
Subject Index 370

viii
List of Tables

1.1 A summary of Baumeister’s (1990) suicide as escape


from self model 06
1.2 Psychological risk and protective factors associated
with suicidal risk 11

2.1 Eight empirically supported areas for suicide risk


assessments 28
2.2 Categories of suicide risk 35
2.3 Suicide risk continuum with indicated responses 37

4.1 UCL dimensions and Their items 78

7.1 Gender differences of suicide attempters (multivariate


logistic regression model) 145

12.1 Factors contributing to increased suicide risk in cases of


comorbid substance abuse disorder and bipolar disorder 261
12.2 A check-list of risk factors for suicidal behaviour in
bipolar disorder 263

13.1 Psychiatric disorders, especially depressive disorders


and suicide in a community-based study using
psychological autopsy 281
13.2 Major depressive disorder and suicide: Comparison of
depressive suicide versus depressive non-suicide controls 285
13.3 Suicidal versus non-suicidal depressed inpatients of the
Weissenau Depression Treatment Unit Therapist´s
assessment at admission (significantly discriminating
variables of a patient’s questionnaire) 287

ix
Suicidal Behaviour

13.4 Symptoms significantly differentiating suicides and


non-suicides in studies with depressive disorder patients 289
13.5 The depressed patient with suicide risk—clinical
psychopathological picture in the presuicidal situation 290

15.1 Suicide rates in alternate years among 15- to 24-year-old


adolescents per 100,000 of the population in selected
Asian countries/regions (1990–2006) 326

x
List of Figures

1.1 An overarching biopsychosocial model of suicidal behaviour 5


1.2 Cry of Pain model 8
1.3 Positive future thinking as a moderator of the SPP–distress
relationship 15

2.1 Biopsychosocial model of suicide 24

3.1 A stress-diathesis model of suicide 58

7.1 Suicide rates by age groups in Hungary


(per 100,000 inhabitants) 146
7.2 Suicide attempts by age groups (per 100,000 inhabitants) 146

xi
xii
List of Abbreviations

ACE Angiotensin converting enzyme


AMP Adenosine monophosphate
APA American Psychiatric Association
APOE4 Apolipoprotein E4
ARISE Adaptive regression in service of ego
ASIQ Adult Suicidal Ideation Questionnaire
AtYS Attitude towards Youth Suicide scale
BD Bipolar disorder
BDI Beck Depression Inventory
BDNF Brain-derived neurotrophic factor
BHS Beck Hopelessness Scale
BPD Borderline personality disorder
BSSI Beck Scale for Suicide Ideation
CAMS Collaborative Assessment and Management of
Suicidality
CAPS Child and Adolescent Perfectionism Scale
CBT Cognitive-behavioural therapy
CCK Cholecystokinin
CDC Centre for Disease Control
CDI Children’s Depression Inventory
CDRS-R Children’s Depression Rating Scale
CDS Collaborative Depression Study
CI Confidence interval
COMT Catechol-O-methyltransferase
CoP Cry of pain
CREB Cyclic AMP response element binding protein
CSF Cerebro-spinal Fluid

xiii
Suicidal Behaviour

CSF-5HIAA Cerebrospinal fluid 5 hydroxyindoleacetic acid


CSPS Child Suicide Potential Scale
CSRP Centre for Suicide Research and Prevention
DBT Dialectical Behaviour Therapy
DHCR7 7-Dehydrocholesterol reductase
DRD2 Dopamine D2 receptor
DSH Deliberate self-harm
DSM Diagnostic and Statistical Manual
DZTs Dizygotic twins
EARS Epigenetic Assessment Rating System
EASQ Extended Attributional Style Questionnaire
ECA Epidemiologic Catchment Area
ECT Electroconvulsive therapy
EFA Ego Function Test
EFA-M Ego Function Assessment Scale-Modified
EPQ Eysenck Personality Questionnaire
EPS Escape Potential Scale
ESEMED European Study on the Epidemiology of Mental
Disorders
FACES Family Adaptability and Cohesion Evaluation Scales
FAST Firestone Assessment of Self-Destructive Thoughts
FDA Food and Drug Administration
FTT Future Thinking Task
FVT Fluid vulnerability theory
GAD Generalised Anxiety Disorder
GSK-3 Glycogen synthase kinase-3-b
HDRS Hamilton Depression Rating Scale
5-HIAA 5-Hydroxy indole acetic acid
HPA axis Hypothalamic-pituitary-adrenal axis
HPLS Hopelessness Scale For Children
HRSD Hamilton Rating Scale for Depression
IASP International Association for Suicide Prevention
ICD-10 International Classification of Diseases 10
IES Impact of Events Scale
LES Life Experience Survey
LHPA Axis Limbic-Hypothalamus-Pituitary-Adrenal Axis
LHPT Axis Limbic-Hypothalamus-Pituitary-Thyroid Axis
LS Lethality Scales
LSI Life Style Index

xiv
List of Abbreviations

MAOA Monoamine oxidase A


MDD Major Depressive Disorder
MDE Major depressive episode
MINI Mini International Neuropsychiatric Interview
MLDA Minimum legal drinking age
MMPI Minnesota Multiphasic Personality Inventory
MPQ Motives for Parasuicide Questionnaire
MPS Multidimensional Perfectionism Scale
MSSI Modified Scale for Suicide Ideation
MUP Multiple Preanalysis
MZTs Monozygotic twins
NIMH National Institute of Mental Health
NOS Nitric oxide synthase
NSSI Non-suicidal self-injurious behaviour
PANAS Positive and Negative Affect Scale
PANSI Positive and Negative Suicide Ideation Inventory
PBI Parental Bonding Instrument
PD Personality Disorder
PET Positron emission tomography
PFC Pre-frontal cortex
PNQ Psychache Needs Questionnaire
PROSPECT
Prevention of Suicide in Primary-care Elderly:
Collaborative Trial
PSI Paykel Suicide Items
PST Group Interpersonal Problem-Solving Skills Training
programme
PTS Post-traumatic stress
PTSD Post-traumatic stress disorder
RADS-2 Reynolds Adolescent Depression Scale-2
RASQ Reasons for Attempting Suicide Questionnaire
RCT Randomised Clinical Trial
RFL Reasons for Living Inventory
RR Relative risk
RRR Risk-Rescue Rating Scale
RSIC Rorschach Suicide Index Constellation
SAT Separation Anxiety Test
SBQ Suicide Behaviours Questionnaire
SCIC Suicide Crisis Intervention Centre

xv
Suicidal Behaviour

SEESA Sequential Emotion and Event Form for Suicidal


Adolescents
SI-IAT Self Injury Implicit Association Test
SIS Suicide Ideation Scale
SIS Suicide Intent Scale
SIS-Q Suicidal Ideation Screening Questionnaire
SIISF Self-Inflicted Injury Severity Form
SMASI Sudden mass assault by a single individual
SMR Standard mortality rate
SMSI Self Monitoring Suicide Ideation Scale
SNP Single nucleotide polymorphisms
SPECT Single Photon Emission Computed Tomography
SPP Socially prescribed perfectionism
SPRC Suicide Prevention Resource Center
SPS Suicide Probability Scale
SSF Suicide Status Form
SSI Scale for Suicide Ideation
SSI-W Scale of Suicide Ideation-Worst
SSRI Selective serotonin reuptake inhibitor
ST Serotonin transporters
TAT Thematic Apperception Test
TAU Treatment as usual
TEMPS-A-Rome Temperament Evaluation of Memphis, Pisa, Paris and
San Diego autoquestionnaire-Rome
TH Tyrosine hydroxylase
TPH1 Tryptophan hydroxylase 1
TSH Thyroid Stimulating Hormone
UCL Utrecht Coping List
VNTR Variable number tandem repeats
WHO World Health Organization
ZDI Zung Depression Scale

xvi
Foreword

T he problem of suicide is important and intriguing. It is important


because it is surprisingly widespread—so widespread, in fact, that it
constitutes a legitimate health hazard. It is also important because it often
involves talented people in key positions in government and industry,
and when it involves these people, key resources disappear forever. The
problem is intriguing because it is counter-intuitive from an existential,
phenomenological and evolutionary perspective—it is a kind of behaviour
that seems to defy the most basic tendencies in human nature, those that
concern survival.
Properly considered, science is a two-step process which I call prediction
and explanation. In the first step, researchers try to detect co-variations in
nature, to determine what goes with what, to establish reliable empirical
relationships. The second step involves trying to explain why the co-
variations occur, what the dynamics are that generate the reliable empirical
relationships. This is a good description of how science actually develops,
but it is a perspective that was badly out of favour among psychologists
and philosophers of science for perhaps 50 years; for the behaviourists
and the logical positivists, prediction and explanation were identical. They
were wrong, and by confusing prediction and explanation, they confused
rather than clarified a lot of problems.
In psychology, prediction is largely accomplished using assessment—
valid psychometric procedures—we use assessment to predict outcomes.
Explanation is accomplished by linking the empirical co-variations (the
predictions) to underlying biological, physiological and neurological
processes.
This book has several strengths, perhaps the first of which is the
degree to which it implicitly honours the distinction between prediction

xvii
Suicidal Behaviour

and explanation. The chapters divide rather neatly into psychometric


assessment methods and explanatory models based on a variety of plausible
perspectives, including neurophysiology, cognitive processes, and cultural
and social influence models.
The second strength of the book is its ecumenical focus; the contributors
come from different countries, cultures and disciplinary backgrounds. They
include psychiatrists, psychologists, sociologists, epidemiologists and
suicidologists. The authors include Americans, Germans, Indians, Italians,
Australians, Norwegians, Chinese, Irish, Hungarians and English scholars.
The book is neither parochial nor partisan.
Third, the contributions are sensitive to epidemiological, cultural and
gender influences. Suicide is examined from a fully rounded perspective.
Finally, the book makes clear that psychological assessment has a
crucial role to play in applied psychological and social research. Modern
measurement research seems often to have lost its way as it pursues item
response theory, structural equations modelling, and latent variable
analyses for their own sake. This book makes it abundantly clear that ‘as-
sessment concerns predicting consequential outcomes, not measuring
abstract entities’. Assessment remains the most important contribution
psychology has so far made to the world of practical affairs.
This book is a very useful, balanced and comprehensive introduction
to the best modern thinking on the prediction, explanation and potential
amelioration of suicide, and the editors and authors are to be congratulated
for putting it together.

Robert Hogan, PhD


Hogan Assessment Systems
Amelia Island, Florida

xviii
Preface

T he increasing suicide rates around the globe have drawn immense


attention of psychologists, mental health professionals, policy-makers
and researchers towards the dynamics of suicidal behaviour. Suicide
constitutes a major public health problem across the world. The United
Nations has estimated that between 500,000 and 1.2 million people die by
committing suicide each year worldwide. According to the World Health
Organization estimates, in the year 2000, approximately 1 million people
died by committing suicide, and 10 to 20 times more people attempted
suicide worldwide. In most countries, suicide is now one of the leading causes
of death among people aged 15–34 years. Indeed, suicide is the most
frequently encountered emergency situation in mental health care pro-
grammes. Moreover, the trauma of suicide does not end with the indi-
vidual, but gets reflected in feelings of loss and suffering in the lives of
family members and friends for years after the suicide.
Innumerable publications are available on the subject of suicidal
behaviour, although most of these revolve around the epidemiological or
theoretical descriptions of suicide. Utilising empirical evidence in active
clinical practice and translating insights gained from clinical experience
into empirical data, however, still remains a challenge. Most of the avail-
able literature focuses on the clinical symptomatology of suicide and its
sociological development and influences. There is a paucity of literature
with a strong psychometric basis that explains suicidal behaviour in normal
population across lifespan. The present volume attempts to fill this void by
following a psychometric approach to outline suicide risk assessment and,
in turn, to comprehensively understand the suicidal personality/behaviour.
This volume includes 15 chapters from experts in the field encompassing
the suicide research carried out globally, and the diverse aspects of this
problem in various socio-cultural contexts.

xix
Suicidal Behaviour

Suicidal behaviour constitutes a vast area for empirical psychological


research. The abundance of research on various aspects of suicide has
led to an improved understanding of the behaviour, although accurate
assessment still remains a challenge. The chapters of the volume are divided
into two sections—‘theoretical issues related to assessment of suicidal
behaviour’ and ‘assessment of people at suicide risk’. The first section
includes nine chapters on topics related to risk assessment: theoretical
issues. The section begins with the chapter by O’Connor that provides
an overview of the ‘psychological perspectives on suicidal behaviour’.
Suicidal behaviour is described as reaction to a situation that has three
components—defeat, no escape and no rescue. The author contextualises
the psychological factors implicated in the aetiology of suicidal behaviour
with a biopsychosocial model. Clinical implications derived from
empirical research are discussed and the need for longitudinal as well
as experimental research is emphasised in order to expand the theoretical
evidence base on the basis of which effective psychological interventions
can be developed.
The overview of psychological perspectives on suicidal behaviour
is followed by the chapter on ‘empirically based assessment of suicide
risk’, which reviews and integrates empirically based strategies for the
assessment of suicide risk. The authors, Morrow, Bryan and Appolonio,
differentiate between risk assessment and prediction and emphasise the
importance of establishing a collaborative relationship with the suicidal
patient. A biopsychosocial model of suicide is also presented, along with
a continuum of suicidality for risk assessment that addresses both the
chronic and acute dimensions of suicide risk. The chapter concludes with
recommendations for clinical decision-making and practice that are based
on current scientific evidence and standards of care.
Trivedi and Varma, in their chapter on ‘neurobiological basis of suicidal
ideation’, point out the inadequacy of the standardised assessment and
prediction scales available and emphasise the need to better understand
and utilise the neurobiological basis of suicidal ideation for prediction
purposes. They delineate upon the neurobiological basis of suicide and
try to identify the biological markers of suicide. They further support
their notions with transmitter non-specific as well as transmitter specific
neuroendocrine studies. A detailed discussion is presented regarding
the genetic basis of suicidal behaviour and role played by different
neurotransmitters.

xx
Preface

The chapter on ‘problem-solving ability and repeated deliberate self-


harm’ by McAuliffe posits that deliberate self-harm has a strong associ-
ation with difficulties in problem-solving. Empirical evidence to establish a
link between repeated acts of deliberate self-harm with motives of escape
and revenge and a passive-avoidant approach to problem-solving is
discussed. The chapter also includes a valuable discussion on the efficacy
of cognitive-behavioural interventions to reduce suicidal behaviour,
including aspects of problem-solving skills training.
The chapter ‘Suicide and Homicide: Theoretical Issues’ discusses vio-
lence as the common thread underlying both suicide as well as homicide.
The authors, Mukherjee, Kumar and Mandal, discuss the evidence in the
light of the recent research findings that bring out the cognitive and per-
sonality factors placing the two behaviours on a continuum. The chapter
highlights the theoretical issues involved and elaborates upon the
neurological, sociological and psychological perspectives. Based on the
theoretical underpinnings, an attempt is made to draw a profile of indi-
viduals likely to engage in such violent behaviours towards self or others.
Erminia Colucci delves into ‘cultural issues in suicide risk assessment’
in the next chapter. She emphasises the significance and relevance of cul-
ture and ethnicity in promoting our understanding of suicidal behaviour,
suicide risk assessment and suicide prevention. The chapter includes an
overview of cross-cultural research on youth suicide, and of the role that
religion and spirituality may play at each step along the suicidal path. The
need to pay more attention to the meaning and interpretation of suicide in
suicide risk assessment and the necessity to establish culture-sensitive
prevention strategies is also addressed.
Gender is a pertinent issue of consideration in the assessment of suicidal
risk. Osvath, Voros and Fekete, in their chapter on ‘gender issues in suicide
risk factor assessment’, discuss the role of gender in determining various
aspects of the suicidal behaviour and summarise epidemiological and
socio-cultural research evidence identifying protective and risk factors for
suicide in males and females. The higher suicide mortality among males
in the Western countries is explained in terms of higher lethality of male
suicide methods, the reluctance of men to seek help, the higher rate of
substance abuse, and some socio-cultural differences. The authors urge
the need for further research in the area of gender differences and point
out how the insights gained can be utilised for suicide prevention in both
genders. The chapter concludes by highlighting gender-specific treatment
possibilities/clinical implications.

xxi
Suicidal Behaviour

Suicidal vulnerability and risk factors pertinent to different life stages


are discussed by Van Orden and Miller in their chapter on ‘developmental
issues in risk factor assessment’. The authors examine the ways in which
developmental issues impact clinical decision-making in the process of
assessing and managing suicide risk across the lifespan ranging from
childhood to late adulthood. They discuss four aspects of suicide risk
assessment where developmental issues may play a role: the content of
information collected during risk assessments, the process of conducting
risk assessments, the context surrounding risk assessments and the
decisions generated regarding crisis management. The factors unique to
particular developmental stages and the way in which these factors manifest
differently at different ages are elaborated upon in a sequential manner.
Research has established that the manners of reporting suicide stories
have a significant influence on suicidal behaviour and suicide-prevention
strategies and policies. Kidwai elaborates upon this emerging area of suicide
research in her chapter ‘Reporting Suicide: Impact on Suicidal Behaviour’.
She discusses the role of cognitive and arousal processes as mediators
that affect suicidal behaviour. The author elaborates upon the contagion
and cultivation effects of suicidal behaviour that are probable after media
coverage of suicidal acts. It is argued that the portrayal of suicides in the
media has an impact on suicidal behaviour, and it is, therefore, essential
to educate reporters, editors, film-television producers about suicide
contagion and related phenomena. The author conclusively assumes that
media’s positive role while reporting suicide stories, based on certain basic
guidelines which she proposes in the end, can go a long way in preventing
such behaviour.
Accurate assessment and prediction of suicidal behaviour is of utmost
importance for effective prevention, which is the primary motive of any
mental health programme. The lacunae often experienced by clinicians in
predicting suicidal behaviour through the use of psychometric assessment
tools are filled up by their clinical experience. The second section of the
present volume—‘Assessment: People-at-Risk’—includes six chapters.
This section opens with Mukhopadhyay’s chapter titled Suicide: Its
Assessment and Prediction. In this chapter, the author argues that the psy-
chological and psychosocial risk factors that lead to suicidality require
systematic psychological evaluation. She lists and critically reviews the
available suicide risk assessment tools for varied parameters ranging from
suicidal thoughts to completed suicide and other related psychological

xxii
Preface

domains. The tools are evaluated in terms of their psychometric properties


and the readers are provided with ample information for choosing the
assessment tool that match their requirement. Given multifactor origin of
suicide, the author argues the use of battery of tests for the evaluation of
direct factors related to suicidal behaviour and indirect potential factors
that contribute to it, so as to ensure the understanding of phenomenon
in totality.
In their chapter on ‘substance use and suicidal behaviour’, Misra,
Sabharwal and Kumar present an overview of the research on the incidence
of suicidal behaviour in conjunction with substance use. The chapter
includes an outline of the possible role of personality, cognitive and
neurobiological factors to predict suicidal behaviour among substance
users. The comorbidity of substance use with other psychiatric disorders,
particularly depression and personality disorders, and the demographic
variables and life events that further complicate the assessment of suicide
risk among substance users are also discussed. Finally, prevention and
intervention measures specific to substance users are listed.
Suicide is also the major source of mortality in the patients of bipolar
disorder. Bipolar disorder is known to increase suicide vulnerability
manifold. Pompili and associates, in their chapter on ‘suicide risk in bipolar
disorder’, comprehensively reviews the relevant literature. The authors
emphasise the high risk of suicide associated with bipolar disorder and
discuss the efficacy of various short and long-term intervention measures
for managing suicidality. Schaller and Wolfersdorf extend the issue on
similar line in their chapter on ‘depression and suicide’. They elaborate
upon the association between depression and suicidal behaviour on the
basis of epidemiological data. With adequate support of community-
based and clinically-based updated studies, the authors discuss the
prevalence of suicides in depressive disorders focusing on clinical signs,
risk factors and symptoms differentiating suicidal and non-suicidal de-
pressed patients. Furthermore, they also discuss about suicide prevention
in depressed patients.
Discussing the issues involved in suicide prevention, Mehlum and
Nrugham describe the emergence, maintenance and prevention of sui-
cidal behaviour in military settings and examine the environmental and
individual risk factors, protective factors and assessment issues. Their
chapter titled The Suicidal Soldier focuses on the suicidal phenomena
in the Norwegian Armed Forces and discusses these in the light of the

xxiii
Suicidal Behaviour

research findings and the personal experiences of the authors with the
armed forces.
The final chapter in this volume focuses on adolescence, a life stage
that is marked with maximum suicide vulnerability. The chapter on ‘sui-
cidal ideation and behaviour among Asian adolescents’, by Leung, Fong
and McBride-Chang, summarises the trends, risk factors, warning signs
and preventive measures for Asian adolescents’ suicidal behaviour.
Although majority of the available researches in the area are from Western
countries, in this chapter, risk factors for adolescent suicide categorised
into psychological, environmental and socio-cultural are discussed in light
of sufficient research evidence across major Asian countries. The authors
also highlight the warning signs for suicide, and propose preventive
measures and treatment options.
The theoretical issues elaborated upon in the first part of this volume
along with the applied and practical issues of the second part are an
effort to put the readers’ insight into the psychometrically sound suicide
risk assessment. Varied paradigms of suicidal behaviour along with the
suggested prevention strategies are discussed in an effort to provide a
scope for widening the horizon of the mental health professionals and
researchers working in this area so as to reduce the suicidal behaviour
across the world. The issues being raised in the volume are supposed to
promote more researches in this area that will certainly prove beneficial
in service of humanity.
Suicide and suicidal behaviour constitute a vast and varied area of
research, and editing a volume on the issue has undoubtedly been a
Herculean task. It would not have been possible to come up with the
volume in the present form without the help and understanding of the
people around us, who provided constant support and encouragement.
We extend our gratitude to one and all who facilitated our endeavours in
however small manner.
We express our gratitude towards our organisation, the Defence Research
and Development Organisation for providing us with infrastructural
support. We are indebted to our mentors, Shri M. Natarajan (Scientific
Advisor to Raksha Mantri, Secretary, Department of Defence Research
and Development, and Director General Research and Development) and
Dr W. Selvamurthy, Distinguished Scientist, Chief Controller Research
and Development (LS & HR) for their encouragement and benevolence.

xxiv
Preface

It has indeed been a rewarding experience to work with researchers


who are dedicated to this field of research. Their varied ideas stemming
from their diverse professional and cultural backgrounds have not only
enhanced our knowledge and understanding of the phenomenon of
suicide, but have also made the experience of editing this volume very
fascinating. We sincerely thank all the authors for their contributions and
look forward to future opportunities to work together. We are also obliged
to acknowledge the painstaking efforts put in by the reviewers, Emeritus
Prof. (Retd.) Sagar Sharma (Punjab University, Chandigarh) and Prof.
(Retd.) G.P. Thakur (Mahatma Gandhi Kashi Vidyapith, Varanasi).
Without their involvement it would not have been possible to come up
with the volume in its present form.
We would also like to extend our gratitude to our colleagues at the
Defence Institute of Psychological Research, Defence Research and Devel-
opment Organisation, for their help and co-operation in our endeavour
to bring out this volume on suicidal behaviour.

Updesh Kumar
Manas K. Mandal

xxv
Suicidal Behaviour

xxvi
SecƟon I

Risk Assessment:
TheoreƟcal Issues
2
1

Psychological PerspecƟves on
Suicidal Behaviour
RÊÙù C. O’CÊÄÄÊÙ

I can’t stop myself thinking, I wish I could turn off, I hate myself, I’m just
not good enough, I am tired of life , I’ve had enough—declares a young
man, aged 19 years, who took his own life (O’Connor, unpublished).
It is generally accepted that suicide is the outcome of a complex inter-
play of aetiological factors which are psychological, biological and social
in origin (e.g., Mann et al., 2005). Indeed, in recent years there has been
a growth in biopsychosocial models including the diathesis-stress model
of suicidal behaviour (e.g., Mann et al., 1999). Exponents of diathesis-
stress perspectives argue that the risk of suicide is determined by the
interaction of predisposing vulnerabilities and the experience of stress
(e.g., Joiner and Rudd, 1995; O’Connor and O’Connor, 2003; Schotte and
Clum, 1987). These vulnerabilities take many forms; they can be biological
(e.g., increased activity of hypothalamic-pituitary-adrenal [HPA] axis,
Mann and Currier, 2007), cognitive (e.g., reduced social problem-solving
capacity, Williams, Barnhofer, Crane and Beck, 2005) or personality/
individual differences factors (e.g., perfectionism, O’Connor et al., 2007).
For the purposes of the present chapter we will focus on some of the
psychological factors and describe how three of the predominant psy-
chological models enhance our understanding of the aetiology and course
of suicidal behaviour.

3
Rory C. O’Connor

AN OVERARCHING BIOPSYCHOSOCIAL MODEL OF


SUICIDAL BEHAVIOUR

In a recent conceptualisation of the role of psychological factors in the


aetiology of mental health problems, Kinderman (2005) argues against a
simple biological reductionist approach. His conceptualisation calls for
greater consideration to be given to the role of psychological processes
in the aetiology and treatment of mental disorders. Within suicidology,
there has also been increased recognition that we need to move beyond the
classic psychiatric diagnostic categories if we are to further understand the
aetiology of suicidal behaviour (van Heeringen, 2001). Therein Kinderman
(2005) restates Engel’s (1980) original formulation that the biological,
psychological and social perspectives which comprise the biopsychosocial
model are equal partners in the aetiology of mental health problems. Indeed,
he argues cogently that the disruption or dysfunction of psychological
processes is a final common pathway in the development of a disorder. He
proposes that biological and social factors together with the individual’s
life experiences (e.g., negative life events) lead to mental health problems
not directly, rather indirectly, to the extent that they disrupt psychological
processes, and that it is this disruption or dysfunction of psychological
processes which leads to the development of a mental disorder. For
example, Kinderman would argue that although there are key biological
substrates involved in depression, it is how these substrates (e.g., low
levels of serotonin) impact on appraisal mechanisms and information
processing which is crucial to the development of depression. From a
cognitive perspective, it is the extent to which biological factors change
your appraisal of yourself, the world and the future which is central to
understanding the aetiology of mental health problems.
For the present purposes, I have extended Kinderman’s (2005) model to
explain suicidal behaviour (Figure 1.1). Consistent with Kinderman, I see
psychological processes as central to understanding this phenomenon.
However, I propose two self-destructive pathways. The first leads from
psychological processes through mental health disorder to suicidal
behaviour. This pathway primarily comprises those individuals who
become depressed as a result of disturbed psychological processes char-
acterised by, for example, low self-worth, a sense of entrapment and
no rescue. Other factors, which include access to the means of suicide,

4
Psychological Perspectives on Suicidal Behaviour

Figure 1.1 An Overarching Biopsychosocial Model of Suicidal Behaviour

Source: Adapted and extended by the author from Kinderman (2005).

cognitive constriction, modelling effects and so on, will determine whether


the depressed individual engages in suicidal behaviour. The second pathway
is direct, leading from psychological processes to suicidal behaviour and
accounts for those suicidal individuals who are not suffering from mental
disorder. An impulsive young man who experienced a sudden loss of
self-esteem following the loss of his job and subsequently became suicidal
would be an example of this second pathway.
Studies to understand the psychological processes which lead indi-
viduals to take their own lives have grown considerably in recent years.
These processes are the focus of the remainder of this chapter. However,
any focus on psychological processes would be incomplete without due
consideration being given to the key psychological models which have
been proffered to explain and treat suicidality. Although Kinderman’s
(2005) model acts as a good overarching framework, which parsimoniously
describes the relationship between biology, psychology and social factors,
a number of other explicitly psychological models have been effective in
describing the how, the why and the when of suicidal risk.

5
Rory C. O’Connor

A description of all such models is beyond the scope of the present


chapter. However, the predominant ones include escape theory (Baumeister,
1990), the cry of pain model (Williams, 2001), the differential activation
model of cognitive reactivity (Williams et al., 2008) and the interpersonal-
psychological model of suicidal behaviour (Joiner, 2005). As escape is a
commonality in many of these models of suicide, I direct our focus on
it now:

ESCAPE AS A CENTRAL SUICIDAL MOTIVE

Escape has long been recognised as a central driving force underpinning


suicidal motives (Shneidman, 1985). Needless to say, escape-motivated
suicides only represent one category of suicides and do not include
ritual/altruistic suicides. Indeed, Baumeister (1990) and others (including
Leenaars, 2004 and Williams, 2001) have developed models and thera-
peutic interventions of suicidal behaviour with escape as a key component.
In his seminal paper, Roy Baumeister (1990) extended the clinical and
sociological models by proposing a psychological model of suicide derived
from social and personality psychology. His main argument was that
suicide is the endpoint of a complex series of causal steps which are driven
by an attempt to escape from painful self-awareness (Table 1.1).

Table 1.1 A summary of Baumeister’s (1990) Suicide as Escape from Self Model

Step Description of step and characteristics


Step 1 During stressful times we fall short of our expectations and standards
resulting from either unrealistic expectations or major setbacks.
Step 2 We attribute the blame (for this shortfall) internally; this leads to blame
and negative implications of the self.
Step 3 These negative self-implications lead to elevated and unpleasant negative
self-awareness such as inadequacy, incompetence or guilt.
Step 4 The negative self-awareness leads to negative affect.
Step 5 To escape this painful self-awareness we engage in cognitive deconstruction
(the removal of meaning from awareness). If deconstruction is not
effective, other means of terminating thoughts/feelings are required.
Step 6 Because we are disinhibited, we see suicide as more acceptable as the
moral/social ‘norms’ (internal barriers to action) are removed.
Source: Compiled by the author from Baumeister (1990).

6
Psychological Perspectives on Suicidal Behaviour

In short, Baumeister’s (1990) model suggests that a negative experi-


ence (or change in severity of existing or chronic negative experiences) is the
catalyst for self and social comparison processes whereby we try to make
sense of why something happened or why circumstances have changed.
If the resultant negative attributions are internal, they lead to blame and
negative self-implications, which in turn lead to negative self-awareness
(feelings of inadequacy and guilt) and negative affect. In an attempt to
escape these painful cognitions and feelings we engage in cognitive decon-
struction, which has been described as lower level awareness/thinking that
results in ‘the removal of higher meanings from awareness’ (Baumeister,
1990: 92). In addition to numbing aversive thoughts, cognitive decon-
struction is also thought to increase disinhibitions such that the social or
moral norms (internal barriers to action) are also attenuated or removed
rendering suicide more likely. A range of other factors, including access to
means, will influence whether suicide is subsequently enacted. A number
of studies with clinical and non-clinical samples have yielded evidence
consistent with escape theory (Dean and Range, 1999; Dean et al., 1996;
Flamenbaum and Holden, 2007; O’Connor and O’Connor, 2003; Tassava
and Ruderman, 1999).

ENTRAPMENT AND THE CRY OF PAIN

Although escape theory is a useful explanatory framework, recent re-


search has extended this model to improve the prediction of suicidal
behaviour—to better identify the circumstances when escape is especially
pernicious. Drawing from the animal behaviour literature and the idea of
‘arrested flight’ Gilbert (2006), Gilbert and Allan (1998), MacLean (1990)
and Williams and colleagues (Williams, 2001; Williams and Pollock, 2000;
Williams, Crane, Barnhofer and Duggan, 2005) have put forward the cry
of pain (CoP) or entrapment model of suicide (Figure 1.2). In essence,
Williams argues that stressful experiences which result in feelings of defeat
and loss can be suicidogenic but that these appraisals are particularly toxic
when we cannot escape from the defeating situation. Escape is determined,
in part, by failure in social problem-solving. In turn, it is this state of
entrapment combined with no opportunities for rescue which Williams
posits to be associated with elevated risk of suicide via the activation of

7
Rory C. O’Connor

Figure 1.2 Cry of Pain Model

Source: Adapted from Williams and Pollock (2001).

the learned helplessness script. Put simply, the latter is the realisation
that there is no relationship between individual action and outcome; in
other words, no matter what I do, I cannot change myself, my future or
my circumstances.
Similar to escape theory, whether we engage in suicidal behaviour is
determined by a number of additional factors including whether we are
modelling others’ behaviour or we have access to the means of suicide.
In short, therefore, Williams and Pollock (2001) proposed that suicidal
behaviour is reactive, the response (‘the cry’) to a situation that has three
components: defeat, no escape and no rescue. The CoP model is garnering
empirical and conceptual support and is attractive not only because it is
parsimonious and intuitive but also because it suggests specific, testable,
moderating and mediating pathways/hypotheses. To this end, we have
completed two clinical case-control type studies which have yielded em-
pirical support for the cry of pain model, specifically demonstrating the
power of the CoP variables in discriminating between self-harm patients

8
Psychological Perspectives on Suicidal Behaviour

and controls, and between first-time and repeat self-harm patients


(O’Connor, 2003; Rasmussen et al., in press).
In addition, these latter studies have found direct evidence for the
hypothesised mediating (defeat–entrapment–suicidality relationship) and
moderating pathways (escape/entrapment–low rescue factors–suicidality
relationship) specified in the CoP model. In the present context, mediating
pathways are particularly important theoretically and clinically as they help
to identify the mechanisms by which psychological factors are translated
into suicide risk. Moderating relationships, on the other hand, specify the
conditions under which suicide risk is high (e.g., when rescue factors are
low) or low. The Rasmussen et al. (in press) study also extended O’Connor
(2003) by employing pure measures of defeat and entrapment to test the
CoP model. The CoP has also received experimental attention (Johnson
et al., 2008) and conceptual consideration in the context of understanding
suicidality in psychosis (Bolton et al., 2007). It is also worth noting that
irrespective of the level of explanation, the CoP is a diathesis-stress model
of suicidality.

‘CRIES OF PAIN’: SOME PERSONAL


COMMUNICATIONS

The CoP constructs are not ethereal psychological constructs; they


are present in the spoken communications of those who attempt sui-
cide. What follows are selected quotations from suicidal patients within
24 hours of a suicidal episode who took part in one of our recent studies
(O’Connor, 2003). They illuminate the human tragedy behind each
suicidal episode. The first example is taken from a 20-year-old woman
who had an overdose history. Her words illustrate the spiral of despair
and ‘overdose-as-problem-solving’ motivation which characterises many
suicidal individuals:

I take them [tablets] just to block everything off but once it’s finished (the
suicidal episode) it’s just there again, so [it’s] a vicious circle, take them again
to take the problems away…take them to take the depression away but because
[I] am taking them [I] get more and more depressed. (O’Connor, Unpublished)

9
Rory C. O’Connor

Defeat, failure and entrapment were communicated explicitly by many


of the suicidal patients. One of the participants, a school pupil, was having
difficulties at school. The last line of her communication is important; the
young person is not able to do what she wanted, she feels trapped:

I was stressed, couldn’t cope with all at school, too much work, too much
pressure…Got results back and [I] failed most of them, just felt depressed. Not
allowed to just sit three of them. (O’Connor, Unpublished)

A 20-year-old unemployed woman reports feeling hopeless and trapped


with no prospect of rescue:

On the day before the overdose I’d been to the GP [doctor] for tests. That night
[I] had nightmares…Feeling fairly down yesterday, hopeless, felt trapped, very
alone, scared of being and feeling alone. Just so much going on I feel I can’t
cope with it all. (O’Connor, Unpublished)

Finally, escape as a motivation for a suicidal episode was succinctly


described by a 38-year-old married woman:

I just wanted to escape. [I] didn’t want to see him. [He] makes me feel
depressed.’ (O’Connor, Unpublished)

Although there is growing qualitative and quantitative evidence for the


CoP constructs, the evidence base requires expansion. It is not yet clear
to what extent these variables predict suicidal behaviour over time, and
to what extent these constructs are modifiable within high risk groups.
Perhaps unsurprisingly, escape/entrapment and defeat are overlapping
constructs (as feelings of defeat may also characterise thoughts of
entrapment). Therefore, it would be helpful to further refine these two
constructs to better specify the distinct active components of defeat and
entrapment, respectively.
A key strength of the CoP model is that it allows for two levels of ex-
planation of suicide risk. At a macro-level (first order), it posits three sets
of judgements concerning defeat, escape and rescue which account for
suicide risk. As noted earlier, these constructs are useful, they seem to be
better proximal discriminators of self-harm than clinical variables and
past behaviour (O’Connor, 2003). In addition, work by Williams and
colleagues have shown that these feelings of defeat and entrapment can be

10
Psychological Perspectives on Suicidal Behaviour

reactivated by negative mood (e.g., Williams et al., 2008). Drawing from


the differential activation literature (e.g., Goldstein and Willner, 2002),
Williams suggests that patterns of cognition and behaviour (including
suicidal ideation and behaviour) become associated when, say, mood is
low and these associations become more closely connected if there is a re-
occurrence of the lower mood. Consequently, in the future, small changes
in mood can give rise to these associations and if suicidal ideation was part
of the network of associations when mood was low, it is likely to emerge
again during a depressive episode. It is the ease with which these changes
in mood reactivate the defeating, entrapping and suicidal thoughts which
is crucial to risk assessment.
The CoP model also suggests a more fine-grained level of explan-
ation (second order), where biases in information processing, memory
deficits, and other individual differences and vulnerabilities affect the
decisions concerning defeat, escape and rescue. This level of explanation
complements the macro-level model by identifying specific fundamental
processes (e.g., future thinking and interpersonal problem-solving) and
individual characteristics (e.g., perfectionism) which should be targeted
for intervention to reduce suicide risk.
A large number of psychological variables have been implicated in
the suicidal process (see Table 1.2 for a list of such variables). However,
as space is limited, I focus our attention on two factors which we have
examined explicitly within the CoP framework: positive future thinking
and socially prescribed perfectionism.

Table 1.2 Psychological Risk and ProtecƟve Factors Associated with Suicidal Risk

Autobiographical memory biases Hopelessness


Interpersonal problem-solving Future thinking
Cognitive rigidity Perfectionism
Impulsivity Self-criticism
Group/social influences Rumination
Goal re-engagement Self-esteem
Attributional/cognitive style Coping
Resilience Optimism/pessimism
Attitudes Attentional biases
Neuroticism Thought suppression
Shame Hardiness
Source: Compiled by the author.

11
Rory C. O’Connor

THOUGHTS OF THE FUTURE AND SUICIDE RISK

In the 1990s, MacLeod and colleagues (1993) were keen to determine


whether different operationalisations of hopelessness (a key proximal
predictor of suicidal behaviour) were functionally equivalent or dif-
ferentially associated with suicide risk. In a series of studies (MacLeod et al.,
1993, 1997, 1998), they determined that suicidal participants consistently
reported significantly fewer positive future thoughts compared to controls
but showed no difference in negative future expectations.
Positive future thinking is usually assessed via the Future Thinking Task
(FTT; MacLeod et al., 1997) which asks participants what they are worried
about (negative future thoughts) and looking forward to (positive future
thoughts) over different future time frames—it generates idiographic, spe-
cific positive future expectancies. In short, future expectations of a posi-
tive valence are particularly associated with suicide risk (MacLeod et al.,
1997) and this finding has been replicated independently (Hunter and
O’Connor, 2003; O’Connor et al., 2000, 2004, 2008; Williams et al., 2008).
This deficit in positive cognitions is even more striking as it cannot be
explained in terms of higher levels of depression among suicidal patients
(e.g., MacLeod et al., 1997) nor can the effect be explained by negative
cognitive style (O’Connor et al., 2000). Moreover, we recently showed
that specific positive future expectancies (as assessed via FTT) were better
predictors of suicidal ideation two and a half months following a suicide
attempt than global hopelessness (measured via the Beck Hopelessness
Scale; Beck et al., 1974; O’Connor et al., 2008).
In terms of the CoP model, the deficit in positive future thinking fits
well: if one finds oneself in a state of entrapment, the model suggests that
the pathway from entrapment to suicidal behaviour is strengthened or
attenuated by the presence or absence of rescue factors. In this case, rescue
factors are operationalised as the relative paucity in positive future think-
ing. Therefore, the CoP posits a strengthening of the entrapment–suicidal
behaviour relationship, and this is what we have found previously (see
Rasmussen et al., in press, for empirical support for this path). Moreover, as
positive thinking is thought to be associated with resilience (Ciarrochi et al.,
2007), the negative impact of its absence on well-being is understandable.
It is encouraging to note that in a pilot Randomised Clinical Trial (RCT)
(brief manual-assisted cognitive behaviour therapy versus treatment as
usual), self-harm patients in the active arm of the trial showed a significant

12
Psychological Perspectives on Suicidal Behaviour

improvement in positive future thinking over the course of the follow-


up period (MacLeod et al., 1998). However, interpretation of the findings
is problematic as the control group also improved their positive future
thinking. Needless to say, future research should be directed at such
interventions.

SOCIALLY PRESCRIBED PERFECTIONISM


AND SUICIDE RISK

Given the importance of impaired positive future thinking in the suicidal


process, it is important to investigate individual differences/personality
factors which may be implicated in the development of this impairment.
One such variable which has been of interest in this regard is socially
prescribed perfectionism (SPP). Socially prescribed perfectionism—a
recognised personality trait—taps beliefs about the excessive expectations
we perceive that our significant others have of us (Hewitt and Flett, 1991)
and it is independently associated with suicide risk (Hewitt et al., 1997,
1998; O’Connor and O’Connor, 2003).
Socially prescribed perfectionism was high up the list of candidate per-
sonality variables given that it is consistent with escape theory and the
CoP model. Indeed, the first step of the causal chain to suicide in escape
theory implicates perfectionistic comparison processes: in stressful times
we fall short of our expectations and standards. Similarly, in the CoP, the
process by which one appraises stressful situations as leading to defeat
and/or entrapment suggests perfectionistic beliefs. It is important to
highlight, though, that the perfectionism literature consistently points to
the specific effects of perceived excessive/high expectations ‘set by others’
(i.e., SPP) as being more pernicious than those set by oneself (O’Connor,
2007; Shafran and Mansell, 2001), hence the focus on SPP here. We believe
that the potency of SPP is due, in part, to the fact that the ‘perceived’
excessive standards set by others are outside one’s personal control, and
loss of control has long been associated with poor psychological adjustment
(O’Connor and Sheehy, 2000). It would also be theoretically and clinically
useful to determine whether SPP increases the ease with which CoP
cognitions are reactivated.

13
Rory C. O’Connor

In a series of studies we have shown not only that SPP is associated


with positive future thinking (e.g., Hunter and O’Connor, 2003) but
that positive future thinking moderates the relationship between SPP
and suicide risk in clinical (O’Connor et al., 2007) and non-clinical
populations (O’Connor et al., 2004). In the latter study, 206 healthy par-
ticipants completed the future thinking task contemporaneously with a
measure of SPP and hopelessness. Although SPP was independently
associated with hopelessness, the relationship was significantly stronger
when considered concomitantly with positive future thinking. In other
words, those individuals with high SPP and low levels of positive future
thinking reported significantly higher levels of hopelessness—a proximal
predictor of suicidal behaviour—than those high on positive future
thinking. In the former study (O’Connor et al., 2007), we followed up 126
suicide attempters two months following a suicide attempt and found that
outcome (i.e., suicidal ideation and hopelessness) was better for those high
on positive future thoughts and low on perfectionism (Figure 1.3). Our
interpretation of these data is that higher levels of positive future thinking
result in (or are akin to) more reasons for living, thereby reducing the sense
of entrapment and suicide risk (O’Connor et al., 2007). Future research
should experimentally test whether higher levels of positive future thinking
decrease perceptions of entrapment.

CLINICAL IMPLICATIONS

There are a number of clinical implications from the research summarised


here. First, the overarching biopsychosocial model recognises the complex
interplay between biology, psychology and social factors in suicidal aetiology
but it proposes that psychological processes are a common pathway to
mental disorder and suicidal behaviour. Whereas the biopsychosocial
model is a generic ‘black box’ model of suicide risk, the escape theory, CoP,
and the differential activation model offer specific details of mechanisms
and processes. Second, it provides a framework on which to develop
treatment interventions and it highlights a number of psychological
processes which ought to be considered when a clinician is assessing suicide
risk. Taking the CoP and the differential activation models together, it is
clear that thoughts of defeat and entrapment are central to the aetiology of
suicidal behaviour but that their impact may be latent until these thoughts

14
Figure 1.3 PosiƟve Future Thinking as a Moderator of the SPP–Distress RelaƟonship

15
Psychological Perspectives on Suicidal Behaviour

Source: O’Connor et al. (2007).


Rory C. O’Connor

are reactivated by a negative mood or the occurrence of a negative life event


and escalate into a suicidal episode over time. So it is important to look at the
conjoint effects of these latent characteristics (e.g., entrapment), trait
factors (including perfectionism) which confer psychological vulnerability,
rescue factors which can act as buffers of psychosocial stressors, as well as
the occurrence of negative life events.
Third, the work on positive future thinking and social perfectionism
highlights specific cognitions and individual characteristics which may
influence risk. Further, it highlights precise cognitive-behavioural mech-
anisms (e.g., entrapment–positive future thinking–suicidal behaviour)
which could be targeted directly through therapy. Specific positive
cognitions embedded in different future-oriented time frames could also
be modified to reduce risk (O’Connor et al., 2008). Indeed, the assessment
of positive future thoughts could be incorporated into treatment protocols,
with their increase/decrease serving as an indicator of risk. A similar
approach with reasons for living has been used successfully by Jobes in his
Collaborative Assessment and Management of Suicidality programme
(CAMS; Jobes, 2006). Our data also add to the growing body of evidence
highlighting the importance of positive cognitions to adaptive self-
regulation and adjustment (e.g., Fredrickson and Losada, 2005).
Fourth, we need to understand further the means by which the presence
of positive future thoughts may buffer against suicide risk. Do they, as
I suspect, reduce entrapment and/or do they impact on our defeat and
rejection appraisals? Finally, relatively little attention has been directed
at conducting and evaluating evidence-based interventions to modify
perfectionism. Consequently, we need to tease out the mechanisms which
are involved in the development and maintenance of perfectionism. In
short, within suicidology, longitudinal and experimental research is urgently
required to extend the theoretical evidence base which will hopefully
inform the development of effective psychological interventions.

REFERENCES

Baumeister, R.F. (1990). Suicide as escape from self. Psychological Review, 97(1), 90–113.
Beck, A.T., A. Weissman, D. Lester and L. Trexler (1974). The measurement of
pessimism: The hopelessness scale. Journal of Consulting and Clinical Psychology,
42(6), 861–65.

16
Psychological Perspectives on Suicidal Behaviour

Bolton, C., P. Gooding, N. Kapur, C. Barrowclough and N. Tarrier (2007). Developing


psychological perspectives of suicidal behaviour and risk in people with a diagnosis
of schizophrenia: We know they kill themselves but do we understand why? Clinical
Psychology Review, 27(4), 511–36.
Ciarrochi, J., P.C.L. Heaven and F. Davies (2007). The impact of hope, self-esteem
and attributional style on adolescents’ school grades and emotional well-being:
A longitudinal study. Journal of Research in Personality, 41(6), 1161–78.
Dean, P.J. and L.M. Range (1999). Testing the escape theory of suicide in an outpatient
clinical population. Cognitive Therapy and Research, 23(6), 561–72.
Dean, P.J., L.M. Range and W.C. Goggin (1996). The escape theory of suicide in college
students: Testing a model that includes perfectionism. Suicide and Life-Threatening
Behavior, 26(2), 181–86.
Engel, G.L. (1980). The clinical application of the biopsychosocial model. American
Journal of Psychiatry, 137(5), 535–44.
Flamenbaum, R. and R.R. Holden (2007). Psychache as a mediator in the relationship
between perfectionism and suicidality. Journal of Counseling Psychology, 54(1),
51–61.
Fredrickson, B.L. and M.F. Losada (2005). Positive affect and the complex dynamics
of human flourishing. American Psychologist, 60(7), 678–86.
Gilbert, P. (2006). Evolution and depression: Issues and implications. Psychological
Medicine, 36(3), 287–97.
Gilbert, P. and S. Allan (1998). The role of defeat and entrapment (arrested flight)
in depression: An exploration of an evolutionary view. Psychological Medicine,
28(3), 585–98.
Goldstein, R.C. and P. Willner (2002). Self-report measures of defeat and entrapment
during a brief depressive mood induction. Cognition and Emotion, 16(5),
629–42.
Hewitt, P.L. and G.L. Flett (1991). Perfectionism in the self and social contexts: Con-
ceptualization, assessment, and association with psychopathology. Journal of
Personality and Social Psychology, 60(3), 456–70.
Hewitt, P.L., J. Newton, G.L. Flett and L. Callander (1997). Perfectionism and suicide
ideation in adolescent psychiatric populations. Journal of Abnormal Child Psych-
ology, 25(2), 95–101.
Hewitt, P.L., G.R. Norton, G.L. Flett, L. Callander and T. Cowan (1998). Dimensions of
perfectionism hopelessness, and attempted suicide in a sample of alcoholics. Suicide
and Life-Threatening Behavior, 28, 395–406.
Hunter, E.C. and R.C. O’Connor (2003). Hopelessness and future thinking in para-
suicide: the role of perfectionism. British Journal of Clinical Psychology, 42(4),
355–65.
Jobes, D.A. (2006). Managing Suicide Risk: A Collaborative Approach. New York:
Guilford Press.
Johnson, J., N. Tarrier and P. Gooding (2008). An investigation of aspects of the cry
of pain model of suicide risk: The role of defeat in impairing memory. Behaviour
Research and Therapy, 46(8), 968–75.

17
Rory C. O’Connor

Joiner, T. (2005). Why people die by suicide. Massachusetts, US: Harvard University Press.
Joiner, T.E. and M.D. Rudd (1995). Negative attributional style for interpersonal events
and the occurrence of severe interpersonal disruptions as predictors of self-reported
ideation. Suicide and Life-Threatening Behavior, 25(2), 297–304.
Kinderman, P. (2005). A psychological model of mental disorder. Harvard Review of
Psychiatry, 13(4), 206–17.
Leenaars, A.A. (2004). Psychotherapy with suicidal people. Chichester: John Wiley & Sons.
MacLean, P.D. (1990). The Triune brain in evolution. New York: Plenum Press.
MacLeod, A.K., B. Pankhania, M. Lee and D. Mitchell (1997). Parasuicide, depression
and anticipation of positive and negative future experiences. Psychological Medicine,
27(4), 973–77.
MacLeod, A.K., G.S. Rose and J.M.G. Williams (1993). Components of hopelessness
about the future in parasuicide. Cognitive Therapy and Research, 17(5), 441–55.
MacLeod, A.K., P. Tata, K. Evans, P. Tyrer, U. Schmidt, K. Davidson, et al. (1998). Recovery
of positive future thinking within a high-risk parasuicide group: Results from
a pilot randomized controlled trial. British Journal of Clinical Psychology, 37(5),
371–79.
Mann, J.J. and D. Currier (2007). A review of prospective studies of biologic predictors
of suicidal behavior in mood disorders. Archives of Suicide Research, 11(1), 3–16.
Mann, J.J., A. Apter, J. Bertolote, A. Beautrais, D. Currier, A. Haas et al. (2005). Suicide
prevention strategies: A systematic review. Journal of American Medical Association,
294(16), 2064–74.
Mann, J.J., C. Waternaux, G.L. Haas and K.M. Malone (1999). Toward a clinical model
of suicidal behavior in psychiatric patients. American Journal of Psychiatry, 156(2),
181–89.
O’Connor, R. C. and N.P. Sheehy (2000). Understanding suicidal behaviour. Chichester:
Wiley Blackwell.
O’Connor, R.C. (2003). Suicidal behaviour as a cry of pain: Test of a psychological
model. Archives of Suicide Research, 7(4), 297–308.
O’Connor, R.C. (2007). The relations between perfectionism and suicidality:
A systematic review. Suicide and Life-Threatening Behavior, 37(6), 698–714.
O’Connor, R.C. (Unpublished). General hospital self-harm: Motives and reasons.
O’Connor, R.C. and D.B. O’Connor (2003). Predicting hopelessness and psychological
distress: The role of perfectionism and coping. Journal of Counseling Psychology,
50(3), 362–72.
O’Connor, R.C., D.B. O’Connor, S.M. O’Connor, J. Smallwood and J. Miles (2004).
Hopelessness, stress and perfectionism: the moderating effects of future thinking.
Cognition and Emotion, 18(8), 1099–120.
O’Connor, R.C., H. Connery and W. Cheyne (2000). Hopelessness: The role of
depression, future directed thinking and cognitive vulnerability. Psychology, Health
and Medicine, 5, 155–61.
O’Connor, R.C., L. Fraser, M.C. Whyte, S. MacHale and G.Masterton (2008). A com-
parison of specific positive future expectancies and global hopelessness as predictors

18
Psychological Perspectives on Suicidal Behaviour

of suicidal ideation in a prospective study of repeat self-harmers. Journal of Affective


Disorders, 110(3), 207–14.
O’Connor, R.C., L. Fraser, M.C. Whyte, S. MacHale and G. Masterton (in press). Self-
regulation of unattainable goals in suicide attempters: the relationship between
goal disengagement, goal reengagement and suicidal ideation. Behaviour Research
and Therapy.
O’Connor, R.C., M.C. Whyte, L. Fraser, G. Masterton, J. Miles and S. MacHale (2007).
Predicting short-term outcome in well-being following suicidal behaviour: The
conjoint effects of social perfectionism and positive future thinking. Behaviour
Research and Therapy, 45(7), 1543–55.
Rasmussen, S., L. Fraser, M. Gotz, S. MacHale, R. Mackie, G. Masterton, S. McConachie
and R.C. O’Connor (in press). Elaborating the Cry of Pain model of suicidality:
Testing a psychological model in a sample of first-time and repeat self-harm
patients. British Journal of Clinical Psychology.
Schotte, D. and G. Clum (1987). Problem-solving skills in suicidal psychiatric patients.
Journal of Consulting and Clinical Psychology, 55(1), 49–54.
Shafran, R. and W. Mansell (2001). Perfectionism and psychopathology: A review of
research and treatment. Clinical Psychology Review, 21(6), 879–906.
Shneidman, E.S. (1985). Definition of suicide. New York: John Wiley & Sons.
Tassava, S.H. and A.J. Ruderman (1999). Application of escape theory to binge eating
and suicidality in college women. Journal of Social and Clinical Psychology, 18(4),
450–66.
van Heeringen, K. (2001). Towards a psychobiological model of the suicidal process. In
K. van Heeringen (ed.) Understanding Suicidal Behaviour (pp. 136–59). Chichester:
John Wiley & Sons.
Williams, J.M.G. (2001). The cry of pain. London: Penguin.
Williams, J.M.G. and L.R. Pollock (2000). The psychology of suicidal behaviour. In
K. Hawton and K. van Heeringen (Eds), The international Handbook of Suicide
and Attempted Suicide (pp. 79–93). Chichester: John Wiley & Sons.
Williams, J.M.G. and L.R. Pollock (2001). Psychological aspects of the suicidal
process. In K. van Heeringen (Ed.), Understanding Suicidal Behaviour (pp. 76–93).
Chichester: John Wiley & Sons.
Williams, J.M.G., A.J.W. van der Does, T. Barnhofer, C. Crane and Z.S. Segal (2008).
Cognitive reactivity, suicidal ideation and future fluency: Preliminary investigation
of a differential activation theory of hopelessness/suicidality. Cognitive Therapy
and Research, 32(1), 83–104.
Williams, J.M.G., C. Crane, T. Barnhofer and D. Duggan (2005). Psychology and suicidal
behaviour: Elaborating the entrapment model. In K. Hawton (Ed.) Prevention and
Treatment of Suicidal Behaviour: From Science to Practice (pp. 71–90). Oxford:
Oxford University Press.
Williams, J.M.G., T. Barnhofer, C. Crane and A.T. Beck (2005). Problem solving
deteriorates following mood challenge in formerly depressed patients with a history
of suicidal ideation. Journal of Abnormal Psychology, 114(3), 421–31.

19
2

Empirically Based Assessment of


Suicide Risk*
C AD E. M ,C A J. B A A D
KA KA A

I t is almost a certainty that mental health professionals will, at some point,


be required to evaluate a patient having some form of suicidality (i.e.,
suicidal thoughts, following a suicide attempt or someone with a history
of multiple attempts). Suicidality is the most frequently encountered
emergency situation in mental health settings (Buzan and Weissberg,
1992) and is the most anxiety-provoking clinical scenario for practitioners
(Pope and Tabachnick, 1993; Rudd, 2006). Approximately one-quarter
of all psychologists will experience suicide by a patient at some point in
their careers (Chemtob et al., 1988a; Pope and Tabachnick, 1993), as will
nearly 50 percent of psychiatrists (Chemtob et al., 1988b) and 23 percent
of counsellors (McAdams and Foster, 2000). Fawcett (1999) has estimated
that up to half of the suicides are by individuals ‘currently in treatment’.
The impact of suicidal behaviour, emotionally and professionally,
on those providing clinical care is profound, with clinicians reporting
shock, self-blame, guilt and shame (Kleespies et al., 1993), which is often

∗ The views expressed in this chapter are those of the authors and do not necessarily reflect
the official policy or position of the Department of Defence, the Department of the Air
Force or the US government.

20
Empirically Based Assessment of Suicide Risk

compounded by the legal expectation that clinicians can predict a patient’s


suicidal behaviour (i.e., the legal concept of ‘foreseeability’). The notion
of suicide prediction is problematic, however, given that it is not possible
to reasonably predict such a low base-rate phenomenon as suicide. In fact,
because death by suicide occurs so infrequently, a clinician would actu-
ally be correct more often if he or she predicted that a patient would not
die by committing suicide, regardless of clinical presentation. Despite
the inherent limitations in predicting suicidal behaviour reliably, this
legal expectation to predict patients’ actions has nonetheless influenced
existing standards of care.
The inability to reliably predict suicidal behaviour does not suggest,
however, that important risk factors associated with increased risk for
suicide have not been identified through research. It is critical for the
clinician to recognise that their task is not to ‘predict’ suicide per se, but
rather to recognise when a patient has entered into a heightened state of
risk (i.e., risk assessment), and to respond appropriately. When conducted
competently, risk assessment both estimates and explains the risk of
suicidal behaviour when used in a consistent manner across all patients,
and provides a template for clinical crisis management, including short-
and long-term treatment targets (Bryan and Rudd, 2006). The recent
publication of core competencies in the assessment and management of
suicide risk (Suicide Prevention Resource Center [SPRC], 2006), and the
American Psychiatric Association’s (APA, 2003) practice guidelines have
clear implications for the nature and process of clinical care. It is important
for clinicians to be familiar with available standards and empirically
supported approaches for managing suicidal risk. Suicide assessment is
a core clinical competency that outpatient mental health clinicians must
have. Unfortunately, less than half of mental health professionals receive
formal training in suicide risk assessment (Bongar and Harmatz, 1991;
Burstein et al., 1973; Feldman and Freedenthal, 2006; Guy et al., 1990),
although suicidality is a clinical scenario frequently, if not uniformly,
encountered in most clinical settings.
In this chapter, suicide risk assessment will be approached from a
clinically balanced and scientifically informed standpoint, translating em-
pirical research into clinical practice. Our risk assessment model is based
on the SPRC’s (2006) core competencies for managing suicidality, and
provides guidelines for assessing suicidal symptoms, directing clinical
decision-making, and adopting a best-practices perspective. We have

21
Chad E. Morrow et al.

organised this discussion into three primary areas: understanding suicide,


managing emotional reactions to suicide and conducting the suicide risk
assessment.

AREA I: UNDERSTANDING SUICIDE

A core area for clinician competency is basic knowledge of suicidal be-


haviours, which includes familiarity with various terms related to suicide,
and statistics and facts about suicide. In particular, it is important that
clinicians are able to effectively differentiate between various suicide-
related behaviours (e.g., suicide threats, self-harm, suicide attempts with
and without injuries), and to articulate an understandable biopsychosocial
model of suicidality that can be related in simple and straightforward terms
to patients, as well as lend itself to clear and straightforward treatment
goals. Perhaps most important, though, is the need for clinicians to have a
solid grounding in, and understanding of, those risk and protective factors
that have demonstrated the strongest empirical associate with suicidal
behaviours. Competency in these areas will increase the effectiveness of
assessing and managing risk, improve communication between clinicians,
and increase the clinician’s ability to clearly document clinical care.

A Biopsychosocial Model of Suicidality

Inherent in any biopsychosocial model of suicide is the recognition of


the relative contributions of biological and genetic (e.g., family history,
inheritability of psychiatric disorders), psychological (e.g., mood states,
impaired problem-solving, hopelessness) and social (e.g., supportive
relationships, access to resources) factors. Unfortunately, the relative lack
of formal training in suicide assessment and intervention among mental
health professionals lends itself to considerable misunderstandings and
misconceptions about suicide, and a general inability to explain suicidal
behaviours to either patients or medical providers. From a clinical per-
spective, this is of grave concern since suicidal patients are often confused
and distressed about their suicidal experience, and look to the clinician to
understand what is happening to them. Clinicians who cannot succinctly

22
Empirically Based Assessment of Suicide Risk

explain suicidal behaviours to patients are unlikely to establish the col-


laborative alliance necessary to positively influence clinical outcomes. An
effective clinician must therefore have a clear and straightforward model
of suicidal behaviour.
One particularly simple and empirically supported biopsychosocial
model of suicidality is ‘fluid vulnerability theory’ (FVT; Rudd, 2006), which
posits that suicide risk exists on two dimensions: baseline risk and acute
risk. According to FVT, baseline risk for suicide varies from individual
to individual and is determined by static variables and personal history
(e.g., multiple attempts, psychiatric diagnoses, biological and genetic
predispositions, history of abuse). Multiple attempters (i.e., individuals
who have attempted suicide twice or more) are more vulnerable to sui-
cide and are therefore at chronically elevated risk. In the presence of an
environmental stressor, usually a perceived loss of some kind (e.g., relation-
ship problem, job loss, financial stress), an acute suicidal episode becomes
activated. The active suicidal episode—what Rudd (2006) has termed the
‘suicidal mode’—consists of cognitive, affective, behavioural, physiological
and motivational systems that interact with each other and sustain the
suicidal state (Figure 2.1). Also central to FVT is the tenant that suicide
risk is inherently dynamic, with fluctuations in intensity and severity
from moment to moment. Acute periods of suicidal crises are therefore
by their very nature time-limited, since most patients cannot maintain the
high level of activation and arousal needed to sustain a suicidal crisis for
more than a few hours or, at most, a few days (Rudd, 2006). A primary
task of the clinician’s risk assessment is to understand how these various
systems interact with each other to sustain the suicidal episode, and to
deliver interventions that serve to ‘deactivate’ the systems that maintain
the suicidal crisis, with the goal being to return the patient to ‘baseline’
functioning.

The Importance of Accurate Language

The language clinicians use when talking about suicidality not only en-
hances communication with other clinicians, but also with suicidal patients
themselves. The issue of inconsistent terminology and language relating to
suicidality has received considerable attention and discussion within the
professional literature, with several groups calling for the adoption of a

23
Chad E. Morrow et al.

Figure 2.1 Biopsychosocial Model of Suicide

Source: Author.

standardised terminology (e.g., O’Carroll et al., 1996; Silverman et al.,


2007). The advantages of using standard terminology include (a) improved
clarity, precision and consistency of a clinician’s practice both over time
and across patients; (b) improved consistency of communication between
clinicians; (c) improved clarity in documentation; (d) elimination of

24
Empirically Based Assessment of Suicide Risk

inaccurate and potentially damaging language from our vocabulary; and


(e) elimination of the unrealistic goal to ‘predict’ suicide (as opposed to
assessing risk) through recognition of the complexity and variability of
suicidal intent in determining ultimate clinical outcome. We, therefore,
recommend that clinicians implement the following terms proposed by
Silverman and colleagues (2007):

z Suicide threat: Any interpersonal action, verbal or non-verbal, without


a direct self-injurious component, that a reasonable person would
interpret as communicating or suggesting that suicidal behaviour
might occur in the near future.
z Suicide plan: A proposed method of carrying out a design that will
lead to a potentially self-injurious outcome.
z Self-harm: A self-inflicted, potentially injurious behaviour for which
there is evidence (either explicit or implicit) that the person did not
intend to kill himself/herself (i.e., had no intent to die). Self-harm
may result in no injuries, non-fatal injuries or death (i.e., self-inflicted
unintentional death).
z Suicide attempt: A self-inflicted, potentially injurious behaviour with
a non-fatal outcome for which there is evidence (either explicit or
implicit) of intent to die. Suicide attempts may result in no injuries or
non-fatal injuries. Suicide attempts that result in death are classified
as ‘suicide’.
z Suicide: A self-inflicted death for which there is evidence (either
explicit or implicit) of intent to die.
z Undetermined suicide-related behaviour: A self-inflicted, potentially
injurious behaviour for which there is an unknown or undetermined
degree of suicidal intent. Undetermined suicide-related behaviours
may result in no injuries, non-fatal injuries or death.

These terms are more thoroughly outlined and described by Silverman


and colleagues (2007), and the reader is encouraged to review these two
articles for additional information about the issue of standardised ter-
minology. We highly recommend clinicians be well-versed in these terms
to improve consistency in care over time, enhance communication with
clinicians and patients, and clarify documentation, especially in clinical
settings with multiple clinicians working together.

25
Chad E. Morrow et al.

AREA II: MANAGING EMOTIONAL REACTIONS


TO SUICIDAL INDIVIDUALS

The importance of establishing a strong relationship with the suicidal


patient cannot be overstated, since even the best therapeutic techniques
are unlikely to be adequately received and implemented by the patient
in the absence of a strong patient–clinician relationship. The literature
is replete with discussions of the central importance of the quality of the
patient–clinician relationship in assessing risk and managing suicidal
patients (e.g., Bongar et al., 1989). Some have even argued that a solid
therapeutic relationship is not just preferable, but rather ‘essential’ to suc-
cessful work with suicidal patients (Maltsberger, 1986; Shneidman, 1981,
1984). Perhaps the best-known work on the topic of relationship issues
in the treatment of suicidal patients is that of Linehan (1993), who has
identified several strategies for approaching the relationship with a sui-
cidal patient, including the targeting of what she calls ‘therapy interfering
behaviours’—problematic interpersonal emotional reactions and be-
haviours of the patient ‘or the clinician’, such as fear, malice, aversion, hate,
anxiety, worry, ending appointments prematurely, being late for ap-
pointments and rescheduling appointments frequently—which serve
to directly interfere with the successful course of treatment. Clinicians
must be alert to these behaviours both within their patients and within
themselves, and possess adequate skills to appropriately respond to them
in the context of the therapeutic relationship.
A relationship dynamic central to clinical work with suicidal patients is
the potential for a conflict between the goals of the patient and the goals
of the clinician. Specifically, the patient’s goal to reduce psychological
suffering through suicide can come into direct conflict with the clinician’s
goal to prevent death by suicide. This conflict must be resolved in order
for the clinician and patient to establish the working relationship ne-
cessary for clinical improvement. Resolution can be accomplished with
a straightforward and simple defining of a common goal: to reduce the
patient’s suffering and emotional pain. Consistent with a functional model
of suicide, as the patient’s pain and suffering resolves, it decreases their risk
for suicide. Defining a common goal of pain remediation therefore lays
the groundwork for the development of a non-adversarial, collaborative
therapeutic stance that facilitates establishing and maintaining a good

26
Empirically Based Assessment of Suicide Risk

working alliance with the patient. Within a collaborative stance, the patient
and clinician work together as a team to target the problem of suicide.
Adopting a collaborative stance in which responsibility for the suicidal
patient’s outcome is shared can help the clinician to manage common emo-
tional reactions to suicidal patients, including fear, anxiety or anger. These
emotional responses can cloud clinician’s judgement and contribute to
suboptimal clinical decision-making.
A hierarchical approach to questioning suicidal patients is therefore
recommended, in which the clinician moves from identifying the
precipitant of the suicidal crisis (e.g., ‘How have things been going for
you recently? Can you tell me about anything in particular that has
been stressful for you?’), to the patient’s symptomatic presentation (e.g.,
‘From what you have shared so far, it sounds like you have been feeling
depressed. Have you been feeling anxious, nervous or panicky lately?’), to
hopelessness (e.g., ‘It is not uncommon when depressed to feel that things
won’t improve and won’t get any better, do you ever feel this way?’), and
finally, to the nature of the patient’s suicidal thinking (e.g., ‘People feeling
depressed and hopeless sometimes think about death and dying; do you
ever have thoughts about death and dying? Have you ever thought about
killing yourself?’). By gradually progressing in the intensity of the interview,
clinicians can manage their own reactions to the suicidal patients while
potentially reducing the patients’ anxiety or agitation at the same time,
which improves rapport and strengthens the therapeutic relationship.
Likewise, by normalising the patient’s hopelessness and suicidal think-
ing within the context of a depressive episode (or other mental disorder),
the clinician can further reduce in-session anxiety, thereby enhancing the
likelihood of honest and more detailed self-disclosure on the part of the
patient, providing a more accurate risk assessment.

AREA III: CONDUCTING THE RISK ASSESSMENT

The following section will provide a brief overview of the most salient em-
pirical findings that directly affect suicide risk, and are therefore central
to accurate and effective risk assessment and management. The reader is
encouraged to review the APA’s (2003) guidelines for a more thorough
review of the scientific literature regarding assessment and management

27
Chad E. Morrow et al.

of suicidality. Several key areas have garnered considerable empirical


support, and are therefore considered to be essential to risk assessment:
predispositions to suicidal behaviour; identifiable precipitants or stressors;
symptomatic presentation; presence of hopelessness; nature of suicidal
thinking; previous suicidal behaviour; impulsivity and self-control; and
protective factors. These areas are listed in Table 2.1. Although this list
is certainly not exhaustive, and other areas could arguably be included,
the literature significantly supports these areas as having the most sound
empirical support, and therefore emerge as clinically meaningful and
critical to the assessment process.
As previously discussed, suicide risk should be considered from two
dimensions: baseline risk and acute risk. Baseline risk is the level of risk
when the patient is not in a state of acute crisis, or in general is at his or her

Table 2.1 Eight Empirically-supported Areas for Suicide Risk Assessments

Static Variables Contributing to Baseline Risk


1. Predispositions to suicide Family history of suicide; history of physical,
emotional or sexual abuse; marital status; previous
psychiatric diagnoses; same-sex orientation; recent
discharge from inpatient hospitalisation
2. Previous suicide attempts Frequency and context of previous suicidal
behaviours; perceived lethality and outcome;
opportunity for rescue and help-seeking; preparatory
or rehearsal behaviours; reaction to survival
3. Impulsivity Subjective reports of self-control (verbiage); objective
reports of self-control behaviours
Aggravating Variables Contributing to Acute Risk
4. Precipitants or stressors Significant loss; social isolation; relationship
problems; health problems; legal problems
5. Symptomatic presentation Depression; anxiety or panic; anger; agitation and
restlessness; psychosis
6. Hopelessness Severity and duration
7. Nature of suicidal thinking Frequency, intensity and duration of suicidal ideation;
presence and specificity of suicide plan; access to
means; preparatory or rehearsal behaviours; sense of
courage or fearlessness
8. Protective factors Reasons for living; presence of children in the home;
positive social support; intact reality-testing; problem-
solving; religious or moral beliefs against suicide
Source: Modified from Rudd (2006), Bryan and Rudd (2006).

28
Empirically Based Assessment of Suicide Risk

relative best. All suicidal individuals have a baseline risk that they return
to during periods of relative calm and remissions of psychopathology, but
baseline risk is not comparable across groups. However, for some patients
(such as multiple attempters), baseline risk level is high and indicates
chronic risk, regardless of any acute crisis. Acute risk, by contrast, is the
level of risk presented during an acute suicidal crisis, when the patient is
symptomatic and at his or her worst. Severity of risk is ‘always’ relative.
Accordingly, the variable nature of suicide risk—even among those at
chronic high risk—can be acknowledged by adding the descriptor
‘acute exacerbation’ when necessary (e.g., chronic high risk with acute
exacerbation).

Baseline Risk

Baseline risk is affected by predispositions to suicidality and historical


factors such as previous suicidal behaviours. These risk factors contribute
to an individual’s overall likelihood to experience or manifest symptoms
of suicidality. Because they are static by nature, they typically cannot be
directly modified through clinical intervention.

PredisposiƟons to suicidality Predispositions to suicidality include


genetic, biological and historical factors associated with increased suicide
risk. Compelling research points to the role of genetic factors in suicidal
behaviours. For example, twin studies have found that 13–19 percent of
monozygotic twin pairs were concordant for death by suicide as compared
to less than 1 percent of dizygotic twin pairs (Roy, 1992). Likewise, family
studies suggest that suicides ‘cluster’ in particular families above and be-
yond psychiatric conditions (Egeland and Sussex, 1985), which might be
due in part to specific chromosomal configurations that have been linked
to suicidal behaviours (Joiner et al., 2002). As such, assessing for family
history of suicide can provide information about a patient’s baseline risk
for suicide. Early life events such as a history of verbal, physical or sexual
abuse (Brown et al., 1999), or emotionally invalidating environments
(Linehan, 1993) also confer vulnerabilities to suicide. A history of psy-
chiatric disorders is associated with increased risk, especially conditions
that are chronic in nature, such as recurrent depression, bipolar disorders
or borderline personality disorder. It is important to note, however, that

29
Chad E. Morrow et al.

almost all psychiatric disorders have been shown to increase risk for suicide
as measured by standardised mortality ratios (Harris and Barraclough,
1998). Clinicians should therefore assess the patient’s psychiatric history
when considering baseline risk.

Previous suicidal behaviours The clinician should also assess the


patient’s history of suicidal thinking, self-injurious behaviours and sui-
cide attempts. Although a history of self-injurious behaviours of any
type raises an individual’s baseline risk for suicide above and beyond any
other risk factor for suicide (Joiner et al., 2005), previous suicide attempts,
in particular, are associated with substantially increased baseline risk
for suicide. A history of two or more suicide attempts (i.e., ‘multiple at-
tempter’ status) therefore indicates chronically elevated risk for suicide
(Rudd et al., 1996; Wingate et al., 2004). The relatively higher risk for sui-
cide associated with multiple suicide attempts, when compared to self-
harm, reflects the importance of differentiating between previous suicide
attempts and non-suicidal self-harm. While it is imperative to assess the
history of multiple suicide attempts, the absence of these behaviours does
not indicate reduced risk. For example, suicide attempters with a history of
recurrent self-harm have a higher level of depression, anxiety, hopelessness
and impulsivity, and tend to underestimate the lethality of their actions
(Stanley et al., 2001), which places them at higher risk for unintended death
by suicide. Given that a history of suicidal behaviours is the singlemost
robust predictor of future suicidal behaviours (Joiner et al., 2005), it is
critical that clinicians obtain this information as a central part of the risk
assessment.
When assessing the history of suicidal behaviours, the clinician should
also attempt to identify behavioural patterns in several dimensions: the
frequency and context of the suicidal behaviour (e.g., ‘How often have you
attempted to kill yourself or hurt yourself in the past? What was going on
at this time in your life?’), perceived lethality (e.g., ‘Why did you choose
that particular method? Did you think it would be enough to successfully
complete suicide?’), opportunity for rescue (e.g., ‘Did you know your
spouse would come home to find you?’), the amount of identifiable
preparations for death (e.g., ‘Had you been putting your will in order in
case of your death? Had you been giving away your possessions?’) and
reaction to survival (e.g., ‘Were you glad to be alive or did you wish you

30
Empirically Based Assessment of Suicide Risk

had died afterwards?’). This should be accomplished for ‘each and every’
episode, with the goal to understand the trajectory of risk over time, and
to identify clues for treatment interventions.
It is recommended that clinicians sequence their questions about
past suicidal behaviours by starting with most distant episodes first and
progressing forward chronologically towards the current situation. Such
an approach can alleviate some of the distress the patient might be ex-
periencing while discussing sensitive and upsetting events, since it can be
easier to talk about distal, historical events than it can be to talk about
more proximal stressors. Sequencing not only provides structure and
order to the clinical interview, but also serves to reduce the likelihood
that important clinical data will be missed.

Impulsivity The clinician should also assess the patient’s subjective sense
of self-control (e.g., ‘Do you consider yourself to be impulsive? Have you
recently felt out of control?’) and compare it with objective identifiers of
self-control including a history of aggression or violence, or engagement
in painful or provocative experiences in life such as high risk activities and
risky behaviours (Van Orden et al., 2008). Use of alcohol and drugs has
consistently been found to be associated with elevated suicide risk (APA,
2003), and can increase suicidality through a variety of ways: impaired
judgement, reduced inhibitions and increased depression. Substance
also correlates with social isolation, and is more likely to be co-morbid
with personality disorders, both of which are independent risk factors for
suicide. Because impulsivity is a fairly stable trait associated with multiple-
attempt status, impulsive multiple attempters should be considered a
chronic suicide risk. In general, a personality style marked by pronounced
impulsivity and aggression describes individuals at risk of suicide attempts
regardless of psychiatric diagnosis (Mann et al., 1999).

Acute Risk

Acute risk involves the level of risk present during an active suicidal crisis,
and is often associated with psychiatric symptom exacerbation. Because
these precipitating risk factors are more dynamic in nature and fluctuate
over time, they are common targets for clinical intervention.

31
Chad E. Morrow et al.

Precipitants and stressors As mentioned previously, patients often


choose to kill themselves following an environmental event or stressor.
When considering what has triggered a suicidal crisis, the clinician should
consider significant stressor, such as a financial, interpersonal or em-
ployment loss (Fu et al., 2002), acute or chronic health problems (Maris
et al., 2000) and family instability (Rubenowitz et al., 2001). Almost
always, the patient has experienced some sort of acute stressor that has
directly contributed to increased symptomatology and distress. A patient
who presents in a suicidal crisis is very likely experiencing one of these
stressors, or may be experiencing an acute escalation of multiple chronic
stressors. One goal of the clinician is to identify the stressor and to assist
the patient in resolving the problem.

SymptomaƟc presentaƟon Clinicians should determine the patient’s


symptomatic picture, along with the severity of these symptoms, includ-
ing both Axis I and Axis II comorbidity, and symptom clusters such as
depression, anxiety, agitation, anger or agitation (APA, 2003). Psychosis
should be assessed and appropriately referred to for evaluation and
treatment, whether on an inpatient or outpatient basis, due to the increased
risk for suicide associated with this condition (Harris and Barraclough,
1998), especially during periods of acute symptomatology (Kaplan and
Harrow, 1999). Shame, guilt, anger, anxiety and depression are particularly
powerful and frequently occurring emotions that can drive and maintain
suicidal episodes. A growing body of research has found that agitation,
restlessness and racing thoughts seem to be a particularly pernicious symp-
tom cluster, especially when they occur in the presence of a depressive
episode (Akiskal and Benazzi, 2005; Benazzi, 2005). A simple strategy for
assessing the intensity or severity of various psychiatric symptoms is to use
a 1-to-10 scale (e.g., 1 being the best the patient has ever felt and 10 being
the worst), which can help to improve clarity in communication between
patient and clinician, as well as offering a simple method for tracking
symptom change over time (cf. Rudd, 2006). Being able to demonstrate
improvement in scores, for example, across clinical care is a very powerful
method for providing the patient with a sense of control by quantifying his
or her emotional experience. Clinicians who can accurately understand the

32
Empirically Based Assessment of Suicide Risk

patient’s unique experience and gear interventions towards the reduction


of psychological distress will be more successful at reducing suicide risk.

Hopelessness A clinician should assess for the presence, severity and


duration of hopelessness with any patient who endorses suicidal be-
haviours. Hopelessness is a well-supported and highly robust predictor of
suicidal behaviour (e.g., Beck et al., 1990; Brown et al., 2004), with many
suicidal patients reporting the presence of severe hopelessness. Consistent
with the functional model of suicidality, relief from the distress associ-
ated with hopelessness is a common motivator driving suicidal behaviours.
Hopelessness is determined by both state (fluctuating) and trait (static)
variables and is greater in suicide attempters than in non-attempters
despite similar rates of objective severity of depression or psychosis (Mann
et al., 1999). The goal of the clinician is not only to ascertain the patient’s
degree of hopelessness, but also to directly target this thought process by
instilling a sense of hope for recovery and problem resolution.

Nature of suicidal thinking When considering the nature of suicidal


ideation, the clinician should assess frequency (e.g., ‘How often do you
think about suicide?’), intensity (e.g., ‘Could you rate the intensity of your
suicidal thoughts on a scale of 0 to 10?’) and duration (e.g., ‘How long do
these thoughts typically last?’) of suicidal thoughts. Intensity and duration
of suicidal ideation are more strongly associated with suicidal behaviours
than frequency of ideation (Joiner et al., 1997), and should therefore be
emphasised in suicide risk assessments. Other dimensions for assessment
include specificity of planning (e.g., ‘Have you thought about how, when
and where you would kill yourself?’), availability of means (e.g., ‘Do you
have access to the means?’), preparatory behaviours (e.g., ‘Have you taken
steps to prepare for suicide such as writing a note, getting financial affairs
in order, giving away possessions?’), explicit intent (e.g., ‘What do you
hope will happen as a result of this behaviour?’) and deterrents to suicide
(e.g., ‘What stops you from killing yourself?’). The most robust predictors
of suicidal behaviours are suicidal thinking, rehearsing and preparing
for suicide—for example, developing a specific suicide plan, acquiring
the means for suicide, counting pills, holding a gun to one’s head, and

33
Chad E. Morrow et al.

other such ‘practice’ activities—and should therefore be carefully assessed


by clinicians (Joiner, 2005; Joiner et al., 1997; Minnix et al., 2007; Pettit
et al., 2004).

Protective Factors

Protective factors, in contrast to risk factors, serve to decrease risk for sui-
cide. Identifying those variables that serve to mitigate risk is a useful
strategy for developing management plans and interventions to target
suicide risk (e.g., ‘What keeps you alive right now? What reasons do you
have to live?’). Examples of protective factors include the presence of
reasons for living (Linehan et al., 1983; Malone et al., 2000), which might
convey a sense of optimism or hope for the future, and strong relationships
with family or friends (Stravynski and Boyer, 2001; Turvey et al., 2002),
including the presence of children in the home (Clark and Fawcett, 1994),
each of which supports the proposition that perceived belongingness to
a social group serves as a buffer to suicide (Joiner, 2005). Even though
risk factors seem to have a stronger empirical relationship with suicidality
than protective factors, suicide interventions that focus on increasing or
strengthening protective factors while simultaneously reducing risk factors
are more effective than focusing on risk factors alone (Bryan and Rudd,
2006). By determining which variables are serving to keep the patient alive,
the clinician can begin to build interventions and strategies that serve to
reduce risk for suicide.

A Word on Access to Lethal Means

Many clinicians spend a considerable amount of time attempting to gauge


the severity of suicidal intent when assessing risk, but overlook the im-
portance of availability of means. Research on the relationship between
intent and death by suicide has demonstrated conflicting results, arguably
due to the confounding variable of availability of means. Intent has been
found to bear little relationship to the lethality of a suicide attempt method
(Brown et al., 2004; Plutchik et al., 1988; Swahn and Potter, 2001), but
availability of means has consistently demonstrated an association with
methodology (Eddleston et al., 2006; Peterson et al., 1985). It is therefore

34
Empirically Based Assessment of Suicide Risk

recommended that clinicians routinely ask about methods and access to


lethal means of suicide. Regular questioning is paramount because suicide
attempts almost always occur during short-term peaks in distress. For
example, among patients who survived life-threatening suicide attempts,
24 percent made the decision within five minutes preceding the attempt
and 70 percent made the decision within the preceding hour (Simon et al.,
2001). Further, when suicide rates by firearms are examined, suicide
rates have been found to be highest immediately following the purchase
of the firearm, with declining risk occurring as time passes: 57 times
higher during the first week following firearm purchase, declining to 30
times higher during the first month and seven times higher after one year
(Wintemute et al., 1999). Because suicide attempts and death by suicide are
commonly impulsive reactions to acute distress, the removal or limitation
of access to lethal means can significantly reduce the probability for a
suicide attempt.

Risk Categories and a Continuum of Risk

We recommend clinicians differentiate between four categories of suicide


risk, which are outlined in Table 2.2. Using these four categories will assist
clinicians in recognising and considering suicidality from both dimensions
of risk. After distinguishing which risk category a suicidal patient falls in,
clinicians should next assess the severity of suicide risk in order to direct
the most appropriate clinical response. A continuum of suicide risk based
Table 2.2 Categories of Suicide Risk

Risk category Criteria


Baseline No significant stressors or prominent symptoms;
only appropriate for ideators and single attempters
Acute Significant stressors and/or prominent symptoms;
only appropriate for ideators and single attempters
Chronic High Risk Baseline risk for multiple attempters; no significant
stressors or prominent symptoms
Chronic High Risk Acute risk category for multiple attempters; presence
of significant
With Acute stressor or prominent symptoms
Exacerbation
Source: Modified from Rudd (2006), Bryan and Rudd (2006).

35
Chad E. Morrow et al.

on a synthesis of risk factors and protective factors is presented in Table


2.3 (cf. Bryan and Rudd, 2006; Somers-Flanagan and Somers-Flanagan,
1995), along with implicated clinical response. It is important to highlight
that suicide risk assessment is complicated by temporal factors in at least
two ways. First, identifiable risk periods are inconsistently defined in the
literature, such that there is no reliable way to determine how long an acute
suicidal episode will endure. Second, chronic suicidality complicates risk
estimates in that multiple attempters have a higher baseline risk for sui-
cide to begin with. In every risk assessment, the clinician should first ask
the question, ‘Is this person a multiple attempter?’ If not, the clinician will
be considering acute risk only. If the patient is a multiple attempter and in
acute distress, however, the clinician should factor in the chronic nature
of the patient’s suicidality, and should automatically consider them to be
at least a moderate risk (Wingate et al., 2004). We recommend, therefore,
that risk assessment be a continuous and routine task throughout the
course of treatment.

Translating Risk Assessment into Effective


Management and Intervention

A distinct risk assessment scheme will ideally translate into straightfor-


ward, clinically informed and effective decisions. Table 2.3 provides a
summary of risk levels with indicated clinical responses or options. Those
determined to be at mild or very low risk require no particular change
in treatment aside from continuous evaluation of ideation and risk fac-
tors. For those at moderate to severe risk, outpatient management can be
accomplished effectively with increases in the intensity of clinical man-
agement strategies. For patients at extreme risk for suicide, there is no
room for debate about the standard of care, which demands for immediate
evaluation for inpatient hospitalisation.
In summary, expectations for assessing suicidal risk have significantly
changed for outpatient mental health clinicians. Although potentially
anxiety-producing for many clinicians at first, clinical decision-making
and management of suicidality are surprisingly straightforward when an
accurate, empirically based risk assessment is completed competently.
Contrary to persistent myths about working with suicidal patients,
outpatient care can be accomplished in a safe and effective manner when
informed by scientific evidence and competency-based practice.

36
Table 2.3 Suicide Risk Continuum with Indicated Responses

Risk Level Description Indicated Response


Very low No identifiable suicidal ideation No particular changes in ongoing treatment evaluation of any
expressed suicidal ideation to monitor change in risk
Mild Suicidal ideation of limited frequency, intensity and duration, no
identifiable plans, no intent, mild dysphoria/symptomatology,
good self-control, few risk factors, and identifiable protective
factors
Moderate Frequent suicidal ideation with limited intensity and duration, 1. Recurrent evaluation of need for hospitalisation
some specific plans, no intent, good self-control, limited 2. Increase in frequency or duration of outpatient visits
dysphoria/symptomatology, some risk factors present, and 3. Active involvement of the family
identifiable protective factors 4. Frequent re-evaluation of treatment plan goals
5. 24-hour availability of emergency or crisis services for patient
6. Frequent re-evaluation of suicide risk, noting specific changes
that reduce or elevate risk
7. Consideration of medication if symptomatology worsens or persists
8. Use of telephone contacts for monitoring
9. Frequent input from family members with respect to indicators
10. Professional consultation as indicated
Severe Frequent, intense and enduring suicidal ideation, specific plans, no Immediate evaluation for inpatient hospitalisation (voluntary or
subjective intent but some objective markers of intent (e.g., choice involuntary, depending on situation)
of lethal method(s), the method is available/accessible, some
limited preparatory behaviour), evidence of impaired self-
control, severe dysphoria/symptomatology, multiple risk factors
present, and few if any protective factors
Extreme Frequent, intense and enduring suicidal ideation, specific plans,
clear subjective and objective intent, impaired self-control, severe
dysphoria/symptomatology, many risk factors, and no protective
factors
Source: Modified from Rudd (2006), Bryan and Rudd (2006).
Chad E. Morrow et al.

REFERENCES

Akiskal, H.S. and F. Benazzi (2005). Psychopathologic correlates of suicidal ideation


in major depressive outpatients: Is it all due to unrecognized (bipolar) depressive
mixed states? Psychopathology, 38(5), 273–80.
American Psychiatric Association (2003). Practice guideline for the assessment and
treatment of patients with suicidal behaviours. Official Journal of the American
Psychiatric Association, 160(Supplement 11), 1–60.
Beck, A.T., G.K. Brown, R.J. Berchick, B.L. Stewart and R.A. Steer (1990). Relationship
between hopelessness and ultimate suicide: A replication with psychiatric
outpatients. American Journal of Psychiatry, 147(2), 190–95.
Benazzi, F. (2005). Suicidal ideation and bipolar-II depression symptoms. Human
Psychopharmacology: Clinical and Experimental, 20(1), 27–32.
Bongar, B. and M. Harmatz (1991). Clinical psychology graduate education in the study
of suicide: availability, resources, and importance. Suicide and Life-Threatening
Behavior, 21(3), 231–44.
Bongar, B., L.G. Peterson, E.A. Harris and J. Aissis (1989). Clinical and legal
considerations in the management of suicidal patients: An integrative overview.
Journal of Integrative and Eclectic Psychotherapy, 8(1), 53–67.
Brown, G.K., G.R. Henriques, D. Sosdjan and A.T. Beck (2004). Suicide intent and
accurate expectations of lethality: predictors of medical lethality of suicide attempts.
Journal of Consulting and Clinical Psychology, 72(6), 1170–74.
Brown, J., P. Cohen, J.G. Johnson and E.M. Samiles (1999). Childhood abuse and neglect:
Specificity of effect on adolescent and young depression and suicidality. Journal of
the American Academy of Child & Adolescent Psychiatry, 38(12), 1490–96.
Bryan, C.J. and M.D. Rudd (2006). Advances in the assessment of suicide risk. Journal
of Clinical Psychology: In Session, 62(2), 185–200.
Burstein, A.G., R.L. Adams and M.B. Giffen (1973). Assessment of suicidal risk by
psychology and psychiatry trainees. Archives of General Psychiatry, 29(6), 792–93.
Buzan, R.D. and M.P. Weissberg (1992). Suicide: Risk factors and prevention in medical
practice. Annual Review of Medicine, 43, 37–46.
Chemtob, C.M., R.S. Hamada, G.B. Bauer, R.Y. Torigoe and B. Kinney (1988a). Patient
suicide: Frequency and impact on psychologists. Professional Psychology: Research
and Practice, 19(4), 416–20.
Chemtob, C.M., R.S. Hamada, G.B. Bauer, R.Y. Torigoe and B. Kinney (1988b). Patient
suicide: Frequency and impact on psychiatrists. American Journal of Psychiatry,
145(2), 224–28.
Clark, D.C. and J. Fawcett (1994). The relation of parenthood to suicide. Archives of
General Psychiatry, 51(2), 160.
Eddleston, M., A. Karunaratne, M. Weerakoon, S. Kumarasinghe, M. Rajapakshe, M.H.
Sheriff, N.A. Buckley and D. Gunnell (2006). Choice of poison for intentional
self-poisoning in rural Sri Lanka. Clinical Toxicology, 44, 283–86.

38
Empirically Based Assessment of Suicide Risk

Egeland, J.A. and J.N. Sussex (1985). Suicide and family loading for affective disorders.
Journal of American Medical Association, 254(7), 915–18.
Fawcett, J. (1999). Profiles of completed suicides. In D.G. Jacobs (Ed.), The Harvard
Medical School Guide to Suicide Assessment and Intervention (pp. 115–24). San
Francisco, CA: Jossey-Bass.
Feldman, B.N. and S. Freedenthal (2006). Social work education in suicide intervention
and prevention: An unmet need? Suicide and Life-Threatening Behavior, 36(4),
467–80.
Fu, Q., A.C. Heath, K.K. Bucholz, E.C. Nelson, A.L. Glowinski, J. Goldberg, M.J. Lyons,
M.T. Tsuang, T. Jacob, M.R. True and S.A. Eisen (2002). A twin study of genetic and
environmental influences on suicidality in men. Psychological Medicine, 32(1),
11–24.
Guy, J.D., C.K. Brown and P.L. Polestra (1990). Who gets attacked? A national survey
of patient violence directed at psychologists in clinical practice. Professional
Psychology: Research and Practice, 21(6), 493–95.
Harris, E.C. and B. Barraclough (1998). Excess mortality of mental disorder. The British
Journal of Psychiatry, 173(1), 11–53.
Joiner, T.E. (2005). Why people die by suicide. Cambridge, MA: Harvard University Press.
Joiner, T.E., F. Johnson and K. Soderstrom (2002). Association between serotonin
transporter gene polymorphism and family history of attempted and completed
suicide. Suicide and Life-Threatening Behavior, 32(3), 329–32.
Joiner, T.E., M.D. Rudd and M.H. Rajab (1997). The modified scale for suicidal
ideation: Factors of suicidality and their relation to clinical and diagnostic variables.
Journal of Abnormal Psychology, 106(2), 260–65.
Joiner, T.E., Y. Conwell, K.K. Fitzpatrick, T.K. Witte, N.B. Schmidt, M.T. Berlim,
M.P.A., Fleck and M.D. Rudd (2005). Four studies on how past and current sui-
cidality relate even when “everything but the kitchen sink” is covaried. Journal of
Abnormal Psychology, 114(2), 291–303.
Kaplan, K.J. and M. Harrow (1999). Psychosis and functioning as risk factors for later
suicidal activity among schizophrenia and schizoaffective patients: A disease-based
interactive model. Suicide and Life-Threatening Behavior, 29(1), 10–24.
Kleespies, P.M., W.E. Penk and J.P. Forsyth (1993). The stress of patient suicidal be-
havior during clinical training: incidence, impact, and recovery. Professional
Psychology: Research and Practice, 24(3), 293–303.
Linehan, M.M. (1993). Cognitive-behavioral treatment of borderline personality disorder.
New York, NY: Guilford Press.
Linehan, M.M., J.L. Goodstein, S.L. Nielsen and J.A. Chiles (1983). Reasons for staying
alive when you are thinking of killing yourself: The reasons for living inventory.
Journal of Consulting and Clinical Psychology, 51(2), 276–86.
Malone, K.M., M.A. Oquendo, G.L. Haas, S.P. Ellis, S. Li and J.J. Mann (2000). Protective
factors against suicidal acts in major depression: Reasons for living. American
Journal of Psychiatry, 157(7), 1084–88.
Maltsberger, J.T. (1986). Suicide risk: The formulation of clinical judgment. NY:
New York University Press.

39
Chad E. Morrow et al.

Mann, J.J, C. Waternaux, G.L. Haas and K.M. Malone (1999). Toward a clinical model
of suicidal behavior in psychiatric patients. American Journal of Psychiatry, 156(2),
181–89.
Maris, R.W., S.S. Canetto, J.L. McIntosh and M.M. Silverman (Eds). (2000). Review of
suicidology. NY: Guilford Press.
McAdams, C.R. and V.A. Foster (2000). Client suicide: Its frequency and impact on
counselors. Journal of Mental Health Counseling, 22(2), 107–21.
Minnix, J.A., C. Romero, T.E. Joiner and E.F. Weinberg (2007). Change in ‘resolved
plans’ and ‘suicidal ideation’ factors of suicidality after participation in an inten-
sive outpatient treatment program. Journal of Affective Disorders, 103(1), 63–68.
O’Carroll, P.W., A. Berman, R.W. Maris and E. K. Moscicki (1996). Beyond the Tower
of Babel: A nomenclature for suicidology. Suicide and Life-Threatening Behavior,
26(3), 237–252.
Peterson, L., M. Peterson, G. O’Shanick and A. Swann (1985). Self-inflicted gunshot
wounds: Lethality of method versus intent. American Journal of Psychiatry, 142(2),
228–31.
Pettit, J.W., T.E. Joiner and M.D. Rudd (2004). Kindling and behavioral sensitization:
Are they relevant to recurrent suicide attempts? Journal of Affective Disorders,
83(2–3), 249–52.
Plutchik, R., H.M. van Praag, S. Picard, H.R. Conte and M. Korn (1988). Is there a rela-
tion between the seriousness of suicidal intent and the lethality of the suicide
attempt? Psychiatry Research, 27(1), 71–79.
Pope, K.S. and B.G. Tabachnick (1993). Therapists’ anger, hate, fear, and sexual
feelings: National survey of therapist responses, client characteristics, critical
events, formal complaints, and training. Professional Psychology: Research and
Practice, 24(2), 142–52.
Roy, A. (1992). Suicide in schizophrenia. International Review of Psychiatry, 4(2),
205–209.
Rubenowitz, E., M. Waern, K. Wilhelmson and P. Allebeck (2001). Life events and
psychosocial factors in elderly suicides—A case-control study. Psychological
Medicine, 31(7), 1193–202.
Rudd, M.D. (2006). The assessment and management of suicidality. Sarasota, FL:
Professional Resource Press.
Rudd, M.D., T.E. Joiner and M.H. Rajab (1996). Relationships among suicide ideators,
attempters, and multiple attempters in a young-adult sample. Journal of Abnormal
Psychology, 105(4), 541–50.
Shneidman, E.S. (1981). Psychotherapy with suicidal patients. Suicide and Life-
Threatening Behavior, 11(4), 341–48.
Shneidman, E. S. (1984). Aphorisms of suicide and some implications for psychotherapy.
American Journal of Psychotherapy, 38(3), 319–28.
Silverman, M.M., A.L. Berman, N.D. Sanddal, P.W. O’Carroll and T.E. Joiner (2007).
Rebuilding the Tower of Babel: A revised nomenclature for the study of suicide
and suicidal behaviors part 2: Suicide-related ideations, communications, and
behaviors. Suicide and Life-Threatening Behavior, 37(3), 264–77.

40
Empirically Based Assessment of Suicide Risk

Simon, T.R., A.C. Swann, K.E. Powell, L.B. Potter, M. Kresnow and P.W. O’Carroll
(2001). Characteristics of impulsive suicide attempts and attempters. Suicide and
Life-Threatening Behavior, 32(Supplement), 49–59.
Somers-Flanagan, J. and R. Somers-Flanagan (1995). Intake interviewing with suicidal
patients: A systematic approach. Professional Psychology: Research and Practice,
26(1), 41–47.
Stanley, B., M.J. Gameroff, V. Michalson and J.J. Mann (2001). Are suicide attempters
who self-mutilate a unique population? American Journal of Psychiatry, 158(3),
427–32.
Stravynski, A. and R. Boyer (2001). Loneliness in relation to suicide ideation and para-
suicide: A population-wide study. Suicide and Life-Threatening Behavior, 31(1),
32–40.
Suicide Prevention Resource Center (2006). Core competencies in the assessment and
management of suicidality. Newton, MA: SPRC.
Swahn, M.H. and L.B. Potter (2001). Factors associated with the medical severity of
suicide attempts in youths and young adults. Suicide and Life-Threatening Behavior,
32(1), 21–29.
Turvey, C.L., Y. Conwell, M.P. Jones, C. Phillips, E. Simonsick, J.L. Pearson and R. Wallace
(2002). Risk factors for late-life suicide: A prospective, community-based study.
American Journal of Geriatric Psychiatry, 10(4), 398–406.
Van Orden, K.A., T.K. Witte, K.H. Gordon, T.W. Bender and T.E. Joiner (2008). Suicidal
desire and the capability for suicide: Tests of the interpersonal-psychological theory
of suicidal behavior among adults. Journal of Consulting and Clinical Psychology,
76(1), 72–83.
Wingate, L.R., T.E. Joiner, R.L. Walker, M.D. Rudd and D.A. Jobes (2004). Empirically
informed approaches to topics in suicide risk assessment. Behavioral Sciences and
the Law, 22(5), 651–65.
Wintemute, G.J., M.A. Wright, C.A. Parham, C.M. Drake and J.J. Beaumont (1999).
Denial of handgun purchase: A description of the affected population and a
controlled study of their handgun preferences. Journal of Criminal Justice, 27(1),
21–31.

41
3

Neurobiological Basis of
Suicidal IdeaƟon
J®ã›Ä—Ùƒ KçÃ٠TÙ®ò›—® ƒÄ— SƒÄÄ®—«ùƒ VƒÙÃ

A ccording to the World Health Report 2002, published by WHO (UC


Atlas of Global Inequality, 2002), suicide was amongst the top 10
leading causes of death across all age groups in the developed nations in
the year 2001. In India, over one lakh lives are lost every year as a result of
suicide and this figure does not take into account the gross underreporting
(Joseph et al., 2003). According to the National Crime Bureau, Ministry
of Home Affairs, Government of India, Accidental Deaths and Suicides
in India, 2007 (National Crime Bureau, 2007), youths (15–29 years) and
lower middle-aged people (30–44 years) were the prime groups taking
recourse to the path of suicides. Around 35.2 percent suicide victims were
youths in the age group of 15–29 years and 34.1 percent were middle-aged
persons in the age group 30–44 years, adversely affecting not just the family
of the victim but also the economic productivity of the nation at large.
Reliable methods of predicting suicide which may enable its prevention by
effective management are therefore energetically being sought after.
In the past it was thought that suicide was related more to defects in
the mind (psychological) than in the brain (neurobiological). Recent ad-
vances in neurosciences are challenging this notion. There is a growing
support for a theory of human emotions that implicates increasingly
well-defined brain regions (Stuss et al., 2001), of which, the frontal lobes

42
Neurobiological Basis of Suicidal Ideation

seem to have the greatest significance due to the central role that they play
in social cognition, aggression and impulse control. This also strengthens
the hypothesis that suicidal behaviour may be due to underlying
neurobiological factors.
Suicide is not an entity of its own but associated with many other
disorders: about 90 percent of the people who commit suicide have a known
psychiatric illness such as:

1. major depressive disorder (15 percent of depressives who are ad-


mitted in a hospital eventually commit suicide) (McIntosh, 2001);
2. schizophrenia (especially in post-psychotic period);
3. bipolar affective disorder (especially bipolar disorder type I);
4. borderline personality disorder and sociopathic personality disorder
in adolescents and young adults;
5. alcohol or other substance abuse are also predictive of suicide (up to
50 percent of the people who commit suicide are intoxicated at the
time of death) (Tanney, 2000); and
6. comorbidity of depressive-mood disorder and substance abuse
greatly raises risk of suicide—70–80 percent of people who commit
suicide have co-morbid diagnoses (Moscicki, 2001).

However, numerous studies controlling for the presence of psychiatric


conditions have indicated that the liability for suicide is independent
from (but conditional on) the genetic factors mediating susceptibility for
major psychiatric disorders (Klempan and Turecki, 2005). The underlying
neurobiological factors for suicide and the associated psychiatric disorders
therefore may be different.
In the past 30 years, a comprehensive and multi-faceted approach to
study has been undertaken to find more accurate methods for predicting
suicide. Even with standardised assessment and prediction scales such as
the Hamilton or Beck depression inventories, it has been difficult to predict
suicide accurately. Finding the neurological basis of suicide will help not
only to better understand this behaviour, but it may also provide us with a
reliable means of predicting suicide accurately. The focus has now shifted
from suicidal ideation being seen as primarily related to depression to
focusing on suicidal acts as primarily being related to biology of aggression
and impulsivity. The following discussion deals with the neurobiological
basis of suicide.

43
Jitendra Kumar Trivedi and Sannidhya Varma

BIOLOGICAL FACTORS

The search for biological markers or aetiology of suicidal behaviour was


spurred by the inability of psychologically oriented studies to adequately
explain and predict suicidal behaviour. The biological changes may be of
two types: (a) structural and (b) neurobiological.
There is little evidence in support of structural changes that may be
specifically linked to suicidal behaviour. Most of the present work is on the
underlying neurobiological factors which are associated with suicide. On
the basis of the pharmaco-challenge studies the ‘neurobiological changes’
were further divided into the following two types:

1. Transmitter Non-specific Neuro-endocrine studies


2. Transmitter Specific Neuro-endocrine studies

Transmi er Non-specific Neuro-endocrine Studies

Limbic-Hypothalamus-Pituitary-Adrenal Axis (LHPA Axis) A few of the


early studies (Bunney and Fawcett, 1965; Krieger, 1974) showed a link
between suicidality and LPHA axis over-activity with increased serum
levels of cortisol and 17-hydroxycorticosteroids in urine. However, this
finding could not be confirmed in other studies. The frontal cortex has
fewer binding sites for corticotrophin-releasing hormone and reduced
plasma adrenocorticotropin and cortisol responsiveness (Pfennig et al.,
2005). Preliminary studies using central corticotrophin-releasing factor
binding (Nemeroff et al., 1988) and preopiomelanacortin mRNA (Lopez
et al., 1990) as markers of LHPA axis activity are showing more definitive
results regarding correlation between axis hyperactivity and suicide.

Limbic-Hypothalamus-Pituitary-Thyroid Axis (LHPT Axis) Studies have


not revealed any conclusive proof of association between LHPT axis activity
and suicide. One group, studying a population with various diagnoses, even
has reported enhanced Thyroid Stimulating Hormone (TSH) response in
patients with suicidal behaviour (Banki and Arato, 1983).

44
Neurobiological Basis of Suicidal Ideation

Transmi er Specific Neuro-endocrine Studies

Serotonin The concept that suicide or suicidal behaviour may arise from
some specific anomaly in the biological system arose from early attempts
to study the role of neurotransmitter serotonin (5-HT) in depression. In
a classic series of papers, Asberg (1997) showed that the Cerebro-spinal
Fluid (CSF) concentration of 5-hydroxy indole acetic acid (5-HIAA) was
reduced in patients with depression. It was observed that a high proportion
of individuals with low CSF 5-HIAA subsequently went on to make suicide
attempts and to kill themselves. Later studies reported that measures of
low 5-HT function were associated with suicidal behaviour, not only in
depression but in schizophrenia and other diagnosis as well (Arango and
Underwood, 1997).
Brain serotonin levels as a predictor of suicide has been the subject of
intense research scrutiny over the past several years, with scientists trying
to find easily accessible markers so that the neurotransmitter’s levels
might someday be readily measured in clinical settings. The reasons for
this approach are:

1. Most effective antidepressant drugs directly/indirectly enhance


5-HT function.
2. Brown and his colleagues (1992) describe a serotonergic trait which
includes sleep difficulties, impulsivity, disinhibition, headaches,
proneness to pain, glucosteroid abnormalities, mood volatility, dis-
order of conduct, poor peer relationships and suicidal behaviours.
Patients with highest aggression have been found to have lowest
serotonin activity.

Serotonergic neurons in brain arise from the raphe nuclei in the brain
stem and from there project to different parts of the brain including the
frontal lobes which are responsible for the integration of sensations, per-
ceptions, consciousness and memory into organised and planned be-
haviours (Fuster, 1997), and the prefrontal cortex which also mediates
prospective cognitive processes.
The primary finding in most of the studies of neurological basis of
suicide is decreased amount of metabolite of serotonin 5-HIAA in the

45
Jitendra Kumar Trivedi and Sannidhya Varma

CSF of suicide attempters/victims than in controls. Low levels of 5-HIAA


in CSF suggests decreased synthesis of serotonin which may be due to
reduced availability of tryptophan or decreased activity of tryptophan
hydroxylase (rate limiting enzyme in the synthesis of serotonin). The
tryptophan hydroxylase has functional genetic polymorphism in the popu-
lation and some studies reveal that the less active forms occur more
commonly among people who attempt to commit suicide recurrently
and impulsively. According to Mann (1998), the extent of the reduction
of the metabolite 5-HIAA correlates with the lethality of the attempt of
suicide. The association of serotonergic dysfunction is particularly strong
for violent and impulsive suicide, but it is not specific to any psychi-
atric diagnosis. Similar findings have been reported for other forms
of violent and impulsive behaviour, such as fire-setting (Linnoila and
Virkkunen, 1992).
It has been found that the number of serotonergic neurons in the
brain of subjects who attempt suicide is increased or remains almost the
same when compared to controls. However, the functional capacity of
the said neurons is less as compared to the controls (Arango et al., 1997).
A more stable index of measuring the serotonin function in the brain is
through the receptor protein. Serotonin transporters (ST) mediate uptake
of serotonin out of the synaptic cleft (pre-synaptic); they are present
in areas of serotonergic activity and low serotonergic activity results in
fewer STs. Ligands like [3H] imipramine, [3H] partoxetine and [3H]
cyanoimipramine are used to study these receptors in vivo. Most studies
show a decreased ST concentration in frontal cortex of suicide victims
(Arango et al., 1995; Mann et al., 1996). However, some studies have con-
tradicted these findings, which may be explained as a result of binding of
the ligands to non-transporter sites (Mann et al., 1999), the function of
which is not known as yet. Depressed patients have a ST deficiency in the
pre-frontal cortex (PFC), but suicidal patients have a distinctive localised
ST deficiency in the ventral or orbital PFC (Mann et al., 2000).
A number of studies found an increase in radioligand binding ([3H]
spiroperidol) to the 5-HT2A receptors in post-mortem brains of suicide
victims (Kamali et al., 2001). In contrast, reduced in vivo radioligand
binding to the 5-HT2A receptors using Single Photon Emission Computed
Tomography (SPECT) has been reported in patients with depression who
were not suicidal (D’Haenan, 2001). The degree of difference between
the suicides and the controls appears to be greater in prefrontal cortex

46
Neurobiological Basis of Suicidal Ideation

(PFC) than in temporal cortex, suggesting regional specificity of the


suicide effect.
In vivo positron emission tomography (PET) imaging studies report
significant reduction in cortical 5-HT1A receptors binding sites in depres-
sion (Sargent et al., 2000). However, numbers are small because of the
expense of the technique and there is no reported link with suicidal be-
haviour or thoughts. Reduced 5-HT1A receptor binding in brains from
depressed suicides has been reported in two studies (Deakin and Ben,
2003). Other serotonin receptor subtypes have barely begun to be in-
vestigated, and studies are ongoing for 5-HT1C, 5-HT1B and 5-HT1D
receptors in suicide victims.
Studies have suggested an abnormality in the signal transduction
mechanism of the above-mentioned receptors in neurobiology of sui-
cide and mood disorders in general, like defect in protein kinase B or
phosphatidylinositol-3-kinase (Akt/PI3) and glycogen synthase kinase-3-β
(GSK-3). In suicide victims, the PI3-K/Akt signalling pathway has been
shown to be blunted in the occipital cortex with a significant decrease in
protein kinase B activity and other upstream effectors (Hsiung et al., 2003).
Glycogen synthase kinase-3-β plays a role in the control of many
regulatory enzymes concerned with cellular processes like cell survival,
apoptosis and embryogenesis. It has been demonstrated that 5-HT
activity regulates the phosphorylation of GSK-3 in mammalian brains in
vivo, a mechanism thought to be achieved by a balance between opposing
actions of 5-HT receptors subtypes (Li et al., 2004). The interest in GSK-3
in psychiatry has been highlighted by the discovery of its inhibition by
lithium, which has further been found to inhibit both forms of GSK-3
(Klein and Melton, 1996; Stambolic et al., 1996). The studies have sug-
gested that activity, rather than quantity of these enzymes is defective. How-
ever, these alterations have so far been shown to be associated with mood
disorders rather than suicide per se (Karege et al., 2007).

5-HT2A Receptors in Platelets and their


Connection with Suicide

Platelets contain a 5-HT2A receptor that is genetically, pharmacologically


and mechanistically the same as the 5-HT2A receptor in the central ner-
vous system. Pandey (1997) found that receptor density was increased

47
Jitendra Kumar Trivedi and Sannidhya Varma

in suicidal subjects independent of their psychiatric diagnosis and


that increased 5-HT2A density was significantly higher in subjects with a
recent attempt compared with those with past suicide attempts. Biegon
and colleagues (1989) reported normalisation of platelet 5-HT2A receptors
after effective anti-depressant activity which leads to the conclusion that
platelet 5-HT2A receptor number might be a state-dependent marker of
suicidality. It was also found that the functioning of platelet 5-HT receptors
(platelet aggregation in response to serotonin) was also decreased in
patients with history of high lethality suicide attempts.

Role of Lipids in Serotonergic System and


Suicidal Behaviour

Another interesting lead that, in spite of much controversy, is supported


by several lines of evidence concerns the relationship between low
cholesterol levels and suicidal behaviour (Kaplan et al., 1997). This is an
intriguing association with unclear mediating mechanisms to explain how
serum cholesterol levels may have an effect on behaviour. In any case, the
investigation of components of the lipid metabolisms in the neurobiology
of suicide and related behaviours has gained renewed interest in light of
the growing evidence demonstrating essential roles for cholesterol in brain
synaptogenesis, as well as evidence suggesting that alterations in brain
sterol composition may mediate this association (Lalovic et al., 2004). It
has been proposed that low cholesterol levels result in reduced serotonin
functioning which predisposes individuals towards impulsive behaviour
like suicide and aggression.

Fenfluramine Challenge Test

Fenfluramine acts on serotonin transporter (ST) and causes serotonin re-


lease, which causes an increase in serotonin concentration in the synaptic
cleft. Amongst factors like oestrogens, thyrotropin-releasing hormone
and others, serotonin also causes a release of prolactin from the anterior
pituitary. An increase in serotonin after giving fenfluramine is reflected
by an increase in serum prolactin levels. Coccaro and Kavoussi (1994)
showed a direct correlation between history of suicide attempt and

48
Neurobiological Basis of Suicidal Ideation

diminished prolactin response to fenfluramine in patients with differ-


ent diagnosis (mood disorder and personality disorder); correlation was
more with suicidal behaviour than depression. Strickland and colleagues
(2002) also showed a correlation between lowered prolactin response
to fenfluramine and particularly lethal suicidal attempt. There was no
correlation between time difference of suicide attempt, sampling and
response to fenfluramine.

Norepinephrine

Studies have shown that people with higher propensity of suicide have
a higher concentration of norepinephrine, tyrosine hydroxylase, α2-
adrenergic receptors, and decreased concentration of post-synaptic
β-adrenergic receptors and norepinephrine transporters (Maris, 2002).
This pattern is similar to that of an excessive stress response that leads
to norepinephrine depletion, perhaps because fewer neurons in locus
coeruleus (in contrast to serotonergic system where the number of neur-
ons remains intact but function is reduced) could mean reduced functional
reserve (Maris, 2002).
The findings in studies of norepinephrine have not been as robust as
those of serotonergic system and it has been more difficult to correlate
different behaviours with the neurochemistry. The findings in these studies
point towards the presence of chronic stress response, which emphasises
its relation with depression, suicide and hypothalamic-pituitary-adrenal
axis. Further studies are warranted along these lines.

GENETICS OF SUICIDE

Over the past 30 years, indirect evidence for the existence of a genetic com-
ponent in the suicidal diathesis has come largely from family, twin
and adoption studies. An adjusted meta-analysis of 21 family studies
(Baldessarini and Hennen, 2004) estimated that close relatives of suicidal
probands have a three times higher risk for engaging in suicidal acts
compared with controls, irrespective of psychiatric history. A substantial
familial component was confirmed by Kim et al. (2005) who compared

49
Jitendra Kumar Trivedi and Sannidhya Varma

suicidal behaviours in relatives of suicide completers with community


controls. After making adjustments for psychopathology, they found
that relatives of people who have committed suicide have about 10 times
higher risk of attempting suicide and an even greater incidence of suicidal
ideation than controls. Family history of suicide was strongly associated
with earlier onset suicide attempts and male gender (Mittendorfer-Rutz
et al., 2007). Overall, family studies have indicated that heritability may
be lowest for suicidal ideation, somewhat higher for suicide attempts, and
highest for suicide completions (Brent et al., 1996; Maris, 2002).
Unlike family-based designs, twin studies allow for a better control of
the influences of the shared environment. After pooling published twin
studies, Baldessarini and Hennen (2004) found a 175 times higher relative
risk among monozygotic twins (MZTs) than dizygotic twins (DZTs).
However, these results must be interpreted cautiously, keeping in mind
the low incidence of suicidal behaviours in twins and lack of control of
post-natal environmental influences.
Adoption studies have been scarce and based primarily on Danish
public health records. These investigations suggested a 7 to 13 times higher
risk for suicidality among biological relatives of adoptees than among
adopted relatives, and stronger heritability for suicide completions than
attempts (Schulsinger et al., 1979; Wender et al., 1986). These estimates
must be considered in light of several limitations, including low number
of adoption studies, poor control of psychiatric confounders, lack of more
recent adoption data and shortage of data from other countries.
Studies for search of genetic markers of suicide have given precedence
to the following neurotransmitter systems:

1. Serotonergic system.
2. Noradrenergic and dopaminergic systems.
3. Hypothalamic-pituitary-adrenal axis.
4. Neurotrophic, GABAergic and glutaminergic systems.

Serotonergic System—Genetic Component

Various components of the serotonergic system have been the subject of


the various studies:

50
Neurobiological Basis of Suicidal Ideation

1. Transport (serotonin transporter).


2. Transmission (receptor genes).
3. Anabolism (tryptophan hydroxylase).
4. Catabolism (mono-amine oxidase A).

Serotonin transporter (SLC6A4) The serotonin transporter is responsible


for determining the duration of serotonergic signal. Genetic studies
focused exclusively on two of its variants, the 44-base promoter deletion/
insertion (LPR) and its associated S/L alleles, respectively, and a variable
number tandem repeats (VNTR) polymorphism in intron 2. Of the 44
studies, approximately 20 have linked the former variant to predisposition
to suicide attempts, finding S allele carriers to have an elevation in risk
ranging between 1.7 and 6.5 times.
An early meta-analysis of 12 studies confirmed the association between
this promoter variant and suicide attempts but not completions (Anguelova
et al., 2003). A subsequent meta-analysis found an overrepresentation of S
genotypes in suicide attempters (P = .004) and violent suicides (P = .0001)
relative to controls (Li and He, 2007).
Only a few studies have reported statistically significant association of
VNTR-2 and suicide (de Lara et al., 2006; De Luca et al., 2006).

Serotonin receptor genes Genes of serotonin receptors have been


studied less often than those of serotonin transporter. Only limited support
was found for the involvement of receptor 5-HT1A (Nishiguchi et al., 2002;
Ohtani et al., 2004; Serretti et al., 2007). Similarly, only one small study
showed an association with 5-HT1B; the remaining investigations found
no association with either suicide or suicide attempts (Hong et al., 2004;
Nishiguchi et al., 2002; Tsai et al., 2004; Videtic et al., 2006). Variants in
other receptors from this class (5-HT1D, 5-HT1E and 5-HT1F) were not asso-
ciated with suicide completions in French Canadian (Turecki et al., 2003)
and Slovenian study samples (Videtic et al., 2006).
The 5-HT2A receptor gene may be the most promising serotonergic
receptor gene. Genetic association studies in 5-HT2A implicated primarily
the T102C polymorphism. However, subsequent meta-analyses based
on nine, and more recently 25 studies failed to confirm its involve-
ment (Anguelova et al., 2003; Li et al., 2006). The latter study suggested

51
Jitendra Kumar Trivedi and Sannidhya Varma

a significant role for the allele A of the A1438G variant, however. Newer
studies are being carried out to find out the role of genetic imprinting in
5-HT2A gene variation.

Tryptophan hydroxylase Tryptophan hydroxylase 1 (TPH1) was the first


gene to be studied for suspected association with suicide. Mutations of this
gene have been found to be associated with repeated attempts, impulsive
and violent suicide in about half of the studies carried out so far. Recent
meta-analysis has shown increased risk (about 60 percent) in carriers of
allele A218C in suicidal behaviour (Bellivier et al., 2004).
These positive findings have, however, been criticised because TPH1
is primarily expressed in the periphery, its significantly associated single
nucleotide polymorphisms (SNP) are intronic and unrelated to splicing
or exon skipping, and it exhibits high exonic homology with TPH2 (Shaltiel
et al., 2005). TPH2 is a recently identified gene, found to be preferentially
expressed in the brain stem and relevant to several psychiatric phenotypes
through its involvement in amygdala-mediated responses to emotional
stimuli (Canli et al., 2005). This gene also seems to have significant allelic,
genotypic and haplotypic relationships with suicidal acts. T and G alleles
of TPH2 have been found to be associated with five-fold increase of risk
of suicide. Similar to other candidate genes, attempts to replicate TPH1
and TPH2 associations have been unsuccessful, encompassing study
populations with different psychiatric diagnoses, such as mood disorders
(Ho et al., 2000), schizophrenia (De Luca et al., 2005), and alcohol
dependence (Zill et al., 2007), and ethnic membership.

Monoamine oxidase A Monoamine oxidase A (MAOA) is an X-linked


gene (Xp11.23) responsible for oxidative deamination of bioamines, such
as central and peripheral serotonin and noradrenaline. Of the nine studies
examining MAOA variation, six focused exclusively on a promoter VNTR.
This variant was linked to suicide and violent suicide attempts in men and
violent suicide attempts in depressed patients. A recent study investigated
Fnu4I, finding its I/I genotypes of relevance in women who had depression
(Nishiguchi et al., 2002).

52
Neurobiological Basis of Suicidal Ideation

Noradrenergic and Dopaminergic Genes—Genetic Component

Tyrosine hydroxylase (TH) is the rate limiting enzyme in the synthesis


of noradrenaline and dopamine. Genetic evidence shows that the TH-K3
allele of the TH gene may be related to the risk for suicide attempts in a
Swedish Caucasian sample with adjustment disorders.
Dopamine D2 receptor (DRD2) is the only dopaminergic gene which
has been linked to suicidal behaviour, with its 141Cdel allele conferring
a 30 percent and its E8 allele a 70 percent increased risk in patients who
had alcohol dependence (Brezo et al., 2008).
The most studied supporting enzyme in adrenergic and dopaminergic
systems has been the catechol-O-methyltransferase (COMT), respon-
sible for metabolising the breakdown of levodopa into 3-O-methyldopa.
Val158Met single nucleotide polymorphism of COMT has been found to
be associated with violent suicide.

Hypothalamic-Pituitary-Adrenal Axis

Variation in only one hypothalamic-pituitary-adrenal (HPA) axis gene


has been studied for association with suicide: the CRCH2 gene’s haplotype
5–2–3 has been found to be positively associated with severity of suicidal
behaviour (De Luca et al., 2007).

Neurotrophic Genes

Brain-derived neurotrophic factor (BDNF), the most abundant neuro-


trophin in the brain, may be lower in the plasma of patients who are suicidal
than in those who are non-suicidal and depressed, although it seems to
be unrelated to the lethality of suicide attempts (Dwivedi et al., 2003). Its
mRNA was, however, reduced in the hippocampus and PFC of individuals
who committed suicide.
Three neurotrophic genes have been studied so far:

53
Jitendra Kumar Trivedi and Sannidhya Varma

1. NOTCH4
2. NGFR
3. BDNF

None of them was found to be associated with major effect on sui-


cidal behaviour. BDNF gene, when considered in context of the environ-
mental factors, was found to be significantly associated with suicide. Val/
Val genotype of BDNF gene was found to be associated with violent suicide
attempts in people who had suffered from childhood sexual abuse.

GABA and Glutaminergic Genes

Although expression data suggests changes in GABA and glutaminergic


systems, very few studies have been carried out so far to examine their gen-
etic variation. None of the four published studies investigating GABAergic
and glutamatergic gene variation showed support for its relevance to
suicidal behaviours. Genes examined include glutamate decarboxylase
1 and 2 (De Luca et al., 2004) and GABA receptor alpha-3 (Baca-Garcia
et al., 2004) and alpha-5 genes (Wasserman et al., 2007), each subject of
one study.

Other Candidate Genes

In the last 30 years, about 20 genes outside the major systems have been
put under the scanner for association with suicidal behaviour. Most of
these candidates are involved in signalling and transport, with a few par-
ticipating in lipid metabolism, deoxidation and gene transcription (Brezo
et al., 2008).
Cyclic adenosine monophosphate (AMP) and phoshpoinositide
signalling systems and their components have been implicated in sui-
cidal behaviour. The cyclic AMP response element binding protein (CREB)
is a transcription factor which is associated with both of the foregoing
systems and an important part of the genes expressed in the neurons

54
Neurobiological Basis of Suicidal Ideation

(Dwivedi et al., 2004). Its mRNA was found to be decreased in PFC of


adolescents who committed suicide (Pandey et al., 2007).
HPA axis dysfunction in suicidal behaviour may be associated with
abnormalities in the renin-angiotensin system by way of corticotrophin-
releasing hormone (Saavedra et al., 2004). Angiotensin converting enzyme
(ACE) gene is one of its chief components. Mutation in intron 16 of ACE
gene is associated with increased suicide risk.
Mutations in nitric oxide synthase (NOS) 1 and 3 are found to be
associated with an increased risk of attempted and completed suicide
respectively.
Genes governing lipid and cholesterol metabolism are now being in-
vestigated as candidates for suicidal behaviour. Deficiency in the 7-
dehydrocholesterol reductase (DHCR7) enzyme involved in cholesterol
biosynthesis may be responsible for a fourfold increase in suicidal acts as
compared with controls (Lalovic et al., 2007). Apolipoprotein E4 (APOE4)
and cholecystokinin (CCK) genes are associated with suicide in geriatric
patients with depression and suicide in Japanese men respectively.
Biological sciences have come a long way in deciphering the mysteries
of mind behind behaviours such as aggression, impulsivity and suicide;
however, most of the studies cited have certain shortcomings which must
be kept in mind before conclusions are drawn. First, many post-mortem
studies have been conducted which have a high number of variables
associated with them, namely, the time after death when the studies were
conducted, the chemicals which have been used to preserve and process
the tissue sample and the manner of suicide (whether by ingestion of poi-
sonous substances or slashing wrists) as it may have an impact on process
of decomposition of the body, etc. Second, not many studies have tried
to take into account the multiple factors that are contributing to cause
suicidal behaviour at one particular moment of time. They usually focus
on one factor which prevents us from developing a multi-dimensional
theory regarding the cause of suicide. Although a lot of information has
been accumulated in the last 30 years or so which will help in our under-
standing of suicidal behaviour, how it will lead to better patient care in
the near future still remains to be seen. A laboratory test to accurately
predict suicide in psychiatric patients could be a vital tool in prevention
of suicides but so far it remains elusive.

55
Jitendra Kumar Trivedi and Sannidhya Varma

NEUROBIOLOGICAL MODELS OF SUICIDE

Requirements of a biological model are that it should be:

1. clinically explanatory,
2. biologically correlated and
3. testable in biological and clinical studies (Mann et al., 1999).

A model meeting the foregoing criteria was proposed by Mann and his
colleagues (1999), based on the following key observations:

1. About 90 percent of the suicide victims are known to have a psy-


chiatric disorder.
2. Most of the patients with psychiatric disorders do not make any at-
tempts to commit suicide.
3. The objective severity of symptoms is not predictive of suicidal
behaviour.
4. In a factor analysis, a combination of aggression and impulsivity
appear to be the most important predicting factors of suicide.
5. Family studies have shown that the inheritance of suicidal behaviour
is independent of the inheritance of any psychiatric disorder.

It has also been found that the people who have attempted or died of
suicide show abnormalities of the (a) PFC which is involved with control
of impulse and aggression, and (b) serotonergic system of the brain which
is also found in aggressive and violent subjects.
Based on the aforementioned findings, the model postulates two
independent components working together in suicidal behaviour:

1. Stressor: The calamity which makes life an ordeal (acute psychiatric


illness, drugs, alcohol, medical illness, family or social stress).
2. Diathesis: Vulnerability to stressor.

With regard to sensitivity to life events, early studies focused on the


hypothesis that a generalised cognitive rigidity (characterised by per-
severation, an inability to tolerate uncertainties, difficulty with changes,
restricted interests, poor judgement and difficulty in taking other persons

56
Neurobiological Basis of Suicidal Ideation

point of view) mediates the relationship between stressful life events and
suicidal behaviour (Brent et al., 1996). However, more recent findings
are consistent with the possibility that among people with depression
those who attempt suicide differ from those who do not on some but not
all neuropsychological tests (King et al., 2000). Using a modified Stroop
task, Becker, Strohbach and Rinck (1999) found that the level of suicidal
ideation in people with depression correlated particularly with biases in
the selective attention (a cognitive process through which a part of the
vast amount of information that we are receiving at a given time is selected
for processing). Another study could not demonstrate any difference in
attention measures between suicide attempters and non-attempters in a
group of people with depression (Becker et al., 1999). Although clearly
much more research is needed, these findings suggest a role of attentional
bias in the development of suicidal ideation—but not suicidal behaviour—
in people with depression.
This diathesis is necessary for suicide but not sufficient to produce it
all by itself. Many patients with increased vulnerability to stress do not
commit suicide. This diathesis can be considered a tendency to take a
decisive action in response to a stressor; an action which is more often than
not aggressive and impulsive, due to a lowered threshold for motor ac-
tivation, decreased inhibitory circuits, or an aggressive style of decision-
making towards self or others.
In spite of this, all psychiatric patients and those having suffered a loss
do not attempt suicide. There probably are powerful protective mech-
anisms that prevent most of us from taking such a step even when we
are faced with great stressors in life. Shakespeare describes this protective
mechanism in his literary work Hamlet where the main character hesitates
in attempting suicide even though he had the stressor of discovering
his father murdered and his succession to the throne of the kingdom
challenged, as follows:

Thus conscience makes cowards of us all;


And thus the native hue of resolution
Is sicklied o’er with the pale cast of thought,
And enterprises of great pith and moment
With this regard their currents turn away
And lose the name of action (Hamlet, Act-III, Scene 1)

57
Jitendra Kumar Trivedi and Sannidhya Varma

In this example from Hamlet, there was a stressor but probably an


absence of diathesis, which prevented Hamlet from attempting suicide.
Suicide tends to occur in circumstances where both the stressor and
diathesis are present (see Figure 3.1).

Figure 3.1 A Stress-diathesis Model of Suicide

Source: Amsel and Mann, 2000; Courtesy of New Oxford Textbook of Psychiatry, p.1045.
(By permission of Oxford University Press.)

CONCLUSION

Finding the biological markers is vital for better detection of at-risk patients
and their subsequent appropriate management. There is some evidence
that indicates that neurobiological factors especially may be associated with
suicide. However, there is as yet nothing concrete in the findings of the
extensive studies that have been carried out till date, which would allow
us to detect and manage people at risk of suicide. Same is true for a large
part of psychiatry, where there is no clear line of demarcation between
different disorders in terms of biology even though their symptoms might
be different. In the end, it can be concluded that though the path to finding
reliable biological markers of suicide is a tortuous one, we must make an
endeavour to tread it as best as we can.

58
Neurobiological Basis of Suicidal Ideation

ACKNOWLEDGEMENTS

The authors would like to thank Dr Mohan Dhyani, MD, Senior Resident,
Department of Psychiatry, Lady Hardinge Medical College, New Delhi,
and Dr Himanshu Sareen, Junior Resident, Department of Psychiatry for
their invaluable inputs.

REFERENCES

Amsel, L. and J.J. Mann (2000). Biological aspects of suicide. In M.G. Gelder, J.J. Lopez-
Ibor and N. Andreasen (Eds), New oxford textbook of psychiatry (p. 1045). Oxford:
Oxford University Press.
Anguelova, M., C. Benkelfat and G. Turecki (2003). A systematic review of association
studies investigating genes coding for serotonin receptors and the serotonin
transporter: II. Suicidal behavior. Molecular Psychiatry, 8(7), 646–53.
Arango, V. and M.D. Underwood (1997). Serotonin chemistry in the brain of suicide
victims. In R. Maris, M. Silverman and S. Canetto (Eds), Review of suicidology
(pp. 237–50). New York: Guilford Press.
Arango, V., M.D. Underwood and J.J. Mann (1997). Biologic alterations in the brainstem
of suicides. The Psychiatric Clinics of North America: Suicide, 20(3), 581–94.
Arango, V., M.D. Underwood, A.V. Gubbi and J.J. Mann (1995). Localized alterations
in pre- and postsynaptic serotonin binding sites in the ventrolateral prefrontal
cortex of suicide victims. Brain Research, 688(1), 121–33.
Asberg, M. (1997). Neurotransmitters and suicidal behaviour: The evidence from
cerebrospinal fluid studies. Annals of the New York Academy of Sciences, 836(1),
158–81.
Baca-Garcia, E., C. Vaquero, C. Diaz-Sastre, L. Jimenez-Trevino, J. de Leon, J. Saiz-Ruiz
et al. (2004). Lack of association between polymorphic variations in the alpha 3
subunit GABA receptor gene (GABRA3) and suicide attempts. Progress in Neuro-
psychopharmacology & Biological Psychiatry, 28(2), 409–12.
Baldessarini, R.J. and J. Hennen (2004). Genetics of suicide: An overview. Harvard
Review of Psychiatry, 12(1), 1–13.
Banki, C.M. and M. Arato (1983). Amine metabolite & neuroendocrine response related
to depression and suicide. Journal of Affective Disorders, 5(3), 223–32.
Becker, E.S., D. Strohbach and M. Rinck (1999). A specific attentional bias in suicide
attempters. Journal of Nervous and Mental Disease, 187(12), 730–35.

59
Jitendra Kumar Trivedi and Sannidhya Varma

Bellivier, F., P. Chaste and A. Malafosse (2004). Association between the TPH gene
A218C polymorphism and suicidal behavior: A meta-analysis. American Journal
of Medical Genetics, Part-B, Neuropsychiatric Genetics, 124(1), 87–91.
Biegon, A., M. Hanau, V. Greenberger and M. Segal (1989). Ageing and brain cholinergic
muscarinic receptor subtypes: An autoradiographic study in the rat. Neurobiology
of Aging, 10(4), 305–10.
Brent, D.A., J. Bridge, B.A. Johnson and J. Connolly (1996). Suicidal behavior runs
in families. A controlled family study of adolescent suicide victims. Archives of
General Psychiatry, 53(12), 1145–52.
Brezo, J., T. Klempan and G. Turecki (2008). The genetics of suicide: A critical review
of molecular studies. Psychiatric Clinics of North America, 31(2), 179–203.
Brown, G.L., L. Gerald, I. Markku, M. Linnoila, K. Frederick and F. Goodwin (1992).
Impulsivity, aggression and associated affects: Relationship to self-destructive
behavior and suicide. In R.W. Maris, A.L. Berman, J.T. Maltsberger and R.I. Yufit
(Eds), Assessment and prediction of suicide (pp. 589–606). New York: Guilford.
Bunney, W. and J. Fawcett (1965). Possibility of a biochemical test for suicide potential.
Archives of General Psychiatry, 13(3), 232–39.
Canli, T., E. Congdon, L. Gutknecht, R.T. Constable and K.P. Lesch (2005). Amygdala
responsiveness is modulated by tryptophan hydroxylase-2 gene variation. Journal
of Neural Transmission, 112(11), 1479–85.
Coccaro, E.F. and R.J. Kavoussi (1994). Neuropsychopharmacologic challenge in
biological psychiatry. Clinical Chemistry, 40(2), 319–27.
de Lara C., A. Dumais, G. Rouleau, A. Lesage, M. Dumont, N. Chawky et al. (2006).
STin2 variant and family history of suicide as significant predictors of suicide
completion in major depression. Biological Psychiatry, 59(2), 114–20.
De Luca, V., D. Voineskos, G.W. Wong, T. Shinkai, C. Rothe, J. Strauss et al. (2005).
Promoter polymorphism of second tryptophan hydroxylase isoform (TPH2) in
schizophrenia and suicidality. Psychiatry Research, 134(2), 195–98.
De Luca, V., G. Zai, S. Tharmalingam, A. de Bartolomeis, G. Wong and J.L. Kennedy
(2006). Association study between the novel functional polymorphism of the
serotonin transporter gene and suicidal behaviour in schizophrenia. European
Neuropsychopharmacology, 16(4), 268–71.
De Luca, V., P. Muglia, M. Masellis, J.E. Dalton, G.W. Wong and J.L. Kennedy (2004).
Polymorphisms in glutamate decarboxylase genes: analysis in schizophrenia.
Psychiatric Genetics, 14(1), 39–42.
De Luca, V., S. Tharmalingam and J.L. Kennedy (2007). Association study between
the corticotrophin releasing hormone receptor 2 gene and suicidality in bipolar
disorder. European Psychiatry, 22(5), 282–87.
Deakin, B. and C.D. Ben (2003). Biological aspects of suicide and suicidal behaviour.
Psychiatry, 1(1), 28–30.
D’Haenan, H. (2001). Imaging the serotonergic system in depression. European Archives
of Psychiatry and Clinical Neurosciences, 251 (Supplement 2), 76–80.

60
Neurobiological Basis of Suicidal Ideation

Dwivedi, Y., H.S. Rizavi, P.K. Shukla, G. Lyons, M. Faludi, A. Palkovits et al. (2004).
Protein kinase A in postmortem brain of depressed suicide victims: Altered
expression of specific regulatory and catalytic subunits. Biological Psychiatry, 55(3),
234–43.
Dwivedi, Y., H.S. Rizavi, R.R. Conley, R.C. Roberts, C.A. Tamminga and G.N. Pandey
(2003). Altered gene expression of brain-derived neurotrophic factor and receptor
tyrosine kinase B in postmortem brain of suicide subjects. Archives of General
Psychiatry, 60(8), 804–15.
Fuster, J.M. (1997). The prefrontal cortex: Anatomy, physiology, and neuropsychology
of the frontal lobe (3rd Ed.). New York: Lippincott-Raven.
Ho, L.W., R.A., Furlong, J.S. Rubinsztein, C. Walsh, E.S. Paykel and D.C. Rubinsztein
(2000). Genetic associations with clinical characteristics in bipolar affective
disorder and recurrent unipolar depressive disorder. American Journal of Medical
Genetics, 96(1), 36–42.
Hong, C.J., G.M. Pan and S.J. Tsai (2004). Association study of onset age, attempted
suicide, aggressive behavior, and schizophrenia with a serotonin 1B receptor
(A-161T) genetic polymorphism. Neuropsychobiology, 49(1), 1–4.
Hsiung, S.C., M. Adlersberg, V. Arango, J.J. Mann, H. Tamir and K.P. Liu (2003).
Attenuated 5HT1a receptor signalling in brains of suicide victims: Involvement of
adenylyl cyclase, phosphatidylinositol 3-kinase, Akt and mitogenactivated protein
kinase. Journal of Neurochemistry, 87(1), 182–94.
Joseph, A., S. Abraham, J.P. Muliyil, K. George, J. Prasad, S. Minz et al. (2003).
Evaluation of suicide rates in rural India using verbal autopsies, 1994-9. British
Medical Journal, 326(7399), 1121–22.
Kamali, M., M.A. Oquendo and J.J. Mann (2001). Understanding the neurobiology of
suicidal behaviour. Depression and Anxiety, 14(3), 164–76.
Kaplan, J.R., M.F. Muldoon, S.B. Manuck and J.J. Mann (1997). Assessing the observed
relationship between low cholesterol and violence related mortality. Annals of the
New York Academy of Sciences, 836, 57–80.
Karege, F., N. Perroud, S. Burkhardt, M. Schwald, E. Ballmann, R. La Harpe et al.
(2007). Alteration in kinase activity but not in protein levels of protein kinase b
and glycogen synthase kinase-3_ in ventral prefrontal cortex of depressed suicide
victims. Biological Psychiatry, 61(2), 240–45.
Kim, C.D., M. Seguin, N. Therrien, G. Riopel, N. Chawky, D. Alain et al. (2005). Familial
aggregation of suicidal behavior: a family study of male suicide completers from
the general population. American Journal of Psychiatry, 162(5), 1017–19.
King, D.A., Y. Conwell, C. Cox, R.E. Henderson, D.G. Denning and E.D. Caine (2000).
A neuropsychological comparison of depressed suicide attempters and non-
attempters. Journal of Neuropsychiatry and Clinical Neurosciences, 12(1), 64–70.
Klein, P.S. and D.A. Melton (1996). A molecular mechanism for the effect of lithium
on development. Proceedings of the National Academy of Sciences, U.S.A, 93(16),
8455–59.

61
Jitendra Kumar Trivedi and Sannidhya Varma

Klempan, T. and G. Turecki (2005). Suicide: A neurobiological point of view. Revista


Brasileira de Psiquiatria, 27(3), 172–73.
Krieger, G. (1974). The plasma level of cortisol as a predictor of suicide. Diseases of Ner-
vous System, 35(5), 237–40.
Lalovic, A., E. Levy, G. Luheshi, L. Canetti, E. Grenier, A. Sequeira et al. (2007).
Cholesterol content in brains of suicide completers. International Journal of Neuro-
psychopharmacology, 10(2), 159–66.
Lalovic, A., L. Merkens, L. Russell, G. Arsenault-Lapierre, M. J. Nowaczyk, F.D. Porter et al.
(2004). Cholesterol metabolism and suicidality in Smith-Lemli-Opitz syndrome
carriers. American Journal of Psychiatry, 161(11), 2123–26.
Li, D. and L. He (2007). Meta-analysis supports association between serotonin
transporter (5-HTT) and suicidal behavior. Molecular Psychiatry, 12(1), 47–54.
Li, D., Y. Duan and L. He (2006). Association study of serotonin 2A receptor (5-HT2A)
gene with schizophrenia and suicidal behavior using systematic meta-analysis.
Biochemical and Biophysical Research Communication, 340(3), 1006–15.
Li, X., Zhu, W., M.S. Roh, A.B. Friedman, K. Rosborough and R.S. Jope (2004). In vivo
regulation of glycogen synthase kinase-3_ (GSK-3_) by serotoninergic activity in
mouse brain. Neuropsychopharmacology, 29(8), 1426–31.
Linnoila, M. and M. Virkkunen (1992). Biologic correlates of suicidal risk and aggressive
behavioural traits. Journal of Clinical Psychopharmacology, 12(2), 19S–20S.
Lopez, J.F., M. Palkovitz, M. Arato, A. Mansour, H. Akil and S.J. Watson (1990). Localiza-
tion and quantification of pro-opiomelanocortin mRNA and glucocorticoid
receptor mRNA in pituitary of suicide victims. Neuroendocrinology, 56(4), 491–501.
Mann, J.J. (1998). The Neurobiology of Suicide. Nature Medicine, 4(1), 25–30.
Mann, J.J., C. Waternaux, G.L. Haas and K.M. Malone (1999). Towards a clinical model
of suicidal behavior in psychiatric patients. American Journal of Psychiatry, 156(2),
181–89.
Mann, J.J., R.A. Henteleff, T.F. Lagattuta, J.A. Perper, S. Li and V. Arango (1996).
Lower 3H-paroxetine binding in cerebral cortex of suicide victims is partly due
to a fewer high- affinity, non-transporter sites. Journal of Neural Transmission,
103(11), 1337–50.
Mann, J.J., Y.Y. Huang, M.D. Underwood, S.A., Kassir, S. Oppenheim, T.M. Kelly et al.
(2000). A serotonin transporter gene promoter polymorphism (5-HTTLPR) and
prefrontal cortical binding in major depression and suicide. Archives of General
Psychiatry, 57(8), 729–38.
Maris, R.W. (2002). Suicide, The Lancet, 360(9329), 319–26.
McIntosh, J.L. (2001). USA Suicide 1999, Official Final Data. Washington, DC: American
Association of Suicidology.
Mittendorfer-Rutz, E., F. Rasmussen and D. Wasserman (2007). Familial clustering of
suicidal behaviour and psychopathology in young suicide attempters: a register-
based nested case control study. Social Psychiatry and Psychiatric Epidemiology,
43, 28–36.

62
Neurobiological Basis of Suicidal Ideation

Moscicki, E. (2001). Epidemiology of suicide. In S. Goldsmith (Ed.). Risk factors for sui-
cide (pp. 1–4). Washington DC: National Academy Press.
National Crime Bureau, Ministry of Home Affairs, Government of India, (2007).
Accidental Deaths & Suicides in India, 2007. Retrieved 5 February 2009 from
http://ncrb.nic.in/ADSI2007/Suicides07.pdf
Nemeroff, C.B., M.J. Owens, G. Bissette, A.C. Andorn and M. Stanley (1988). Reduced
corticotropin releasing factor binding sites in the frontal cortex of suicide victim.
Archives of General Psychiatry, 45(6), 577–79.
Nishiguchi, N., O. Shirakawa, H. Ono, A. Nishimura, H. Nushida, Y. Ueno et al. (2002).
Lack of an association between 5-HT1A receptor gene structural polymorphisms
and suicide victims. American Journal Medical Genetics, 114(4), 423–25.
Ohtani, M., S. Shindo and N. Yoshioka (2004). Polymorphisms of the tryptophan
hydroxylase gene and serotonin 1A receptor gene in suicide victims among
Japanese. Tohoku Journal of Experimental Medicine, 202(2), 123–33.
Pandey, G.N. (1997). Altered serotonin function in suicide. Evidence from platelet &
neuroendocrine studies. Annals of the New York Academy of Sciences, 836, 182–201.
Pandey, G.N., Y. Dwivedi, X. Ren, H.S. Rizavi, R.C. Roberts, R.R. Conley et al. (2007).
Cyclic AMP response element-binding protein in post-mortem brain of teenage
suicide victims: specific decrease in the prefrontal cortex but not the hippocampus.
International Journal of Neuropsychopharmacology, 10(5), 621–29.
Pfennig, A., H.E. Kunzel, N. Kern, M. Ising, M. Majer, B. Fuchs et al. (2005).
Hypothalamus-pituitary-adrenal system regulation and suicidal behavior in
depression. Biological Psychiatry, 57(4), 336–42.
Saavedra, J.M., H. Ando, I. Armando, G. Baiardi, C. Bregonzio, M. Jezova et al. (2004).
Brain angiotensin II, an important stress hormone: Regulatory sites and therapeutic
opportunities. Annals of New York Academy of Science, 1018, 76–84.
Sargent, P.A., K.H. Kjaer, C.J. Bench, R.A. Eugenii, M. Cristina, J. Meyer et al. (2000).
Brain serotonin 1A receptor binding measured by PET with [11C] WAY-100635:
effects of depression and antidepressant treatment. Archives of General Psychiatry,
57(2), 174–80.
Schulsinger, F., S.S. Key, D. Reoshental and P.H. Wender (1979). A family study of
suicide. In M. Schou and E. Stromgren (Eds), Origins, prevention, and treatment
of affective disorders (pp. 277–87). New York: Academic Press.
Serretti, A., L. Mandelli, I. Giegling, S. Schneider, A. M. Hartmann, A. Schnabel et al.
(2007). HTR2C and HTR1A gene variants in German and Italian suicide attempters
and completers. American Journal Medical Genetics, Part-B, Neuropsychiatric
Genetics, 144(3), 291–299.
Shaltiel, G., A. Shamir, G. Agam and R.H. Belmaker (2005). Only tryptophan hy-
droxylase (TPH)-2 is relevant to the CNS. American Journal Medical Genetics,
Part-B, Neuropsychiatric Genetics, 136(1), 106.
Stambolic, V., L. Ruel and J.R. Woodgett (1996). Lithium inhibits glycogen synthase
kinase-3 activity and mimics wingless signalling in intact cells. Current Biology,
6(12), 1664–68.

63
Jitendra Kumar Trivedi and Sannidhya Varma

Strickland, P.L., J.F.W. Deakin, C. Percival, J. Dixon, R.A. Gator and D.P. Goldberg (2002).
Biosocial origins of depression in the community. Interaction between social
adversity, cortisol and serotonin neurotransmission. British Journal of Psychiatry,
180(2), 168–73.
Stuss, D.T., G. Gallup and M.P. Alexander (2001). The frontal lobes are necessary for
‘theory of mind’. Brain, 124(2), 279 –86.
Tanney, B.L. (2000). Psychiatric diagnoses and suicide. In R.W. Maris, A.L. Berman
and M.M. Silverman (Eds), Comprehensive textbook of suicidology (pp. 311–41).
New York: Guilford.
Tsai, S.J., C.J. Hong, Y.W. Yu, T. J. Chen, Y.C. Wang and W.K. Lin (2004). Associ-
ation study of serotonin 1B receptor (A-161T) genetic polymorphism and sui-
cidal behaviors and response to fluoxetine in major depressive disorder.
Neuropsychobiology, 50(3), 235–38.
Turecki, G., A. Sequeira, Y. Gingras, M. Séguin, A. Lesage, M. Tousignant et al. (2003).
Suicide and serotonin: study of variation at seven serotonin receptor genes in
suicide completers. American Journal Medical Genetics, Part-B, Neuropsychiatric
Genetics, 118(1), 36–40.
UC Atlas of Global Inequality (2002). Cause of Death, Leading Causes of Death in 2001.
Retrieved 13 February from http://ucatlas.ucsc.edu/cause.php
Wasserman, D., T. Geijer, M. Sokolowski and J. Wasserman (2007). Genetic variation
in the hypothalamic-pituitary- adrenocortical axis regulatory factor, T-box 19, and
the angry/hostility personality trait. Genes, Brain and Behavior, 6(4), 321–28.
Wender, P.H., S.S. Kety, D. Rosenthal, F. Schulsinger, J. Ortmann and I. Lunde (1986).
Psychiatric disorders in the biological and adoptive families of adopted individuals
with affective disorders. Archives of General Psychiatry, 43(10), 923–29.
Videtic, A., G. Pungercic, I.Z. Pajnic, T. Zupanc, J. Balazic, M. Tomori et al. (2006). Asso-
ciation study of seven polymorphisms in four serotonin receptor genes on suicide
victims. American Journal Medical Genetics, Part-B, Neuropsychiatric Genetics,
141(6), 669–72.
Zill, P., U.W. Preuss, G. Koller, B. Bondy and M. Soyka (2007). SNP- and haplotype
analysis of the tryptophan hydroxylase 2 gene in alcohol-dependent patients and
alcohol-related suicide. Neuropsychopharmacology, 32(8), 1687–94.

64
4

Problem-solving Ability and Repeated


Deliberate Self-harm
C RM M A

S uicidal behaviour including deliberate self-harm (DSH) can be


conceptualised as a maladaptive coping response (Sakinofsky, 2000)
in which the person’s overarching motive is to escape from a problem
situation (Baumeister, 1990; Williams and Pollock, 2001) to solve a prob-
lem (Applebaum, 1963; Chiles and Strosahl, 2004), or to avoid or relieve
unpleasant emotions (Bancroft et al., 1976, 1979; Chapman et al., 2006;
McAuliffe et al., 2007). Deliberate self-harm and its repetition pose a major
challenge to mental health and social services internationally (Madge et al.,
2008; Sakinofsky, 2000). Repetition accounts for a significant proportion of
hospital-treated DSH episodes and is regarded as the central characteristic
of self-harm (Kerkhof, 2000). More recent evidence shows that repetition
is also common among adolescents who do not present to medical services
following a DSH episode (Madge et al., 2008). Repeated DSH increases
the risk of subsequent suicide (Hawton and Fagg, 1988; Maris, 1992;
Nordentoft et al., 1993; Zahl and Hawton, 2004) and there is evidence that
rates of repetition are increasing (Hawton, Harris, Hall, Simkin, Bale and
Bond, 2003; Henriques et al., 2004; National Suicide Research Foundation,
2008; Schmidtke et al., 2004). A pragmatic response to this problem has
been to identify the characteristics of individuals who engage in repeated
suicidal behaviour in order to better inform the development of effective

65
Carmel McAuliffe

interventions for its prevention and treatment. Although the association


between problem-solving difficulties and repeated self-harm is poorly
understood (Pollock and Williams, 1998; Sakinofsky, 2000) there
is increasing evidence that treatment interventions which incorporate
problem-solving skills training are effective in the prevention of repeated
self-harm (Arensman et al., 2001; Brown et al., 2005; Hawton et al., 1998;
McLeavey et al., 1994; Salkovskis et al., 1990; Slee et al., 2008). The current
chapter will review the existent research evidence of the role of problem-
solving difficulties in self-harming behaviour and its repetition and
describe a randomised controlled trial of group problem-solving skills
training which was designed to prevent repeated DSH.
In order to understand the association between problem-solving dif-
ficulties and repetition of DSH it is important to first consider the problem-
solving process and how problem-solving ability acts as a vulnerability
factor for DSH. Given the considerable variation in terminology used to
describe different forms of suicidal behaviour in the scientific literature
(O’Carroll et al., 1996), in this chapter the terms ‘deliberate self-harm’
(DSH) and ‘self-harm’ will be used to reflect the diversity of motives
associated with this behaviour, without making any assumptions about
suicide intent.

THE PROBLEMͳSOLVING PROCESS

What role does problem-solving ability play in vulnerability to DSH? The


following early definition of the problem-solving process (D’Zurilla and
Goldfried, 1971: 108) gives some clues: ‘… a behavioural process, whether
overt or cognitive in nature, which a) makes available a variety of potentially
effective response alternatives for dealing with the problematic situation
and b) increases the probability of selecting the most effective response
from among these various alternatives.’ This process enables people to re-
spond appropriately in situations they have not previously experienced or
to problems where no immediate solutions are apparent. D’Zurilla and
colleagues (D’Zurilla, 1986; D’Zurilla and Goldfried, 1971; D’Zurilla and
Nezu, 1990) in their stage sequential model of social problem solving dis-
tinguish two main processes: (a) problem orientation and (b) problem
solving. Problem orientation refers to a set of relatively stable schemata

66
Deliberate Self-harm

governing how people relate to problems in living and their thoughts


and feelings with regard to their own problem solving ability. Problem-
solving relates to a person’s application of problem-solving skills aimed
at finding the ‘best’ or most appropriate solution for a specific problem
(D’Zurilla and Chang, 1995). The five stage model outlined by D’Zurilla
and Nezu (1990) includes problem orientation, generation of alterna-
tive solutions, decision-making, solution implementation and solution
verification. As Pollock and Williams (1998) explain, these stages do not ne-
cessarily illustrate how healthy people engage in problem solving but they
do provide an explanation for the particular parts of this process that can
fail. People who engage in DSH have cognitive characteristics which under-
mine this process, including greater cognitive rigidity (McLeavey et al.,
1987; Neuringer, 1964; Patsiokas et al., 1979), greater dichotomous think-
ing (Neuringer and Lettieri, 1971) and field dependence (Patsiokas et al.,
1979); overgeneral autobiographical memory (Evans et al., 1992; Pollock
and Williams, 2001); poorer performance on means–ends thinking
(Dieserud et al., 2001; Linehan et al., 1987; McLeavey et al., 1987; Pollock and
Williams, 2004) and lower self-appraised problem-solving ability (Dieserud
et al., 2001; McLeavey et al., 1987).

PROBLEMͳSOLVING ABILITY AS A VULNERABILITY


FACTOR FOR SUICIDAL BEHAVIOUR

Clearly, not all individuals with poor problem-solving ability engage in


DSH. As Williams and Pollock (2001) point out, problem-solving dif-
ficulties in and of themselves are not important, but when they signify
to the person that there is ‘no escape’ they become important. There is
substantial evidence that problem-solving ability mediates the relationship
between stress and DSH, whereby individuals with poor problem-solving
ability under chronic stress are more likely to become hopeless and/or
suicidal (Sandin et al., 1998; Schotte and Clum, 1982, 1987). There is also
evidence that good problem solving—as a stable trait characteristic—
protects against DSH, independently of depression or hopelessness
levels (Dieserud et al., 2001). Problem-solving ability is important in the
association between environmental factors—for example, poor parenting
and early exposure to stressful life events—and subsequent suicidal

67
Carmel McAuliffe

behaviour, as it helps to distinguish those who are more likely to engage in


DSH. Furthermore, early exposure to chronic stress may adversely affect
the development of problem-solving ability (Carriss et al., 1998; Clum
et al., 1979; Yang and Clum, 2000).
A number of models based on the original interactional model of
suicidal behaviour by Braucht (1979), incorporating both individual and
environmental factors, outline how problem-solving ability influences
other associated risk factors for DSH in setting the context in which DSH
can occur. For example, Schotte and Clum (1982, 1987) developed their
seminal diathesis-stress-hopelessness model of suicidal behaviour in which
cognitive rigidity mediates the relationship between stress and suicidal
ideation, based on the finding that college students who reported high
levels of negative life stress and were poor interpersonal problem-solvers
also had the highest levels of hopelessness and suicidal ideation. They
retested this model with a sample of suicidal psychiatric patients (Schotte
and Clum, 1987) and found that they generated fewer than half as many
potential solutions to interpersonal problems as matched controls. Levels
of stress were positively correlated with levels of hopelessness and suicide
ideation.
Some interactional models emphasise the formative influence of early
life experiences including parenting and adverse childhood experiences
on problem-solving ability. Clum and colleagues (1979) focused on the
interactions between early learning, the development of certain response
sets conducive to DSH and precipitating environmental stressors. They
described a developmental path analysis model based on a review of the
empirical evidence, outlining multiple paths to DSH. One of their ex-
amples includes a parental role model for suicidal behaviour and a period
of stress during adolescence leading to DSH. Another path they describe
involves reward of avoidant behaviour combined with low reinforcement
of problem solving, leading to cognitive rigidity, which—they argue—in
the face of chronic or accumulated stress combined with an environment
lacking in supports, increases the likelihood of DSH. Clum and colleagues
point out that these paths are overlapping rather than mutually exclusive.
Yang and Clum (2000) found that the impact of early negative life events
on suicidal behaviour in adulthood was considerably stronger when
examined through their impact on an individual’s cognitive function-
ing (including self-esteem, locus of control, hopelessness and problem-
solving difficulties) than through their direct impact on suicidal behaviour.

68
Deliberate Self-harm

They also found that early negative life events affected cognitive difficulties,
whereas current life events did not. They hypothesised that high levels of
stress early in life may interfere with the development of effective problem-
solving skills, leading to an internalised model of lower perceived personal
efficacy and increased helplessness and hopelessness. Sandin and colleagues
(1998) describe a stress process model, outlining the impact of psychosocial
stress (major life events, daily hassles and chronic stressors) on suicidal
behaviour by integrating mediating variables (including negative appraisal,
coping, problem solving and hopelessness) and moderating variables
(including social supports and individual characteristics). They suggest
that coping, which includes problem solving along with emotion-focussed
and appraisal-focussed strategies, and hopelessness are critical end points
of a causal mechanism leading to suicidal behaviour. Carriss and colleagues
(1998) tested a mediational model of family rigidity, adolescent problem-
solving difficulties and suicidal ideation. They found that family rigidity
affects adolescent suicidal ideation indirectly through its effect on ado-
lescent problem-solving ability as measured by the problem-solving
inventory (Heppner, 1988). However, the study was cross-sectional, and
family rigidity, adolescent problem solving and adolescent suicidal idea-
tion were assessed concurrently.
Problem-solving ability appears to be an important mediator of the
relationship between psychosocial stress and suicidal behaviour. How-
ever, the effect of low mood on problem-solving ability is an important
consideration and we now turn our attention to the possible effects of
hopelessness on problem solving among those who engage in suicidal
behaviour.

HOPELESSNESS

One explanation for the observed problem-solving difficulties among


self-harm patients is that they are brought about by low mood at the time
of a self-harm episode. However, hopelessness seems to operate inde-
pendently of problem-solving difficulties. For example, Dieserud and
colleagues (2001) found evidence that hopelessness and problem-solving
ability are part of two separate processes in a two-path model of DSH
in a clinical sample. The first path was one of depression/hopelessness,

69
Carmel McAuliffe

in which low self-esteem, loneliness and divorce collectively increased


vulnerability to the development of depression which was further mediated
by hopelessness and suicidal ideation in its relationship with DSH. The
second path was one of problem-solving difficulties, separate from the
path mediated by depression, hopelessness and suicidal ideation. This
cognitive path was based on low self-esteem and a low sense of self-efficacy.
In this second separate path they found that negative self-appraisal of
problem-solving ability and poor interpersonal problem-solving skills
mediates the relationship between low self-esteem/low self-efficacy and
DSH. There were no significant associations between the problem-solving
variables and the depression/hopelessness variables when the distal vari-
ables (low self-esteem, low self-efficacy, loneliness and separation/divorce)
were controlled for. Earlier studies have also found that correlations between
hopelessness and interpersonal problem-solving skills are mainly non-
significant (MacLeod and Williams, 1992; Schotte and Clum, 1987).
Another study (Pollock and Williams, 2004) comparing a group who
had engaged in DSH with a non-suicidal psychiatric control group found
poorer social problem-solving ability among the DSH group; and this
difference persisted while levels of depression, hopelessness and suicide
ideation reduced following an episode of DSH.
Hopelessness may be more important for problem orientation—the
aspect of problem solving governing how people approach problems
rather than problem-solving performance. When hopelessness is high, the
range of alternatives that the individual perceives to be available to him
or her may be restricted. In one study, depressed patients suffering from
severe hopelessness—compared with those who had lower hopelessness
scores—indicated that more life areas were presenting as problems to
them (Nekanda-Trepka et al., 1973). They were also inclined to hold more
negative expectations of the prospect of solving their problems, and of a
poorer outcome if preoccupying problems did not improve. They also
felt less competent about effecting an improvement (poorer self-efficacy)
than did those who scored lower on hopelessness. Rudd, Rajab and Dahm
(1994) report that suicide ideators similarly tend to focus on potentially
negative consequences of implementing alternative solutions. Hopelessness
may therefore influence problem orientation in such a way that the person
not only tends to perceive more life issues as problematic but also feels
unable to alleviate them and predicts negative consequences of their
attempts to solve problems. Psychosocial treatment interventions that use
problem-solving skills training need to address hopelessness by getting

70
Deliberate Self-harm

clients to set realistic goals and by reinforcing clients’ efforts to solve


their problems.
To summarise, hopelessness may influence problem orientation
among those engaging in DSH as distinct from problem-solving per-
formance. This is important because it suggests that poor problem-solving
ability is a stable cognitive characteristic among those engaging in DSH
as opposed to compromised coping due to high levels of hopelessness.
However, there is evidence that suicide ideators with a history of depression
are vulnerable to deterioration in problem-solving performance as a result
of low mood (Williams et al., 2005) while in a separate study (Schotte
et al., 1990) a sample of suicide ideators and attempters showed significant
improvements between time of admission and one week later in levels of
depression, anxiety, hopelessness and suicide intent, which were associated
with improvements in problem-solving skills in one week post admission.
Based on these studies it appears that those with a history of depression and
suicidal ideation are a sub-group for whom mood may influence problem-
solving performance. In both studies, however, a large proportion of the
samples examined were suicide ideators without a history of DSH.

PROBLEMͳSOLVING ABILITY AS A MAINTENANCE


FACTOR FOR SUICIDAL BEHAVIOUR

Early exposure to adverse life events and specific styles of parenting, as


previously stated, are important in the development of problem-solving
skills (Carriss et al., 1998; Clum et al., 1979; Yang and Clum, 2000). This
is all the more relevant given the evidence that compared with first evers,
repeaters suffer greater psychosocial disadvantage from early in life and
chronic in nature (Arensman and Kerkhof, 2004b). Repeaters also have
a greater range of psychiatric disorders (Rudd et al., 1996) including co-
morbid psychiatric disorders (Hawton, Houston, Haw, Townsend and
Harriss, 2003) and more severe symptoms of depression (Rudd et al.,
1996), hopelessness (Arensman and Kerkhof, 2004a) and suicidal ideation
(Rudd et al., 1996). The risk factors related to repeated DSH therefore
comprise a heterogeneous set, of which the only common factor may
be the inclination to respond to a wide variety of stressful experiences
with a repeated act of self-harm (Clum et al., 1979; Sakinofsky, 2000).

71
Carmel McAuliffe

More recent work has shown that among female self-poisoners in contrast
to those presenting with a first DSH episode, repeated episodes of self-
harm are more autonomous and are less determined by the occurrence
of a specific stressful event (Crane et al., 2007).
While previous self-harm was included in the model tested by Dieserud
and colleagues (2001) the interactional models generally do not offer
explanations for repeated self-harm and have not tested the association
between problem-solving ability and DSH prospectively. One of the
few prospective studies of problem-solving difficulty in repeated DSH
found that repeaters view their problems as more insurmountable or
overwhelming and themselves as relatively powerless over their lives.
A tendency to perceive problems as more severe was the factor most pre-
dictive of repetition at three months in one prospective study of 228
consecutive DSH patients who were treated in hospitals (Sakinofsky and
Roberts, 1990). Based on a separate analysis from the same study, non-
repeaters reported a significantly greater number of improvements in
terms of personal change, financial situation, marriage, family and work,
but had no fewer reports of experiencing new stressful events (Sakinofsky
et al., 1990). Taken together these findings suggest that orientation to
problems—rather than the specific problem or event—is particularly
important in the case of repeaters of DSH and that a positive approach to
problems is likely to buffer against the effects of new emerging problems
in the aftermath of a self-harm episode.

EFFICACY OF PROBLEMͳSOLVING THERAPY IN


TREATING PEOPLE WHO SELFͳHARM

In one review of cognitive-behavioural therapy (CBT) interventions to


reduce suicidal behaviour, Tarrier and colleagues (2008) report that
14 of the 28 CBT studies reviewed included some aspect of problem-
solving training. Five of these trials focused uniquely on problem-solving
therapy. Early trials using problem-solving skills training (McLeavey et
al., 1994; Salkovskis et al., 1990) found promising results in reducing
repetition of DSH (Arensman et al., 2001; Hawton et al., 1998). How-
ever, these early studies suffered from the limitations of a small sample
size, making it impossible to detect any significant treatment effects

72
Deliberate Self-harm

on repetition. Townsend and colleagues (2001) reported on out-


comes other than repetition from a meta-analysis of six trials that tested
problem-solving therapy for those engaging in DSH. They found that in
comparison with control treatments, problem-solving therapy showed a
significantly better effect in lowering levels of depression and hopelessness
and significantly more of those assigned to problem-solving therapy
reported an improvement in problems. Trials of Dialectical Behaviour
Therapy for patients with Borderline Personality Disorder, which
incorporates elements of interpersonal problem-solving skills training,
have reported a significant reduction in rates of repeated DSH (Linehan
et al., 1991, 1993, 2006). In many of the more recent larger trials, problem-
solving interventions have more often been delivered as part of a broader
treatment package of CBT. One trial comparing usual care with cognitive
therapy based on 10 outpatient cognitive-therapy sessions that included a
problem-solving component reported a significantly lower re-attempt rate
in the cognitive-therapy group (Brown et al., 2005). The investigators also
reported a significant improvement in self-reported levels of depression
and hopelessness. There was no difference between treatment groups on
rates of suicidal ideation. Overall, 96.7 percent of participants received at least
one treatment session. In a separate study, Slee and colleagues (2008) exam-
ined a 12-session CBT programme. The programme was based on a
model of maintenance factors of DSH including negative thinking and
problem-solving difficulties to be modified by CBT. The authors reported
a significant reduction in the number of repeated DSH episodes in the
CBT group. There was full compliance with treatment sessions by all
participants assigned to the CBT treatment arm. The failure of an earlier
trial using a manual-assisted cognitive-therapeutic approach to demon-
strate a reduction in repeat episodes (Tyrer et al., 2003), in which over
one-third of the active treatment sample received a treatment manual
alone without any treatment sessions, suggests that reliance purely on a
self-help approach among repeaters of DSH is ineffective in reducing
repetition, in particular among DSH patients with a history of previous
DSH episodes (Arensman et al., 2004).
Overall, despite promising findings from trials that incorporate problem-
solving skills training, the number of studies focussing uniquely on a
problem-solving approach remains small. In studies where significant
effects have been found for CBT in reducing repeated suicidal behaviour,
there is still uncertainty regarding the therapeutic processes within CBT
that may be effective.

73
Carmel McAuliffe

METHODOLOGICAL ISSUES

A major methodological problem hindering research into the identifi-


cation of risk factors for repeated DSH is the use of heterogeneous samples,
combining for example, suicide ideators with deliberate self-harmers, or
repeaters with first evers (Rudd et al., 1996). Most studies that examine
the association between problem solving and repeated DSH use ‘process’
measures—which assess attitudes and skills in problem solving—as
opposed to ‘outcome’ measures of problem-solving ability (Rudd et al.,
1996) which assess problem-solving performance based on solution
efficacy. The use of a retrospective design makes it difficult to establish
if ‘risk’ factors are causal or consequential. In terms of the evidence base
for the efficacy of problem-solving therapy, large randomised controlled
trials of problem-solving therapy are lacking.
For these reasons we now turn our attention to three studies in which
we have investigated the association between problem-solving ability and
repeated DSH. A fourth study, examining the efficacy of a programme
of problem-solving therapy in preventing repeat episodes of DSH is also
described. These studies were carried out to address the following four
research questions:

1. Motives for repeated DSH: Do repeaters have distinct motives for


engaging in DSH; and a greater number of motives, indicative of
greater problem-solving difficulties, when compared with first evers?
2. Problem-solving orientation of repeaters of DSH: Do repeaters of
DSH have a different orientation to problems when compared with
non-repeaters and is their problem-solving style characterised by a
more negative problem-solving orientation generally?
3. Optional thinking ability: Can deliberate self-harmers who sub-
sequently repeat be identified prospectively on the basis of their
performance in problem solving?
4. Problem-solving skills training: Can a treatment programme based
specifically on problem-solving skills training in addition to standard
care significantly reduce repeated DSH compared with stand-
ard care alone?

In all four studies ‘deliberate self-harm’ was defined according to the


definition of parasuicide/attempted suicide devised by the World Health

74
Deliberate Self-harm

Organization (WHO) Working Group of the WHO/EURO Multicentre


Study on Suicidal Behaviour:

…an act with non-fatal outcome, in which an individual deliberately


initiates a non-habitual behavior that, without intervention from others,
will cause self-harm, or deliberately ingests a substance in excess of the pre-
scribed or generally recognised therapeutic dosage, and which is aimed
at realizing changes which the subject desired via the actual or expected
physical consequences. (Platt et al., 1992)

This definition includes acts that are interrupted before DSH is inflicted,
for example, a person removed from a bridge before jumping off, but
excludes episodes by individuals who do not understand the meaning or
the outcome of their act, for example, due to a learning disability or severe
mental disorder (Bille-Brahe et al., 1994). The terms ‘parasuicide’, ‘at-
tempted suicide’ and ‘deliberate self-harm’ were used interchangeably by
the WHO/EURO Multicentre Study on Suicidal Behaviour.

MoƟves for Repeated Deliberate Self-harm

We examined the motives reported for engaging in DSH by medically


treated DSH patients interviewed at one centre participating in the
Repetition-Prediction part of the WHO/EURO Multicentre Study on
Suicidal Behaviour (McAuliffe et al., 2007). Previous history of DSH
was established on the basis of specific questions asked during the initial
interview. Repeaters were identified as those who engaged in more than
one episode of self-harm before their initial interview—whether treated in
hospitals or not. First evers were defined as patients whose index
episode was their only known act of DSH at the time of the EPSIS I
interview (European Parasuicide Study Interview Schedule). The first
objective of this study was to establish if repeaters of DSH have a greater
number of motives for engaging in DSH compared with first ever
patients, as measured by the motives for parasuicide questionnaire (MPQ;
Kerkhof et al., 1993) administered at index episode. A greater number of
motives would be indicative of greater problem-solving difficulties in a
range of areas.
As expected, compared with first ever patients, repeaters reported sig-
nificantly more motives at the time of the index DSH episode, suggesting

75
Carmel McAuliffe

a wider range of difficulties than first evers. This is consistent with


Sakinofsky’s (2000) observation that what may be the only factor common
to all repeaters is a tendency to repeat acts of DSH in response to a wide
variety of non-specific stresses.
The second objective was to establish whether repeaters have distinct
motives for engaging in DSH compared with first ever patients. We found
this to be the case. There were important differences in the reported
motives for DSH according to the repeater status of DSH patients at index
episode. Repeaters significantly more often reported motives aimed at
escape from an unbearable mental state or an unbearable situation, by
death or in order to make things easier for others. Repeaters were more
likely to report motives of revenge against others—to make others pay and
to make others feel guilty—and an appeal motive to show someone how
much they loved them, to get help from someone, to know if someone
cared or to change someone’s mind. Self-harm patients with high scores
on the ‘escape’, ‘appeal’ and ‘revenge’ motive factors were at significantly
increased risk of being a repeater. Controlling for the effects of the other
motives, high scores on ‘escape’ and ‘revenge’ motives remained signifi-
cantly associated with previous DSH.

Clinical implicaƟons A large number of motives were reported by


DSH patients generally and by repeaters in particular, indicating coping
difficulties for the group as a whole, but with an even wider range of coping
difficulties among those who repeat. Routine risk assessment of DSH
patients should include specific questions regarding a broad range of
motives for the DSH episode to effectively assess the extent of problem-
solving difficulties. Psychotherapeutic interventions should address self-
harm patients’ underlying problems across a broad range of areas. As an
escape motive reflecting an inability to tolerate distressing thoughts or
situations is independently associated with a history of DSH, therapeutic
approaches with repeaters should include problem-solving skills training.
Acceptance-based behavioural therapy or mindfulness-based cognitive
therapy may be useful for improving emotion regulation and distress
tolerance and for offering alternatives to avoiding or escaping difficult
situations. The importance of the revenge motive for repeaters indicates
considerable communication and interpersonal difficulties which also
need to be focused on in treatment programmes.

76
Deliberate Self-harm

Problem-solving OrientaƟon of Repeaters of


Deliberate Self-harm

We investigated problem-solving orientation among medically treated


DSH patients (McAuliffe, Corcoran, Keeley et al. 2006) based on responses
to the Utrecht Coping List (UCL; Schreurs et al., 1988), a process measure
of problem-solving ability administered as part of the WHO/EURO
Multicentre Repetition-Prediction study. The main objectives of the
study were:

1. to examine the association between problem-solving orientation


and previous history of DSH and
2. to establish if those reporting previous DSH are characterised by a
more negative problem-solving orientation (a negative approach
to problem solving—most typically a pessimistic attitude and a
passive response to problems).

Previous history of DSH was established on the basis of specific ques-


tions asked during the initial and follow-up interviews (EPSIS I and II).
Repeaters were identified as patients who engaged in more than one
episode of DSH before their follow-up interview—whether treated in
hospital or not.
Analysis of responses to the UCL was carried out from follow-up EPSIS II
interviews at 12 participating centres of the WHO/EURO Multicentre
study to explore the association between problem-solving orientation
and previous history of DSH. Greater passivity was independently asso-
ciated with increased risk of a previous history of DSH, when considered
alongside gender and age. In addition, greater avoidance was independ-
ently associated with a previous history of DSH. The problem-solving
style represented by this new factor, passive-avoidance (Table 4.1) is best
described as a negative problem-solving orientation in which the individual
entertains negative feelings, gets caught up in problems, worries about the
past, and feels hopeless and helpless. In terms of behavioural response,
there is a greater likelihood of giving in so as to avoid further difficult
situations, a greater tendency to resign oneself to the situation and to try
to avoid problems. For the repeater, the effect of this approach to problem
solving is that he or she stops trying to solve problems.

77
Carmel McAuliffe

Table 4.1 UCL Dimensions and Their Items

Problem-solving dimension N items Items


Passive-Avoidance 7 Being totally pre-occupied with the problems
Feeling unable to do anything
Worrying about the past
Taking a gloomy view of the situation
Giving in—in order to avoid difficult situations
Resigning oneself to the situation
Trying to avoid difficult situations as much as
possible
Active Handling 7 Finding out all about the problem
Making several alternative plans for handling a
problem
Considering different solutions to the problem
Making a direct intervention when problems occur
Using a direct approach in order to solve the
problem
Considering problems as a challenge
Realising every cloud has a silver lining
Source: McAuliffe, McLeavey, Corcoran, et al. (2006); originally compiled by the author.

The association between passive-avoidance and previous history of


DSH was no longer significant when the remaining four problem-solving
dimensions along with general health, problem drinking and self-esteem
were included in a regression. It was principally the inclusion of self-esteem
that weakened the association between passive-avoidance and previous
DSH. Self-esteem has elsewhere been described as a moderating individual
characteristic of the relationship between psychosocial stress and sui-
cidal behaviour. Low self-esteem can prime the individual to appraise prob-
lems negatively in situations of high stress (Sandin et al., 1998), whereas
at high levels, self-esteem may protect against repeated self-harm (Petrie
et al., 1988) by exerting a positive impact on the attitudes of people who
do self-harm to their ability to deal with problems.
High scores on the active handling dimension (Table 4.1) were sig-
nificantly associated with reduced likelihood of previous DSH, when
considered alongside age and gender. However, the association was weaker
than that of passive-avoidance. Male repeaters were significantly more
inclined to engage in negative emotional expression (showing anger or
annoyance). Males were also significantly less inclined to share problems,

78
Deliberate Self-harm

particularly male non-repeaters. However, while these differences were


statistically significant, their magnitude in real terms was relatively small.

Clinical implicaƟons Psychotherapeutic interventions for those who


repeatedly self-harm should directly address passive and avoidant
problem orientations. The characteristic negative thoughts and feelings
evoked in DSH patients (e.g., hopelessness, helplessness) when confronted
with problems should be explored in order to increase awareness of the
impact of negative cognitions in responding to problems. Training in
mindfulness-based cognitive therapy or acceptance-based behavioural
therapies to enhance tolerance of negative cognitions, together with
training a more active approach to problem-solving are key components
of these interventions.
The characteristic problem-solving style of repeaters of DSH is a com-
bination of a passive and an avoidant orientation to problems. Level of
self-esteem is a negative confounder of the association between passive-
avoidance and repeated self-harm. This means that preventing repeti-
tion of DSH in repeaters is unlikely to be achieved by addressing a passive
and avoidant problem-solving orientation without addressing low
self-esteem. This indicates the need for intensive therapeutic input and
follow-up for patients with repeated self-harm to develop a positive
approach to problems through a structured problem-solving skills
training programme; and to involve clients in practicing skills with their
own personal problems as opportunities to increase self-efficacy and
self-esteem.

OpƟonal Thinking Ability and Repeated Deliberate Self-harm

A recent study examined the association between optional thinking


ability—one of the principal difficulties in problem-solving performance
among those who deliberately self-harm (McLeavey et al., 1987; Neuringer,
1964; Patsiokas et al., 1979)—and prospective repetition of DSH within
12 months (McAuliffe et al., 2008). The main objective was to identify
DSH patients who subsequently repeat within one year on the basis of
an outcome measure of problem-solving ability (the optional thinking
test) administered at index episode, controlling for previous DSH, and

79
Carmel McAuliffe

other relevant sociodemograhic and clinical variables including level of


education, levels of suicide intent and hopelessness.
Poorer optional thinking ability on interpersonal problems was asso-
ciated with an increased risk of repeated DSH within one year among
patients medically treated for a first DSH episode. This means that, the
first ever patients who had difficulty in generating alternative solutions to
interpersonal problems were more likely to repeat DSH within one year
of their index episode. The association between optional thinking ability
and further self-harm was not significant for those who were repeaters at
index episode, as they were significantly more likely to have repeated
again by 12-month follow-up, independently of their optional thinking
ability. In fact, the association between history of self-harm and a further
repeat episode within 12 months was independent of the effects of all the
other variables. Taken together these results suggest that optional thinking
difficulties increase the risk of a first repeat DSH episode but for those
who have two or more previous episodes, a history of repeated self-harm
is more important than optional thinking in identifying those most likely
to repeat again.

Clinical implicaƟons For patients presenting with a first DSH episode,


problem-solving skills training to enhance optional thinking ability may
help to prevent a first repeat episode of DSH. For DSH repeaters, optional
thinking ability may be less important as a single factor in determining
outcome compared with first ever patients. As a corollary, problem-solving
interventions for first episode self-harmers should be delivered as soon as
possible following the DSH episode as this is the optimal time to interrupt
the development of DSH as an automatic stress response.

Problem-solving Skills Training to Prevent


Repeated Deliberate Self-harm

A large randomised controlled trial was carried out to determine the


efficacy of a structured Group Interpersonal Problem-Solving Skills
Training programme (PST) as an intervention approach to DSH in
addition to treatment as usual (TAU) as offered by the mental health

80
Deliberate Self-harm

services, compared with TAU alone (McAuliffe, McLeavey, Corcoran


et al., 2006; McLeavey et al., in press. The PST consisted of six two-hour
sessions, held weekly, of structured group interpersonal problem-solving
skills training, facilitated by a trained therapist and a cotherapist. A
manualised skills training programme (McLeavey et al., 2001) based on
the original five-stage model of problem solving (D’Zurilla and Goldfried,
1971; D’Zurilla and Nezu, 1990) was strictly adhered to. Participants were
encouraged to carry out homework assignments on a structured practice
journal between sessions using their own interpersonal problems.
During the recruitment phase patients aged between 18 and 64 years
with a recent episode of DSH were screened for the trial. Just under two-
thirds of those recruited had a previous history of DSH at index episode.
A total of 221 patients were randomised to group interpersonal problem-
solving therapy, while a further 222 were randomised to treatment as usual.
Over half of the patients screened were ineligible for the trial, the most
common reason for ineligibility being alcohol dependence, followed by
age outside trial range, followed by a diagnosis of psychosis and living
outside the trial area.

Outcome measures at baseline and follow-up With the exception of


impulsivity, which was lower among participants in PST, there were
no significant differences between treatment conditions on any of the
outcome measures at baseline. Comparing baseline to six weeks follow-
up, a significant reduction was observed for suicidal ideation, level of
depression, hopelessness and anxiety; while a significant increase was
observed for self-efficacy and self-rated problem-solving ability, in both
treatment conditions. A significant improvement was also observed on
perceived problem-solving ability. At six months follow-up these gains
were maintained, with no further changes on these outcome measures.
Compared to those in the TAU condition, DSH patients who had
completed PST did not show a significantly greater change on any of the
outcome measures at either follow-up stage with the exception of the
practical support subscale from the social life scale on which they showed
significant improvement at six weeks. The practical support subscale is a
measure of the extent to which a person both needs and receives practical
help from the person closest to them.

81
Carmel McAuliffe

Similar proportions of repeaters (hospital treated and non-hospital


treated episodes) were identified in each treatment condition at all three
follow-up periods (6 weeks, 6 months and 12 months). A similar number of
repeat episodes was reported by participants in both treatment conditions
at six weeks and at six months follow-up. There were no significant dif-
ferences between treatment conditions in terms of the number of repeat
hospital treated episodes at 12 months follow-up based on manual checks
of hospital records. Compared to those who received standard care
alone, participants assigned to PST evaluated the treatment programme
significantly more positively in terms of: satisfaction with the treatment
programme, relevance of the treatment programme to their problems,
usefulness of the treatment programme for coping with problems, their
understanding of the content of sessions and application of skills to their
own problems. Furthermore the level of compliance with PST was high.
The experimental treatment (PST) condition targeted the development
of improved interpersonal problem-solving skills in participants. How-
ever, no significant differences were found between those in PST and those
in TAU on any of the outcome measures of problem solving. Both treat-
ment groups improved to a similar extent on these outcomes between
baseline and follow-up, which suggests that the experimental treatment
programme was not sufficiently long to effect a significantly greater change
in the problem-solving skills of those in PST.

Clinical implicaƟons No significant differences were found at follow-up


between DSH patients in the problem-solving therapy (PST) and TAU
conditions on any of the outcome measures examined, with the exception
of needing and receiving practical support from a significant other, on
which participants in the PST condition showed a significantly greater
improvement. Although there was a significantly higher level of satisfac-
tion with the PST treatment programme among the patients assigned,
it is likely that it was too short to effect significantly greater changes in
problem-solving ability than TAU, particularly in light of the fact that the
majority were already repeaters at intake. The use of a group-based format
alone in the PST intervention may not have sufficiently addressed the
needs of DSH patients with a history of previous DSH episodes.

82
Deliberate Self-harm

INTEGRATING THE FINDINGS: KEY CONCLUSIONS


WITH REGARD TO PROBLEM SOLVING

There is a dearth of studies investigating the association between problem-


solving ability and repeated DSH. Based on the studies described earlier,
it was possible to distinguish repeaters of DSH as a sub-group of the self-
harming population on the basis of their problem-solving characteristics.
The following conclusions can be drawn:

1. Important associations exist between the problem-solving char-


acteristics of DSH patients and repetition.
2. Repeaters’ acts of DSH are motivated by a wider range of factors,
which implies a self-harm response that is more indiscriminate and
generalised.
3. Escape and revenge motives are significantly associated with
increased risk of repeated DSH.
4. Repeaters suffer greater impairment both in terms of how they ap-
proach problems and how they perform in problem solving. Their
passive-avoidant orientation to problems locks them into a cycle
of repeated failure to solve problems, which is likely to be one of
the factors contributing to their lower self-esteem.
5. Different aspects of problem-solving ability may be relevant at
different stages of the suicidal career: Poor optional thinking ability
is particularly important in the transition from a first to a repeat
DSH episode but once a first repeat episode occurs previous DSH
history has a more important association with repetition than does
optional thinking ability.

These findings could be thought to provide an evidence base for the use
of structured interpersonal problem-solving skills training programmes
to reduce the likelihood of repetition in DSH patients. However, in the
randomised controlled trial of structured group interpersonal problem-
solving skills training described earlier, no differences in outcomes were
observed between the PST experimental treatment condition and standard
care. This leads us to conclude that brief programmes of structured
group interpersonal problem-solving skills training probably do not

83
Carmel McAuliffe

allow sufficient time for the acquisition of skills and the development of
higher levels of self-efficacy, particularly where individuals have already
established a pattern of self-harming behaviour in response to the problems
they encounter. A longer treatment intervention based on structured
problem-solving skills training should be tested:

1. A six-session programme of structured group interpersonal


problem-solving skills training probably does not allow sufficient
time for the acquisition of skills and the development of higher
levels of self-efficacy, particularly in a sample of self-harmers where
the majority have already established a pattern of self-harming
behaviour in response to the problems they encounter.
2. A longer treatment intervention based on structured problem-
solving skills training should be tested in a large randomised con-
trolled trial to allow adequate time for acquisition of problem-solving
skills, including emotion regulation skills, and opportunities to
increase levels of self-esteem in their successful application to the
problems encountered by this client group. This longer programme
of problem-solving therapy should incorporate initial individual
sessions followed by group therapy sessions.

ACKNOWLEDGEMENT

The author wishes to acknowledge the important contribution of Dr


Ella Arensman, Director of Research at the National Suicide Research
Foundation, Cork, Ireland.

REFERENCES

Applebaum, S. (1963). The problem solving aspect of suicide. Journal of Project


Technology, 27, 259–68.
Arensman, E. and A.J.F.M. Kerkhof (2004a). Classification of attempted suicide:
A review of empirical studies, 1963–1993. Suicide and Life-Threatening Behavior,
26(1), 46–64.

84
Deliberate Self-harm

Arensman, E. and A.J.F.M. Kerkhof (2004b). Negative life events and non-fatal suicidal
behaviour. In D. De Leo, U. Bille-Brahe, A. Kerkhof and A. Schmidtke (Eds),
Suicidal behaviour: Theories and research findings (pp. 93–109). Göttingen:
Hogrefe and Huber.
Arensman, E., C. McAuliffe, P. Corcoran and I.Perry (2004). Correspondence. Psychological
Medicine, 34, 1143–44.
Arensman, E., E. Townsend, K. Hawton, S. Bremner, E. Feldman, R. Goldney et al.
(2001). Psychosocial and pharmacological treatment of patients following deliberate
self-harm: The methodological issues involved in evaluating effectiveness. Suicide
and Life-Threatening Behaviour, 31(2), 169–80.
Bancroft, J., K. Hawton, S. Simkin, B. Kingston, C. Cumming and D. Whitwell (1979). The
reasons people give for taking overdoses: A further inquiry. British Journal of
Medical Psychology, 52(4), 353–65.
Bancroft, J., A.M. Skrimshire and S. Simkin (1976). The reasons people give for taking
overdoses. British Journal of Psychiatry, 128(6), 538–48.
Baumeister, R.F. (1990). Suicide as escape from self. Psychological Review, 97(1),
90–113.
Bille-Brahe, U., A. Schmidtke, A.J.F.M. Kerkhof, D. De Leo, J. Lönnqvist and S. Platt
(1994). Background and introduction to the study. In A.J.F.M. Kerkhof, A.
Schmidtke, U. Bille-Brahe, D. De Leo and J. Lönnqvist (Eds), Attempted suicide in
Europe: Findings from the multicentre study on parasuicide by the WHO regional
office for Europe (pp. 3–15). Leiden: DSWO Press.
Braucht, G.N. (1979). Interactional analysis of suicidal behaviour. Journal of Consulting
and Clinical Psychology, 47(4), 653–69.
Brown, G.K., T. Ten Have, G.R. Henriques, S.X. Xie, J.E. Hollander and A.T. Beck
(2005). Cognitive therapy for the prevention of suicide attempts: A randomised
controlled trial. Journal of the American Medical Association, 294(5), 563–70.
Carriss, M.J., L. Sheeber and S. Howe (1998). Family rigidity, adolescent problem-
solving deficits, and suicidal ideation: A mediational model. Journal of Adolescence,
21(4), 459–72.
Chapman, A.L., K.L. Gratz and M.Z. Brown (2006). Solving the puzzle of deliberate
self-harm. The experiential avoidance model. Behavioural Research and Therapy,
44(3), 371–94.
Chiles, J. and K. Strosahl (2004). Clinical manual for assessment and treatment of
suicidal patients. London: American Psychiatric Publishing.
Clum, G.A., A.T. Patsiokas and R.L. Luscomb (1979). Empirically based comprehensive
treatment program for parasuicide. Journal of Consulting and Clinical Psychology,
47(5), 937–44.
Crane, C., J.M.G. Williams, K. Hawton, E. Arensman, H. Hjelmeland, U. Bille-Brahe et al.
(2007). The association between life events and suicide intent in self-poisoners
with and without a history of deliberate self-harm: A preliminary study. Suicide
and Life-Threatening Behaviour, 37(4), 367–78.

85
Carmel McAuliffe

Dieserud, G., E. Røysamb, Ø. Ekeberg and P. Kraft (2001). Toward an integrative model
of suicide attempt: A cognitive psychological approach. Suicide and Life-Threatening
Behaviour, 31(2), 153–68.
D’Zurilla, T. (1986). Problem-solving therapy: A social competence approach to clinical
intervention. New York: Springer.
D’Zurilla, T.J. and E.C. Chang (1995). The relations between social problem solving
and coping. Cognitive Therapy and Research, 19(5), 547–62.
D’Zurilla, T.J. and M.R. Goldfried (1971). Problem solving and behaviour modification.
Journal of Abnormal Psychology, 78(1), 107–26.
D’Zurilla, T.J. and A.M. Nezu (1990). Development and preliminary evaluation of the
Social Problem-Solving Inventory (SPSI). Psychological Assessment: A Journal of
Consulting and Clinical Psychology, 2, 156–63.
Evans, J., J.M.G. Williams, S. O’ Loughlin and K. Howells (1992). Autobiographical memory
and problem-solving strategies of parasuicide patients. Psychological Medicine,
22(2), 399–405.
Hawton, K., E. Arensman, E. Townsend, S. Bremner, E. Feldman, B. Goldney et al. (1998).
Deliberate self-harm: Systematic review of efficacy of psychosocial and pharma-
cological treatments in preventing repetition. British Medical Journal, 317(7156),
441–47.
Hawton, K. and J. Fagg (1988). Suicide, and other causes of death, following attempted
suicide. British Journal of Psychiatry, 152(3), 359–66.
Hawton, K., L. Harriss, S. Hall, S. Simkin, E. Bale and A. Bond (2003). Deliberate self-
harm in Oxford 1990–2000: A time of change in patient characteristics. Psychological
Medicine, 33(6), 987–95.
Hawton, K., K. Houston, C. Haw, E. Townsend and L. Harriss (2003). Comorbidity of
axis I and axis II disorders in patients who attempted suicide. American Journal of
Psychiatry, 160(8), 1494–500.
Henriques, G.R., G.K. Brown, M.S. Berk and A.T. Beck (2004). Marked increases in
psychopathology found in a 30-year cohort comparison of suicide attempters.
Psychological Medicine, 34(5), 833–41.
Heppner, P. (1988). Manual for the problem solving inventory. Palo Alto, CA: Consulting
Psychologists Press.
Kerkhof, A. (2000). Attempted suicide: Patterns and trends. In K. Hawton and K. van
Heeringen (Eds), The international handbook of suicide and attempted suicide
(pp. 49–64). Chichester: Wiley.
Kerkhof, A., W. Bernasco, U. Bille-Brahe, S. Platt and A. Schmidtke (1993). European
Parasuicide Study Interview Schedule (EPSIS I, Version 6.1). In WHO/EUR/ICP/
PSF 018. Copenhagen.
Linehan, M.M., H.E. Armstrong, A. Suarez, D. Allmon and H.L. Heard (1991). Cognitive-
behavioral treatment of chronically parasuicidal borderline patients. Archives of
General Psychiatry, 48(12), 1060–64.
Linehan, M.M., P. Camper, J.A. Chiles, K. Strosahl and E.L. Shearin (1987). Inter-
personal problem-solving and parasuicide. Cognitive Therapy and Research, 11(1),
1–12.

86
Deliberate Self-harm

Linehan, M.M., K.A. Comtois, A.M. Murray, M.Z. Brown, R.J. Gallop, H.L. Heard
et al. (2006). Two-year randomised controlled trial and follow-up of dialectical
behaviour therapy vs therapy by experts for suicidal behaviours and borderline
personality disorder. Archives of General Psychiatry, 63(7), 757–66.
Linehan, M.M., H.L. Heard and H.E. Armstrong (1993). Naturalistic follow-up of a
behavioural treatment for chronically parasuicidal borderline patients. Archives
of General Psychiatry, 50(12), 971–74.
MacLeod, A.K. and J.M.G. Williams (1992). Cognitive psychology of parasuicidal be-
haviour. In P. Crepet, G. Ferrari, S. Platt and M. Bellini (Eds), Suicidal behaviour in
Europe: Recent research findings (pp. 217–24). Rome: Libbey.
Madge, N., A. Hewitt, K. Hawton, E. Jan de Wilde, P. Corcoran, S. Fekete et al. (2008).
Deliberate self-harm within an international community sample of young people.
Comparative findings from the Child and Adolescent Self-harm in Europe (CASE)
study. The Journal of Child Psychology and Psychiatry, 49(6), 667–77.
Maris, R.W. (1992). The relationship of non-fatal suicide attempts to completed suicide.
In R.W. Maris, A.L. Berman, J.T. Maltsberger and R.I. Yufit (Eds), Assessment and
prediction of suicide. New York: Guilford.
McAuliffe, C., E. Arensman, H.S. Keeley, P. Corcoran and A.P. Fitzgerald (2007).
Motives and suicide intent underlying hospital treated deliberate self-harm and
their association with repetition. Suicide and Life-Threatening Behavior, 37(4),
397–408.
McAuliffe, C., P. Corcoran, P. Hickey and B.C. McLeavey (2008). Optional thinking
ability among hospital treated deliberate self-harm patients: A one-year follow-up
study. British Journal of Clinical Psychology, 47(1), 43–58.
McAuliffe, C., B.C. McLeavey, P. Corcoran, B. Carroll, B. O Keeffe, M. O’ Regan et al.
(2006). Baseline characteristics and comparative treatment satisfaction of deli-
berate self-harm patients recruited in a randomised controlled trial of group inter-
personal problem-solving skills training compared with standard care. Psychiatrica
Danubina, 18(1), 90.
McAuliffe, C., P. Corcoran, H.S. Keeley, E. Arensman, U. Bille-Brahe, D. De Leo, et al.
(2006). Problem-solving ability and repetition of deliberate self-harm: A multi-
centre study. Psychological Medicine, 36(1), 45–55.
McLeavey, B. C., R. J., Daly, J. W., Ludgate and C. M. Murray (1994). Interpersonal problem-
solving skills training in the treatment of self-poisoning patients. Suicide and Life-
Threatening Behavior, 24(4), 382–394.
McLeavey, B.C., R.J. Daly, C.M. Murray, J. O’ Riordan and M.Taylor (1987). Inter-
personal problem-solving deficits in self-poisoning patients. Suicide and Life-
Threatening Behavior, 17(1), 33–49.
McLeavey, B., C. McAuliffe, E. Arensman, P. Corcoran, B. Caroll and L. Ryan (in press).
Problem-solving skills training for patients who deliberately self-harm:
A randomised controlled trial. (Submitted for publication in 2009.)

87
Carmel McAuliffe

McLeavey, B.C., C. McAuliffe and R.J. Daly (2001). Interpersonal problem-solving


skills training manual. Department of Psychiatry, University College Cork and
the Southern Health Board.
National Suicide Research Foundation. (2008). National registry of deliberate self-harm
Ireland: Annual Report 2006/7. Cork: National Suicide Research Foundation.
Nekanda-Trepka, C.J.S., S. Bishop and I.M. Blackburn (1973). Hopelessness and
depression. British Journal of Clinical Psychology, 22(1), 49–60.
Neuringer, C. (1964). Rigid thinking in suicidal individuals. Journal of Consulting
Psychology, 28(1), 54–58.
Neuringer, C. and D.J. Lettieri (1971). Cognition, attitude and affect in suicidal indi-
viduals. Life-Threatening Behaviour, 1(2), 106–24.
Nordentoft, M., L. Breum, L.K. Munck, A.G. Nordestgaard, A. Hunding and P.A.
Laursen Bjaeldager (1993). High mortality by natural and unnatural causes:
A 10-year follow-up study of patients admitted to a poisoning treatment centre
after suicide attempts. British Medical Journal, 306(6893), 1637–41.
O’Carroll, P.W., A.L. Berman, R.W. Maris, E.K. Moscicki, B.L. Tanney and M.M.
Silverman (1996). Beyond the Tower of Babel: A nomenclature for suicidology.
Suicide and Life-Threatening Behavior, 26(3), 237–52.
Patsiokas, A.T., G.A. Clum and R.L. Luscomb (1979). Cognitive characteristics of suicide
attempters. Journal of Consulting and Clinical Psychology, 47(3), 478–84.
Petrie, K., K. Chamberlain and D. Clarke (1988). Psychological predictors of future
suicidal behaviour in hospitalized suicide attempters. British Journal of Clinical
Psychology, 27(3), 247–57.
Platt, S., U. Bille-Brahe, A. Kerkhof, A. Schmidtke, T. Bjerke, P. Crepet et al. (1992). Parasui-
cide in Europe. The WHO/EURO multicentre study on parasuicide I. Introduction
and preliminary analysis for 1989. Acta Psychiatrica Scandinavica, 85(2), 97–104.
Pollock, L.R. and J.M.G. Williams (1998). Problem solving and suicidal behaviour.
Suicide and Life-Threatening Behavior, 28(4), 375–87.
Pollock, L. and J.M.G. Williams (2001). Effective problem solving in suicide attempters
depends on specific autobiographical recall. Suicide and Life Threatening Behavior,
31(4), 386–96.
Pollock, L.R. and J.M.G. Williams (2004). Problem-solving in suicide attempters.
Psychological Medicine, 34(1), 163–67.
Rudd, M.D., T. Joiner and M.H. Rajab (1996). Relationships among suicide ideators,
attempters and multiple attempters in a young adult sample. Journal of Abnormal
Psychology, 105(4), 541–50.
Rudd, M.D., M.H. Rajab and P.F. Dahm (1994). Problem-solving appraisal in suicide
ideators and attempters. American Journal of Orthopsychiatry, 64(1), 136–49.
Sakinofsky, I. (2000). Repetition of suicidal behaviour. In K. Hawton and K. van
Heeringen (Eds), The international handbook of suicide and attempted suicide
(pp. 385–404). Chichester: Wiley.
Sakinofsky, I. and R.S. Roberts (1990). Why parasuicides repeat despite problem
resolution. British Journal of Psychiatry, 156(3), 399–405.

88
Deliberate Self-harm

Sakinofsky, I., R.S. Roberts, Y. Brown, C. Cumming and P. James (1990). Problem re-
solution and repetition of parasuicide. A prospective study. British Journal of
Psychiatry, 156(3), 395–99.
Salkovskis, P., C. Atha and D. Storer (1990). Cognitive-behavioural problem solving in
the treatment of patients who repeatedly attempt suicide: A controlled trial. British
Journal of Psychiatry, 157(6), 871–76.
Sandin, B., P. Chorot, M.A. Santed, R.M. Valiente and T.E. Joiner (1998). Negative life
events and adolescent suicidal behavior: A critical analysis from a stress process
perspective. Journal of Adolescence, 21(4), 415–26.
Schmidtke, A., U. Bille-Brahe, D. De Leo, A. Kerkhof, C. Löhr, B. Weinacker et al. (2004).
Sociodemographic characteristics of suicide attempters in Europe: Combined
results of the monitoring part of the WHO/EURO Multicentre Study on Suicidal
Behaviour. In A. Schmidtke, U. Bille-Brahe, D. De Leo and A. Kerkhof (Eds),
Suicidal behaviour in Europe (pp. 29–43). Göttingen: Hogrefe & Huber.
Schotte, D.E. and G.A. Clum (1982). Suicide ideation in a college population: A test of
a model. Journal of Consulting and Clinical Psychology, 50(5), 690–96.
Schotte, D.E. and G.A. Clum (1987). Problem solving skills in suicidal psychiatric
patients. Journal of Consulting and Clinical Psychology, 55(1), 49–54.
Schotte, D.E., J. Cools and S. Payvar (1990). Problem-solving deficits in suicidal
patients. Trait vulnerability or state phenomenon? Journal of Consulting and
Clinical Psychology, 58(5), 562–64.
Schreurs, P.J.G., G. van de Willige, B. Tellegen and J.F. Brosschot (1988). De Utrechtse
Copinglijst: Handleiding. Lisse: Swets En Zeitlinger.
Slee, N., N. Garnefski, R. van der Leeden, E. Arensman and P.H. Spinhoven (2008).
Cognitive-behavioural intervention for self-harm: Randomised controlled trial.
British Journal of Psychiatry, 192(3), 202–11.
Tarrier, N., K. Taylor and P. Gooding (2008). Cognitive behavioural interventions to
reduce suicide behaviour: A systematic review and meta-analysis. Behaviour
Modification, 32(1), 77–108.
Townsend, E., K. Hawton, D.G. Altman, E. Arensman, D. Gunnell, P. Hazell, A. House
and K. Van Heeringen (2001). The efficacy of problem-solving treatments after
deliberate self-harm: Meta-analysis of randomized controlled trials with respect to
depression, hopelessness and improvement in problems. Psychological Medicine,
31(6), 979–88.
Tyrer, P., S. Thompson, U. Schmidt, V. Jones, M. Knapp, K. Davidson 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(6), 969–76.
Williams, J.M., T. Barnhofer, C. Crane and A.T. Beck (2005). Problem-solving
deteriorates following mood challenge in formerly depressed patients with a history
of suicidal ideation. Journal of Abnormal Psychology, 114(3), 421–31.

89
Carmel McAuliffe

Williams, J.M.G. and L.R. Pollock (2001). Psychological aspects of the suicidal process.
In K. van Heeringen (Ed.), Understanding suicidal behaviour (pp. 76–93).
Chichester: Wiley.
Yang, B. and G.A. Clum (2000). Childhood stress leads to later suicidality via its effect
on cognitive functioning. Suicide and Life-Threatening Behavior, 30(3), 183–98.
Zahl, D.L. and K. Hawton (2004). Repetition of deliberate self-harm and subsequent
suicide risk: Long term follow-up study of 11,583 patients. British Journal of
Psychiatry, 185(1), 70–75.

90
5

Suicide and Homicide: TheoreƟcal Issues


Sóƒã® M绫›Ù¹››, U֗›Ý« KçÃ٠ƒÄ— MƒÄƒÝ K. MƒÄ—ƒ½

T he twentieth century has been termed as the century of violence. It is


not uncommon to talk about the ‘culture of violence’ prevailing in
almost all the societies and across cultures. Based on the resolution adopted
at the Forty-ninth World Health Assembly (WHO, 1996) about violence
prevention, Krug, Dahlberg, Mercy, Zwi and Lozano (2002: 5) define
violence as ‘the intentional use of physical force or power, threatened or
actual, against oneself, another person, or against a group or community,
that either results in or has a high likelihood of resulting in injury, death,
psychological harm, maldevelopment or deprivation’. According to World
Health Organization (WHO, 2002), each year more than a million people
lose their lives as a result of self-inflicted, interpersonal or collective
violence. Almost half of these are attributed to suicide and almost one-
third to homicide. Violence is a major public health problem in most coun-
tries that incurs a huge cost not only in terms of the financial loss and
infrastructural requirements for its prevention and management, but also
in terms of the pain, suffering and trauma faced by the victims.
The word suicide was coined by Sir Thomas Browne in his Religio
Medici in the year 1642 based on Latin sui (of oneself) and caedere (to kill)
intending to distinguish between the homicide of oneself and the killing
of another. Freud (1915/1957) had hypothesised suicide as an expression
of anger towards a love object that the individual turns back on the self.

91
Swati Mukherjee et al.

Menninger (1938) described suicide as the gratification of self-destructive


tendencies, which include a wish to kill, a wish to be killed and a wish to
die. The distinction between suicidal tendencies and violence directed
towards the other has often been emphasised by researchers (e.g., Breed,
1963; Henry and Short, 1954; Smith and Parker, 1980) based on epi-
demiological, aetiological and socio-cultural data. Recent studies, however,
have highlighted the commonality between the two types of violence,
though there is a very small epidemiological overlap reported between
the two. van Praag, Plutchik and Apter (1990) have reported that nearly
30 percent of violent individuals also have a history of self-destructive
behaviour, conversely 10–20 percent of those who are suicidal also have
a history of violence.

HOMICIDE: FATAL VIOLENCE TURNED TO THE OTHER

The word homicide originates from the Latin homicidium, composed of


homo, meaning ‘man’, and cidium derived from the verb caedo, meaning ‘to
cut’ or ‘to kill’. Homicide is a complex phenomenon constituting ‘a com-
plex set of behaviours with distinct and varying aetiologies, clinical
courses, and prognoses’ (Schlesinger, 2007: 708). As it is true for sui-
cide, each homicide is unique, defined by specific characteristics of the
perpetrator, the victim and the circumstances. Despite the dynamic and
interdisciplinary nature of homicidal behaviour, most research in the area
of homicide studies has been carried out with a criminological perspective.
The complexity of the phenomenon makes it difficult to explain a homicide
solely on the basis of a psychopathology, except in the cases where psychosis
(usually paranoid) or brain pathology leading to aggression is involved
(Schlesinger, 2007). A dynamic approach to explain homicidal behaviour,
involving biological, psychiatric and socio-cultural factors is warranted.
Attempts have been made to classify homicides using various criteria,
such as on the basis of the relationship between perpetrator and victim,
like uxoricide (killing the spouse or consort), filicide (killing one’s chil-
dren), familicide (killing one’s family members), siblicide (killing one’s
sibling) or fratricide (killing one’s brother, which has acquired a wider
connotation in the military context, and is used to denote killing of
one’s own companions). Certain other classifications use motive as the

92
Suicide and Homicide

criteria and provide categories such as gang-related homicide, homicide


in retaliation to prior homicide and family-intimate partner homicide
(Loeber et al., 2005). Heide (2003) has offered a classification based on
the circumstances leading to homicide, for example, psychotic episodes,
conflicts and crime leading to homicide. However, such classifications are
limited in the sense that these add very little to the understanding of the
aetiology of violence and homicidal behaviour (Loeber et al., 2005).
Another approach to classifying homicides depends on the character-
istics of the perpetrators, specifically on their ‘developmental pathways’
(Loeber et al., 2005). The developmental approach takes into account the
past history of the individual for aggression and involvement in violent acts
other than homicide. However, based on a review of homicide offenders,
Heide (2003) has found mixed evidence to support the fact that homicide
offenders have had a lengthy history of violence and aggression. This
implies definite difficulties involved in predicting homicidal behaviour
based on developmental indicators. Loeber et al. (2005) provide a brief
overview of the theories explaining the origin of violent behaviour at the
individual level, and classify these into ontogenic, sociogenic and mixed
theories. The ontogenic theories propose that violent behaviour can be
explained by the initial proneness of an individual (determined by factors
like neurobiological deficits, personality, temperament and parenting)
that stays fairly stable over time. The sociogenic theories postulate that life
events have a dynamic impact on the individual, increasing or decreasing
the probability of occurrence of violent behaviours. The mixed theories
resolve the seemingly oppositional propositions made by the ontogenic and
sociogenic theories, by combining the impact of early individual differences
model with adulthood life circumstances. Lober and colleagues (2005)
propose the mixed model as being useful for explaining and predicting
homicidal behaviours.

Understanding the Overlap with suicide

Psychiatric studies One common way of exploring the overlap between


suicide and violence has been to study the occurrence of these behav-
iours among psychiatric patients. It has been noted that an underlying
psychiatric illness might be the root cause of both the destructive be-
haviours (Nock and Marzuk, 2000). The psychiatric illnesses most often

93
Swati Mukherjee et al.

associated with suicidal behaviours are mood disorders, substance abuse,


schizophrenia and personality disorders. There is a plethora of research
available exploring these associations that have been discussed elsewhere
in the present book. Most of these disorders have also been found to be
associated with violence. Alcohol and substance abuse have been reported
to increase the prevalence of violence in psychiatric inpatients as well as
community samples (Steadman et al., 1998). Schizophrenia is long known
to be associated with violence, self directed or directed at others (Eronen
et al., 1996; Link et al., 1992; Swanson et al., 1990). Violent individuals
have often been diagnosed with personality disorders, especially anti-
social and borderline personality disorder (Eronen et al., 1996; Volavka,
1995). Prevalence of mood disorders among the violent individuals has
been reported to be three times higher than the non-violent individuals
(Swanson et al., 1990). Diagnostic and Statistical Manual-IV (DSM-IV)
categories that include violence and aggression have been reported by Lion
(1995). These include substance abuse disorders, personality disorders,
conduct disorder, dementia, post-traumatic stress disorder, mental
retardation, sexual disorders and schizophrenia. The psychiatric disorder
most intensively studied in association with aggression and violence is
intermittent explosive disorder (Lion, 1995), which is characterised by
aggressive impulses out of proportion to any precipitating psychological
stressor (Burt, 1995).

Evidence from jail and prison data Another opportunity to understand


the overlap between suicidal and homicidal behaviours is given by the
research exploring suicidal tendencies among criminals, especially among
those convicted of homicide (Nock and Marzuk, 2000). Many studies have
reported unusually high rates of suicide among prison and jail inmates
(Joukamaa, 1997; Kerkhof and Bernasco, 1990). It has also been reported
that being charged with murder/manslaughter significantly raises the risk
for suicidal behaviours (DuRand et al., 1995; Kerkhof and Bernasco, 1990).
However, both kinds of research studies have limited applicability as
both focus on inmate population and not on general community samples.
Moreover, the labels of prison inmate or psychiatric inmate might be
superficial, as the labeling usually depends on which system the individual
encounters first—legal or psychiatric (Nock and Marzuk, 2000). Though
an overview of such studies suggests the underlying psychopathology
in both the behaviours (increased impulsiveness, affective lability,

94
Suicide and Homicide

disinhibition, problems with reasoning and decision-making leading to


increased aggressiveness; Nock and Marzuk, 2000), it also highlights the
need for further exploring the dynamics of violent behaviours among the
general population.
The dynamic nature of violence has been emphasised by the ecological
model (Bronfenbrenner, 1979; Garbarino and Crouter, 1978). The eco-
logical model explores the relationship between individual and contextual
factors, and considers violence as the product of multiple levels of influence
on behaviour. It seeks to understand, predict and prevent violence on the
basis of individual, relationship, community and societal level factors.
Extending the ecological model, the research studies exploring the link
between different types of violence can be utilised to further the knowledge
of the overlap between suicidal and homicidal behaviours. It is accepted
that exposure to violence in home during childhood is associated with being
a victim or perpetrator of violence in adolescence and adulthood (Maxfield
and Widom, 1996). Associations have been reported between suicidal
behaviour and different types of violent behaviours. Paolucci, Genuis and
Violato (2001) have reported associations of childhood maltreatment,
the experience of being rejected and abused with violent and suicidal be-
haviours during adulthood. Intimate-partner violence (Stark and Flitcraft,
1995), sexual assault (Paolucci et al., 2001) and abuse of the elderly
(Bristowe and Collins, 1989; Pillemer and Prescott, 1989) have also been
found to be associated with suicidal behaviours.

Homicide–suicide An extreme manifestation of the link between homi-


cidal and suicidal behaviours (Nock and Marzuk, 2000) is the occurrence of
both the destructive acts in close succession involving the same individual.
Homicide–suicide is a relatively rare phenomenon in which the perpetrator
kills another individual, usually an intimate partner, and commits suicide
shortly thereafter, usually within minutes or hours (Cohen, 2000). The
defining aspect of homicide–suicide is considered to be the intrinsic link
between the two acts (Nock and Marzuk, 1999). Both homicide and suicide
are seen as ‘antagonistic expressions of human aggression’ (Liem et al.,
2007), and the occurrence of homicide–suicide poses a unique opportunity
to understand the dynamics of both. Though most of the studies in the
area do not focus on the clinical aspects of the phenomenon, researchers
have shown the uniqueness of homicide–suicide based on epidemiological
data, demographic characteristics of the victims and perpetrators, and the

95
Swati Mukherjee et al.

modus operandi of the act (Liem et al., 2007). The typical psychological
profile of the perpetrator in cases of homicide–suicide appears to be ‘that
of a passive–aggressive, young adult, with poor self-esteem, insecure, and
socially inadequate, who occasionally uses drugs and alcohol and exhibits
proneness to explosive behaviour. The background frequently includes a
dysfunctional family, including sexual abuse as a child’ (Palermo, 2007: 10).
Liem and her colleagues (2007) have distinguished three subtypes of
homicide–suicide and have described the dynamics involved. According to
her an incidence of homicide–suicide can be distinguished by the under-
lying motive, either primarily suicidal or primarily homicidal. She has
proposed a model that describes suicide, homicide and homicide–suicide as
the culmination of aggressive intent, directed either against the self or the
other, depending upon the attribution made by the individual for his/her
frustrations, and emotional dependence on the victim. Based on a review
of statistical data regarding the phenomenon and placing these in the
context of sociological, psychiatric and psychoanalytic theories, Palermo
(1994) prefers to rename homicide–suicide as ‘extended suicide’. He
profiles the perpetrator as a fragile, dependent, ambivalent and aggressive
individual who hides behind a facade of self-assertion. Unable to withstand
the rejection by an intimate partner, on whom he is dependent, he kills
himself after killing his ‘extended self’.
Though most homicides involve a single victim, it is not uncommon to
find multiple homicides in most societies. Palermo (2007) has described
three types of multiple homicides—spree killings, mass killings and serial
murders. Various similar behaviours occurring in different cultures,
involving a violent outburst have been described by Cooper (1934), such as
‘amok’ in Malaysia, ‘Wihtico psychosis’ among the Cree Indians, ‘jumping
Frenchman’ in Canada and ‘imu’ in Japan. Most of these are culture-bound
syndromes that involve inappropriate and grossly exaggerated response
to sudden or loud stimuli, high suggestibility, echolalia, echopraxia and
violent behaviour that results in causing injuries/fatalities to others and
subsequently to the individual himself (Colman, 2001). These disorders,
despite the probability of a similar aetiology and underlying pathology, are
manifested in culture-specific ways. Determined by larger socio-cultural
environment and opportunities for expressing the aggression, these unique
manifestations seem to highlight the role of sociogenic factors. However,
Hempel and colleagues (2000) found more similarities than differences
between oriental and occidental cases of running amok. They compared
a nonrandom sample of North American cases of sudden mass assault by

96
Suicide and Homicide

a single individual (SMASI, n = 30) with a nonrandom sample of Laotian


amok cases (n = 18) and other amok studies, and reported perpetrators
showing evidence of social isolation, loss, depression, anger, pathological
narcissism and paranoia, across the cultures.

CULTURE SPECIFICITY OF HOMICIDAL BEHAVIOUR

It is an accepted fact that homicide rates vary in different countries and


across societies. In the United States, homicide rate varies from 12 to 140
per million depending on the social group involved and the area studied
(Cordess, 1995). European rates are generally lower, with rates in England
and Wales being 12 per million (Home Office, 1993). According to the
WHO report (Krug et al., 2002) the region-wise rates of violent deaths (both
homicidal as well as suicidal deaths) based on Global Burden of Disease
project for 2000, Version 1 show wide disparity across regions. According
to WHO, the rates of violent death vary according to country income levels.
In 2000, the rate of violent death in low- to middle-income countries was
32.1 per 100,000 population, more than twice the rate in high-income
countries (14.4 per 100,000). Large differences in suicidal and homicidal
rates are also reported among different countries within a region, between
urban and rural populations, between rich and poor groups, and between
different racial and ethnic groups. This clearly shows that much more is
involved in homicidal occurrences than mere economic disparities. Daly
and Wilson (2003) provide a provoking discussion on the involvement
of cultural issues in homicide, and debate the proposition of attributing
high prevalence of violence and consequent high rates of homicides in
certain groups to a ‘subculture of violence’. Seeking more ‘utilitarian ex-
planations’ than mere transmission of values that legitimise violence
they call for exploring the possible role played by the current economic,
demographic and social circumstances of the individuals involved.

TheoreƟcal Underpinnings

Each homicide is a complex conglomeration of factors at varying levels,


and cannot be explained by resorting to psychogenic factors or sociogenic

97
Swati Mukherjee et al.

factors in isolation. Scholars have propagated different theoretical


propositions in an attempt to elaborate upon the phenomenon.

Neurobiological perspecƟve Neurobiological evidence has shown the


involvement of amygdale, hippocampus, hypothalamic nuclei and pre-
frontal lobes in violent behaviour (Weiger and Bear, 1988). The limbic
system and mid-temporal zones, which are the sites of emotional trigger
zones, are also reported as having involvement in homicidal syndrome
(Palermo, 2007). Impulsivity, lack of control, poor objectivity, poor
discriminating capacity and improper situational appraisal are some of
the manifestations of prefrontal lobe dysfunction found in majority of
homicidal individuals (Palermo, 2007).
Studies exploring neuro-chemical basis of violence have found ab-
normalities in the metabolism of inhibitory neurotransmitters (e.g.,
serotonin and cerebro-spinal fluid 5 hydroxyindoleacetic acid [CSF-
5HIAA]), as well as the metabolism of excitatory neurotransmitters
(e.g., norepinephrine) (Lion, 1995). The knowledge gained through animal
experimentation applied to human studies has led to enhanced under-
standing of both suicidal and homicidal behaviours. Diminished CSF-
5HIAA levels have been reported to be associated with impulsivity and
aggressive dyscontrol. Lion (1995) reports significantly lower CSF-5HIAA
levels in suicide survivors who had used violent methods as compared to
those who had used less violent methods. The proposition that suicide and
homicide are variants of the same underlying aggressive intent seems to
be supported by neurological evidence.

Sociological perspecƟve Sociological explanations for violence, spe-


cifically homicide, assert that the violent tendencies manifest because of
a subculture of violence (Wolfgang and Ferracuti, 1967). Certain others
attribute violence to presence of social vacuum (Palermo, 2004). Societal
factors that have been reported to enhance the probability of homicidal be-
haviours are disinhibition due to drugs or alcohol, socially dysfunctional
family structure, substandard economic conditions and social emargination
(Palermo, 2004; Wolfgang and Ferracuti, 1967). Poor achievement, lack
of specific skills and lack of steady employment have also been associated
with homicidal behaviour (Palermo, 2007). The sociological theories of
violence resort to the macro-level socio-cultural and economic dynamics
for explaining individual and interpersonal behavioural phenomena.

98
Suicide and Homicide

Durkheim (1897/1951) has elaborated upon the influence of sociogenic


factors upon individual behaviour. Construing homicide and suicide as
interconnected and interdependent phenomena, he asserted that occur-
rence of both these violent behaviours is determined by the degree of
integration or cohesion in a society and the degree of social regulation. Ac-
cording to him impulsivity increases in individuals as a result of social
disorganisation, especially in times of rapid social change. The social
anomie in such societies causes ‘a state of irritated weariness’ (p. 257) at
the individual level that is either turned towards self (resulting in suicide),
or towards the other (resulting in homicide). Within the typology of
suicide (egoistic, altruistic, anomic and fatalistic) developed by Durkheim
(1897/1951), fatalism has often been used to explain suicide as well as
homicide (e.g., Cavan, 1928; Peck, 1979). Cavan provides an explanation
for homicide-suicide describing it as an act of fatal resignation, where the
individual finds no solution to his problem than suicide, but also blames
the other for his unhappiness and kills that person in anger, revenge or
jealousy. Durkheim’s (1897/1951) propositions provide a theoretical
framework to explore the sociological and socio-psychological roots of
violent behaviours. Summarily, the factors contributing to the homicidal
profile from sociological perspective are proposed as relative deprivation
(ascribed and illegitimate economic disparity), oppression (injustice,
real or perceived, by government/other group), social learning (violence
follows observation) and cultural dissociation (in-group, out-group).

Psychological perspecƟve A review of the available research suggests


that the homicidal behaviour, ‘short of those cases in which there is pre-
meditation, organisation, and clear planning, is viewed today as the out-
come of an individual’s disorganisation and his or her incapacity to
control basic dangerous impulses, internal or external’ (Palermo, 2007: 6).
Freud (1920/1961) viewed both suicide and homicide as expressions
of aggression, as alternative responses to common cause. According to
him, suicide represents an impulse to kill another turned inward on the
self, and it has two roots, the death wish (Thanatos instinct), and sexual
frustration or repression. Homicide occurs when the life instinct (Eros)
counters the Thanatos and turns its impulse to kill one self to the other.
Conversely, suicide happens when frustration is coupled with a blocked
desire to commit murder. Freud (1920/1961) attributed all goal-directed
behaviour to libidinal force and opined that as libido drives an individual

99
Swati Mukherjee et al.

to achieve good goals, it may also direct one to destructive aggression, in-
cluding homicide. Disinhibited aggressive behaviour, according to Freud,
results due to instinctual forces overcoming the control imposed by ego
and superego and allowing for an uninhibited expression of basic negative
emotions.
One of the most influential theories based on Freudian propositions
is the frustration–aggression hypothesis, proposed by Miller (1941) in
association with Robert R. Sears, O.H. Mowrer, Leonard W. Doob and
John Dollard. The original propositions of the frustration–aggression
hypothesis (Dollard et al., 1939) presuppose existence of frustration in all
cases of aggression, and conversely construe aggression as the only con-
sequence of frustration. Dollard and colleagues (1939) also said that absence
of overt aggression subsequent to frustration was only due to inhibition
caused by threat of punishment to self or to loved ones. Over the years
the theory has generated much research and debates. Many modifications
have been incorporated beginning with Miller (1941) and colleagues, who
accepted that frustrations (i.e., the inability to attain desired goal due to exter-
nal thwarting) can have non-aggressive behavioural consequences as well,
though, they said that if the thwarting to goal-directed behaviour continue,
the aggressive responses would eventually become dominant over non-
aggressive responses.
Another relevant proposition is of the distinction between hostile
aggresion and instrumental aggression (Feshbach, 1964). Aggression is
defined as a behaviour that is aimed at causing harm or injury to the target
(Dollard et al., 1939). Hostile aggression is primarily aimed at causing
harm whereas instrumental aggression is primarily oriented at attain-
ment of some other objective, like money, status or power. The frustration–
aggression hypothesis has relevance in the context of understanding
the genesis of hostile aggression, manifested in violence towards self or
towards others.
The frustration–aggression hypothesis also proposes that the aggres-
sion generated by the frustration is directed at the agent perceived to be
the source of frustration (Dollard et al., 1939: 39), or its displacement
to substitute targets having appropriate stimulus characteristics. On the
basis of empirical evidence, Berkowitz (1989) concluded that attributional
interpretation of the aggression-provoking situation (the intentionality
and perceived legitimacy of the thwarting) is significant in determining
the emotional reaction of the individual. Berkowitz (1989) also concluded

100
Suicide and Homicide

that aggression facilitating cue do not generate aggression, but play an


important role in reinforcing or intensifying the aggressive reaction
generated by goal thwarting.
The role of social rules and inhibitions against aggressive reactions
(Cohen, 1955), role of prior learning (Bandura, 1973), individual differ-
ences like ego-strength (Block and Martin, 1955), or personality type
(Strube et al., 1984), and cognitive processes like thoughts about the goal
(Folger, 1986; Thibaut and Kelly, 1959), or thoughts about the reasons for
frustration (Averill, 1982; Zillmann, 1978) have been shown to determine
the emotional reactions to frustrations.

CONCLUDING REMARKS

The evidence from neurological, sociological and psychological per-


spectives points towards negating the dichotomy of suicide and homicide.
It is suggested, therefore, to view both behaviours as variations of the same
underlying aggressive intent. Heightened aggressiveness has been report-
ed among suicidal individuals, and the research involving individuals
with a history of violent behaviours shows high incidence of suicidal
behaviours. The causal factors for suicidal and homicidal behaviours seem
to be involved in a complex dynamic pattern. Though, impulsivity and
heightened aggressiveness form the core of both behaviours, the dynamics
of specific manifestations are not adequately explained by the theoretical
perspectives or the empirical evidence gathered so far. The directionality
(inward bound or outward bound) and strength of the aggressive intent
are key issues in explaining the suicidal and homicidal behaviours.
Current body of knowledge does not provide an adequate explanation to
the question why or how the same aggressive intent causes two different
manifestations.
An understanding of the theoretical perspectives is undoubtedly help-
ful in gaining a deeper knowledge of the dynamics involved in destructive
behavioural manifestations at the individual, interpersonal and group
levels. However, an overview of the various perspectives reveals the limita-
tions of each to adequately explain the destructive behavioural phenomena.
The neurological perspective provides an explanation that is idiosyncratic
in nature and looks for roots of violence in genetic, neuro-chemical and

101
Swati Mukherjee et al.

neurobiological anomalies. It elaborates upon the biological processes


that create behavioural dispositions and potential for aggressive be-
havioural manifestations. These potential behavioural dispositions are
manifested in various specific manners mediated by social and cultural
facilitations or reinforcements. The neurological perspective, thus, pro-
vides an answer to the ‘what’ of destructive behaviours, and seeks the
aid of socio-cultural perspectives to explicate upon the ‘how’ and ‘why’ of
specific behavioural manifestations. Keeping the caution to guard against
regarding the biological dispositions as deterministic, the neurobiological
perspective proves its worth by providing with an important criterion
to identify the people-at-risk, that is, individuals with known genetic,
hormonal or neuro-chemical anomalies that make them vulnerable to
violent behaviour.
A closely related position is taken by psycho-pathological perspective
that predicts aggressive behaviour based on the known psychopathology
of the individual. Research has shown that certain psychiatric disorders
are specifically linked to aggressive behaviours. Lion (1995) has provided
a summary of these disorders with known aggressive manifestations.
Though, it is known and accepted that lowered inhibitions and heightened
aggressiveness due to excitatory dyscontrol are at the base of violent be-
haviours, establishing a specific aetiology still remains. Moreover, the
dynamics of the process leading from anger to aggression, aggression to
violence, and ultimately to homicidal violence are yet to be explored.
Though, major mental disorders have been associated with an increased
probability of homicidal behaviour (Schamda et al., 2004), it usually occurs
when the individual is under the influence of delusions or hallucinations
(Palermo, 2007). Frustration, fear and general behavioural immaturity
have been associated with homicidal behaviour (Palermo, 2007), but
the threshold that turns the ‘normal’ individuals into ‘time bombs’ that
‘suddenly explode, and their destructive fury kills both those known and
unknown to them’ (Palermo, 2007: 8) remains to be defined in operational
terms for assessment and prediction purposes.
The socio-cultural perspective provides the larger canvas on which
the homicidal behaviour unfolds. The facilitations, reinforcements and
learning opportunities provided by the socio-cultural environmental
context, not only influence the manifestations of aggressive behaviour, but
also decide the social acceptability of particular behaviour patterns. The
cultural variations reported in the literature underline the need to validate

102
Suicide and Homicide

the theories and models developed to explain destructive behaviours


across cultures and develop predictive criteria based on a profile of the
people-at-risk.

REFERENCES

Averill, J.R. (1982). Anger and aggression: An essay on emotion. New York: Springer-
Verlag.
Bandura, A. (1973). Aggression: A social learning analysis. Eaglewood Cliffs, NJ:
Prentice-Hall.
Berkowitz, L. (1989). Frustration-aggression hypothesis: Examination and refor-
mulation. Psychological Bulletin, 106(1), 59–73.
Block, J. and B.C. Martin (1955). Predicting the behaviour of children under frustration.
Journal of Abnormal and Social Psychology, 51(2), 281–85.
Breed, W. (1963). Occupational mobility and suicide among white males. American
Sociological Review, 28(2), 179–88.
Bristowe, E. and J.B. Collins (1989). Family-mediated abuse of non-institutionalised
elder men and women living in British Columbia. Journal of Elder Abuse and Neglect,
1(1), 45–54.
Bronfenbrenner, V. (1979). The ecology of human development: Experiments by nature
and design. Cambridge, MA, Harvard University Press.
Burt, V.K. (1995). Impulse Control Disorder not elsewhere classified and adjustment
disorders. In H.I. Kaplan and B.J. Sadock (Eds), Comprehensive textbook of
psychiatry VI, Vol. 2. (pp. 1409–18). Baltimore, Maryland: Williams & Wilkins.
Cavan, R. (1928). Suicide. Chicago: University of Chicago Press.
Cohen, A.R. (1955). Social norms, arbitrariness of frustration, and status of the agent
of frustration in frustration-aggression hypothesis. Journal of Abnormal and Social
Psychology, 51(1), 222–26.
Cohen, D. (2000). Homicide-suicide in older people. Psychiatric Times, January 2000,
17, 1. Retrieved 4 March 2009 from http://www.baylor.edu/ content/ services/
document. php/28830.pdf
Colman, A.M. (2001). ‘latah.’ A Dictionary of Psychology. Retrieved 4 March 2009
from Encyclopedia.com: http://www.encyclopedia.com/doc/1O87-latah.html
Cooper, J.M. (1934). Mental disease situation in certain cultures: A new field for
research. Journal of Abnormal Social Psychology, 29, 10–17.
Cordess, C. (1995). Crime and mental disorder. In D. Chiswick and R. Cope (Eds),
Seminars in practical forensic psychiatry (pp. 14–51). London: Gaskell.
Daly, M. and Wilson, M. (2003). Why are homicide rates so variable between times
and places ? Presented at Symposium on Cultural & Ecological Foundations of the
Mind, Hokkaido University June, 2003. Retrieved 9 March 2009 from http://lynx.
let.hokudai.ac.jp/COE21/presentation/1stcefom21/Daly&Wilson.pdf

103
Swati Mukherjee et al.

Dollard, J., L. Doob, N. Miller, O. H. Mowrer and R. Sears (1939). Frustration and
aggression. New Haven, CT: Yale University Press.
DuRand, C.J., G.J. Burtka, E.J. Federman, J.A. Haycox and J.W. Smith (1995). A quarter
century of suicide in a major urban jail: Implications for community psychiatry.
American Journal of Psychiatry, 153(7), 1077–80.
Durkheim, E. (1951). Suicide: A study in sociology (J.A. Spaulding and G. Simpson, Trans.).
New York, NY: Free Press. (Original work published 1897).
Eronen, M., P. Hakola and J. Tiihonen (1996). Mental disorders and homicidal be-
haviour in Finland. Archives of General Psychiatry, 53(6), 497–501.
Feshbach, S. (1964). The function of aggression and the regulation of aggressive drive.
Psychological Review, 71(4), 257–72.
Folger, R. (1986). A referent cognitions theory of relative deprivation. In J.M. Olson,
C.P. Herman and M.P. Zanna (Eds), Relative deprivation and social comparison
(pp. 33–55). Hillsdale, NJ: Earlbaum.
Freud, S. (1957). Mourning and melancholoia. In J. Strachey (Trans. and Ed.). The
standard edition of the complete psychological works of Sigmund Freud (pp. 243–58).
London: Hogarth. (Original work published 1915.)
Freud, S. (1961). Beyond the pleasure principle. In J. Strachey (Trans. and Ed.). The
standard edition of the complete psychological works of Sigmund Freud: Vol. 18
(pp. 7–64). London: Hogarth. (Original work published 1920.)
Garbarino J. and A. Crouter (1978). Defining the community context for parent–
child relations: The correlates of child maltreatment. Child Development, 49(3),
604–16.
Heide, K.M. (2003). Youth Homicide: A review of the literature and blueprint for action.
International Journal of Offender Therapy and Comparitive Criminology, 47(1),
6–36.
Hempel, A.G., R.E. Levine, J.R. Meloy and J. Westermeyer (2000). A cross-cultural review
of sudden mass assault by a single individual in the oriental and occidental cultures.
Journal of Forensic Sciences, 45(3), 582–88.
Henry, A. and J. Short (1954). Suicide and homicide: Some economic, sociological and
psychological aspects of aggression. Glencoe, IL: Free Press.
Home Office (1993). Criminal statistics England and Wales. London: HMSO.
Joukamaa, M. (1997). Prison suicide in Finland, 1969–1992. Forensic Science Inter-
national, 89(3), 167–74.
Kerkhof, J.F.M. and W. Bernasco (1990). Suicidal behaviour in jails and prisons in The
Netherlands: Incidence, characteristics, and prevention. Suicide and Life-
Threatening Behaviour, 20(2), 123–37.
Krug, E.G., L.L. Dahlberg, J.A. Mercy, A.B. Zwi and R. Lozano (Eds) (2002). World
report on violence and health. Geneva: World Health Organization.
Liem, M., I. Deerenberg and P. Nieuwbeerta (2007). Homicide followed by Suicide:
A comparison with both homicide and suicide. Presented at the 8th Annual Con-
ference of the European Society of Criminology, Edinburgh, September 2008.
Retrieved 9 March 2009 from http://www.aic.gov.an/conferences/ 2008-homicide/
liem.pdf

104
Suicide and Homicide

Link, B.G., H. Andrews and F.T. Cullen (1992). The violent and illegal behaviour of
mental patients reconsidered. American Sociological Review, 57(3), 275–92.
Lion, J.R. (1995). Aggression. In H.I. Kaplan and B.J. Sadock (Eds), Comprehensive
textbook of psychiatry VI, Vol. 2. (pp. 310–17). Baltimore, Maryland: Williams &
Wilkins.
Loeber, R., E. Lacourse and D.L. Homish ( 2005). Homicide, violence, and developmental
trajectories. In R.E. Tremblay, W.W. Hartrup and J. Archer (Eds), Developmental
Origins of Aggression. (pp. 202–22). New York, NY: The Guilford Press.
Maxfield M.G. and C.S. Widom (1996). The cycle of violence: Revisited 6 years later.
Archives of Pediatrics and Adolescent Medicine, 150(4), 390–95.
Menninger, K. (1938). Man against himself. New York: Harvest.
Miller, N.E. (1941). The frustration-aggression hypothesis. Psychological Review, 48(4),
337–42.
Nock, M.K. and P.M. Marzuk (1999). Murder-suicide: Phenomenology and clinical
implications. In D.G. Jacobs (Ed.), Harvard Medical School guide to suicide assess-
ment and intervention. (pp. 188–209). San Francisco: Jossey Bass.
Nock, M.K. and P.M. Marzuk (2000). Suicide and violence. In K. Hawton and K. van
Heeringen (Eds), The International Handbook of Suicide and Attempted Suicide
(pp. 437–56). Chichester, England: John Wiley & Sons.
Palermo, G.B. (1994). Murder-suicide: An extended suicide. International Journal of
Offender Therapy and Comparative Criminology, 38(3), 205–16.
Palermo, G.B. (2004). The faces of violence. Springfield, IL: Charles C. Thomas.
Palermo, G.B. (2007). Homicidal syndromes: A clinical psychiatric perspective. In
Richard N. Kocsis (Ed.), Criminal profiling: International theory, research, and
practice (pp. 3–26). Totowa, NJ: Humana Press.
Paolucci, E.O., M.L. Genuis and C. Violato (2001). A meta-analysis of the published re-
search on the effects of child sexual abuse. Journal of Psychology, 135(1), 17–36.
Peck, D.L. (1979). Fatalistic suicide. Palo Alto, CA: R and E Research Associates.
Pillemer, K.A. and D. Prescott (1989). Psychological effects of elder abuse: a research
note. Journal of Elder Abuse and Neglect, 1(1), 65–74.
Schamda, G., G. Knecht, D. Schreinzer, T. Stompe, G. Ortwein-Swoboda and
T. Waldhoer (2004). Homicide and major mental disorders: A 25-year study.
Acta Psychiatrica Scandinavica, 110(2), 98–107.
Schlesinger, L.B. (2007). Psychopathology of homicide. In A.M. Goldstein (Ed.).
Forensic Psychology: Emerging topics and expanding roles (pp. 708–33). Hoboken,
New Jersey: John Wiley & Sons, Inc.
Smith, M.D. and R.N. Parker (1980). Type of homicide and variation in regional rates.
Social Forces, 59(1), 136–47.
Stark, E. and A. Flitcraft (1995). Killing the beast within: Woman battering and female
suicidality. International Journal of Health Services, 25(1), 43–64.
Steadman, H.J., E.P. Mulvey, J. Monahan, P.C. Robbins, P.S. Applebaum, T. Grisso
et al. (1998). Violence by people discharged from acute psychiatric inpatient
facilities and others in the same neighbourhoods. Archives of General Psychiatry,
55(5), 393–401.

105
Swati Mukherjee et al.

Strube, M.J., C.W. Turner, D. Cerro, J. Stevens and F. Hinchey (1984). Interpersonal
aggression and the Type A coronary-prone behaviour pattern: A theoretical
distinction and practical implications. Journal of Personality and Social Psychology,
47(4), 839–47.
Swanson, J.W., C.E. Holzer, V.K. Ganju and R.T. Jano (1990). Violence and psychiatric dis-
orders in community: Evidence from the Epidemiologic Catchment Area Survey.
Hospital and Community Psychiatry, 41(7), 761–70.
Thibaut, J.W. and H.H. Kelly (1959). The social psychology of groups. New York: Wiley.
van Praag, H.M., R. Plutchik and A. Apter (Eds) (1990). Violence and suicidality: Per-
spectives in clinical and psychobiological research. New York: Brunner Mazel.
Volavka, J. (1995). Neurobiology of violence. Washington, DC: American Psychiatric
Press.
Weiger, W.A. and D.M. Bear (1988). An approach to the neurology of aggression.
Journal of Psychiatric Research, 22(2), 85–98.
WHO Global Consultation on Violence and Health. (1996).Violence: a public health
priority. Geneva, World Health Organization, (document WHO/EHA/ SPI.POA.2).
WHO (2002). World report on violence and health. Geneva: World Health Organization.
Wolfgang, M. and F. Ferracuti (1967). The subculture of violence. London: Social
Science Paperbacks.
Zillmann, D. (1978). Attribution and misattribution of excitatory reactions. In J.H.
Harvey, W.J. Ickes and R.F. Kidd (Eds), New directions in attribution research:
Vol. 2. Hillsdale, NJ: Erlbaum.

106
6

Cultural Issues in Suicide Risk Assessment*


EÙM®Ä®ƒ Cʽ瑑®

… no one who commits suicide does so without reference to the prevailing


normative standards and attitudes of the cultural community. Therefore,
cause of suicide can be understood only with reference to the socio-cultural
norms and attitudes that govern suicide in each cultural community. (Boldt,
1988: 106)

THE IMPACT OF CULTURE ON SUICIDAL BEHAVIOUR

O verall, suicide rates of different countries tend to be relatively


stable over time and very different from one another. For example,
Lester (1987, cited in Zonda and Lester, 1990) found that suicide rates of
European countries in 1975 were strongly associated with the suicide rates
of 100 years earlier. The different suicide rates persist when immigrants

∗ Acknowledgements: Thanks to Anne O’Hanlon for granting permission to reproduce parts


of Colucci (2006).
Parts of this chapter were previously published in Colucci, E. (2006). The cultural facet of
suicidal behaviour: Its importance and neglect. Australian e-Journal for the Advancement
of Mental Health, 5(3), 1–13. Retrieved December 25, 2006, from
www.auseinet.com/journal/vol5iss3/colucci.pdf. Permission of reproduction has been
granted.

107
Erminia Colucci

from these countries are examined in the United States, Canada and
Australia (De Leo, 2002; Dusevic et al., 2002; Lester, 1994). Similar con-
siderations led to Zonda and Lester’s (1990: 381) conclusion that ‘these
national and regional variations in suicide rates point to the possible role
of cultural factors’. In addition, De Leo (2002), interpreting the World
Health Organization (WHO) rates of suicide in different countries, noted
that epidemiological studies provide evidence that social and cultural
dimensions amplify any biological and psychological aspect. In particular,
the male/female ratio appears to be particularly influenced by the cultural
context (De Leo, 2002).
Similarly, other researchers have noticed cultural differences in the
epidemiology of suicidal behaviour among a range of countries. For
example, Mayer and Ziaian (2002) and Vijayakumar (2005) pointed out
different suicide patterns in Asian compared to Western countries. For
instance, the age distribution and male to female ratio are different: rates
are highest in the elderly in Western countries, but in young people in Asia.
In the former, the male to female ratio is greater at 3 (or more):1 whereas in
the latter the ratio is smaller at 2:1, with some countries like India show-
ing a very similar ratio (1.4:1) and China showing higher suicide in
women (Vijayakumar, 2005). Emphasising further the presence of import-
ant socio-cultural differences among countries, the selective review of
Vijayakumar, John, Pirkis and Whiteford (2005) pointed out that in some
developing countries (e.g., India) being female, living in a rural area and
holding religious beliefs that sanction suicide, may be of more relevance
to suicide risk than the same factors in developed countries. On the other
hand, being single or having a history of mental illness may be of less
significance. Similar findings and reflections indicate how important it is
for researchers to identify which findings have cross-cultural generality
and which are culturally specific (Lester, 1992–93; Mishara, 2006).
Considerations of this kind led various scholars to recognise that sui-
cide is a phenomenon that needs to be studied and understood in its social
and cultural milieu. For instance, Tseng (2001: 392) stated that ‘suicide,
even though it is a personal act, is very much socio-culturally shaped and
susceptible to socio-cultural factors’ and Kazarian and Persad (2001)
affirmed that the embrace of culture and life-enhancing perspective to
research and practice are likely to contribute to better understanding of
suicidal behaviour and to improved individual, family and community
well-being. Range and colleagues (1999), after examining suicide among
African Americans, Hispanic Americans, Native Americans and Asian

108
Cultural Issues in Suicide

Americans, declared that suicide must be studied from all angles, and
ethnic origin is one of the characteristics that must be recognised and con-
sidered in assessing risk and designing interventions.
In spite of the well-established and long-term interest in socio-
cultural aspects of suicidal behaviour, the research in this area is still in an
embryonic stage and, as Kral (1998: 225) underlined, ‘we are only beginning
to look seriously at the power of cultural ideas like suicide’. Furthermore,
as pointed out by Lester (1992–93), although culture may influence the
incidence of suicide, the circumstances and the methods, the reasons and
meanings of suicide, most researchers have focused on the association
between culture and incidence of suicide. This was also underlined by
Marsella (2000).
This partially finds its reason in the fact that, even though some re-
searchers attempted to study the way in which culture influences suicidal
behaviour, the conceptual consideration (i.e., theorisation) of the inter-
face between culture and suicide has been, with few exceptions (e.g.,
Durkheim, 1897/1997), an overall recent phenomenon (Kazarian and
Persad, 2001).
As an example of a theoretical explanation, Cohen, Spirito, Apter and
Saini (1997) hypothesised that culture affects the development of psy-
chopathology which, in turn, affects suicide rates. Similarly, Tseng (2001)
applied his theorisation of the effects of culture on psychopathology
to suicidal behaviour, indicating various effects of culture on suicide,
although suggesting an arguable application of the pathological frame to
suicidal behaviour:

1. Culture contributes to the nature and severity of the distress that


people may suffer (e.g., China and Korea prohibit the union of cer-
tain couples), which may then contribute to the occurrence of their
suicidal behaviour (pathogenetic effects of culture).
2. Culture demonstrates pathoselective effects in a person’s choice
of suicide over other possible solutions to problems (e.g., facing
bankruptcy).
3. The pathoplastic effects of culture on suicide are well illustrated
by the manifestation of special forms of suicidal behaviour in add-
ition to individual personal suicide, such as family suicide, group
suicide and mass suicide or seppuku (traditionally observed in
Japan) and suttee (practised in India in the recent past).

109
Erminia Colucci

4. A pathoelaborating effect may be shown in various terminologies


that have evolved to recognise and distinguish different forms of
suicide, such as Japan, where laypersons use many terms to refer
to different kinds of suicide.
5. The pathofacilitative effects are well proven by the variation in
frequencies and rates of suicide among different societies.
6. Many societies have a very negative attitude towards suicidal be-
haviour: Muslims see it as an unforgivable sin and Indians still
see it as crime, whereas Japanese have a more sympathetic view. Atti-
tudes and stigmas show the pathoreactive effects of culture on
suicidal behaviour.

The impact that culture has on suicide has been the focus of research
by a number of scholars although, as mentioned before, compared with
other aspects of suicidal behaviour, this has been a rather neglected area of
study, considering its importance, as observed also by other scholars (e.g.,
Eskin, 1999; Kral, 1998; Leenaars et al., 2003; Shiang, 2000; Tortolero and
Roberts, 2001; Trovato, 1986). Watt and Sharp (2002) noted that there are
relatively few available cross-cultural studies of suicide, and they are mainly
on adults; usually young people are not examined separately. Captivated
by this observation, Colucci and Martin (2007a, 2007b) reviewed all the
trans/cross-cultural studies on youth suicide. The findings from the 82
references matching the review criteria were published in two papers: one
on suicide rates and methods (Colucci and Martin, 2007a) and the other on
risk and precipitating factors, and attitudes towards suicide (Colucci and
Martin, 2007b). The main findings and considerations from this review
will be summarised in the following section (readers are referred to Colucci
and Martin [2007a, 2007b] for a more detailed review and to Leach [2006]
for a review of the literature also on other age groups).

ETHNOͳCULTURAL ASPECTS OF YOUTH SUICIDE

Roberts, Chen and Roberts (1997: 209) pointed out: ‘In general, ethnicity
has been little studied in relation to suicidal behaviours; results from the
few studies that have examined ethnic differences have been equivocal.’
Also the results of the first part of the literature review (Colucci and Martin,

110
Cultural Issues in Suicide

2007a) were not homogenous and many studies seemed to provide dis-
cordant results on the epidemiology of suicide among youth belonging
to different ethnic groups. However, some trends can be traced: for
instance, the increase of suicide for young African Americans, particularly
in young males (which are levelling the White versus Black suicide rates;
see McIntosh, 1989), a pattern of extremely high peaks of suicide in young
Pacific Islanders (Tseng et al., 1982) and aboriginal peoples (e.g., Kirmayer,
1994), and more frequent suicide attempts among Asian females compared
to females of other ethnic groups (e.g., Bhugra et al., 1999). Suicide rates
seem to be particularly high (Raleigh, 1996) and increasing (Bhugra et al.,
1999) in young people from the Indian subcontinent. Hispanics appear to
have higher rates of self-harmful behaviour than Whites (e.g., Gutierrez
et al., 2001).
The effect of migration on suicide statistics was the focus of a few studies
but the data are ambiguous. For instance, Sorenson and Shen (1996)
showed that, in California in the period 1970–92, foreign-born Whites
were at higher suicide risk, foreign-born Blacks and Asian/Others were
at similar risk, but foreign-born Hispanics were at lower risk. Like the
prevalence of suicidal behaviour, the method chosen for suicide is also
(at least partially) culturally determined (e.g., Colucci, 2008a; Sorenson
and Shen, 1996).
The second part of the review (Colucci and Martin, 2007b) covered
the cross-cultural literature regarding youth suicide risk and precipitat-
ing factors, and attitudes towards suicide (it also provided suggestions
for future research on cultural aspects of suicide). Overall, cross-cultural
studies in young people have demonstrated many of the same risk
factors for suicidal behaviour as found in more general research; for
instance, depression, exposure to suicide, previous suicidal behaviour and
interpersonal problems. However, some differences between ethnic groups
emerged as well. For instance, while the cross-cultural literature showed
that previous suicidal behaviour is a strong predictor of future suicidal
behaviour, Heisel and Fusé (1999) found higher levels of suicidality over
time in Japanese previous suicide attempter students but not in Canadians.
Exposure to suicidal attempts and suicides of relatives, parents and
friends is also known to promote the same kind of behaviour, but few pub-
lications have analysed the impact of exposure on different ethnic groups.
Although this research is sparse, it does appear that exposure to suicidal
behaviour may be a universal risk factor across cultures.

111
Erminia Colucci

With regards to interpersonal factors, cross-cultural research showed


the impact of low family support, unsatisfying parental relations, or
parental derogation on suicidal ideation and behaviour (e.g., Perkins and
Hartless, 2002). However, here ethnicity also seems to play a role. For
instance, of adolescents presenting to a UK accident and emergency depart-
ment for self-poisoning (Biswas, 1990), significantly more White than
Asian boys reported parental conflict with possibility of divorce. Also
Handy and colleagues’ (1991) study showed that more self-poisoned
English youth came from a disrupted family than did young Asian
people. On the other side, in a similar UK emergency department study,
Kingsbury (1994) showed that Asian adolescents, with lower suicidal
intent but higher rates for suicide risk factors, experienced parents as
more controlling than their Caucasian counterparts.
In a study by Vega, Gil, Zimmerman and Warheit (1993) parent dero-
gation was an important risk factor for African American and Nicaraguans
but not for White boys. Another important suicide risk factor is a rela-
tional problem with peers and friends, although here also there are some
cultural (and gender) differences (Colucci and Martin, 2007b). School
connectedness emerged as a significant protective factor for suicide
thinking (Resnick et al., 1997). This has also been shown in cross-cultural
studies of suicide attempts in Hispanic and White students but not in
Blacks by Borowsky and colleagues (2001), whereas Rew and colleagues
(2001) demonstrated protection in all three ethnic groups.
Physical and sexual abuse are risk factors for suicidal behaviour among
young people has been demonstrated among the Black, Hispanic and
White representative US sample investigated by Borowsky et al. (2001)
and Rew et al. (2001). However, in a sample of White and Black students
(Thatcher et al., 2002), sexual and physical abuse were associated with
increased risk of attempted suicide in White males and females but not
in Blacks.
A few studies have examined personality aspects, such as coping style,
locus of control, anxiety and hopelessness (Colucci and Martin, 2007b).
Eshun (1999) demonstrated a similar significant positive correlation
between suicidal ideation and hopelessness in both Americans and
Ghanaians.
A number of comparative studies have demonstrated relationships
between depression and suicidal thinking or suicide attempts across
samples comprising Indian, Chinese, Malaysian, Kuwaiti, Turkish,

112
Cultural Issues in Suicide

Mexican American, Filipinos, and in Northern Plains and Pueblo Indian


tribes (Colucci and Martin, 2007b). However, in the model developed by
Gutierrez and colleagues (2001), depression had an impact on suicidal
ideation in all groups (White, Black and Hispanic students), but the
influence of depressive symptoms on history of suicide attempts was
higher in Blacks than Whites.
Another area extensively studied is alcohol and drug use and abuse in
suicidal people, but these risk factors are also influenced by cultural deter-
minants (e.g., Vega, Gil, Warheit, Apospori and Zimmerman, 1993). As
observed in Colucci and Martin (2007b), suicide-precipitating factors are
also culturally dependant.
Besides risk and protective factors outlined earlier, Colucci and Martin
(2007b) reported the findings on several studies examining attitudes
towards suicide in different ethnic groups. In summary, young Americans
showed more positive attitudes towards suicide than Ghanaian, New
Zealander, Nigerian and Mexican American youth, but Canadians and
Japanese were more accepting of suicide than Americans. Indian students
manifested more negative attitudes towards suicide than both Dutch and
Austrian students, and Singaporeans more than Australians. Other groups
that have been compared on this aspect are Zambians, Swedish, Chinese
and Turkish. More positive attitudes seem to be correlated with suicidal
ideation (Domino et al., 2001–02; Eshun, 2003; Zhang and Jin, 1996), but
more studies are needed to clarify this relationship.
Basically, as shown by the review, there are some risk factors that might
be generalisable to almost all ethnic groups, and others more specific to a
given culture or ethnic group. Furthermore, the same factors may influ-
ence but to a greater or lesser degree, as demonstrated by Perkins and
Hartless (2002) in European American and African American students. For
example, hopelessness was a significant predictor for African Americans
and European Americans but the association was significantly greater for
European Americans, as was also physical abuse. Hard drug use was a
stronger predictor for European Americans. Also in Vega, Gil, Zimmerman
and Warheit (1993), the same risk factors had different importance for
suicide attempts across the ethnic group: for African American and Hispanic
boys, low self-esteem as well as negative parent and teacher perception
were more important, while for White boys personal deviancy and de-
linquent behaviour were more important in predicting suicide attempts.
To date, however, little research has been conducted examining whether
the extent of risk factors in typically increasing suicide risk differs based

113
Erminia Colucci

on cultural groups, as also observed by Leach (2006). Beyond the specific


variable importance in the determination of suicidal behaviour, risk factors
interact in different ways depending on the specific culture examined. For
example, Yuen, Nahulu, Hishinuma and Miyamoto (2000) estimated that
suicide attempts were best predicted for Hawaiian students by depression,
substance abuse, grade level, Hawaiian culture affiliation and the main
wage earner’s education, compared to substance abuse, depression and
aggression for non-Hawaiians.
Concluding, both parts of the literature review underlined (also in light
of often discordant results) the urgent need to extend and deepen the cross-
cultural research on suicide, to more exhaustively understand the influence
of cultural dimensions on the personal choice to end life (Colucci and
Martin, 2007a).1 Following this understanding, more culturally adequate
suicide risk assessments and prevention strategies may be possible.
While the literature review mainly explored epidemiological studies,
other scholars reflected on the way culture affects the particular mean-
ing attributed to suicidal behaviour in various cultures. This will be
briefly examined in the following section (for further discussion, see
Colucci, 2006).

THE CULTURAL MEANINGS OF SUICIDE

As argued by Leenaars, Maris and Takahashi (1997: 2):

Individuals live in a meaningful world. Culture may give us meaning in


the world. It may well give the world its theories/perspectives. This is true
about suicidology. Western theories of suicide, as one quickly learns from
a cultural perspective, may not be shared. Suicide has different meanings
for different cultures.

1
It must be noted that, at the end of the review, I was critical of the current cross-cultural
literature on youth suicide for different reasons (see Colucci and Martin (2007a, 2008), for
more details). In particular, I criticised the fact that the majority of the studies have been
conducted in Western developed countries, especially in the United States. Furthermore,
the majority of youth suicide cross-cultural research is epidemiological (at the time of the
review, I could not find any qualitative cross-cultural study on youth suicide) and cross-
national instead of cross-cultural.

114
Cultural Issues in Suicide

Shneidman cautioned us, when making cross-cultural comparisons, to


not make the error of assuming that ‘a suicide is a suicide’ (1985, cited in
Leenaars et al., 1997). Lester (1997) too recognised that suicidal behaviour
may be quite differently determined and have different meanings in
different cultures. In Suicide in Different Cultures, Farberow (1975) also
affirmed that suicide is viewed very differently by different cultural groups
and that culture influences form, meaning and frequency of suicide. Of
the same opinion are Maris and Lazerwitz (1981) and Hendin (1964, cited
in Boldt, 1988) who made the point that suicide varies culturally and that
differences in meaning may have an influence on suicide. Also Durkheim,
as Boldt (1988) revealed, explicitly recognised the potential influence of
cultural meanings on suicide rates, but excluded meanings from his
analysis because he believed the Protestants and Catholics, comprising
his sample, espoused the same meanings.
Discourses on meaning of suicide may be confused with discussion on
the description of the word ‘suicide’ but the ‘meaning’ of suicide, as Boldt
(1988) pointed out, must be differentiated from the ‘definition’ of suicide.
He stated that meaning goes beyond universal criteria for certifying and
classifying self-destructive deaths, to how suicide is conceptualised in terms
of cultural normative values. He then listed some examples of peculiar
socio-cultural conceptualisations of suicide: suicide as an unforgivable sin,
a psychotic act, a human right, a ritual obligation, an unthinkable act.
Despite the number of scholars who have underlined the relevance of
what suicide means to people belonging to different socio-cultural back-
grounds, the study of meaning is an unjustifiably missing area in suicide
research. To date studies analysing this aspect are very rare and Meng’s
(2002) paper on suicide as a symbolic act of rebellion and revenge for some
Chinese women is one of very few exceptions. Everall (2000), in her study
about the meaning of suicide in young people, noted that despite the
amount of research conducted in suicidology, surprisingly little is
known about the experience of being suicidal and argued that ‘while
demographic variables may be useful in identifying at-risk groups, they
provide little in the way of meaningful understanding of the suicidal
individual’ (p. 111). In a similar way, Boldt (1988: 95) showed concern
about the scarce consideration manifested for the study of the meaning of
suicide: ‘Suicidologists use the term “suicide” as though there is no need
to understand its meaning. This neglects the fact that meaning precedes
ideation and action, and that individuals who commit suicide do so with
reference to cultural-normative specific values and attitudes.’

115
Erminia Colucci

Boldt (1988) and Douglas (1967) tried to find some reasons for this
neglect and both of them concluded that perhaps the main reason resides
in the scholars’ tendency to not pay attention to fundamental (but taken
for granted, obvious) things. Another reason for the few studies to date on
the cultural meaning of suicide might be linked also to the arduousness
of this kind of study, for the difficulty to get in contact with meanings—
not only for the researcher but for the subject under study as well—as
stated by both Boldt and Douglas. But, of course, the difficulty of fully
understanding the meaning of suicide should not become a justification
to not dedicate as much effort and resources as possible to this import-
ant topic. On the contrary, the recognition and study of cultural relativity
in the meaning of suicide is an urgent need in the present phase of social
scientific suicide research. Only by differentiating as precisely as possible
the culture-dependant meanings of suicide, and by systematically bringing
these into the research paradigm, can the development of valid theories
of ‘cause’, prevention and treatment begin.
Trying to amplify this field of knowledge, I explored the cultural mean-
ings of youth suicide among university students of 18–24 years of age
in three different countries—Italy, India and Australia—using a com-
bination of qualitative and quantitative methods (Colucci, 2008a). Some
of the findings from this study are presented in the next section, as an
exemplification of the way culture may influence several aspects of suicidal
behaviour.

THE CULTURAL MEANINGS OF SUICIDE: A COMPARISON


BETWEEN ITALIAN, INDIAN AND AUSTRALIAN YOUTHS

In this study (Colucci, 2008a), data was collected through a multi-method


approach, using a questionnaire with structured and semi-structured
questions (such as case scenarios, word association, attitude scale, open-
ended questions) and tape-recorded focus groups2 (research methods are
partially described in Colucci, 2007, 2008b).

2
Quantitative data was analysed using SPSS 13.0. Qualitative data was analysed separately
and then discussed by two bilingual psychologists and myself. The categories so developed
were compared with those of a third psychologist, to create a final list of codes. The coding
process was supported by the software for qualitative analysis ATLAS.ti 5.0.

116
Cultural Issues in Suicide

Almost 700 students across countries took part in the first phase of the
study (i.e., survey) and 96 participated in the following focus groups (two
sessions for each group for a total of 24 sessions).
The comparisons highlighted differences and similarities across cul-
tures in meanings and social representations of suicide. First, there were
differences on prevalence: more than half of the total sample reported sui-
cide ideation but this was higher among Italian and Australian participants,
compared to Indians. In contrast, the latter reported more suicide attempts,
followed by Australians and then Italians.
Other questions investigated reasons for young people to attempt
suicide or to indeed suicide. There were statistically significant differences
on almost all suicide attempt reasons between cultures. For example,
Indians showed higher agreement that youth at times attempt suicide to
force others to do what they want. Compared with the other two samples,
Italian participants showed higher disagreement that youth who attempt
suicide are mentally ill. Another question asked to rank a list of 14 reasons
for youth suicide. Participants presented statistically significant differences
on all of them. For example, financial problems were among the most
important reasons for Indians. Mental illness, depression and anxiety were
more important for Australians and loneliness or interpersonal problems
were so for Italians.
The questionnaire also included a 21-item attitudes scale. Both mean
scores on the single items and subscales scores showed cultural differences.
For example, Indians, followed by Australians, had more negative attitudes
towards youth suicide compared to Italians.
The open-ended section of the questionnaire was composed of various
parts investigating participants’ mental associations with the word ‘suicide’
and interpretations of both this word and ‘attempted suicide’, feelings
about death, stereotypes of the ‘kind of’ youth who attempt suicide or
kill themselves, reasons for living and suicide prevention strategies. For
instance, when asked for which reasons they would not suicide, participants
from the three countries wrote similar motivations, referring to the value
and love for life, loved ones and the belief that difficulties are part of life
and can be overcome. But there were also differences. In India, for example,
participants more frequently mentioned God as a deterrent against suicide
compared to participants in Italy and Australia. Italians rarely expressed
negative judgements towards suicide (e.g., suicide is selfish) to justify
the choice not to suicide, whereas this was quite frequent in Indians,
followed by Australians. Furthermore, Australians more often expressed

117
Erminia Colucci

the hope that they would get some help and support compared with the
other groups.
In relation to help-seeking, overall the majority of students reported that,
if they were thinking about killing themselves, they would talk to no one or
talk to friends, followed by someone in the family. Some students, especially
in Australia, referred to professional help. Focus group transcripts helped
to further understand questionnaire answers and pointed out issues
such as altruistic suicide (i.e., suicide to not be a burden on the family) and
expected/forced suicide in India, the pressure to be ‘macho’ thus not
expressing emotions and sharing problems and the conflict of expectations
between friends and adults in Australian men, and the ‘involvement’ in
other people’s life in Italian youth.
Although gender issues were not the specific aim of the study, the cross-
cultural comparison revealed several differences based on participants’
sexes (e.g., Indian females thought, planned and attempted suicide sig-
nificantly more than Indian males). Most importantly, it was evident that
gender issues were central in several participants’ beliefs and narratives
about youth suicide, particularly among Indians, followed by Australians.
For example, there was generally a slightly more accepting attitude in
females compared to males and more negative attitude in males, especially
Indian males. In each country, a lower propensity to ask for help was re-
ported by males, and this was amplified in the Australian sample where it
was also stressed a greater social pressure towards males to conform to the
‘male image’ (i.e., macho-man) repressing emotions and feelings. These sort
of findings highlights that it is also critical to consider gender differences
when exploring the cultural meanings of suicidal behaviour.3 In summary,
a number of culture-related issues emerged in this study which emphasise
the importance of developing culturally sensitive suicide risk assessments
and prevention strategies.

3
For example, Counts (1988, cited in Lester, 1992–93) illustrated the ways in which a culture
can determine the meaning of the individual suicidal act in her account of suicide among
females in Papua New Guinea, where female suicide is a culturally recognised way of imposing
social sanctions, with political implications for the kin and for those held responsible for the
events leading to the act. A similar study of accounts of suicidal behaviour showed that Sri
Lankan participants associated essentialist accounts with women’s suicides and contextual
accounts with men’s suicides (Marecek, 1998). Canetto and Lester (1998) also suggested
that narratives of suicidal behaviours can be examined through the lens of gender-specific
cultural scripts.

118
Cultural Issues in Suicide

As part of understanding the cultural aspects of suicide, I also in-


vestigated participants’ spiritual beliefs, which later became one of my main
research interests. The inner experience of spiritual and religious feelings
is an integral part of the everyday lives of many individuals. Therefore,
when considering ethno-cultural issues in the suicide risk assessment, we
certainly cannot avoid (or should not) investigating our client’s (religious
and/or not religious) spiritual beliefs. In this regard, Bhugra and Osbourne
(2004: 6) recommended that the role of religion/spirituality ‘in individual’s
cultural identity must be taken into account when ascertaining mental state
and formulating management strategies’. Shafranske and Malony (1996,
cited in Johnson and Hayes, 2003) suggested that religion/spirituality needs
to be considered like any other client’s cultural characteristic and mental
health professionals have an ethical obligation to consider spirituality a part
of a standard assessment. At the opposite, to date mental health sciences
have ‘excluded spirituality apart from seeing it as a form of pathology or
pathological response’ (Dein, 2005: 538).
For over 100 years scholars from several disciplines have contributed
to the study of the role of religion as a deterrent to suicidal behaviour. The
following section offers a synthesis of existing literature investigating
the relationship between religion/spirituality and suicide keeping the
‘suicide continuum’ as a reference: from suicidal ideation to non-lethal
suicidal behaviour to lethal suicidal behaviour. Studies giving indications
of religious beliefs that might lead to suicide and spiritually oriented
intervention strategies for suicidal patients and suicide survivors are also
discussed. Particular emphasis is given to the importance of routinely
taking spiritual issues into account when assessing people at risk of sui-
cide and placing a more significant emphasis on spirituality in suicide
prevention and intervention strategies: ‘[S]uicidology cannot continue to
turn a blind eye to the central role that spirituality often plays in the ex-
perience of and recovery from suicidality’ (Webb, 2003: 5).

RELIGION AND SPIRITUALITY IN SUICIDE


RISK ASSESSMENT

Scholars and clinicians might have their own beliefs regarding the im-
pact of spirituality/religion on suicidal behaviour and its role in suicide

119
Erminia Colucci

prevention and intervention. However, if they desire to provide evidence


(pros or cons) for their ideas, they might find it difficult to ‘disentangle’
themselves from the literature which, in one way or the other, addressed
this subject. For this reason, I attempted to give an aid for this difficult task
by systematising the existing literature investigating the relationship
between religion/spirituality and suicide (Colucci and Martin, 2008). This
article began with an overview of the attitudes of the main religions (e.g.,
Catholicism, Islam and Buddhism) towards suicide, followed by the
mainstream theories developed to ‘explain’ the link between religion and
suicide. The core of the article was the description of the main findings from
studies on religion/spirituality distributed in three parts of the ‘suicidal
path’: suicidal attitudes and ideation, non-lethal suicidal behaviour and
lethal suicidal behaviour. In the article, attention was also given to the
(few) studies presenting indications of religious beliefs as a risk factor for
suicidal behaviour, studies on suicidal behaviour prevention and treatment
and on survivors. In this chapter, I summarise the main findings from
the cited literature review but readers are referred to Colucci and Martin
(2008) for the complete review.
When the literature was analysed along the suicidal path, the main
findings were as follows:4

Spirituality in Thought: Suicidal IdeaƟon and Aƫtudes

Overall, studies have shown that religious factors are associated with lower
suicidal ideation/plan and with more negative attitudes towards suicide.
For instance, in the study mentioned earlier (Colucci, 2008a), I showed
that Indian students who defined themselves as religious/spiritual report-
ed lower suicidal ideation compared to those who were non-religious/
spiritual. Furthermore, participants’ specific religious preference was

4
Inconsistent findings are present in the literature investigating the spiritual/religious
variables and suicide (Colucci and Martin, 2008). A possible reason for the disparity in
results is the plethora of indicators used to study the impact of religion on suicidal behaviour.
Another issue that must be taken into consideration is that the impact of religion/spirituality
changes in different cultural and socio-political contexts and during historical periods.

120
Cultural Issues in Suicide

associated with presence of suicidal thoughts. There was an association


also between students’ self-reported religiosity/spirituality and attitudes
towards youth suicide as measured by the 21-items Attitude towards
Youth Suicide scale (AtYS; Colucci, 2008a). In particular, whilst there
were no statistically significant differences in India between religious/
spiritual and non-religious/spiritual students, in Italy religious/spiritual
students believed more that suicide is preventable and agreed less that it
is acceptable. In Australia, religious/spiritual students manifested more
negative attitudes and less acceptability. Scores on the subscales were also
analysed looking at the specific religious affiliations, which significantly
impacted the scores on each of the four attitudes subscales.
However, other scholars have shown that religiosity is only weakly
associated (Lester and Francis, 1993) or not associated (Dervic et al., 2004;
Eshun, 2003; Kamal and Loewenthal, 2002; Loewenthal et al., 2003) with
suicidal ideation. On the other hand, participants with distress and strain
related to religious or spiritual concerns reported more suicidal ideation
(Exline et al., 2000; Johnson and Hayes, 2003). In the former study, it
was the belief of having committed a sin too big to be forgiven which was
primarily associated with suicidal thoughts.
An observation from this first part of the review (Colucci and Martin,
2008) was that research has rarely focused on non-religious forms of
spirituality and meaning in life or has put religion/spirituality in relation
to the ethno-cultural context.

Spirituality in Non-lethal Suicidal Behaviour

First of all, of the three, this was the part of the suicide path least investigated
by scholars. Although a few studies have shown that religious/spiritual
persons have lower rates of suicide attempts, a few others did not find
any association between religiosity and suicidal behaviour. For instance,
whilst in Italy and Australia a similar percentage of religious/spiritual
and non-religious/spiritual students attempted suicide, Indian students
who recognised themselves as non-religious/spiritual attempted suicide
more often than religious/spiritual participants; the specific religion was
also associated with suicide attempts (Colucci, 2008a). At the oppos-
ite, Loewenthal et al. (2003) found no association between participants’

121
Erminia Colucci

religiosity and suicide attempts. On the other side, a couple of studies


suggested that in some circumstances religion might be a risk factor for
such behaviour (e.g., Grossoehme and Springer, 1999).
It is important to note that research suggested that only some aspects
of religiosity (e.g., importance of religion) or some spiritual variables (e.g.,
sense of connectedness or coherence) might be associated with suicidal
behaviour (Colucci and Martin, 2008).

Spirituality in Lethal Suicidal Behaviour

The majority of research on lethal suicidal behaviour has focused on mor-


tality statistics and indicated that variables such as religious affiliation
and church attendance offer some protection against suicide. Breault and
Barkey (1982) not only found an association between religious integration
(as measured by religious book and newspapers production), but that
this relationship was exponential: past a certain point or threshold, even
a small decrease in integration was associated with a sharp increase in
suicide rates.
However, some studies have shown that the association between religion
and suicide is mediated or influenced by other variables underlining an
overall weak association between religion and suicide, for instance, level of
urbanisation (Faupel et al., 1987), whereas others have reported an overall
absence of a significant association (e.g., Wasserman and Stack, 1993).
The epidemiological data have been questioned and Sorri, Henriksson
and Lonnqvist’s study (1996) showed that in some cases religiosity can
support a suicide decision. Also in this ‘path section’, research on non-
religious spirituality or on essential aspects of religiosity/spirituality (i.e.,
meaning/purpose in life) is scarce.

HOW CAN CLINICIANS CONSIDER CULTURAL ISSUES


IN SUICIDE RISK ASSESSMENT?

Leach (2006: VII), like previous scholars, argued that ‘understanding cul-
tural nuances can assist with typical suicide assessment procedure to

122
Cultural Issues in Suicide

offer greater breath and depth to the evaluation, thus assisting clinicians
with their decision-making processes and interventions’. But how can we
consider cultural issues in the assessment?
When considering the client’s ethno-cultural background and spiritual
beliefs while assessing their risk for suicide, we might assume that there will
be a scale or similar psychometric measure readily available for this task.
This would be a wrong assumption because there is no tool developed for
a particular ethnic/racial/cultural group and no instrument might ever be
able to measure such multi-faceted and complex constructs. We can find
scales measuring concepts such as ethnic identity (see, for instance, Yamada
et al., 2002) but ‘understanding’ cultural aspects of suicide is a life-long
task, which requires much more than a list of items. Familiarising oneself
with the major faith traditions and cultural groups at least in the district
where the clinician works is a first step, but above all it is important to ask
clients’ explanations for things we cannot make sense of within our own
culture (even if sometimes clients might be surprised that we ask questions
for taken-for-granted facts) and to learn to listen ‘for understanding’.
Rumbold (2007: 61) stated that ‘spiritual assessment must be a process,
not merely an event’ and this applies, in a broader sense, to any cultural
assessment.
Having said this, some countries (e.g., India) have developed their own
suicide risk scales and clinicians might want to consider using them, al-
though these are generally published in local literature and only those
familiar with the language of publication might be able to use them.
If we look specifically at spiritual issues, in Kehoe and Gutheil’s (1994)
evaluation of suicide assessment instruments, the authors noted that,
although the psychiatric literature suggests that religion and spirituality
are significant and meaningful forces in suicidal patients, the number
of religious items included on assessment scales approaches zero. For
this reason, they criticise designers of suicide scales, which appear to seek
factors that may help to identify people at risk of suicide but ignore the
possible impact of what ‘a person, on the brink of life itself, believes about
life and about life after death’ (p. 368). As concluded by these authors,
front-line clinicians do not regularly investigate the religious area of a
person’s life as a factor in assessing suicidal risk. But, for those mental
health professionals sensitive to patients’ spiritual needs, scales are available

123
Erminia Colucci

which could potentially become part of the clinical assessment of a suicidal


person.5 A few also measure non-strictly religious aspects of spirituality
(e.g., Paloutzian and Ellison, 1982 cited in Ellison, 1983; Underwood
and Teresi, 2002; WHOQOL SPRB Group, 2002). Although rare, some
scales (e.g., Crumbaugh, 1977; Ryff and Keyes, 1995) address meaning or
a sense of life’s purpose, a primary component of spirituality which has
been neglected by the suicide literature (Colucci, 2008c).
However, like Swinton (2001), I am very critical of the use of quantitative
methods to study religion and spirituality and any reductionistic and
simplistic approaches to the investigation of these constructs. In the
cross-cultural journals reviewed by Tarakeshwar, Stanton and Pargament
(2003), these authors criticised that religious dimension is assessed through
a few global indicators (such as church affiliation, church attendance and
prayer) which do not reflect the multi-dimensional nature of religion.
The point, which was made also by Oldnall (1996), is that spirituality does
not lend itself to measurement in a natural science sense. On the contrary,
as postulated also by Burkhardt (1989), spirituality lends itself more to
qualitative measures, where the subjectivity of response is valued. Therefore
clinicians who prefer to use a psychometric measure to investigate the
cultural milieu in which their client’s life is embedded should not delimit
their assessment to these kind of tools. Talking specifically about religion,
clinicians should consider that it might be more important to ask ‘how’ a
person is religious rather than ‘whether’ a person is religious (Allport, 1950,
in Hill and Pargament, 2003). Regardless of the method used, clinicians
should investigate if their client’s spiritual/religious beliefs changed over
time or if they are different from the spiritual/religious beliefs held by
the people around them. Assessing clients’ beliefs about the afterlife is an

5
Readers interested in such scales might find it useful to consult the book Measures of
Religiosity by Hill and Hood (1999), which classifies more than one hundred scales on
religious development, beliefs, values, attitudes, attribution, orientation, practices, coping
and problem-solving, commitment and fundamentalism. Scales on spirituality, mysticism,
God concept, views of death/after life, forgiveness and others are represented in the book as
well. Furthermore, the Fetzer Institute and the National Institute of Aging (1999) convened a
panel of scholars with expertise in religiousness/spirituality and health/well-being to develop
items in order to assess health-relevant domains of religiousness and spirituality. The resulting
instrument ‘Multidimensional Measurement of Religiousness/Spirituality’ is composed of
various scales representing different domains of religiousness and spirituality (e.g., meaning,
values, beliefs, private religious practice, religious/spiritual coping).

124
Cultural Issues in Suicide

important part of suicide risk assessment, as it was also observed by Leach


(2006). However, clinicians must also bear in mind that, although religion
is generally regarded as a protective factor against suicide, a person who
defines him/herself as strongly religious might still be at risk of suicide.
They should also note that a few studies have suggested that religious beliefs
may expose individuals or groups of believers to a greater risk of suicidal
behaviour (e.g., Mancinelli et al., 2002; Zhang and Jin, 1996). Furthermore,
clinicians should not ask only about a specific religious affiliation but need
to be aware that people might hold multi-faith beliefs or might express a
non-religious form of spirituality. In particular, the spiritual dimension of
‘meanings and purpose in life’ must be investigated as part of the suicide
risk assessment (see Colucci, 2008c, for further discussion on spirituality
and meanings, and instruments).
Lastly, when undertaking suicide risk assessment, mental health pro-
fessionals need to be aware that more research is required to define which
aspects of religiosity and spirituality are protective against suicide, because
religious affiliation or simply attending church are not ‘necessary and
sufficient’ conditions to prevent suicidal behaviour (Colucci, 2008c).

CONCLUSIONS

Range and Leach (1998) remarked that research methodology in Sui-


cidology has historically developed from philosophical roots in logical
positivism and structural determinism. This had led to research based on
assumptions of cause–effect relationships, reductionistic analysis and focus
on the individual as the primary unit under study, which might explain why
relatively little research has addressed socio-cultural aspects of suicide. On
the other side, some mental health experts recognise that culture functions
as a lens through which we construct, define and interpret reality (Marsella
and Kaplan, 2002) and a growing number of scholars have underlined the
need to consider culture during suicide risk assessment and in planning
suicide prevention/intervention strategies. But the path to inclusion of
ethno-cultural considerations in mainstream mental health sciences and
Suicidology is still a lengthy and arduous one.
Kral (1998: 230) concluded his essay posing the question: ‘Is it time to
ask different questions in Suicidology?’ My answer is ‘definitely yes’ and

125
Erminia Colucci

my hope is that this chapter will act as an invitation for a larger number
of researchers, clinicians and policy makers to consider the socio-cultural
milieu of their participants, clients and communities when assessing and
treating suicidal ideation and behaviour.
In order to improve our ability to assess the risk of suicide it is not
enough to know what the principal suicide risk factors are. As observed
by Leach (2006: 3), ‘[I]t is through culture that we begin to understand
personal meaning, because culture offers the lens through which suicide
factors such as coping styles, buffers, emotional expressions, family struc-
tures, and identity can be viewed.’ As it has been highlighted in this chapter,
we need to understand the prevailing norms, meanings, social represen-
tations and attitudes regarding suicide in the many cultural (and sub-
cultural) communities of the world, even though this is a difficult task,
where no ‘true’ answer should be expected and no ‘right’ instrument
should be assumed. All people involved in suicide prevention—health
professionals, policy makers, spiritual guides, suicide survivors and so
on—are required to understand what the act of suicide symbolises and
represents for that person and that cultural group if we really want to help
them find a different way—constructive and not destructive—to express
and manifest those meanings.
In conclusion, the following are a few points that clinicians should bear
in mind during suicide risk assessment:

1. Although the literature on cultural influences on suicide is sparse


at best, we have some evidence to suggest that a number of sui-
cide risk and protective factors seem to be shared across cultural
groups whereas others are more relevant or unique to a cultural
group. However, the same risk and protective factors might have a
different impact or act differently in different cultures and between
genders in the same cultural group. For example, as reported by
Leach (2006), previous suicidal attempts is a stronger predictor of
suicide completion among European Americans than among
African Americans, and depression seems to be a stronger predictor
of suicide completion among African American females than
among African American males. Thus, when assessing the risk of
suicide, we should be aware that even the best-known risk factors
for suicide, such as previous suicide attempts and depression, might
show differences among cultural groups and genders and should

126
Cultural Issues in Suicide

not assume cultural invariability or universality for anything. This


also includes suicide warning signs.6
2. As there is the risk of mistakenly assuming universality, there is
always the risk, when considering suicidal behaviour in its cultural
milieu, to discount similarities in favour of dissimilarities whereas—
as shown by Marshall and Yazdani’s (1999) research on construals
of self-harm amongst Asian young women—it is important to re-
cognise the commonalities across ethno-cultural groups rather
that starting with the expectation of cultural differences. This
was criticised also by Mishara (2006: 3) who pointed out that
‘Suicidology research tends to either ignore cultural differences
entirely or focus upon a specific culture without examining possible
commonalities across cultures’. Therefore, the author further stated,
it is important to explore and understand the frontier between
universal aspects of suicide and its cultural specificity. Leach (2006:
IX) also commented that there is overlap among cultural groups, for
example ‘hopelessness leading to suicide manifests itself similarly
regardless of culture’ and this was also partially confirmed by the
literature review earlier summarised (Colucci and Martin, 2007a,
2007b). However, as also argued by Leach, the nuances of culture
need to be examined further with regard to many suicide factors,
such as risk and protective factors, cultural views on illness and
treatments and so on.
As it is important to consider the similarities between cultures in
suicide risk assessment the existence of within-group differences
cannot be overlooked.
3. The relationship between religiosity and suicide has been scien-
tifically investigated extensively since Durkheim (1897/1997)
formulated the concept of religious integration. Still, the topic
needs further and deeper examination, because data is inconsistent.
Nevertheless, the majority of the studies seem to suggest that reli-
giosity exercises a certain degree of protection against suicide. How-
ever, more research is needed that addresses the spiritual dimension

6
Suicide First Aid guidelines for Japan, India and Philippines, and the mental health
professionals involved in the Delphi study reported some cultural variations also on suicide
warning signs (Colucci et al., in press).

127
Erminia Colucci

of human beings, not only spirituality in its religious form—as it


has been the tendency until now—but also in non-religious forms,
as also observed by Swinton (2001).
Scholars (e.g., Greening and Stoppelbein, 2002; Leach, 2006;
Marion and Range, 2003) have suggested taking religious and
spiritual issues into account routinely when assessing people at
risk of suicide. I also invite scholars and clinicians to place a more
important role on spirituality in suicide risk assessment and sui-
cide prevention/intervention strategies, as discussed in Colucci
(2008c). In this regard, Sexson (2004: 45) commented, ‘Integrating
questions about spiritual and religious beliefs into the routine
history taking and assessment … demystifies the subject and
establishes its importance along with other aspects of the social
history.’ While doing this, clinicians must note that, although
less frequently found, there is some clinically important evidence
of harmful consequences of religiosity, thus religious beliefs may
expose individuals or groups of believers to a greater risk of suicidal
behaviour. Furthermore, holding religious beliefs might decrease
but not necessarily eliminate such risk.
4. Knowing risk and protective factors is not enough for suicide
risk assessment but it is also necessary to understand the cultural
meanings of suicide: mental health professionals must not take
for granted that the meanings, interpretations and mental repre-
sentations of suicidal behaviour remain the same in different
(sub)cultures.
5. Discussing cultural issues of suicide prevention strategies is beyond
the scope of this chapter (other scholars in this book have written
about this topic). Nonetheless, I would like to conclude this
chapter by arguing that suicide prevention strategies need to be
developed from within the cultural milieu, rather than merely be
adapted, that is making use of what has been done elsewhere. This
point has also been made by the present International Association
for Suicide Prevention (IASP) president (Mishara, 2006) while
addressing current challenges for the association. He wrote that,
to date, the emphasis in IASP has been upon commonalities in
suicide, with adaptations to specific cultures and settings, and that
the association’s activities—while allowing for cultural diversity—
have usually assumed a universal perspective. Some steps in this

128
Cultural Issues in Suicide

direction have been taken but the cultural perspective of suicide


intervention continues to be in an embryonic stage and more needs
to be done.

REFERENCES

Bhugra, D. and T.R. Osbourne (2004). Spirituality and Psychiatry. Indian Journal of
Psychiatry, 46(I), 5–6.
Bhugra, D., M. Desai and D.S. Baldwin (1999). Attempted suicide in west London,
I. Rates across ethnic communities. Psychological Medicine, 29(5), 1125–30.
Biswas, S. (1990). Ethnic differences in self poisoning: A comparative study between an
Asian and White adolescent group. Journal of Adolescence, 13(2), 189–93.
Boldt, M. (1988). The meaning of suicide: Implications for research. Crisis, 9(2),
93–108.
Borowsky, I.W., M. Ireland and M.D. Resnick (2001). Adolescent suicide attempts:
Risks and protectors. Pediatrics, 107(3), 485–93.
Breault, K.D. and K. Barkey (1982). A comparative analysis of Durkheim’s theory of
egoistic suicide. Sociological Quarterly, 23(3), 321–31.
Burkhardt, M.A. (1989). Spirituality: An analysis of the concept. Holistic Nursing Prac-
tice, 3(3), 69–77.
Canetto, S.S. and D. Lester (1998). Gender, culture and suicidal behavior. Transcultural
Psychiatry, 35(2), 163–90.
Cohen, Y., A. Spirito, A. Apter and S. Saini (1997). A cross-cultural comparison of
behavior disturbance and suicidal behavior among psychiatrically hospitalized
adolescents in Israel and the United States. Child Psychiatry & Human Development,
28(2), 89–102.
Colucci, E. (2006). The cultural facet of suicidal behaviour: Its importance and neglect.
Australian e-Journal for the Advancement of Mental Health, 5(3), 1–13. Retrieved
25 December 2006 from www.auseinet.com/journal/vol5iss3/colucci.pdf
Colucci, E. (2007). ‘Focus groups can be fun’: The use of activity-oriented questions
in focus group discussions, Qualitative Health Journal, 17(10), 1422–33.
Colucci, E. (2008a). The cultural meaning of suicide: A comparison between Italian,
Indian and Australian students. Unpublished doctoral dissertation, The University
of Queensland, Australia.
Colucci, E. (2008b). On the use of focus groups in cross-cultural research. In
P. Liamputtong (Ed.), Doing cross-cultural research: Ethical and methodological
considerations (pp. 233–52). Dordrecht, The Netherlands: Springer.
Colucci, E. (2008c). Recognizing spirituality in the assessment and prevention of sui-
cidal behaviour. World Cultural Psychiatry Research Review (WCPRR), Special issue:

129
Erminia Colucci

Suicide and Culture, 3(2), 77–95. Retrieved on 31 December 2008 from http://www.
wcprr.org/index-2003_2002.htm
Colucci, E. and G. Martin (2007a). The ethno-cultural aspects of youth suicide:
Rates and methods of youth suicide. Suicide & Life-Threatening Behavior, 37(2),
197–221.
Colucci, E. and G. Martin (2007b). The ethno-cultural aspects of youth suicide: Risk
factors, precipitating agents and attitudes towards suicide. Suicide & Life-
Threatening Behavior, 37(2), 222–37.
Colucci, E. and G. Martin (2008). Spirituality and religion along the suicidal path.
Suicide & Life-Threatening Behaviour, 38(2), 229–44.
Colucci, E.I., C. Kelly, H.I. Minas and A.F. Jorm (in press). Mental Health First
Aid guidelines for helping a suicidal person: A Delphi consensus study in India
International Journal of Mental Health Systems.
Crumbaugh, J.C. (1977). The Seeking of Noetic Goals test (SONG): A complementary
scale to the Purpose in Life test (PIL). Journal of Clinical Psychology, 33(3), 900–07.
De Leo, D. (2002). Struggling against suicide: The need for an integrative approach.
Crisis, 23(1), 23–31.
Dein, S. (2005). Spirituality, Psychiatry and participation: A cultural analysis. Trans-
cultural Psychiatry, 42(4), 526–44.
Dervic, K., M.A. Oquendo, M.F. Grunebaum, S. Ellis, A.K. Burke and J.J. Mann (2004).
Religious affiliation and suicide attempt. American Journal of Psychiatry, 161(12),
2303–08.
Domino, G., A. Su and S. Lee Johnson (2001–2002). Psychosocial correlates of suicide
ideation: A comparison of Chinese and U.S. rural women. Omega: Journal of Death
and Dying, 44(4), 371–89.
Douglas, J. D. (1967). The social meaning of suicide. New Jersey: Princeton University
Press.
Durkheim, E. (1997). Il suicidio, Studio Italiano di Suicidologia (R. Scramaglia, Trans.).
Milano: BUR. (Original work published in 1897.)
Dusevic, N., P. Baume and A.E. Malak (2002). Cross-cultural suicide prevention:
A framework. Sydney: Transcultural Mental Health Centre.
Ellison, C.W. (1983). Spiritual well-being: Conceptualization and measurement. Journal
of Psychology and Theology, 11(4), 330–40.
Eshun, S. (1999). Cultural variations in hopelessness, optimism, and suicidal ideation:
A study of Ghana and U.S. college samples. Cross-Cultural Research: The Journal
of Comparative Social Science, 33(3), 227–38.
Eshun, S. (2003). Sociocultural determinants of suicide ideation: A comparison between
American and Ghanaian college samples. Suicide & Life-Threatening Behavior,
33(2), 165–71.
Eskin, M. (1999). Gender and cultural differences in the 12-month prevalence of
suicidal thoughts and attempts in Swedish and Turkish adolescents. Journal of
Gender, Culture, and Health, 4(3), 187–200.

130
Cultural Issues in Suicide

Everall, R.D. (2000). The meaning of suicide attempts by young adults. Canadian
Journal of Counselling, 34(2), 111–25.
Exline, J.J., A.M. Yali and W.C. Sanderson (2000). Guilt, discord, and alienation: The
role of religious strain in depression and suicidality. Journal of Clinical Psychology,
56(12), 1481–96.
Farberow, N.L. (1975). Suicide in Different Cultures. Baltimore: University Park Press.
Faupel, C.E., G.S. Kowalski and P.D. Starr (1987). Sociology’s one law: Religion and
suicide in the urban context. Journal for the Scientific Study of Religion, 26(4),
523–34.
Fetzer Institute and the National Institute of Aging. (1999). Multidimensional
Measurement of Religiousness/Spirituality for Use in Health Research. Retrieved 29
July 2003 from http://www.fetzer.org/Resources/resources_multidimens.htm
Greening, L. and L. Stoppelbein (2002). Religiosity, attributional style, and social
support as psychosocial buffers for African American and white adolescents’
perceived risk for suicide. Suicide & Life-Threatening Behavior, 32(4), 404–17.
Grossoehme, D.H. and L.S. Springer (1999). Images of God used by self-injurious burn
patients. Burns, 25(5), 443–48.
Gutierrez, P.M., P.J. Rodriguez and P. Garcia (2001). Suicide risk factors for young
adults: Testing a model across ethnicities. Death Studies, 25(4), 319–40.
Handy, S., R.N. Chithiramohan, C.G. Ballard and W.R. Silveira (1991). Ethnic
differences in adolescent self-poisoning: A comparison of Asian and Caucasian
groups. Journal of Adolescence, 14(2), 157–62.
Heisel, M.J. and T. Fusé (1999). College student suicide ideation in Canada and Japan.
Psychologia: An International Journal of Psychology in the Orient, 42(3), 129–38.
Hill, P.C. and R.W. Hood (Eds). (1999). Measures of religiosity. Birmingham, Alabama:
Religious Education Press.
Hill, P.C. and K.I. Pargament (2003). Advances in the conceptualization and measure-
ment of religion and spirituality: Implications for physical and mental health
research. American Psychologist, 58(1), 64–74.
Johnson, C.V. and J.A. Hayes (2003). Troubled spirits: Prevalence and predictors of
religious and spiritual concerns among university students and counseling center
clients. Journal of Counseling Psychology, 50(4), 409–19.
Kamal, Z. and K.M. Loewenthal (2002). Suicide beliefs and behavior among young
Muslims and Hindus in the UK. Mental Health, Religion & Culture, 5, 111–18.
Kazarian, S.S. and E. Persad (2001). Cultural issues in suicidal behavior. In S.S. Kazarian
and D.R. Evans (Eds). Handbook of Cultrual Health Psychology (pp. 267–302). San
Diego, CA: Academic Press.
Kehoe, N.C. and T.G. Gutheil (1994). Neglect of religious issues in scale-based assessment
of suicidal patients. Hospital and Community Psychiatry, 45(4), 366–69.
Kingsbury, S. (1994). The psychological and social characteristics of Asian adolescent
overdose. Journal of Adolescence, 17(2), 131–35.

131
Erminia Colucci

Kirmayer, L.J. (1994). Suicide among Canadian Aboriginal peoples. Transcultural


Psychiatric Research Review, 31(1), 3–58.
Kral, M.J. (1998). Suicide and the internalization of culture: Three questions.
Transcultural Psychiatry, 35(2), 221–33.
Leach, M.M. (2006). Cultural diversity and suicide. Binghamton, NY: The Haworth Press.
Leenaars, A.A., J. Haines, S. Wenckstern, C. L. Williams and D. Lester (2003). Suicide
notes from Australia and the United States. Perceptual and Motor Skills, 96 (3, pt 2),
1281–82.
Leenaars, A.A., R. Maris and Y. Takahashi (Eds). (1997). Suicide: Individual, cultural,
international perspectives. New York: Guilford Press.
Lester, D. (1992–1993). Suicide and culture. Homeostasis, 34(1–2), 96–102.
Lester, D. (1994). Suicide in immigrant groups as a function of their proportion in the
country. Perceptual & Motor Skills, 79(2), 994.
Lester, D. (1997). Suicide in America: A nation of immigrants. Suicide & Life-
Threatening Behavior, 27(1), 50–59.
Lester, D. and L.J. Francis (1993). Is religiosity related to suicidal ideation after per-
sonality and mood are taken into account? Personality & Individual Differences,
15(5), 591–92.
Loewenthal, K.M., A.K., Macleod, S. Cook, M. Lee, and V. Goldblatt (2003). The suicide
beliefs of Jews and Protestants in the UK: How do they differ? Israel Journal of
Psychiatry & Related Sciences, 40(3), 174–181.
Mancinelli, I., A. Comparelli, P. Girardi and R. Tatarelli (2002). Mass suicide: Historical
and psychodynamic considerations. Suicide & Life-Threatening Behavior, 32(1),
91–100.
Marecek, J. (1998). Culture, gender, and suicidal behavior in Sri Lanka. Suicide & Life-
Threatening Behavior, 28(1), 69–81.
Marion, M.S. and L.M. Range (2003). Do extenuating circumstances influence African
American women’s attitudes toward suicide? Suicide & Life-Threatening Behavior,
33(1), 44–51.
Maris, R.W. and B. Lazerwitz (1981). Pathways to suicide: A survey of self-destructive
behaviors. Baltimore, MD: Johns Hopkins University Press.
Marsella, A.J. (2000). Socio-cultural considerations in the assessment of suicidal
behavior: In pursuit of cultural and community indices, dimensions, and perceptions.
(Unpublished technical report.) Geneva, Switzerland: WHO/ Mental Health.
Marsella, A.J. and A. Kaplan (2002). Cultural considerations for understanding,
assessing, and treating depressive experience and disorder. In M.A. Reinecke and
M.R. Davison (Eds), Comparative treatments of depression (pp. 47–78). New York:
Springer.
Marshall, H. and A. Yazdani (1999). Locating culture in accounting for self-harm
amongst Asian young women. Journal of Community & Applied Social Psychology,
9(6), 413–33.
Mayer, P. and T. Ziaian (2002). Suicide, gender, and age variations in India. Are women
in Indian society protected from suicide? Crisis, 23(3), 98–103.

132
Cultural Issues in Suicide

McIntosh, J.L. (1989). Trends in racial differences in U.S. suicide statistics. Death
Studies, 13(3), 275–86.
Meng, L. (2002). Rebellion and revenge: The meaning of suicide of women in rural
China. International Journal of Social Welfare, 11, 300–09.
Mishara, B.L. (2006). Cultural specificity and universality of suicide. Challenges for the
International Association for Suicide Prevention. Crisis, 27(1), 1–3.
Oldnall, A. (1996). A critical analysis of nursing: Meeting the spiritual needs of patients.
Journal of Advanced Nursing, 23(1), 138–44.
Perkins, D.F. and G. Hartless (2002). An ecological risk-factor examination of suicide
ideation and behavior of adolescents. Journal of Adolescent Research, 17(1),
3–26.
Raleigh, V.S. (1996). Suicide patterns and trends in people of Indian subcontinent and
Carribean origin in England and Wales. Ethnicity & Health, 1(1), 55–63.
Range, L.M. and M.M. Leach (1998). Gender, culture, and suicidal behavior: A feminist
critique of theories and research. Suicide & Life-Threatening Behavior, 28(1),
24–36.
Range, L.M., M.M. Leach, D. McIntyre, P.B. Posey-Deters, M.S. Marion, S. H. Kovac et
al. (1999). Multicultural perspectives on suicide. Aggression and Violent Behavior,
4(4), 413–30.
Resnick, M.D., P.S. Bearman, R.W. Blum, K.E. Bauman, K.M. Harris, J. Jones et
al. (1997). Protecting adolescents from harm: Findings from the National
Longitudinal Study on Adolescent Health. Journal of the American Medical
Association, 278(10), 823–32.
Rew, L., N. Thomas, S.D. Horner, M.D. Resnick and T. Beuhring (2001). Correlates
of recent suicide attempts in a triethnic group of adolescents. Journal of Nursing
Scholarship, 33(4), 361–67.
Roberts, R.E., R. Chen and C.R. Roberts (1997). Ethnocultural differences in prevalence
of adolescent suicidal behaviors. Suicide and Life-Threatening Behavior, 27(2),
208–17.
Rumbold, B.D. (2007). A review of spiritual assessment in health care practice. The
Medical Journal of Australia, 186(10), 60–62.
Ryff, C.D. and C.L. Keyes (1995). The structure of psychological well-being revisited.
Journal of Personality and Social Psychology, 69(4), 719–27.
Sexson, S.B. (2004). Religious and spiritual assessment of the child and adolescent.
Child and Adolescent Psychiatric Clinics, 13(1), 35–47.
Shiang, J. (2000). Considering cultural beliefs and behaviors in the study of suicide.
In R.W. Maris and S.S. Canetto (Eds), Review of Suicidology (pp. 226–41). NY,
US: Guilford Press.
Sorenson, S.B. and H. Shen (1996). Youth suicide trends in California: An examination
of immigrant and ethnic group risk. Suicide and Life-Threatening Behavior, 26(2),
143–54.
Sorri, H., M. Henriksson and J. Lonnqvist (1996). Religiosity and suicide: Findings
from a nationwide psychological autopsy study. Crisis, 17(3), 123–27.

133
Erminia Colucci

Swinton, J. (2001). Spirituality and mental health care: Rediscovering a ‘forgotten’ dimen-
sion. London: J. Kingsley Publishers.
Tarakeshwar, N., J. Stanton and K.I. Pargament (2003). Religion: An overlooked
dimension in cross-cultural psychology. Journal of Cross Cultural Psychology, 34(4),
377–94.
Thatcher, W.G., B.M. Reininger and J.W. Drane (2002). Using path analysis to examine
adolescent suicide attempts, life satisfaction, and health risk behavior. Journal of
School Health, 72(2), 71–77.
Tortolero, S.R. and R.E. Roberts (2001). Differences in nonfatal suicide behaviors
among Mexican and European American middle school children. Suicide & Life-
Threatening Behavior, 31(2), 214–23.
Trovato, F. (1986). Suicide and ethnic factors in Canada. International Journal of Social
Psychiatry, 32(3), 55–64.
Tseng, W.S., J. Hsu, A. Omori and D.G. McLaughlin (1982). Suicidal behaviour in
Hawaii. In K.L. Peng and W.S. Tseng (Eds), Suicidal behaviour in the Asia-Pacific
region (pp. 231–48). Singapore: Singapore University Press.
Tseng, W.S. (2001). Handbook of cultural psychiatry. San Diego, CA: Academic Press.
Underwood, L.G. and J. A. Teresi (2002). The Daily Spiritual Experience Scale: Devel-
opment, theoretical description, reliability, exploratory factor analysis, and
preliminary construct validity using health-related data. Annals of Behavioral
Medicine, 24(1), 22–33.
Vega, W.A., A. Gil, G. Warheit, E. Apospori and R. Zimmerman (1993). The relationship
of drug use to suicide ideation and attempts among African American, Hispanic,
and White non-Hispanic male adolescents. Suicide and Life-Threatening Behavior,
23(2), 110–19.
Vega, W.A., A.G. Gil, R.S. Zimmerman and G.J. Warheit (1993). Risk factors for suicidal
behavior among Hispanic, African-American, and non-Hispanic White boys in
early adolescence. Ethnicity and Disease, 3(3), 229–41.
Vijayakumar, L. (2005). Suicide and mental disorders in Asia. International Review of
Psychiatry, 17(2), 109–14.
Vijayakumar, L., S. John, J. Pirkis and H. Whiteford (2005). Suicide in developing
countries (2): Risk factors. Crisis, 26(3), 112–19.
Wasserman, L. and S. Stack (1993). The effect of religion on suicide: An analysis of cultural
context. Omega: Journal of Death & Dying, 27(4), 295–305.
Watt, T.T. and S.F. Sharp (2002). Race differences in strains associated with suicidal be-
havior among adolescents. Youth and Society, 34(2), 232–56.
Webb, D. (2003). Self, soul and spirit-Suicidology’s blind spots? New Paradigm.
Retrieved 29 October 2003 from http://www.vicserv.org.au/publications/
new_para/pdf/webbd2.pdf
WHOQOL SPRB Group (2002). World Health Organization Quality of Life, Spirituality,
Religiousness and Personal Beliefs (WHOQOL SRPB) Field-Test Instruments.
Retrieved 29 April 2007 from http://www.who.int/msa/qol

134
Cultural Issues in Suicide

Yamada, A.M., A.J. Marsella and H.R. Atuel (2002). Development of a Cultural
Iden-tification Battery for Asian and Pacific Islander Americans in Hawaii. Asian
Psychologist, 3(1), 11–20.
Yuen, N.Y., L.B., Nahulu, E.S. Hishinuma and R.H. Miyamoto (2000). Cultural iden-
tification and attempted suicide in Native Hawaiian adolescents. Journal of the
American Academy of Child and Adolescent Psychiatry, 39(3), 360–67.
Zhang, J. and S. Jin (1996). Determinants of suicide ideation: A comparison of Chinese
and American college students. Adolescence, 31(122), 451–67.
Zonda, T. and D. Lester (1990). Suicide among Hungarian Gypsies. Acta Psychiatrica
Scandinavica, 82(5), 381–82.

135
7

Gender Issues in Suicide Risk


Factor Assessment
P›T›Ù OÝòƒT«, V®»TÊÙ VÊÙÊÝ ƒÄ— SƒÄ—ÊÙ F›»›T›

I n the developed Western countries the number of men committing


suicide is three times more than that of women. However, women are
more likely to attempt suicide than men (Kaplan and Sadock, 2003). This
is not a recent phenomenon, as in the nineteenth century Durkheim (1897)
pointed out a similar gap in suicide mortality among men and women; but
the difference became more apparent in the last decades of the twentieth
century. There are many ways to explain why completed suicides are more
prevalent in men; however, they do not explain the relatively low female
suicide rates all over the world. Two approaches arise about gender dif-
ferences in suicide: ‘why suicide rates are so high among males?’ or ‘why
they are significantly lower in females?’ The answer to these questions lies
within the multi-dimensional quality of suicidal behaviour, namely that
completed and attempted suicides are not different phenomena, and many
times often the reasons, the mechanisms and the methods are the same in
both cases, only the outcome differs.
Since gender is one of the most frequently replicated predictors for
suicide, gender differences in suicidal behaviour have been analysed in a
number of recent studies (Canetto and Sakinofsky, 1998; Hawton, 2000;
Moscicki, 1994). Identifying variables which indicate greater risk of suicidal

136
Gender Issues in Suicide

behaviour and investigating whether risk factors associated with suicide


differ by gender are among the principal tasks of the epidemiological
and socio-cultural research. According to these studies numerous socio-
economic, demographic, psychiatric, familial and help-seeking differences
can be recognised in the various protective and risk factors between males
and females (Mortensen et al., 2000; Qin et al., 2000). The explanations for
the higher suicidal mortality of males remain insufficient, which may be
caused by the complexity of factors involved in suicidal behaviour.
The main hypotheses—aimed to explain the higher suicide mortality
in men—are the following: the higher lethality of male suicide methods;
the reluctance of men to seek help, the higher rate of substance (especially
alcohol) abuse; and some socio-cultural differences. There has been a sig-
nificant rise in the number of suicide rates among young males and a
decline in rates among—especially elderly—females in numerous Western
countries in recent years (Hawton, 1998).

DIFFERENCES IN EPIDEMIOLOGY

In most countries, more men die because of suicide than women (male/
female ratio: 2–4/1); however, China is one of the few but most important
countries of the exceptions (Cheng and Lee, 2000). According to the ‘gender
paradox’ (Canetto and Sakinofsky, 1998), suicide attempts are more com-
mon among females compared to males (female/male ratio: 4–6/1), which
lead the researchers to believe that male gender can be classified as a special,
tertiary risk factor for suicide in the hierarchical classification of suicide
risk factors characterised by Rihmer (Rihmer et al., 2002). Paradoxically,
the higher number of suicide attempts tend to lower the risk of fatal
outcome among women.
In the past couple of years most countries have experienced a signifi-
cant decrease in their overall suicide rates (mainly in older females), despite
slight increase in the rates of younger age groups, particularly among
males. Social factors—mostly linked to changes in gender roles—and the
fact that men respond more strongly to the changes in social and economic
conditions, seem to be the two most likely explanations for this phenom-
enon (Hawton, 1998).

137
Peter Osvath et al.

DIFFERENCES IN SOCIO ECONOMIC RISK FACTORS

Many studies indicated that socio-economic risk factors for suicide also
differ among males and females (Canetto and Sakinofsky, 1998). Con-
cerning socio-economic factors, unemployment, retirement, single lifestyle
and absence from work due to sickness are the most significant risk factors
for suicide, mainly among younger males (Qin et al., 2000). Economic
stressors, such as employment status, income and wealth, work as more
significant triggers for suicide among males than in females. This fact sup-
ports the hypothesis that men respond more strongly to changes in social
and economic conditions than women (Varnik et al., 1998).
Occupational factors are also particularly important for males. Increased
occupational instability has been proposed as one factor behind the recent
increase in young male suicides (Hawton, 1998). Absence from work due
to illness was significantly associated with a higher suicide risk, but only for
males. This indicates that physical weakness might more easily lead to lack
of self-esteem and confidence (Qin et al., 2000). Given the occupational
content of male gender role stereotypes, it seems likely that unemploy-
ment and uncertainties at work would have a stronger impact on the male
population’s self-esteem and mental stability, while women have more
possibilities to retain other status and domestic and caring responsibilities
(Payne et al., 2008). Male status is more often dependent on success at the
workplace and control over their work and financial background, so they
may be more sensitive to deprivation and more vulnerable to the basic
gender-role distress. We can observe quite a similar situation in the case of
European (especially males) adolescents’ suicides, who are more vulnerable
to social changes (e.g., unemployment), so suicide may be a response
to their problems with work, which in many countries is considered a
‘masculine’ response and behaviour (Mittendorfer-Rutz, 2006).
From a socio-biological viewpoint there are some age-dependent gender
differences, which might be in connection with the diminished capacity to
reproduce and to get social support. In many countries (e.g., United States,
Hungary) suicide rates in women tend to peak around middle age (the
years of the menopause and the ‘middle life crisis’), while male suicide rates
are much more higher among the elderly (Fekete et al., 2005). In old age,
men become less fit physically and the reproductive capacity diminishes
with isolation and deteriorated social support (Maris, 2002). The region of
living also has a special gender-specific effect on suicide risk. Urban living

138
Gender Issues in Suicide

is linked to a higher rate of female suicide, mostly because women living


in cities feel more alienated and socially isolated (Murphy, 1998).

DIFFERENCES IN CHOICE OF METHODS

Concerning the differences between various methods, males tend to use


more violent methods both in completed and attempted suicide, while
women are more likely to choose self-poisoning (Kaplan and Sadock,
2003). A recent European multicentre study on youth suicide showed
that males have a significantly higher risk for using firearms, hanging
and poisoning by other means in their suicide attempts and lower risk of
poisoning by drugs and jumping (Varnik et al., 2008). The difference
between the rates of suicide attempts and completed suicides among
women may reflect a lesser degree of intent, but it also shows a tendency
among women not to use highly violent or particularly lethal methods.
There is, however, an important exception, because in China—where self-
poisoning with pesticides is a common lethal method (Zhang et al.,
2008)—the rate of suicide is equal in males and females (or higher in the
rural female population).
Greater suicidal intent, aggression, knowledge regarding violent means,
less concern about bodily disfigurement are all likely explanations for the
excess of violent suicide in males (Hawton, 2000). Strong cultural beliefs
of suicide considered as ‘masculine’ and surviving a suicide as being
culturally unacceptable might influence young males to use more violent
or lethal methods (Mittendorfer-Rutz, 2006). This hypothesis is supported
by the fact that men are more likely to attempt suicide through violent
methods. For men surviving the suicidal act is perceived as inappropriate,
and—from the viewpoint of traditional masculinity—death by suicide
among men is viewed as ‘less wrong’ than in women (Canetto, 1997). Lethal
suicide among men may be seen as an act of masculine expression or as an
attempt to escape the negative consequence of surviving a suicide attempt.
According to another explanation, the gender difference in choosing the
method also relates to the communicative aspect of suicidal behaviour.
By using a less violent method women might seek to protect others. So
they tend to choose methods without regarding their attractiveness (Payne
et al., 2008).

139
Peter Osvath et al.

DIFFERENCES IN MENTAL DISORDERS

Mental illnesses are the most replicated predictors for suicides among
both genders but especially for women (Rihmer et al., 2002). Psycho-
logical autopsy studies clearly demonstrate that affective disorders carry
the highest risk—both to males and females—often comorbidly with
personality-disorders and with other mental disorders (Hawton, 2000).
Rates of schizophrenia and addictions are higher among males, while eating
disorders—especially anorexia nervosa—are more common among female
suicide victims (Harris and Barraclough, 1997; Mortensen et al., 2000).
In a longitudinal study, hospitalised mental illness (particularly recent
discharge from a psychiatric hospital) appeared to be the most prominent
risk factor for suicide with both genders (Qin et al., 2000).
These differences may be viewed as artifacts of men’s lower likelihood
to seek help or because the symptoms of male depression are different
from women’s. If the symptoms of a mental disorder are perceived as in-
consistent with masculinity, men try to hide such symptoms (as signs of
weakness) and do not ask for treatment (Payne et al., 2008). Men in line
with norm-congruent behaviour drink more and more alcohol to combat
depression instead of seeking professional help. Furthermore, alcohol and
substance abuse, in its own right, has positive associations with suicide,
especially among women (Payne et al., 2008).

DIFFERENCES IN PROTECTIVE FACTORS

In spite of the fact that little research attention has been paid to factors
which protect against suicide, there are some differences even among the
protective factors. According to a Danish study, parenthood appears to
explain an apparent protective effect of marriage for women, rather than
the marriage itself per se, whereas among men marriage appears to be a
protective factor in its own right (and single status is a risk factor) (Qin
et al., 2000). According to another study, pregnancy has been found to be
a protective factor for women (Appleby, 1996).
Single men have higher risk for suicide than single women, and divorce
is a significant risk factor for men, but not with women (Louma and Pearson,
2002; Qin et al., 2003). Interestingly, marital status seems to be more of a

140
Gender Issues in Suicide

protective factor for men rather than for women which can be explained
with the socially constructed gender role. Marriage offers emotional and
social integration, which are particularly important for men because they
have fewer alternatives to having closer human relationships, and the gender
role stereotypes (particularly of the need to be independent) diminish the
capacity of males to develop social networks. Divorce or the death of a
spouse can cause significantly more stress for males and might lead them
to suicidal behaviour, while women are more likely to have extended and
rooted social networks which might help them to cope with interpersonal
losses (Payne et al., 2008).
For women, the social construction of femininity includes family roles
and the caring for children which offer them benefits to fulfil the socio-
cultural stereotypes based on traditional gender roles (Payne et al., 2008).
There is another protective factor for women, namely to hold religious
beliefs and negative attitudes towards suicide (Steen and Mayer, 2004).

DIFFERENCES IN HELP SEEKING BEHAVIOUR

While being a male is an important risk factor for suicide, the presenta-
tion of suicidal behaviour is generally more common among women.
Females are also more likely to seek help from general practitioners for their
mental health problems (Osvath, Michel and Fekete, 2003). There are also
special gender differences in various cultures; men often view help-seeking
as a sign of weakness (Murphy, 1998). Men rarely ask for professional
help and they are also more reluctant to ask for support from family and
friends (Biddle et al., 2004). This reluctance and the special features of
male depression may contribute to the fact that depression is more often
undetected and untreated among men (Rihmer et al., 2002). This may—at
least partially—explain the striking paradox: major depression (which has
the strongest association with suicide among mental disorders) is about
twice as common among women than men, but men are four times more
likely to commit suicide than women.
In the Gotland study the decrease in depressive suicide has almost
entirely been the result of a decrease in female depressive suicides, while
male suicidal rate has not changed. This probably explains why the apparent
benefits of the educational programme in detection and treatment of

141
Peter Osvath et al.

depression for general practitioners on the Swedish island of Gotland were


confined to females who were treated for depression (Rutz, 2001).
Several studies and summarised clinical impressions suggest that the
compliance for the therapy and prevention of male patients is significantly
poorer than for females, and there is also some indication from treatment
studies that fewer male attempters actually benefit from the treatments
offered to them (Hawton, 1998, 2000). While this may reflect differences
in the overall attitude towards help, it could also result from the available
therapy type. Gender differences in verbal abilities and the reluctance of
many males to share emotional problems may make some of the usual talk-
ing therapies less attractive to some males, at least initially (Hawton, 2000).
These data suggest that gender-specific suicide prevention could be fur-
ther improved by diagnosing and treating the male patients’ mental
problems in a more effective manner, since they often mask their problems
with impulsive, aggressive behaviour or alcohol abuse and their behaviour
is frequently characterised by a lack of help-seeking attitude and non-
compliance (Rihmer et al., 2002).

SOCIO CULTURAL ASPECTS OF GENDER DIFFERENCES


IN SUICIDAL BEHAVIOUR

Despite the mentioned main differences in suicidal behaviour, gender


is mostly treated as a descriptive and causal factor in suicidal behaviour.
But gender is not one of an array of individual, social and demographic
characteristics (e.g., household composition, employment status, edu-
cation). It is rather an inter-dependent variable that connects with, and
impacts on, other influences and it also has special effects on suicidal
behaviour (Payne et al., 2008). These effects could be explained with a com-
plex socio-cultural theory about socially constructed masculinities and
femininities, which may impact on a suicide-related behaviour and help
explain the gender difference. We emphasised the fact that suicide is a
result of a very complex interaction of a number of precipitating factors
and that the socio-cultural theories focus on the social determinants of
suicidal behaviour.

142
Gender Issues in Suicide

In the complex socio-cultural model the traditional gender roles have


a close relationship with suicide-related behaviour and the socio-cultural
factors impact and reinforce these. Male gender role (characterised by
dominance, aggressiveness, invulnerability) can help to explain the reason
why men tend to choose more lethal methods, why they are struggling
with asking for help and support for psychological problems and mental
disorders, and why they tend to misuse alcohol and other substances as
inappropriate self-medication. In contrast, traditional female role typically
includes fragility, emotionality, expressiveness and family orientation,
which may explain women’s help-seeking behaviour and their tendency
to use less lethal methods. Additionally women have more opportun-
ities to cope with negative life-events in a more effective manner (Payne
et al., 2008).
Houle, Mishara and Chagnon (2008) found some empirical evidence
of a mediation model, in which the masculine gender role increased the
suicide risk indirectly while being exercised through various mediator
variables. In this model the most important variables were of negative
influence on the mental state, social support and help-seeking. According
to the model, men with traditional roles often experience a greater feeling
of shame because of their problems which can diminish their self-esteem
and undermine their mental stability. The masculine gender role deprives
men from important sources of social support (due to the higher demands
of autonomy). Some forms of emotional expression are less valued or
denied (especially emotionality and weakness), while others, such as anger
and aggression are accepted. The traditional masculine gender role may
result in increasing the risk of suicide indirectly via mediator factors, since
it tends to undermine the stability of mental state and inhibit the protective
effect of help-seeking and social support.
An interesting socio-cultural explanation to gender differences had
been found in China. Researchers found a significantly higher rate of
suicide among young females living in the rural areas. The findings are
mostly due to the fact that women hold quite a low status in the Chinese
society which can cause their deeply frustrating constraints. For young
Chinese women the suicide act is therefore a form of protest and an escape
from social distress, maltreatment and their lack of freedom (Mitra and
Shroff, 2008).

143
Peter Osvath et al.

GENDER DIFFERENCES IN SUICIDE ATTEMPTERS

The excess rate of suicide attempts in females, and the stronger associ-
ation between attempted and completed suicide in males (Hawton, 1998)
refer to the fact that attempts by females are more often based on non-
suicidal, but communicative motivation, while in males the attempts are
often associated with greater suicidal intent. In a cross-national survey,
the risk of suicidal behaviour (suicidal ideation, plan and attempt) was
found significantly related to women compared to men (OR: 1,4 for suicide
ideation, 1,4 for suicide plan and 1,7 for attempt) (Nock et al., 2008).
Since gender is one of the most significant risk factors for suicide, and
in a great number of completed suicides there are preceding suicide at-
tempts in the medical history of suicide victims, it is important to study
suicide attempters in relation to gender differences. Additionally, nu-
merous reliable data are available on the characteristics of suicide victims,
but the proportion of studies on attempted suicide is relatively low.
Therefore, a collaborative study was conducted to detect differences in
the suicidal behaviour between males and females by examining a large
sample of medically treated suicide attempters in Hungary, a country with
one of the highest suicide rates in the world. This analysis was performed
within the frameworks of the WHO/Euro Multicentre Study on Suicidal
Behavior (Osvath, Kelemen, Erdõs, Vörös and Fekete, 2003; Schmidtke
et al., 1996). The aim of the European collaborative study was to identify
epidemiological, socio-demographic features of suicide attempters, to find
protective and risk factors of suicide behaviour. Registration of attempted
suicides was carried out on consecutive episodes at university clinics in
Pecs. In the data collection a standardised monitoring form developed for
this multicentre study was used (Schmidtke et al., 1996). Out of 1158
medically treated suicide attempters, 63 percent (n = 728) were females.
The highest rates of suicide attempts belonged to the adolescent and
middle-aged population in both genders. More than half of the attempts
were repeated suicide attempts, both in males (53.3 percent) and in females
(52.1 percent). The statistical analysis (logistic regression model) separated
male and female attempters quite sharply (Table 7.1).
A ‘typical’ female suicide attempter can be characterised as follows:
retirement or economical inactivity, widowhood, divorce and depression
in personal history. Female attempters were mainly repeaters, using the
method of self-poisoning, mostly with benzodiazepines. Among males

144
Gender Issues in Suicide

Table 7.1 Gender Differences of Suicide A empters


(Mul variate Logis c Regression Model)

Male Female
Employment status Economical inactivity Economical
(unemployment) inactivity (retirement)
Marital status and Living alone, never married Divorced or widowed
household composition
Mental disorder Alcohol abuse Depression
Method of attempt Violent (cutting, jumping) Self-poisoning
Medication type in self- Meprobamate, carbamazepine Benzodiazepines
poisoning
Other — Repeated attempts
Source: Fekete, Voros and Osvath (2005).

‘unemployment, living alone, never been married, addictive problems and


the use of violent methods’ were the main characteristics. In case of self-
poisoning, males are more likely to take meprobamate or carbamazepine
(Fekete et al., 2005).
A couple of years prior to this study, Hungary had the highest suicide
rate in the world. Nevertheless, it is peculiar that in spite of the more than
30 percent decrease (from 45/100,000 to under 27/100,000) of suicides
in the last two decades, the number of suicide attempts of both genders
stayed fairly stable.
While comparing rates of suicide (Figure 7.1) and attempted suicide
(Figure 7.2)—according to gender and age groups—it is well recognised
that suicides show almost a linear increase in both genders. The number of
attempts does not a show a continuous decline—as indicated from some
previous data—but it has three different peaks parallel in both sexes.
The first peak is the young-aged group (around the twenties), at the
time of adolescent crisis; the second is the middle-aged group (around the
forties), in the years of the mid-life crisis; and the third peak is the old-
aged group (around the eighties), at the time of life-end period. The asso-
ciation of significant life periods of the psychosocial developmental crisis
(Erikson, 1950) and the age curve of suicide attempts emphasise the im-
portance of psycho-socio-cultural origin of suicidal behaviour.
Suicide attempts are very high in the groups of young males, and
young and middle-aged females, thus indicating a more important com-
municative aspect (asking for protection from others in the crisis situation)
of these acts. Another possibility is the inadequate knowledge of methods,

145
Peter Osvath et al.

Figure 7.1 Suicide Rates by Age Groups in Hungary (per 100,000 inhabitants)

Source: Hungarian Central Statistical Office (2002).

Figure 7.2 Suicide A empts by Age Groups (per 100,000 inhabitants)

Source: Data from Pecs Center of WHO/Euro Multicenter Study on Suicidal Behaviour
2001 (Fekete et al. 2005).

146
Gender Issues in Suicide

or it also might be the case that these people are physically more resistant.
There were no significant differences in mental illnesses between males
and females, except for the rates of affective disorders and addiction.
The former was almost twice as much in females than in males, while
the latter was much higher among males, which may support the theory
(similar to the suicide victims) that alcohol abuse among men might be a
symptom of an affective disorder, or rather alcohol abuse in men marks
latent or masked depression and it might be an inadequate method of
self-medication. From another perspective, it is also possible to suggest
that females ‘resort to’ depression instead of turning to alcohol.
Significant differences were found concerning methods of self-
poisoning according to age, gender and repeated attempters. Regarding
self-intoxication, benzodiazepines were the most chosen drugs; besides,
men tended to use meprobamate and carbamazepine as well, which in
fact might be linked to the high prevalence of alcohol-related disorders in
Hungary (Osvath, Kelemen, Erdõs, Voros and Fekete, 2003). Among re-
peated attempters, a higher rate of taking an antidepressant, carbamazepine
or antipsychotic medicine was found compared to benzodiazepines, which
might indicate a higher prevalence of mental illnesses, or a disappointment
of the therapy among repeaters. In the group of first attempters using
a benzodiazepine rather than an antidepressant might indicate that
many attempters only get symptomatic therapy for anxiety and sleeping
disturbances, and the depressive disorders remain concealed (Osvath,
Michel and Fekete, 2003). This phenomenon is mainly characteristic to
men, whose help-seeking behaviour and compliance for therapy is poorer.
The importance of the study lies in the fact that—by studying a great
sample group—significant gender differences were found in suicide at-
tempters, in line of former results on suicide victims. The results supported
the significance of socio-cultural factors in association with the gender roles
and suicide attempts, considering age, marital status, choice of method
and mental disorders.

GENDER SPECIFIC TREATMENT POSSIBILITIES:


CLINICAL IMPLICATIONS

Since suicide is a phenomenon with multiple causes, its therapy and


prevention should be complex and gender differences should be taken

147
Peter Osvath et al.

into consideration while building up helping strategies. Investigations of


genetic and biological factors and epidemiological studies related to risk
of suicidal behaviour are in their infancy but should be conducted from
a gender perspective. Improved detection and management of psychi-
atric disorders are undoubtedly key factors in the prevention of suicidal
behaviour among men as well as among women (Rihmer et al., 2002).
Therefore, the awareness of gender differences in the clinical manifestation
of depression is important in successful suicide prevention in depressed
patients. We emphasise the significance of preventive efforts such as
adequate diagnosing of mental disorders (especially affective disorders,
alcohol abuse/dependency and personality disorders) and the effective
treatment of these disturbances (rather antidepressants, instead of sedato-
hypnotics), both of which are crucial factors—particularly among men—in
the prevention of suicide attemps in both genders (Osvath, Michel and
Fekete, 2003).
There is increasing evidence that alterations in socio-economic and
socio-cultural conditions are also relevant to suicide prevention in males.
Suicide prevention strategies reinforced that clinicians and other health
care professionals should have adequate risk factor assessment skills.
Although the predictive power of these risk factors is not clear (Goldney,
2000), some studies suggest that separate risk-assessment schedules are
required for the two genders, meaning that based on gender differences
initiatives to develop gender-specific approaches may be indicated. It is
important to train health care professionals for a better understanding of
the psychological characteristics of males, based on traditional masculine
role, to improve their competencies for adequate and effective intervention
in the mental problems of males (Houle et al., 2008). Investigating the
socio-economic status, finding the risk and protective factors according
to gender, and a marked concern for the high-risk groups, for example
repeaters, elderly, chronically (physically or mentally) ill, could also help
to prevent suicidal acts in everyday practice (Rihmer et al., 2002).
Based on socio-cultural models it would be important from a primary
prevention perspective to encourage prevention programmes for young
boys focusing on the development of effective coping mechanism and on
learning to express their emotions and to seek help without a feeling of
shame. By doing this they can get a chance to solve difficult or stressful situ-
ations more successfully, without feeling shame or experiencing a decrease
in self-esteem and efficiency (Houle et al., 2008).

148
Gender Issues in Suicide

Little research attention has been paid to possible gender differences in


response to treatment in people at risk of suicidal acts. Studies on gender
differences suggest that treatment programmes—which have more of a
practical emphasis on problem-solving—can prove more successful in
engaging males-at-risk. Gender differences in suicidal behaviour clearly
merit more research attention to generate information that can guide
clinical practice and prevention strategies in ways that will prove most
effective for preventing suicidal behaviour in both genders.

REFERENCES

Appleby, L. (1996). Suicidal behaviour in childbearing women. International Review


of Psychiatry, 8(1), 107–15.
Biddle, L., D. Gunnel, D. Sharp and J.L. Donovan (2004). Factors influencing help
seeking in mentally distressed young adults: A cross sectional survey. British Journal
General Practice, 54(501), 248–53.
Canetto, S.S. (1997). Meanings of gender and suicidal behavior during adolescence.
Suicide and Life-Threatening Behavior, 27(4), 339–51.
Canetto, S.S. and L. Sakinofsky (1998). The gender paradox in suicide. Suicide and
Life-Threatening Behavior, 28(1), 1–23.
Cheng, A.T.A. and C.S. Lee (2000). Suicide in Asia and the Far East. In K. Hawton and
K. Van Heeringen (Eds), The International Handbook of Suicide and Attempted
Suicide (pp. 121–35). Chicester: John Wiley and Sons.
Durkheim, E. (1897). Le Suicide. Etude de Sociologie. Paris: Alcan.
Erikson, E. (1950). Childhood and society. New York: WW Norton.
Fekete, S., V. Voros and P. Osvath (2005). Gender differences in suicide attempters in
Hungary: Retrospective epidemiological study. Croatian Medical Journal, 46(2),
288–93.
Goldney, R.D. (2000). The privilege and responsibility of suicide prevention. Crisis,
21(1), 8–15.
Harris, E.C. and B. Barraclough (1997). Suicide as an outcome for mental disorders.
A meta-analysis. British Journal of Psychiatry, 170 (3), 205–28.
Hawton, K. (1998). Why has suicide increased in young males? Crisis, 19(3), 119–24.
Hawton, K. (2000). Sex and suicide: Gender differences in suicidal behaviour. British
Journal of Psychiatry, 177 (6), 484–85.
Houle, J., B.L. Mishara and F. Chagnon (2008). An empirical test of a mediation model
of the impact of the traditional male gender role on suicidal behavior in men.
Journal of Affective Disorders, 107(1–3), 37–43.
Hungarian Central Statistical Office (2002). Hungarian Statistical Bulletin, 2001.
Hungarian Central Statistical Office, Budapest, Hungary.

149
Peter Osvath et al.

Kaplan, H.I. and B.J. Sadock (2003). Kaplan & Sadock’s synopsis of psychiatry: Be-
havioural sciences/clinical psychiatry. Philadelphia, USA: Lippincott Williams &
Wilkins.
Louma, J.B. and J.L. Pearson (2002). Suicide and marital status in the United States,
1991–1996: Is widowhood a risk-factor? American Journal of Public Health, 92(9),
1518–22.
Maris, R.W. (2002). Suicide. Lancet, 360 (9329), 319–26.
Mitra, S. and S. Shroff (2008). What suicides reveal about gender bias. Journal of Socio-
Economics, 37(5), 1713–23.
Mittendorfer-Rutz, E. (2006). Trends of youth suicide in Europe during the 1980s and
1990s – Gender differences and implications for prevention. Journal of Mental
Health, 3(3), 250–57.
Mortensen, P.B., E. Agerbo, T. Erikson, P. Qin and N. Westergaard-Nielsen (2000).
Psychiatric illness and risk factors for suicide in Denmark. Lancet, 355(9197),
9–12.
Moscicki, E.K. (1994). Gender differences in completed and attempted suicides. Annuals
of Epidemiology, 4(2), 152–58.
Murphy, G.E. (1998). Why women are less likely than men to commit suicide.
Comprehensive of Psychiatry, 39(4), 165–75.
Nock, M.K., G. Borges, E.J. Bromet, J. Alonso, M. Angermeyer, A. Beautrais et al.
(2008). Cross-national prevalence and risk factors for suicidal ideation, plans and
attempts. British Journal of Psychiatry, 192(2), 98–105.
Osvath, P., G. Kelemen, B.M. Erdõs, V. Voros and S. Fekete (2003). The main factors of
repetition. Review of some results of the Pecs Center in the WHO/EURO Multi-
centre Study on Suicidal Behaviour. Crisis, 24(4), 151–54.
Osvath, P., K. Michel and S. Fekete (2003). Contacts of suicide attempters with
healthcare services in Pecs and Bern in the WHO/EURO Multicentre Study on
Parasuicide. International Journal of Psychiatry in Clinical Practice, 7(1), 3–8.
Payne, S., V. Swami and D.L. Stanistreet (2008). The social construction of gender and
its influence on suicide. A review of the literature. Journal of Men’s Health, 5(1),
23–35.
Qin, P., P.B. Mortensen, E. Agerbo, N. Westergaard-Nielsen and T. Eriksson (2000). Gender
differences in risk factors for suicide in Denmark. British Journal of Psychiatry,
177(6), 546–50.
Qin, P., E. Agerbo and P.B. Mortensen (2003). Suicide risk in relation to socioeconomic,
demographic, psychiatric and familial factors: A national register-based study of
all suicides in Denmark, 1981–1997. American Journal of Psychiatry, 160(4), 765–72.
Rihmer, Z., N. Belsõ and K. Kiss (2002). Strategies for suicide prevention. Current
Opinion in Psychiatry, 15(1), 83–87.
Rutz, W. (2001). Preventing suicide and premature death by education and treatment.
Journal of Affective Disorders, 62(1–2), 123–29.
Schmidtke, A., U. Bille Brahe, D. De Leo, A. Kerkhof, T. Bjerke, P. Crepet et al. (1996).
Attempted suicide in Europe: Rates and sociodemographic characteristics of suicide

150
Gender Issues in Suicide

attempters during the period 1989-1992. Results of the WHO/EURO Multicentre


Study on Parasuicide. Acta Psychiatrica Scandinavica, 93(5), 327–38.
Steen, M.D. and P. Mayer (2004). Modernization and the male-female suicide ration
in India 1967–1997: Divergence or convergence? Suicide and Life-Threatening Be-
havior, 34(2), 147–58.
Varnik, A., D. Wasserman, M. Dankowicz and G. Eklund (1998). Age specific suicide
rates in the Slavic and Baltic regions of the former USSR during perestroika, in
comparison with 22 European countries. Acta Psychiatrica Scandinavica,
394(Supplement), 20–25.
Varnik, A., K. Kolves, J. Allik, E. Arensman, E. Aromaa, C. van Audenhove et al.
(2008). Gender issues in suicide rates, trends and methods among youths aged
15–24 in 15 European countries. Journal of Affective Disorders, doi: 10.1016/
j.jad.2008.06.004.
Zhang, X., H.S. Li, Q.H. Zhu, J. Zhou, S. Zhang, L. Zhang et al. (2008). Trends in sui-
cide by poisoning in China 2000–2006: Age, gender, method, and geography.
Biomedical and Enviromental Sciences, 21(3), 253–56.

151
8

Developmental Issues in
Risk Factor Assessment
KIM A. V O A L. MI

I ndividual biology, cognitions, emotions, social interactions and life


stressors impact risk for suicidal behaviour (Maris et al., 1992). Indi-
viduals at different points in human development differ across these five
domains suggesting that adopting a developmental perspective—and
attending to developmental issues—may help to explain and prevent
suicidal behaviour across the lifespan (Lester, 1991). An examination
of US suicide rates across the lifespan (Gould et al., 2003) indicates the
following characteristics: first, suicide is uncommon in children and early
adolescents (i.e., up to age 14); second, the incidence of suicide increases
at a high rate starting in later adolescence (i.e., ages 15–18) and continuing
through early adulthood (i.e., early twenties); suicide rates remain at a
relatively stable level from the early twenties through the fifties; suicide
rates are markedly elevated among the elderly (i.e., 60 years and older; in
the United States elevated rates among the elderly are accounted for by
White men).
The current chapter considers developmental issues in risk assess-
ment; we focus our discussion on four broad ‘developmental stages’, each
of which corresponds to one of the age-related patterns delineated earlier:
childhood, adolescence and early adulthood, middle adulthood, late
adulthood. We examine the ways in which developmental issues impact

152
Risk Factor Assessment

clinical decision-making in the process of assessing and managing suicide


risk across the lifespan. Our discussion is organised around four aspects
of suicide risk assessment where developmental issues may play a role: the
‘content’ of information collected during risk assessments, the ‘process’ of
conducting risk assessments, the ‘context’ surrounding risk assessments
and the ‘decisions’ generated regarding crisis management.

CONTENT OF RISK ASSESSMENTS

In the following section, we examine how the content of risk assessments—


the risk and protective factors assessed by clinicians—can be informed
by a consideration of developmental issues. A useful distinction is
between long-standing risk factors (i.e., distal factors that may predispose
individuals to suicidal behaviour) and more proximal risk factors and
warning signs (e.g., acutely stressful life events; Rudd et al., 2006). First, we
consider distal risk factors that are common across the lifespan, including
those that evidence differential predictive power at different ages. Next,
we consider the same for proximal risk factors. Subsequently, we consider
risk factors and warning signs that are common across the lifespan but
manifest differently at different ages. Finally, we consider factors unique
to particular developmental stages.

Risk Factors Common across the Lifespan

Across the lifespan, the presence of a prior history of suicidal behaviour1


is one of the strongest distal risk factors for future suicidal behaviour
(Brown et al., 2000; Joiner et al., 2005; Maser et al., 2002; Pfeffer et al.,
1993). Suicidal ideation and suicide attempts in childhood elevate risk for
suicide attempts in adolescence (Pfeffer et al., 1993). A previous suicide

1
The definition of suicidal behaviour varies across suicidologists. We consider suicidal
behaviour to include suicidal ideation, suicide attempts and completed suicide. Non-suicidal
self-injurious behaviour (NSSI; any acute deliberate destruction of body tissue without
intent to die) is an independent risk factor for suicide; however, we reference this behaviour
(NSSI) separately.

153
Kimberly A. Van Orden and Alec L. Miller

attempt is one of the strongest predictors of death by suicide in adolescents


(Brent et al., 1999; Shaffer et al., 1996), adults (Harris and Barraclough, 1997;
Mann et al., 1999) and the elderly (Beautrais, 2002; Conwell et al., 2002).
The presence of ‘multiple past attempts’ is an especially pernicious predictor
of future suicidal behaviour in both adolescents (Kotila and Lonnqvist,
1987) and adults (Rudd et al., 1996).
Although the presence of past attempts is a potent predictor of suicidal
behaviour, most individuals who die by suicide will do so on their first
attempt (Brown et al., 2000; Simon, 2006). This finding is stronger for elders,
as the lethality of suicide attempts is higher among the elderly: a greater
proportion of suicide attempts result in death among elders compared to
younger individuals (Crosby et al., 1999; McIntosh et al., 1994). ‘Non-
suicidal self-injury’ also confers risk for non-lethal and lethal attempts in
adolescents and adults (Lipschitz et al., 1999; Stanley et al., 2001). ‘Family
history of suicide’ has also been found to increase risk for death by suicide
in youth (Agerbo et al., 2002) and adults (Runeson and Asberg, 2003).
Other distal risk factors include ‘impulsivity and aggression’, which
show a documented relationship with suicidal behaviour among children
(Dervic et al., 2008), adolescents (Witte et al., 2008) and adults (Baca-
Garcia et al., 2005). ‘Problem-solving deficits’ are a feature among suicidal
children (Cohen-Sandler et al., 1982; Orbach et al., 1987), adolescents
(Klimes-Dougan et al., 1999) and adults (Wingate et al., 2005). Targeting
these deficits has been found to be helpful in reducing suicidal ideation in
younger adults (Wingate et al., 2005) and in increasing quality of life for
older adults (Gellis et al., 2007).
Several forms of stressful life events, considered proximal risk factors,
have also been found to increase risk for suicide across the lifespan.
‘Bereavement’ is one such risk factor: maternal death in youth elevates risk
(Agerbo et al., 2002) and loss of a spouse elevates risk for adults, though
the effect is stronger for younger compared to older adults (i.e., at a time
when such losses are less expected; Duberstein et al., 1998). ‘Incarceration’
in jail or prison also elevates risk for suicide in adolescents (Penn et al.,
2003; Sanislow et al., 2003) and adults (Metzner and Hayes, 2006). Certain
‘physical illnesses’—acquired immune deficiency syndrome (AIDS), brain
cancer and multiple sclerosis—elevate risk for suicide and this relation-
ship is moderated by age, with older adults having greater risk (for a review,
see Hughes and Kleespies, 2001).

154
Risk Factor Assessment

Finally, limiting ‘access to lethal means’ is one of the few interventions


that has been demonstrated to have a population-level impact on suicide
rates (Kreitman and Platt, 1984; Ludwig and Cook, 2000), indicating that
access to lethal means is a powerful risk factor for death by suicide. Data
from the United States suggests that limiting access to handguns may
be a particularly effective strategy for suicide prevention in the elderly:
initiation of the ‘Brady Act’ was associated with lowered suicide rates only
among individuals over age 55 (Ludwig and Cook, 2000). Limiting access
to handguns may also be particularly indicated for the youth given the data
that the availability of a handgun in the home is associated with a four-fold
increase in adolescents’ risk for suicide (Brent et al., 1993).

Warning Signs Common across the Lifespan

Indicators of (the likely presence of) current suicidal crises are termed
warning signs. The American Association of Suicidology created an
evidence-based list of suicide warning signs (Rudd et al., 2006), which can be
remembered with the mnemonic IS PATH WARM?—‘I’ is for ideation (as
in suicidal ideation); ‘S’ is for substance abuse; ‘P’ is for purposelessness;
‘A’ is for anxiety and agitation; ‘T’ is for ‘trapped’ (as in feeling trapped);
‘H’ is for hopelessness; ‘W’ is for withdrawal; ‘A’ is for anger; ‘R’ is for
recklessness; and ‘M’ is for mood fluctuations. We focus our discussion of
these warning signs on five of the most potent predictors of suicidal crises
across the lifespan—suicidal intent, hopelessness, social isolation (listed as
withdrawal), agitation and sleep disturbances (not included in the list).
‘Intent to die’ at the time of self-injury (i.e., the degree to which an indi-
vidual wishes to die when engaging in self-injurious behaviours) is a
warning sign for suicidal crises and has been shown to predict death by
suicide in adults (Harriss et al., 2005). Current intent for suicide is a key
component of suicide risk assessment protocols for adolescents (Reynolds,
1991) and adults (Jobes, 2006; Joiner et al., 1999; Linehan et al., 2000; Shea,
1999; Simon, 2006). Pfeffer (2003) indicates that assessing intent for suicide
in children is difficult because children may not be able to verbalise intent
or may deny intent even when significant risk for suicide exists. Children
as young as three years have concepts of death and even though these
conceptions may be inaccurate (e.g., to die is to go to sleep), self-destructive
behaviour has been observed in children with intent to achieve their

155
Kimberly A. Van Orden and Alec L. Miller

idiosyncratic conceptualisations of death (Pfeffer, 2003). To address these


complications, Pfeffer advises clinicians to gather collateral information
from parents, teachers and others in a child’s life, and to consider both
the outcome the child expects and the likely objective outcome when
ascertaining the degree of dangerousness inherent in a child’s level of
intent to engage in self-harm behaviours.
The relation between ‘hopelessness’ and suicidality is one of the most
consistent and robust findings in the suicidality literature. Hopelessness
predicts future suicidal behaviour in children (Nock and Kazdin, 2002) and
adolescents (Huth-Bocks et al., 2007). In adult samples, hopelessness has
been shown to prospectively predict death by suicide one year later (Brown
et al., 2000) and over a 13 years follow-up (Wen-Hung et al., 2004). Less
research has been conducted on hopelessness in the elderly, but available
data indicates that hopelessness predicts higher levels of suicidal ideation
in elders (Heisel and Flett, 2005). In contrast, the ability to identify ‘reasons
for living’ is a strong protective factor against suicidal behaviour in both
youth (Osman et al., 1998) and adults (Strosahl et al., 1992).
‘Social isolation’ is a strong risk factor for suicide across the lifespan.
Loneliness predicts suicidal ideation and behaviour among adolescents
(Roberts et al., 1998), adults (Dieserud et al., 2001; Koivumaa-Honkanen
et al., 2001) and the elderly (Waern et al., 2003). Social isolation was one
of the most frequently cited problems among a sample of children and
younger adolescents (ages 8–15) who died by suicide (Hawton et al., 1996).
Among adults, rates of suicide are lower among those who are married
(Stack, 2000) and those who have children (Hoyer and Lund, 1993; Qin
and Nordentoft, 2005), suggesting a protective effect of social connections.
Social isolation has also been found to be a pernicious risk factor for suicide
among the elderly (Conwell and Heisel, 2006). Having a greater number
of friends and family members to confide in is protective against suicide
in the elderly (Duberstein et al., 2004; Turvey et al., 2002), while living
alone is a risk factor for suicidal behaviour among the elderly (Conner
et al., 1999; De Leo et al., 2001).
‘Severe anxiety/agitation’ is a strong predictor of imminent risk for
suicide. In a study of inpatients ranging in age from 16 to 72 who died by
suicide, 79 percent demonstrated severe anxiety/agitation within one
week of death (Busch et al., 2003). In addition, agitation has been posited
as a potential mechanism whereby antidepressant side effects may increase
risk for suicidal behaviours in youth (Smith, forthcoming).

156
Risk Factor Assessment

‘Sleep disturbances’ are a warning sign for suicidal ideation and behav-
iours across the lifespan. In a sample of depressed children and adolescents,
the presence of insomnia was found to differentiate between those youth
who presented with current (or past) suicidal ideation with a plan from
non-suicidal youth (Barbe et al., 2005). Dysregulated sleep differentiated
between a group of adolescents who died by suicide and healthy controls
even when controlling for mood disorders (Goldstein et al., 2008). Finally,
reduced sleep quality was found to predict suicide in a prospective study
of older adults (Turvey et al., 2002).

Differen al Manifesta ons of Risk across the Lifespan

In this section, we review several risk factors common to all ages that present
differently across the lifespan. One such risk factor is the ‘experience of
physical and sexual abuse’. Childhood physical and sexual abuse are risk
factors for suicidal behaviour among young children (Rosenthal and
Rosenthal, 1984), death by suicide in childhood and adolescence (Dervic
et al., 2008; Wagner, 1997), suicide attempts in adolescence (Martin et al.,
2004), a greater number of lifetime suicide attempts as an adult (Brown
et al., 1999; Joiner et al., 2007). The experience of sexual violence as an
adult increases risk for suicidal ideation and suicide attempts as an adult
(Stepakoff, 1998). Relatedly, the experience of intimate partner violence
as an adult has been found to predict suicidal ideation in a sample of
African American females (Leiner et al., 2008), history of suicide attempts
in female psychiatric inpatients (Sansone et al., 2007), and death by suicide
in a psychological autopsy study in Sri Lanka (a country with a markedly
elevated suicide rate; Samaraweera et al., 2008). Finally, although definitive
data are needed, elder abuse has been posited as a risk factor for suicidal
behaviour among the elderly (Conwell, 1995).
The vast majority of individuals of all ages who die by suicide (i.e., ap-
proximately 95 percent) suffer from ‘mental disorders’ (Cavanagh et al.,
2003; Gould et al., 2003), and it is quite possible that the remaining 5 percent
suffer from subclinical variants of mental disorders. However, it is also the
case that the vast majority of individuals of all ages who suffer from mental
disorders do not die by committing suicide (Conner et al., 2001), indicating
that the presence of psychopathology in itself does not confer a high
degree of specificity with regards to an individual’s level of risk for suicide.

157
Kimberly A. Van Orden and Alec L. Miller

An increase in specificity can be achieved by focusing on mental disorders


that confer the greatest risk for different age groups. Among children who
die by committing suicide, mood disorders and disruptive behaviour
disorders are the most common diagnoses (Dervic et al., 2008). Major de-
pression, bipolar mixed state, substance abuse and conduct disorder are
the forms of psychopathology most common among adolescents who
die by suicide (Brent et al., 1993; Moskos et al., 2005; Shaffer et al., 1996).
Among adults, the mental disorders that confer the greatest risk for
suicide are major depressive disorder, bipolar disorder, anorexia nervosa,
schizophrenia and borderline personality disorder (Joiner et al., 2009).
Among older adults, depression stands out as the mental disorder with
the strongest relation with suicide (Conwell et al., 2002). Although
research on the relationship between dementia and suicide has been
mixed (Conwell et al., 2002), it appears that when dementia is diagnosed
during hospitalisation it is a strong predictor of suicide among elders
(Erlangsen et al., 2008). Finally, personality traits that involve rigidity
and low openness to experience have been found to predict suicide risk in
the elderly (Duberstein, 2001). This personality style may increase elders’
risk for suicide by decreasing capabilities to adapt to aging in healthy ways
and by making psychological distress more difficult for others to notice
(Duberstein, 2001).
Life stressors that involve ‘interpersonal conflict’ are also robust pre-
cipitants of suicidal behaviour across the lifespan. Adolescents report
70 percent of their suicidal behaviour is precipitated by interpersonal
conflicts (Miller and Glinski, 2000). For children and adolescents, discord
with parents is a predictor of lethal suicidal behaviour (Brent et al., 1994;
Dervic et al., 2008), while family cohesion and higher levels of parental
involvement function as protective factors against suicidal behaviour in
children and adolescents (Flouri and Buchanan, 2002; Rubenstein et al.,
1998). Peer conflict predicts suicidal behaviour in adolescents: loss of
friends after disclosing sexual orientation is a strong predictor of non-
lethal attempts among gay, lesbian and bisexual youth (Hershberger et al.,
1997); severe suicidal ideation has been found among adolescents who are
bullied and those who bully others (Kaltiala-Heino et al., 1999). Among
adults, interpersonal conflict also manifests in romantic relationships: it
is a well-replicated finding that suicide rates are elevated among divorced
individuals in the United States (Stack, 2000). Finally, family conflict is
also a predictor of suicidal behaviour in the elderly (Beautrais, 2002;
Rubenowitz et al., 2001).

158
Risk Factor Assessment

Life stressors that involve ‘perceptions of lowered social competence/


contribution’ are also robust precipitants of suicidal behaviour across
the lifespan. Among adults, perceptions of burdensomeness on family
members (and other significant others) have been found to predict suicidal
ideation (de Catanzaro, 1995; Van Orden et al., 2008) and to differentiate
between non-lethal and lethal suicide attempt status (Joiner et al., 2002). In
youth, this risk factor may manifest as perceptions of ‘expendability in the
family’: Sabbath (1969: 272–73) describes the ‘expendable child’ construct
as a ‘parental wish, conscious or unconscious, spoken or unspoken, that
the child interprets as [the parents’] desire to be rid of him, for him to die’.
Research has borne out this relationship between feelings of expendability
and suicidal behaviour in children. In a sample of preschoolers, suicidal
children were significantly more likely to be the product of ‘unwanted’
pregnancies (Rosenthal and Rosenthal, 1984). In addition, Woznica and
Shapiro (1990) demonstrated that suicidal adolescents scored higher
on a measure of perceived expendability than non-suicidal adolescent
outpatients. Lowered competence may also manifest in adolescence as
academic difficulties, a life stressor that has been shown to predict suicidal
behaviour in this age group (Mazza and Eggert, 2001). Finally, functional
decline—related to such indices as retirement and physical illness—places
elders at increased risk for suicide, but only in the presence of other risk
factors (e.g., depression) as all elders experience functional decline, but
most do not die by committing suicide (Conwell and Heisel, 2006).

Factors Unique to Developmental Stages

The phenomenon of ‘suicide contagion’ involves the interpersonal trans-


mission of elevated risk for suicidal behaviour through mechanisms such
as modeling. A related concept is suicide clustering—deaths by suicide
occurring in a greater concentration than would be expected by random
chance. The preponderance of evidence in support of contagion/modeling
effects (including media depictions) has been for youth (Insel and Gould,
2008). In addition, evidence suggests that these effects may be strongest
among younger adolescents and lose their potency in young adulthood
(Gould et al., 1994). Adolescents with exposure to friends, relatives,
classmates or other significant others, who attempt or commit suicide are
at elevated risk for engaging in suicidal behaviour. Therefore, it is critically

159
Kimberly A. Van Orden and Alec L. Miller

important for family members and school personnel to assess and monitor
a teen whose peer from school has committed suicide or when a teen icon
(e.g., Kurt Cobain) has killed himself or herself.
An additional factor unique to adolescents and children is the possible
association between ‘antidepressant prescription’ and increased risk for
suicidal behaviour. The ‘black box’ advisory issued by the Food and
Drug Administration (FDA) in October 2004 on the potential danger
of treating children and adolescents with antidepressants (US Food and
Drug Administration Public Health Advisory, 2004) was primarily based on
findings from a meta-analysis of 23 randomised trials that found higher
rates of suicidality (no lethal attempts occurred) among youth prescribed
with selective serotonin reuptake inhibitors (SSRIs; Hammad et al., 2006).
Subsequent investigations have raised doubts about the FDA’s conclusion
that SSRIs cause suicidality in youth (Kaizar et al., 2006) and some sug-
gest that the risk-to-benefit profile indicates that benefits outweigh risks
in the treatment of youth with SSRIs (Bridge et al., 2007). Regarding
potential mechanisms for increased risk for suicidality in youth prescribed
with SSRIs, Smith (2009) found that increased risk is partially attributable
to medication half-life, with shorter half-life medications conveying
greater risk, possibly due to greater risk for side effects such as akathisia.
Concerning effects of the FDA regulations, data indicate that SSRI pre-
scription rates for children and adolescents decreased subsequent to the
advisory, with associated increases in suicide rates (Bridge et al., 2008;
Gibbons et al., 2007). These data indicate a need for ongoing assessment of
suicide risk in children and adolescents who are prescribed SSRIs, including
monitoring of side effects and clear communication with youth and parents
that they must make contact with the physician at the first signs of any
troublesome changes in mood or behaviour (Brent, 2004).

PROCESS AND CONTEXT OF RISK ASSESSMENTS

The majority of elders who commits suicide present for services in primary
care settings: these elders do not, in most cases, present to clinics for mental
health services (Luoma et al., 2002). In fact, a recent review of studies
reported that 77 percent of older adults (across 40 studies) were seen by
a primary care physician in the year before their deaths by committing

160
Risk Factor Assessment

suicide (100 percent of the women, 78 percent of the men; Luoma et al.,
2002). These data indicate that risk assessment efforts focused solely in
mental health clinics are unlikely to reach most elders, whereas efforts
targeting both mental health clinics and primary clinics have the potential
to reach the majority of elders at high risk for suicide. Thus, a consideration
of the context of risk assessment for elders must include primary care
as well as mental health settings. The process of risk assessment in primary
care differs from the process used in mental health clinics because of the
greater heterogeneity in risk levels. To address this issue, a hierarchical
stepped care model of assessment and intervention can be used, with more
in-depth assessment and intervention for elders who present with elevated
risk on an initial screening. Prevention of Suicide in Primary-care Elderly:
Collaborative Trial (PROSPECT) is an evidence-based suicide prevention
programme for the elderly, based in primary care settings, that utilises a
stepped care model and has demonstrated efficacy for older adults in the
reduction of suicidal ideation (Bruce et al., 2004) and in the treatment of
residual symptoms of depression (Alexopoulos et al., 2005).
Screening for suicide risk outside of mental health clinics is also rele-
vant to suicide prevention in children and adolescents. Estimates suggest
that many (if not most) of the adolescents who commit suicide might not
have received mental health treatment (Blumenthal, 1990). Gatekeeper
training as well as screening for risk factors represent potentially effective
methods of suicide prevention (Shaffer and Pfeffer, 2001). Gatekeeper
training involves educating laypersons in direct contact with youth (e.g.,
teachers, parents, clergy and peers) about warning signs for suicide (e.g.,
IS PATH WARM acronym) so that the ‘gatekeepers’ can refer these youth
to mental health professionals. Research has demonstrated that school-
wide screening for suicide risk can be efficacious in the prevention of
adolescent suicide (Reynolds, 1991; Shaffer and Craft, 1999) but school
officials are often hesitant to do so because they believe that asking about
suicide will encourage youth to engage in suicidal behaviour (Miller et al.,
1999). A randomised controlled trial of the effect of youth screening for
suicide on suicidal behaviour did not find iatrogenic effects of screening
(Gould et al., 2005), and the study authors concluded that screening for
suicide in youth is a safe component of prevention programmes. Several
suicide-risk screening protocols for adolescents have been developed and
tested, including the Columbia University Teen Screen Program (Shaffer
et al., 2004).

161
Kimberly A. Van Orden and Alec L. Miller

The process of risk assessment with children and adolescents differs


from the process used with adults because of a stronger emphasis on the
role of collateral information from family members. While gathering
collateral information from significant others can be valuable when work-
ing with adults—an important component of some adult risk assessment
protocols (e.g., Shea, 1999)—the process of including family members in
risk assessment is especially important when working with children and
adolescents because data indicate that youth often minimise their degree
of intent for recent suicidal behaviour (Kingsbury, 1993), or often have
difficulties expressing suicidal ideas and intent (Pfeffer, 2003), including
having differential responses to interviewer-administered versus self-report
questions regarding suicide (Velting et al., 1998), and may misjudge the
lethality of suicidal methods (Pfeffer, 2003). Ability to consent to treatment
and issues of confidentiality surrounding sharing information with parents
are additional issues that concern risk assessment with youth.
Studies suggest that both suicidal adolescents and elders may demon-
strate less willingness to disclose suicidal behaviours compared to adults.
Regarding adolescents, Gould and colleagues (2004) reported data
indicating that many youth believe they should be able to handle problems
on their own and should keep suicidal behaviours a secret—and thus
believe they should not seek help for mental health problems. Regarding
elders, recall the research described earlier demonstrating that elder
adults with personality styles characterised by rigidity and low openness
to experience may be less likely to disclose suicidal ideation (Duberstein,
2001). Thus, clinicians working with elders and youth are advised to
use interviewing techniques designed to increase willingness to disclose
and to obtain collateral information on other relevant risk factors.
Clinicians working with adolescents may wish to consider the Evaluation
of Suicide Risk Among Adolescents and Imminent Danger Assessment
(Rotheram, 1987), a clinician-administered interview that addresses risk
factors unique to, or especially relevant for, adolescents, including ex-
posure to suicidal behaviour in family and friends and conduct disorder
symptomatology. Clinicians working with elder adults may wish to
consider the Chronological Assessment of Suicide Events (Shea, 1999)—a
risk assessment framework that emphasizes the use of rapport-building
techniques designed to enhance individuals’ willingness to disclose
information about suicidal behaviours—or the Collaborative Assessment
and Management of Suicidality (Jobes, 2006), a risk assessment framework

162
Risk Factor Assessment

that emphasises collaboration as well as the importance and value of the


therapeutic alliance.

PULLING IT TOGETHER: CLINICAL DECISION MAKING

We now turn to the fourth aspect of risk assessment that should be in-
formed by developmental considerations, crisis management—clinical
decision-making as how to best manage risk and prevent a suicide at-
tempt. Involvement of family members to promote a safe environment
is essential when working with suicidal children and adolescents. For
example, clinicians should meet with parents/guardians to ensure that
firearms, lethal medications or any means for suicide are removed from
the home or, at the very least, inaccessible to the youth (Shaffer and Pfeffer,
2001). Research suggests that an explicit discussion of this precaution with
parents is necessary to ensure that it is implemented: parents/guardians
do not tend to take these precautions unless explicitly instructed to do so
(McManus et al., 1997). Removal of firearms from the home is an action
relevant across the lifespan, but may be particularly indicated for youth,
given that the presence of a firearm is a strong predictor of suicide in
adolescents (Brent et al., 1999) and elders, given that most elders who
die by suicide use firearms, and suicidal behaviour in the elderly involves
a greater degree of planning and the use of more lethal means (Conwell
and Heisel, 2006).
Our discussion of developmental considerations relevant to suicide
risk assessment has emphasised the commonalities in risk assessment
across the lifespan while highlighting unique features when appropriate.
A final feature of suicide risk common across the lifespan is the fact that
the vast majority of individuals exhibiting the risk factors and warning
signs presented in this chapter will (fortunately) not engage in suicidal
behaviour. Research has yet to definitively elucidate which factors are
necessary and sufficient to result in suicidal behaviour. It is our hope that
the material presented in this chapter will serve two inter-related functions:
first, to assist clinicians in skillfully assessing and managing suicide risk
for individuals of all ages; and second, to assist researchers in identifying
commonalities in suicidal behaviour that may provide clues as to those
factors most likely to cause suicidal behaviour.

163
Kimberly A. Van Orden and Alec L. Miller

REFERENCES

Agerbo, E., M. Nordentoft and P.B. Mortensen (2002). Familial, psychiatric, and
socioeconomic risk factors for suicide in young people: Nested case-control study.
British Medical Journal, 325(7355), 74.
Alexopoulos, G.S., I.R. Katz, M.L. Bruce, M. Heo, T.T. Have, P. Raue et al. (2005). Remis-
sion in depressed geriatric primary care patients: A report from the PROSPECT
study. American Journal of Psychiatry, 162, 718–24.
Baca-Garcia, E., C. Diaz-Sastre, E.G. Resa, H. Blasco, D.B. Conesa, M.A. Oquendo
et al. (2005). Suicide attempts and impulsivity. European Archives of Psychiatry
and Clinical Neuroscience, 255, 152–56.
Barbe, R.P., D.E. Williamson, J.A. Bridge, B. Birmaher, R.E. Dahl, D.A. Axelson et al.
(2005). Clinical differences between suicidal and nonsuicidal depressed children
and adolescents [Comparative Study; Research Support, Non-U.S. Gov’t; Research
Support, U.S. Gov’t, P.H.S.]. The Journal of clinical psychiatry, 66, 492–98.
Beautrais, A.L. (2002). A case control study of suicide and attempted suicide in older
adults. Suicide and Life-Threatening Behavior, 32(1), 1–9.
Blumenthal, S.J. (1990). Youth suicide: Risk factors, assessment, and treatment of
adolescent and young adult suicidal patients. Psychiatric Clinics of North America,
13, 511–56.
Brent, D.A. (2004). Antidepressants and pediatric depression — The risk of doing
nothing. New England Journal of Medicine, 351(16), 1598–601.
Brent, D.A., M. Baugher, J. Bridge, T. Chen and L. Chiappetta (1999). Age- and sex-
related risk factors for adolescent suicide. Journal of the American Academy of Child &
Adolescent Psychiatry, 38, 1497–505.
Brent, D.A., J.A. Perper, G. Moritz, C. Allman, A. Friend, C. Roth et al. (1993). Psychi-
atric risk factors for adolescent suicide: A case-control study. Journal of the American
Academy of Child & Adolescent Psychiatry, 32, 521–29.
Brent, D.A., J.A., Perper, G. Moritz, M. Baugher, J. Schweers and C. Roth (1994).
Suicide in affectively ill adolescents: A case-control study. Journal of Affective
Disorders, 31(3), 193–202.
Bridge, J.A., J.B. Greenhouse, A.H., Weldon, J.V. Campo and K.J. Kelleher (2008). Sui-
cide trends among youths aged 10 to 19 years in the United States, 1996–2005.
Journal of the American Medical Association, 300(9), 1025–26.
Bridge, J.A., S. Iyengar, C.B. Salary, R.P. Barbe, B. Birmaher, H.A. Pincus et al. (2007).
Clinical response and risk for reported suicidal ideation and suicide attempts in
pediatric antidepressant treatment: A meta-analysis of randomized controlled
trials. Journal of the American Medical Association, 297, 1683–96.
Brown, G.K., A.T. Beck, R.A. Steer and J.R. Grisham (2000). Risk factors for suicide
in psychiatric outpatients: A 20-year prospective study. Journal of Consulting and
Clinical Psychology, 68, 371–77.

164
Risk Factor Assessment

Brown, J., P. Cohen, J.G. Johnson and E.M. Smailes (1999). Childhood abuse and
neglect: Specificity of effects on adolescent and young adult depression and
suicidality. Journal of the American Academy of Child & Adolescent Psychiatry,
38(12), 1490–96.
Bruce, M.L., T.R. Ten Have, C. F. Reynolds, III, I.I. Katz, H.C. Schulberg, B.H. Mulsant
et al. (2004). Reducing suicidal ideation and depressive symptoms in depressed
older primary care patients: A randomized controlled trial. JAMA: Journal of
the American Medical Association, 291, 1081–91.
Busch, K.A., J. Fawcett and D.G. Jacobs (2003). Clinical correlates of inpatient suicide.
Journal of Clinical Psychiatry, 64(1), 14–19.
Cavanagh, J.T., A.J. Carson, M. Sharpe and S.M. Lawrie (2003). Psychological autopsy
studies of suicide: A systematic review. Psychological Medicine, 33(3), 395–405.
Cohen-Sandler, R., A.L. Berman and R.A. King (1982). Life stress and symptomatology:
Determinants of suicidal behavior in children. Journal of the American Academy
of Child Psychiatry, 21(2), 178–86.
Conner, K.R., P.R. Duberstein and Y. Conwell (1999). Age-related patterns of factors
associated with completed suicide in men with alcohol dependence. The American
Journal on Addictions, 8(4), 312–18.
Conner, K.R., P.R. Duberstein, Y. Conwell, L. Seidlitz and E.D. Caine (2001). Psy-
chological vulnerability to completed suicide: A review of empirical studies. Suicide
and Life-Threatening Behavior, 31(4), 367–85.
Conwell, Y. (1995). Elder abuse—A risk factor for suicide? Crisis: The Journal of Crisis
Intervention and Suicide Prevention, 16(3), 104–05.
Conwell, Y., P.R. Dubertstein and E.D. Caine (2002). Risk factors for suicide in later
life. Biological Psychiatry, 52(3), 193–204.
Conwell, Y. and M. J. Heisel (2006). The elderly. In R.I. Simon and R.E. Hales (Eds),
Textbook of Suicide Assessment and Management (pp. 57–76). Arlington, VA: The
American Psychiatric Publishing.
Crosby, A.E., M.P. Cheltenham and J.J. Sacks (1999). Incidence of suicidal ideation
and behavior in the United States, 1994. Suicide and Life-Threatening Behavior,
29(2), 131–40.
de Catanzaro, D. (1995). Reproductive status, family interactions, and suicidal ideation:
Surveys of the general public and high-risk groups. Ethology & Sociobiology, 16(5),
385–94.
De Leo, D., W. Padoani, P. Scocco, D. Lie, U. Bille-Brahe, E. Arensman et al. (2001).
Attempted and completed suicide in older subjects: Results from the WHO/
EURO Multicentre Study of Suicidal Behaviour. International Journal of Geriatric
Psychiatry, 16, 300–10.
Dervic, K., D.A. Brent and M.A. Oquendo (2008). Completed suicide in childhood.
Psychiatric Clinics of North America, 31(2), 271–91.
Dieserud, G., E. Roysamb, O. Ekeberg and P. Kraft (2001). Toward an integrative model
of suicide attempt: A cognitive psychological approach. Suicide & Life- Threatening
Behavior, 31(2), 153–68.

165
Kimberly A. Van Orden and Alec L. Miller

Duberstein, P.R. (2001). Are closed-minded people more open to the idea of killing
themselves? Suicide and Life-Threatening Behavior, 31(1), 9–14.
Duberstein, P.R., Y. Conwell, K.R. Conner, S. Eberly, J.S. Evinger and E. D. Caine
(2004). Poor social integration and suicide: Fact or artifact? A case-control study.
Psychological Medicine, 34(7), 1331–37.
Duberstein, P.R., Y. Conwell and C. Cox (1998). Suicide in widowed persons: A psy-
chological autopsy comparison of recently and remotely bereaved older subjects.
American Journal of Geriatric Psychiatry, 6(4), 328–34.
Erlangsen, A., S.H. Zarit and Y. Conwell (2008). Hospital-diagnosed dementia and sui-
cide: A longitudinal study using prospective, nationwide register data. American
Journal of Geriatric Psychiatry, 16(3), 220–28.
Flouri, E. and A. Buchanan (2002). The protective role of parental involvement in
adolescent suicide. Crisis: The Journal of Crisis Intervention and Suicide Prevention,
23(1), 17–22.
Gellis, Z.D., J. McGinty, A. Horowitz, M.L. Bruce and E. Misener (2007). Problem-
solving therapy for late-life depression in home care: A randomized field trial.
American Journal of Geriatric Psychiatry, 15(11), 968–78.
Gibbons, R.D., C.H. Brown, K. Hur, S.M. Marcus, D.K. Bhaumik, J.A. Erkens et al.
(2007). Early evidence on the effects of regulators’ suicidality warnings on SSRI pre-
scriptions and suicide in children and adolescents. American Journal of Psychiatry,
164, 1356–63.
Goldstein, T.R., J.A. Bridge and D.A. Brent (2008). Sleep disturbance preceding com-
pleted suicide in adolescents [Peer Reviewed]. Journal of Consulting and Clinical
Psychology, 76(1), 84–91.
Gould, M.S., F.A. Marrocco, M. Kleinman, J.G. Thomas, K. Mostkoff, J. Cote et al. (2005).
Evaluating Iatrogenic Risk of Youth Suicide Screening Programs: A Randomized
Controlled Trial. JAMA: Journal of the American Medical Association, 293, 1635–43.
Gould, M.S., K. Petrie and M.H. Kleinman (1994). Clustering of attempted suicide: New
Zealand national data. International Journal of Epidemiology, 23(6), 1185–89.
Gould, M.S., D. Shaffer and T. Greenberg (2003). The epidemiology of youth suicide. In
Robert A. King and A. Apter (Eds), Suicide in children and adolescents (pp. 1–40).
United Kingdom: Cambridge University Press.
Gould, M.S., D. Velting, M. Kleinman, C. Lucas, J.G. Thomas and M. Chung (2004).
Teenagers’ Attitudes About Coping Strategies and Help-Seeking Behavior for
Suicidality. Journal of the American Academy of Child & Adolescent Psychiatry,
43(9), 1124–33.
Hammad, T.A., T. Laughren and J. Racoosin (2006). Suicidality in pediatric patients
treated with antidepressant drugs. Archives of General Psychiatry, 63(3), 332–39.
Harris, E.C. and B. Barraclough (1997). Suicide as an outcome for mental disorders.
British Journal of Psychiatry, 170(3), 205–28.
Harriss, L., K. Hawton and D. Zahl (2005). Value of measuring suicidal intent in the
assessment of people attending hospital following self-poisoning or self-injury.
British Journal of Psychiatry, 186(1), 60–66.

166
Risk Factor Assessment

Hawton, K., J. Fagg and S. Simkin (1996). Deliberate self-poisoning and self-injury
in children and adolescents under 16 years of age in Oxford, 1976–1993. British
Journal of Psychiatry, 169(2), 202–08.
Heisel, M.J. and G.L. Flett (2005). A psychometric analysis of the Geriatric Hopelessness
Scale (GHS): Towards improving assessment of the construct. Journal of Affective
Disorders, 87(2–3), 211–20.
Hershberger, S.L., N.W. Pilkington and A.R. D’Augelli (1997). Predictors of suicide
attempts among gay, lesbian, and bisexual youth. Journal of Adolescent Research,
12(4), 477–97.
Hoyer, G. and E. Lund (1993). Suicide among women related to number of children
in marriage. Archives of General Psychiatry, 50(2), 134–37.
Hughes, D. and P. Kleespies (2001). Suicide in the medically ill. Suicide and Life-
Threatening Behavior, 31 (1 Supplement), 48–59.
Huth-Bocks, A.C., D.C.R. Kerr, A.Z. Ivey, A.C. Kramer and C.A. King (2007).
Assessment of psychiatrically hospitalized suicidal adolescents: Self-report
instruments as predictors of suicidal thoughts and behavior. Journal of the
American Academy of Child & Adolescent Psychiatry, 46(3), 387–95.
Insel, B.J. and M.S. Gould (2008). Impact of modeling on adolescent suicidal behavior.
Psychiatric Clinics of North America, 31(2), 293–316.
Jobes, D.A. (2006). Managing suicidal risk: A collaborative approach. New York, NY:
Guilford Press.
Joiner, T., J.W. Pettit, R.L. Walker, Z.R. Voelz, J. Cruz, M.D. Rudd et al. (2002).
Perceived burdensomeness and suicidality: Two studies on the suicide notes
of those attempting and those completing suicide. Journal of Social & Clinical
Psychology, 21, 531–45.
Joiner, T.E., Jr., Y. Conwell, K.K. Fitzpatrick, T.K. Witte, N.B. Schmidt, M.T.Berlim
et al. (2005). Four studies on how past and current suicidality relate even when
“Everything But the Kitchen Sink” is covaried. Journal of Abnormal Psychology,
114, 291–303.
Joiner, T.E., Jr., N.J. Sachs-Ericsson, L.R. Wingate, J.S. Brown, M.D. Anestis and E.A.
Selby (2007). Childhood physical and sexual abuse and lifetime number of suicide
attempts: A persistent and theoretically important relationship. Behaviour Research
and Therapy, 45(3), 539–47.
Joiner, T.E., Jr., R.L. Walker, M.D. Rudd and D.A. Jobes (1999). Scientizing and
routinizing the assessment of suicidality in outpatient practice. Professional
Psychology: Research and Practice, 30(5), 447–53.
Joiner, T.E., K.A. Van Orden, T.K. Witte and M.D. Rudd (2009). The interpersonal
theory of suicide: Guidance for working with suicidal clients. Washington, D.C.:
American Psychological Association.
Kaizar, E.E., J.B. Greenhouse, H. Seltman and K. Kelleher (2006). Do antidepressants
cause suicidality in children? A Bayesian meta-analysis. Clinical Trials, 3(2), 73–90;
discussion 91–78.

167
Kimberly A. Van Orden and Alec L. Miller

Kaltiala-Heino, R., M. Rimpela, M. Marttunen, A. Rimpela and P. Rantanen (1999).


Bullying, depression, and suicidal ideation in Finnish adolescents: School survey.
British Medical Journal, 319(7206), 348–51.
Kingsbury, S.J. (1993). Clinical components of suicidal intent in adolescent overdose.
Journal of the American Academy of Child & Adolescent Psychiatry, 32(3),
518–20.
Klimes-Dougan, B., K. Free, D. Ronsaville, J. Stilwell, C.J. Welsh and M. Radke-Yarrow
(1999). Suicidal ideation and attempts: A longitudinal investigation of children of
depressed and well mothers. Journal of the American Academy of Child & Adolescent
Psychiatry, 38(6), 651–59.
Koivumaa-Honkanen, H., R. Honkanen, H. Viinamaki, K. Heikkila, J. Kaprio and M.
Koskenvuo (2001). Life satisfaction and suicide: A 20-year follow-up study. The
American Journal of Psychiatry, 158(3), 433–39.
Kotila, L. and J. Lonnqvist (1987). Adolescents who make suicide attempts repeatedly.
Acta Psychiatrica Scandinavica, 76(4), 386–93.
Kreitman, N. and S. Platt (1984). Suicide, unemployment, and domestic gas detoxi-
fication in Britain. Journal of Epidemiology & Community Health, 38(1), 1–6.
Leiner, A.S., M.T. Compton, D. Houry and N.J. Kaslow (2008). Intimate partner
violence, psychological distress, and suicidality: A path model using data from
African American women seeking care in an urban emergency department. Journal
of Family Violence, 23(6), 473–81.
Lester, D. (1991). A brief introduction to the stages of development. In A.A. Leenars
(Ed.), Life span perspectives of suicide: Time lines in suicide process (pp. 17–24).
New York: Plenum.
Linehan, M.M., K.A. Comtois and A. Murray (2000). The University of Washington
Risk Assessment Protocol (UWRAP). University of Washington.
Lipschitz, D.S., R.K. Winegar, A.L. Nicolaou, E. Hartnick, M. Wolfson and S.M.
Southwick (1999). Perceived abuse and neglect as risk factors for suicidal behavior
in adolescent inpatients. Journal of Nervous and Mental Disease, 187 (1), 32–39.
Ludwig, J. and P.J. Cook (2000). Homicide and suicide rates associated with
implementation of the Brady Handgun Violence Prevention Act. JAMA: Journal
of the American Medical Association, 284(5), 585–91.
Luoma, J.B., C.E. Martin and J.L. Pearson (2002). Contact with mental health and
primary care providers before suicide: A review of the evidence. American Journal
of Psychiatry, 159(6), 909–16.
Mann, J.J., C. Waternaux, G.L. Haas and K.M. Malone (1999). Toward a clinical
model of suicidal behavior in psychiatric patients. American Journal of Psychiatry,
156(2), 181–89.
Maris, R.W., A.L. Berman and J.T. Maltsberger (1992). Summary and conclusions:
What have we learned about suicide assessment and prediction? In R.W. Maris,
A.L. Berman, J.T. Maltsberger and R.I. Yufit (Eds), Assessment and Prediction of
Suicide (pp. 640–72). New York: Guilford Press.

168
Risk Factor Assessment

Martin, G., H.A. Bergen, A.S. Richardson, L. Roeger and S. Allison (2004). Sexual abuse
and suicidality: Gender differences in a large community sample of adolescents.
Child Abuse & Neglect, 28(5), 491–503.
Maser, J.D., H.S. Akiskal, P. Schettler, W. Scheftner, T. Mueller, J. Endicott et al. (2002).
Can temperament identify affectively ill patients who engage in lethal or near-
lethal suicidal behavior? A 14-year prospective study. Suicide and Life-Threatening
Behavior, 32, 10–32.
Mazza, J.J. and L.L. Eggert (2001). Activity involvement among suicidal and nonsuicidal
high-risk and typical adolescents. Suicide and Life-Threatening Behavior, 31(3),
265–81.
McIntosh, J.L., J.F. Santos, R.W. Hubbard and J.C. Overholser (1994). Elder suicide:
research, theory and treatment. Washington, DC, US: American Psychological
Association.
McManus, B.L. M.J. Kruesi, A.E. Dontes, C.R. Defazio, J.T. Piotrowski and P.J.
Woodward (1997). Child and adolescent suicide attempts: An opportunity for
emergency departments to provide injury prevention education. American Journal
of Emergency Medicine, 15(4), 357–60.
Metzner, J.L. and L.M. Hayes (2006). Suicide prevention in jails and prisons. In
R.I. Simon and R.E. Hales (Eds), Textbook of suicide assessment and management
(pp. 139–58). Arlington, VA: The American Psychiatric Publishing.
Miller, A.L. and J. Glinski (2000). Youth suicidal behavior: Assessment and intervention.
Journal of Clinical Psychology, 56(9), 1131–52.
Miller, D.N. T.L. Eckert, G.J. DuPaul and G.P. White (1999). Adolescent suicide pre-
vention: Acceptability of school-based programs among secondary school
principals. Suicide and Life-Threatening Behavior, 29(1), 72–85.
Moskos, M., L. Olson, S. Halbern, T. Keller and D. Gray (2005). Utah youth suicide
study: Psychological autopsy. Suicide and Life-Threatening Behavior, 35(5),
536–46.
Nock, M.K. and A.E. Kazdin (2002). Examination of affective, cognitive, and behavioral
factors and suicide-related outcomes in children and young adolescents. Journal
of Clinical Child and Adolescent Psychology, 31(1), 48–58.
Orbach, I., E. Rosenheim and E. Hary (1987). Some aspects of cognitive functioning
in suicidal children. Journal of the American Academy of Child & Adolescent
Psychiatry, 26(2), 181–85.
Osman, A., W.R. Downs, B.A. Kopper, F.X. Barrios, M.T. Baker, J. R. Osman et al. (1998).
The Reasons for Living Inventory for Adolescents (RFL-A): Development and
psychometric properties. Journal of Clinical Psychology, 54, 1063–78.
Penn, J.V., C.L. Esposito, L.E. Schaeffer, G.K. Fritz and A. Spirito (2003). Suicide
attempts and self-mutilative behavior in a juvenile correctional facility. Journal of
the American Academy of Child & Adolescent Psychiatry, 42(7), 762–69.
Pfeffer, C.R. (2003). Assessing suicidal behavior in children and adolescents. In
R.A. King and A. Apter (Eds), Suicide in children and adolescents (pp. 211–26).
New York: Cambridge University Press.

169
Kimberly A. Van Orden and Alec L. Miller

Pfeffer, C.R., G.L. Klerman, S.W. Hurt, T. Kakuma et al. (1993). Suicidal children grow
up: Rates and psychosocial risk factors for suicide attempts during follow-up.
Journal of the American Academy of Child & Adolescent Psychiatry, 32, 106–13.
Qin, P. and M. Nordentoft (2005). Suicide risk in relation to psychiatric hospitalization.
Archives of General Psychiatry, 62(4), 427–32.
Reynolds, W.M. (1991). A school-based procedure for the identification of adolescents
as risk for suicidal behaviors. Family and Community Health, 14(3), 64–75.
Roberts, R.E., C.R. Roberts and Y.R. Chen (1998). Suicidal thinking among adolescents
with a history of attempted suicide. Journal of the American Academy of Child and
Adolescent Psychiatry, 37(12), 1294–300.
Rosenthal, P.A. and S. Rosenthal (1984). Suicidal behavior by preschool children.
American Journal of Psychiatry, 141(4), 520–25.
Rotheram, M.J. (1987). Evaluation of imminent danger for suicide among youth.
American Journal of Orthopsychiatry, 57(1), 102–10.
Rubenowitz, E., M. Waern, K. Wilhelmson and P. Allebeck (2001). Life events and
psychosocial factors in elderly suicides—A case-control study. Psychological
Medicine, 31(7), 1193–1202.
Rubenstein, J.L., A. Halton, L. Kasten, C. Rubin and G. Stechler (1998). Suicidal behavior
in adolescents: Stress and protection in different family contexts. American Journal
of Orthopsychiatry, 68(2), 274–84.
Rudd, M.D., A.L. Berman, T.E. Joiner, M.K. Nock, M.M. Silverman, M. Mandrusiak
et al. (2006). Warning signs for suicide: theory, research, and clinical appli-
cations. Suicide and Life-Threatening Behavior, 36, 255–62.
Rudd, M.D., T. Joiner and M.H. Rajab (1996). Relationships among suicide ideators,
attempters, and multiple attempters in young-adult sample. Journal of Abnormal
Psychology, 105(4), 541–50.
Runeson, B. and M. Asberg (2003). Family history of suicide among suicide victims.
American Journal of Psychiatry, 160(8), 1525–26.
Sabbath, J.C. (1969). The suicidal adolescent: The expendable child. Journal of the
American Academy of Child Psychiatry, 8(2), 272–85.
Samaraweera, S., A. Sumathipala, S. Siribaddana, S. Sivayogan and D. Bhugra (2008).
Completed suicide among Sinhalese in Sri Lanka: A psychological autopsy study.
Suicide and Life-Threatening Behavior, 38(2), 221–28.
Sanislow, C.A., C.M. Grilo, D.C. Fehon, S.R. Axelrod and T.H. McGlashan (2003). Cor-
relates of suicide risk in juvenile detainees and adolescent inpatients. Journal of the
American Academy of Child & Adolescent Psychiatry, 42(2), 234–40.
Sansone, R.A., J. Chu and M.W. Wiederman (2007). Suicide attempts and domestic vio-
lence among women psychiatric inpatients. International Journal of Psychiatry in
Clinical Practice, 11(2), 163–66.
Shaffer, D. and L. Craft (1999). Methods of adolescent suicide prevention. Journal of
Clinical Psychiatry, 60(2 Supplement), 70–74.
Shaffer, D., M.S. Gould, P. Fisher, P. Trautman, D. Moreau, M. Kleinman et al.
(1996). Psychiatric diagnosis in child and adolescent suicide. Archives of General
Psychiatry, 53, 339–48.

170
Risk Factor Assessment

Shaffer, D. and C.R. Pfeffer (2001). Practice parameter for the assessment and treatment
of children and adolescents with suicidal behavior. Journal of the American Academy
of Child & Adolescent Psychiatry, 40(7 Supplement), 24S–51S.
Shaffer, D., M. Scott, H. Wilcox, C. Maslow, R. Hicks, C.P. Lucas et al. (2004). The Columbia
suicide screen: validity and reliability of a screen for youth suicide and depression.
Journal of the American Academy of Child and Adolescent Psychiatry, 43, 71–79.
Shea, S.C. (1999). The practical art of suicide assessment: A guide for mental health pro-
fessionals and substance abuse counselors. Hoboken, NJ, US: John Wiley & Sons.
Simon, R.I. (2006). Suicide risk: Assessing the unpredictable. In R.I. Simon and
R.E. Hales (Eds), Textbook of Suicide Assessment and Management (pp. 1–32).
Washington, DC: American Psychiatric Publishing, Inc.
Smith, E.G. (2009). Association between antidepressant half-life and the risk of suicidal
ideation or behavior among children and adolescents: Confirmatory analysis and
research implications. Journal of Affective Disorders, 114(1–3), 143–48.
Stack, S. (2000). Suicide: A 15-year review of the sociological literature. Part II:
modernization and social integration perspectives. Suicide & Life-Threatening
Behavior, 30(2), 163–76.
Stanley, B., M.J. Gameroff, V. Michalsen and J.J. Mann (2001). Are suicide attempters
who self-mutilate a unique population? American Journal of Psychiatry, 158 (3),
427–32.
Stepakoff, S. (1998). Effects of sexual victimization on suicidal ideation and behavior in
U.S. college women. Suicide and Life-Threatening Behavior. Special Issue: Gender,
culture and suicidal behavior, 28(1), 107–26.
Strosahl, K., J.A. Chiles and M. Linehan (1992). Prediction of suicide intent in hos-
pitalized parasuicides: Reasons for living, hopelessness, and depression. Compre-
hensive psychiatry, 33(6), 366–73.
Turvey, C.L., Y. Conwell, M.P. Jones, C. Phillips, E. Simonsick, J.L. Pearson et al. (2002).
Risk factors for late-life suicide: A prospective community-based study. American
Journal of Geriatric Psychiatry. Special Issue: Suicidal behaviors in older adults, 10,
398–406.
US Food and Drug Administration Public Health Advisory (2004). Worsening de-
pression and suicidality in patients being treated with antidepressant medications.
Retrieved 15 November 2004 from http://www.fda.gov/cder/drug/antidepressants/
AntidepressanstPHA.htm
Van Orden, K.A., T.K. Witte, K.H. Gordon, T.W. Bender and T. E. Joiner, Jr. (2008).
Suicidal desire and the capability for suicide: Tests of the interpersonal-
psychological theory of suicidal behavior among adults. Journal of Consulting and
Clinical Psychology, 76(1), 72–83.
Velting, D.M., J.H. Rathus and G.M. Asnis (1998). Asking adolescents to explain
discrepancies in self-reported suicidality. Suicide and Life-Threatening Behavior,
28(2), 187–96.
Waern, M., E. Rubenowitz and K. Wilhelmson (2003). Predictors of suicide in the old
elderly. Gerontology, 49(5), 328–34.

171
Kimberly A. Van Orden and Alec L. Miller

Wagner, B.M. (1997). Family risk factors for child and adolescent suicidal behavior.
Psychological Bulletin, 121(2), 246–98.
Wen-Hung, K., J.J. Gallo and W.W. Eaton (2004). Hopelessness, depression, substance
disorder, and suicidality: A 13-year community-based study. Social Psychiatry and
Psychiatric Epidemiology, 39(6), 497–501.
Wingate, L.R., K.A. Van Orden, T.E. Joiner, Jr., F.M. Williams and M.D. Rudd (2005).
Comparison of Compensation and Capitalization Models When Treating Sui-
cidality in Young Adults. Journal of Consulting and Clinical Psychology, 73(4),
756–62.
Witte, T.K., K.A. Merrill, N.E. Stellrecht, R.A. Bernert, D. L. Hollar, C. Schatschneider
et al. (2008). “Impulsive” youth suicide attempters are not necessarily all that
impulsive. Journal of Affective Disorders, 107(1–3), 107–16.
Woznica, J.G. and J.R. Shapiro (1990). An analysis of adolescent suicide attempts: The
expendable child. Journal of Pediatric Psychology, 15(6), 789–96.

172
9

ReporƟng Suicide: Impact on


Suicidal Behaviour
FƒÙƒ« K®—óƒ®

S uicide, widely considered as the most tragic way of ending one’s life,
poses a major challenge to civil society. Worldwide, there is an esti-
mated 850,000 deaths due to suicide and well beyond 15 million suicide
attempts every year. Suicidal behaviour is now acknowledged as a major
global public health problem. It hits particularly the young, and currently
worldwide deaths from suicide are among the top three causes of death
among people aged 15–35 years, for both males and females. Social
scientists have discovered that the majority of people who consider sui-
cide are ambivalent. They are not sure that they want to die. One of the
key factors leading a vulnerable individual to suicide could be publicity
about suicides.
Although suicide accounts for only 1 percent of all deaths, yet, when
these occur they frequently attract disproportionate media interest. It has
long been thought that widespread coverage of a suicide by the media is
capable of triggering copycat suicides in the mass public. According to the
social learning theory, the greater the amount of coverage of suicide in the
media, the greater is the increase in suicide rate. Research has established
that when media, that is newspapers, film and television, report suicidal
deaths, additional suicides may result by virtue of contagion or copycat
effects (Etzendorfer et al., 1992; Gould, 2001; Stack, 2000a, 2000b, 2003).

173
Farah Kidwai

It also shows that an increased number of suicides result from media ac-
counts of suicide which romanticise or dramatise the description of sui-
cidal deaths (Cheng, Hawton, Lee and Chen, 2007).
The work of David Phillips in the 1970s initiated the systematic scien-
tific investigations on copycat suicide. The suicide of the well-known
movie star Marilyn Monroe resulted in the largest possible copycat effects.
There were an additional 303 suicides, an increase of 12 percent, during
the month of her suicide in August 1962. However, in general, highly pub-
licised stories increase the national suicide rate by only 2.51 percent in
the month of media coverage (Stack, 2000b).

SUICIDE CONTAGION

In the present era, when suicides involving the young have assumed another
dimension of political and global terror, identification and efficient
interventions for suicidal behaviour pose a more daunting challenge
before societies across the nations. The role of media too has come under
closer scrutiny as it was often seen as glorifying or legitimatising suicides
for a ‘cause’. There is a paucity of media research on impact of media on
suicide in India. However, in Western academic sphere, the existence of
‘suicide contagion’ is ably recognised by all those working in the field of
mental health, social sciences and mass media. Suicide contagion refers to
a process by which exposure to the suicide or suicidal behaviour of one or
more persons influences others to commit or attempt suicide (Davidson
and Gould, 1989). It implies the exposure to suicide or suicidal behaviours
within one’s family, one’s peer group, or through media reports of suicide
and can result in an increase in suicide and suicidal behaviours. Media
reports may encourage vulnerable individuals, who may have had some
predispositions towards suicide ideation but normally would not have
carried out a suicidal attempt, to act on their suicidal impulses.
Considerable evidence has been accumulated for imitation effects
from suicide reported via newspaper and television (Etzendorfer et al.,
1992; Stack, 2000a). Imitation effect tends to be particularly strong
when newspaper stories about suicides are featured prominently. Imita-
tion is more likely among audience members who can identify with the
suicide victim in some way; for example by age, gender or nationality.

174
Reporting Suicide

Young people and elderly people appear to be more vulnerable than those
in their middle years to media-related suicide contagion. Stack (2000b)
has affirmed that such imitation effects are particularly likely to affect
young people. Increased numbers of television news reports of suicides
have been found associated with a significant increase in suicides for those
under the age of 25 (Romer et al., 2006). Greater numbers of newspaper
reports on suicide were associated with suicide deaths across age groups.
A substantial increase in deaths by suicide has been observed in Hong
Kong, following the death of a well-known Hong Kong pop singer who
jumped from a high building (Yip et al., 2006). This again underscored
the importance of influence of extensive and dramatic media coverage
of suicides.
Media studies conducted mostly in the United States, the United
Kingdom, etc., have documented that the risk for suicide contagion as a
result of media reporting can be minimised by factual and concise media
reports of suicide (Hawton and Coulter, 2003). These research findings
have shown that reports of suicide should not be repetitive, as prolonged
exposure can increase the likelihood of suicide contagion. However,
suicide risk can be minimised by having family members, friends, peers
and colleagues of the victim evaluated by a mental health professional.
Alarmed by a high rate of suicide in its ranks, the United States’ army
has prepared a unique prevention tool—an interactive video ‘Beyond the
Front’ (2008) which is set to be mandatory viewing army-wide, in which
soldiers play the role of an anguished infantryman and make virtual choices
that lead the character to get help or, in the worst case, shoot himself in
the head. In an article in Washington Post, Scott (2008) analysed a video
‘Beyond the Front’ which has a specialist Kyle Norton narrating experi-
ence of a 19-year-old after a bomb-clearing mission in Iraq. ‘Beyond the
Front’ leads the viewer through a detailed drama in which Norton is hit by
relationship troubles, financial problems and scrapes with the law—what
US Army research shows are major events that precipitate suicide. Norton
is blindsided by an e-mail from his fiancée, who has become pregnant by
another man. He is devastated further when one of his best friends is killed
in an ambush. Questions pop onto the screen at key moments, prompting
the viewer to decide whether to get help by opening up with buddies, the
sergeant, clergy or the counsellor. Depending on the choices, Norton
edges towards recovery or sinks deeper into suicidal thoughts. The goal
of the video preparation is to immerse the viewer into Norton’s life in a

175
Farah Kidwai

way that makes preventive lessons stick, claimed army officials and the
video’s producers. The video is one of several initiatives launched by the
US Army to try to stem the suicide rate among active-duty soldiers. That
rate increased from 12.4 per 100,000 in 2003, when the Iraq war started,
to 18.1 per 100,000. In all of 2007, 115 soldiers committed suicide. Suicide
attempts by soldiers have also increased since 2003.
Berman (as cited by Scott, 2008), executive director of the American
Association of Suicidology, who viewed part of the video, said that ‘it’s
obviously done in a much more realistic fashion’ than previous interactive
prevention efforts. Nevertheless, he warned that it is risky to widely
distribute such a programme without scientific evaluation to determine
its impact on a suicidal person. ‘Some media presentations about suicide
can increase the likelihood of suicidal behaviour, so there is a potential
danger,’ he said. According to this school, the media constantly pro-
vides opportunity for transmission of suicide contagion. This means of
influence is potentially more far reaching than direct person-to-person
propagation.
Suicide contagion should be viewed within the larger context of
behavioural contagion, which has been described as the situation in which
the same behaviour spreads quickly and spontaneously through a
group. Behavioural contagion has also been conjectured to influence the
transmission of conduct disorder, drug abuse and teenage pregnancy.
According to behavioural contagion theory, an individual has a pre-
existing motivation to perform a particular behaviour, which is offset by
an avoidance gradient, so as that an approach-avoidance conflict exists.
As media and violence studies have shown, the coverage of suicides in the
media may serve to reduce the avoidance gradient—the observer’s internal
restraints against performing the behaviour.
Evidence clearly establishes that the media may affect method-
specific suicide rates. Ashton and Donnan (1981) revealed that in Britain,
an excess of about 60 suicides by burning occurred in the 12 months after
the widely publicised political suicide by burning of a woman in Geneva.
Bhattacharya (2003) has reported that an ‘alarming escalation’ in people’s
use of burning charcoal to commit suicide followed after detailed media
accounts of a woman who took her own life by starting a charcoal fire in her
cramped apartment and suffocating in the carbon monoxide gas produced.
Increase in suicide rates, following the reporting of real life suicide, has
been described both in Britain and the United States (Barraclough et al.,

176
Reporting Suicide

1977; Phillips and Cartensen, 1986). Schmidtke and Hafner (1988) have
produced more robust evidence by examining suicide rates after two
separate broadcasts of the fictional portrayal of a young man’s suicide on
a railway line. An imitation effect was observed leading to methods specific
and absolute increases in the number of suicides. The imitation effects
were greatest in those of the same age and sex as the fictional character,
and the numbers of suicides closely corresponded with the audience figures
for the two broadcasts. Effects were observed for up to 70 days after the
broadcast; an estimated overall excess of 60 suicides occurred (Schmidtke
and Hafner, 1988). The effect of a television series, dramatising the work of
the Samaritans, on suicide rates has also been studied. Although the series
led to a rise in new client referrals, no effect was seen on the number of
suicides (Holding, 1975).
It is argued that only the choice of method is influenced by publicity but
suicides occur only among those who are already suicidal. A century ago
Durkheim argued that although media attention may precipitate clusters of
suicide, these occur only among those who would commit suicide sooner
or later anyway, the publicity merely acting as a precipitant to an inevitable
event. Schmidtke and Hafner (1988) observed greatest increases among
those most similar to the ‘model’ portrayed, but Ashton and Donnan
(1981) did not. It is difficult to disentangle these conflicting hypotheses
as suicide is a rare event and the particular methods examined constitute
only a small fraction of all suicides.
There is an ample amount of evidence in Western academic circles
to show that the magnitude of the increase in suicidal behaviour after
newspaper coverage is related to the amount of publicity given to the story
and the prominence of the placement of the story in the newspaper.
Imitation appears more likely when the suicide is covered on the front page,
in large headlines, and is heavily publicised, suggesting a ‘dose–response’
relationship. Phillips, Lesyna and Paight (1992) have argued that repetition
is a key factor for news stories’ imitative potential.
In contrast to the ‘structural’ elements of the story, there is less infor-
mation on what characteristics of the ‘models or content’ of the story have
imitative effects. One characteristic of the model that has been studied is
the ‘celebrity status’ of the suicide victim (Cheng, Hawton, Chen et al.,
2007). Wasserman (1984) found that a significant rise in the national
suicide rate occurred only after celebrity suicides were covered on the
front page of the New York Times. Stack (1987) replicated this study, but

177
Farah Kidwai

upon correcting substantial measurement error, a later analysis found that


non-celebrity stories also had a significant impact, although not as great as
publicised celebrity stories. Gundlach and Stack (1990) reported that
non-celebrity stories yielded imitative suicides if they received enough
publicity. In another study, Weimann and Fishman (1995) conducted a
content analysis of more than 430 suicide cases published in the two leading
daily newspapers in Israel in which suicide report was analysed for the
form of its coverage—including space allocation, placement in the paper
and inclusion of picture—and for the content of its coverage, including
demographic characteristics of the victim, mode of suicide, attribution of
responsibility and general attitude towards the act or person. They found
that the space devoted to suicide stories and the prominence of the stories
increased steadily during the 1980s and 1990s. Newspaper reports focused
on the more violent modes of suicide. An economic/financial motive was
attributed mainly to males, while romantic motives or problems with
a partner were attributed mainly to females. Most of the reports were
neutral, but among those that did express an attitude, approximately
18 percent were positive and 8 percent were negative. Positive coverage was
more likely when external causes were presented and when suicides were
committed during military service (Weimann and Fishman, 1995).

CULTIVATION THEORY

In the context of media and suicide, Gerbner and Gross (1976) had pro-
posed a ‘cultivation theory’ which argued that humans cultivate under-
standings of the world around them through indicators found within
television programming. Concerning the relationship between media
portrayal of suicide and suicidal behaviour, the evidence has established
a causal association between non-fictional media reporting of suicide
and suicidal behaviour, and between fictional media portrayal and actual
suicide. These studies have been based on social learning theory and
emphasised on the effects of television viewing on the attitudes rather
than the behaviour of viewers. According to this theory, most human
behaviour is learned observationally through modelling. Imitative learning
is influenced by a number of factors, including the characteristics of the
model and the consequences or rewards associated with the observed

178
Reporting Suicide

behaviour. Consequences or rewards, such as public attention, may lower


behaviour restraints and lead to the disinhibition of otherwise ‘frowned
upon’ behaviour (Gerbner and Gross, 1976).
Heavy watching of television is seen as ‘cultivating’ attitudes which are
more consistent with the world of television programmes than with the
everyday world. Watching television may tend to induce a general mindset
about violence in the world, quite apart from any effects it might have
in inducing violent behaviour. Cultivation theorists (Gerbner and Gross,
1976) distinguish between ‘first order’ effects (i.e., general beliefs about the
everyday world) and ‘second order’ effects (i.e., specific attitudes, such as to
law and order or to personal safety). Gerbner and Gross (1976) argue that
the mass media cultivate attitudes and values which are already present in a
culture: the media maintain and propagate these values amongst members
of a culture, thus binding it together. They have argued that television tends
to cultivate middle-of-the-road political perspectives. They considered
that ‘television is a cultural arm of the established industrial order and as
such serves primarily to maintain, stabilize and reinforce rather than to
alter, threaten or weaken conventional beliefs and behaviours’ (Gerbner
and Gross, 1976). Boyd-Barrett and Braham (1987) observed that such a
function is conservative, but heavy viewers tend to regard themselves as
‘moderate’.
Youth exposed to media, such as television programmes, posters that
glorify previous martyrs and Internet websites, are being cultivated into
becoming martyrs themselves. Gerbner and Gross (1976: 173) believe
that television’s ‘system of messages, with its storytelling functions, makes
people perceive as real and normal and right that fits the established social
order.’ Moreover, television offers frames of reference and the means to
make sense of the world around us. As youth are exposed to messages about
their lives, the programmes they watch define the range of possibilities
for their lives. The BBC Producers’ Guideline (BBC, 2003) in its chapter
on Values, Standards and Principles, has observed, ‘Suicide is a legitimate
subject for news reporting but the factual reporting of suicides may
encourage others. Reports should avoid glamorizing the story, provid-
ing simplistic explanations, or imposing on the grief of those affected.
They should also avoid graphic or technical details of a suicide method
particularly when the method is unusual. Sensitive use of language is also
important.’

179
Farah Kidwai

In her landmark film The Making of a Martyr, Goldstein (2006) makes


it clear that youth who are persuaded into suicide bombing attacks have
limited resources, both symbolic and material, available to them. When
traumatised by acts of violence, they look for frameworks to make sense
of the chaotic world around them. Explanations such as the demonisation
of the enemy or the hope of paradise in life after death become frames
that assign meanings to events and to the corresponding actions that a
youngster could take.
In the case of media images, Goldstein (2006) documents a variety of
images that are used to instil a particular view of reality in the youth of
ravaged areas. Television, posters and graffiti, all glorify the acts and agents
of martyrdom, reinforcing a system of bodily sacrifice that adversely affects
the children. These symbolic resources do not provide other avenues for
children to broaden their horizons and tend to limit their vision. Goldstein
(2006) believes that censorship practices would be one means by which
these messages could be combated. In addition to her call, alternative
message strategies that promote different possibilities for youth in the
Middle East could offer youngsters other options than being enticed
into martyrdom. In other words, they need to cultivate other ideas about
the importance of their lives. If presented with a variety of images and
values, these youngsters may be less likely to believe that martyrdom is
an acceptable end to their lives.

IMPACT OF MEDIA ON COGNITIVE AND


AROUSAL PROCESSES

In psychological terms, it is pertinent to understand the effect of media


portrayal of suicide on cognitions, attitude and emotions of an individual.
Factors that mediate between viewing media coverage of suicide and
suicidal behaviour have also been subject matter of intense debate and dis-
cussion among the social scientists. As it is understood, cognitive and
arousal processes are the two major types of mediating process that affect the
acquisition, maintenance and emission of suicidal behaviour. At cognitive
level, reaction to exposure depends on the observer’s interpretations of
the witnessed suicidal behaviour and the thought activated by viewing

180
Reporting Suicide

suicide. In viewing suicidal coverage by media, people form a representation


or cognitive structure comprised general social knowledge about the
positive value that can be attached to viewing suicide and how it can be
committed. The depth and exact content of such a cognitive structure also
take in account factors such as attention, comprehension, attitude, moral
evaluation and attributions.

Atten on and Comprehension

It has been observed that while viewing media, observers form some mental
representation. What is gleaned from the media depends on attentional
process and processes of comprehension. For example, the salience and
complexity of the programme will determine the degree of his or her
attention, which will in turn affect the rate and degree of comprehension.
In addition, media also affects observer acceptability of the event (i.e.,
attitude), attribution and moral evaluations. The nature of the mental
representation that a person forms in viewing media is partially dependent
on attentional process. These processes determine what is selectively
observed and extracted from the observed material. Material that is not
salient or is too complex is not likely to be remembered.
Huston and Wright (1983) have found that certain perceptually salient
formal features (e.g., action, high pace and sound effects), which are
characteristics of much television fare, attract and hold the attention
of observers/viewers. More importantly, these studies have shown
that attention is elicited and maintained more by formal features (e.g.,
pacing, action) than by the content of the films. Huston and Wright
(1983) suggested that salient formal characteristics along with the
content of programmes would hold a viewer’s attention more, thereby
facilitating comprehension and the formation of an enduring mental
representation.
Research studies concerning comprehension have shown that it is im-
portant for understanding the impact of media because of two reasons.
First, a person’s comprehension of an event is related to his/her attitude
towards the event or character. Second, comprehension is also related to
the viewer’s tendency to identify with the characters. Thus, attitude and
identification are the two important factors that determine the impact of
media on individual’s behaviour.

181
Farah Kidwai

Attribu ons and Moral Evalua ons

Some authors have also suggested that attributions and moral evaluations
contribute to the extent to which an unwanted behaviour will be inhibited
(Berkowitz, 1984; Ferguson and Rule, 1983; Rule and Ferguson, 1984).
Attributions pertain to the perceived causes of, or reasons for, a particular
behaviour. For example, an instance of humiliation may have been
perceived as having been produced intentionally but for justifiable reasons,
or intentionally but for unjustifiable reasons. When an individual perceives
that he has been intentionally subjected to humiliation for an unjusti-
fied cause, it affects self-esteem and develops a feeling of worthlessness
which may lead a person to extreme behaviour. Similarly, moral evaluations
pertain to the perceived praiseworthiness or blameworthiness of an action.
Thus, an unjustifiably intended action would be seen as more blameworthy
than justifiably intended action. As a result there would be less inhibition
of unwanted behaviour against expression of justified behaviour than
unjustified behaviour.

Attitude

The most common source of information about suicide tends to be


media. In a study conducted by Beautrais, Horwood and Fergusson (2004),
it was observed that young people tend to hold mixed attitudes towards
suicide, having both liberal and conservative views. Those with lifetime
histories of suicidal ideation or suicide attempt and those with family
histories of suicide or suicide attempt tended to hold more liberal attitudes.
Attitudes towards suicide were unrelated to gender and to knowledge
about suicide. Moreover, the younger generation overestimated the pre-
valence of youth suicide and the fraction of suicides accounted for by youth
deaths, and held both conservative and liberal attitudes towards suicide
(Beautrais et al., 2004). Their primary source of information about suicide
is the media. These findings raise concerns about the potential for media
coverage of youth suicide issues to normalise suicide as a common, and
thereby acceptable, response among young people, and suggest the need

182
Reporting Suicide

for careful dissemination of accurate information about suicide by know-


ledgeable, respectable and reputable sources.

Emo onal Arousal

Media influences a viewers’ arousal state. There has been considerable


interest in the role of arousal as a mediator of the link between exposure
to media and a particular behaviour. Arousal refers to an energiser of be-
haviour. It has been found that media exposure not only elevates excitation,
it also maintains a particular level of pre-exposure arousal. Exposure to
media violence is assumed to affect suicidal tendency through emotional
arousal. Observing suicide may arouse feelings associated with suicidal
thoughts, but at the same time it may reduce emotional reactions to the
negative consequences of suicide for the self and the people affected by it.
Thus, people may get immune to suicide viewing and their tendency to
commit suicide may increase. The major factor that determines the impact
of media coverage of suicide is emotional habituation.

Emo onal Habitua on/Desensi sa on

People who initially experience negative emotional responses while ob-


serving suicidal behaviour may respond less emotionally after repeated
exposure to its viewing. The reduced negative effect associated with
increased exposure to suicidal behaviour may increase the likelihood of
committing suicide or toying with its idea. Moreover, such emotional
habituation (‘adaptation’, ‘desensitisation’) may reduce concern for
others’ suicidal attempts. In two studies of desensitisation conducted
by Björkqvist (1985) and Linz, Donnerstein and Penrod (1984), it was
found that repeated exposure decreases emotional responsiveness. Various
other researches using both correlational and experimental designs show
that repeated exposure leads to decline in physiological arousal and also
decreases the intensity of self-reported emotion.
While there is growing consensus that some types of media reporting
and portrayal of suicide increases suicide risk among others, ongoing

183
Farah Kidwai

and future research projects are focusing on the need for enhancing safe
reporting and portrayal of suicide by the media.

HOW CAN MEDIA BE MORE RESPONSIBLE


IN REPORTING SUICIDES?

In the context of the substantial evidence for suicide contagion, some so-
cieties tried to work out a ‘prevention strategy’ that sought to educate
reporters, editors and film and television producers about contagion in
order to yield media stories that minimise harm. It also took in account
the media’s positive role in educating the public about risks for suicide.
In the United States, the Centres for Disease Control published a set
of recommendations on reporting of suicide that emerged from a na-
tional workshop. The American Association of Suicidology adopted
these as their official guidelines for journalists in an attempt to minimise
contagious effects from news reports of suicides. Guidelines for media
reporting now exist in several countries, including Australia, Austria,
Canada, Germany, Japan, New Zealand and Switzerland. Additional guide-
lines have been developed by the World Health Organization and the
American Foundation for Suicide Prevention (2001). Although the
media guidelines that were developed have so far not put to any empirical
validation, adopting these guidelines certainly seems to be effective in
avoiding suicide contagion behaviour. Prior to reporting, media should
properly analyse whether any act of suicide is newsworthy or not. Media
must avoid misrepresenting suicide as a mysterious act by an otherwise
‘healthy’ or ‘high achieving’ person. It must necessarily be indicated that
suicide is most often a fatal complication of different types of mental
illness, many of which are treatable. It would be of prime importance to
highlight that suicide is not a reasonable way of solving problems. Media
must always keep in mind that suicide is not portrayed in a heroic or
romantic fashion. Proper care is required to be exercised with publishing
pictures of the victim and/or grieving relatives and friends to avoid
fostering over identification with the victim and inadvertently glorifying
the death. The coverage must be minimised to only necessary content
and detailed description of adopted methods must positively be avoided.
Media is required to limit the prominence, length and number of stories

184
Reporting Suicide

about a particular suicide and front-page coverage and sensational and


inappropriate headlines must be avoided. Furthermore, local treatment
resource information must also be provided.
Closer home, in Sri Lanka and Hong Kong, the Hong Kong Journalists
Association (2002) have established similar guideline on suicide report-
ing. According to their manual, professional journalists should prevent
reporting suicide in a way that could have potential negative effect on the
vulnerable group of suicide or youth readers. Media professionals can play
a role in publicising the warning signs and to convey the message of help-
seeking. Some of their suggestions include: (a) report news with public
interest; (b) minimise harm and (c) appropriate reporting. In specific
terms of placement of news and headline, they advise against publishing
the suicide news on the front page unless the reporting involves public
interest; avoid using large font headline; avoid mentioning the suicide
method or suspected cause such as ‘jumping’ or ‘charcoal burning’ in
the headline. In contents such as wordings, they suggest that phrases like
‘successful suicide’, ‘unsuccessful suicide’ or ‘suicide-prone person’ should
be avoided altogether.
As per World Health Organization (WHO, 2000) guidelines, the media
worldwide should observe certain amount of restrain and address specific
issues that need to be addressed when reporting on suicide. For instance,
statistics about suicide should be interpreted carefully and correctly and
only authentic and reliable sources should be used. The WHO (2000) wants
that media practitioners should avoid making impromptu comments
in spite of time pressures and generalisations based on small figures.
Expressions such as ‘suicide epidemic’ or ‘the place with the highest suicide
rate in the world’ should be avoided under most circumstances. It also
calls for reporting on suicidal behaviour as a responsible social behaviour
where degradation should be resisted. In what it describes as ‘Do’s
and Don’ts, the WHO (2000) urges media to work closely with health
authorities in presenting the facts. Media has been suggested to (a) refer to
suicide as a completed suicide, not a successful one; (b) highlight alternatives
to suicide; (c) publicise risk indicators and warning signs and (d) provide
information on help lines and community resources. In addition, WHO
(2000) has further suggested media to stick to the following:

1. Do not publish photographs or suicide notes.


2. Do not report specific details of the method used.

185
Farah Kidwai

3. Do not give simplistic reasons.


4. Do not glorify or sensationalise suicide.
5. Do not use religious or cultural stereotypes.
6. Do not apportion blame.

The following points (WHO, 2000) should be borne in mind while


‘reporting on a specific suicide’:

1. The coverage of suicides should be minimised to the extent possible.


Sensational coverage of suicides should be conscientiously avoided,
particularly when a celebrity is involved. Every effort should be
made to avoid overstatement. Any mental health problem the
celebrity may have had should also be acknowledged. Photographs
of the deceased, of the method used and of the scene of the suicide
are to be avoided. Front-page headlines are never the ideal location
for suicide reports. Research has shown that media coverage of sui-
cide has a greater impact on the method of suicide adopted than
the frequency of suicides. Added publicity with certain locations
(e.g., bridges, cliffs, tall buildings, railways) that are traditionally
associated with suicide increases the risk that more people will use
them. Therefore, detailed descriptions of the method used and how
the method was procured should be avoided.
2. Suicide is never the result of a single factor or event. It is usually
caused by a complex interaction of many factors such as mental and
physical illness, substance abuse, family disturbances, interpersonal
conflicts and life stressors. Hence, suicide should not be reported as
unexplainable mystery or in a simplistic way. Acknowledging that
a variety of factors contributes to suicide would be helpful. Further-
more, media coverage reports should take account of the impact of
individual’s suicide on families and other survivors in terms of both
stigma and psychological suffering. Suicide should not be depicted
as a method of coping with personal problems such as bankruptcy,
failure to pass an examination or sexual abuse.
3. The emphasis of media report should be on mourning the person’s
death. Glorifying suicide victims as martyrs and objects of public
adulation may suggest to susceptible persons that their society
honours suicidal behaviour. Describing the physical consequences
of non-fatal suicide attempts (brain damage, paralysis, probability
of being disable/handicapped, etc.) can act as a deterrent.

186
Reporting Suicide

PROACTIVE ROLE OF MEDIA


IN PREVENTION OF SUICIDES

Media can play a proactive role in helping to prevent suicide by publishing


some necessary information along with news on suicide. Media should
list available mental health services and help lines with their up-to-
date telephone numbers and addresses so that a person can find it easy
to approach support services in times of problem. Media should also
publicise the warning signs of suicidal behaviour so that the people close
to the suicidal person can identify and provide help to him. Media must
convey the message that depression, which is often associated with suicidal
behaviour, is a treatable condition. A message of sympathy must be offered
to the survivors in their hour of grief and they should be provided the
telephone numbers of support groups, if available. This increases the like-
lihood of intervention by mental health professionals, friends and family
in suicidal crises. Concentrating on all the discussed issues media can also
play an active role in the prevention of suicides.

REFERENCES

American Foundation for Suicide Prevention. (2001). Reporting on suicide: Recom-


mendations for the media. American Association of Suicidology and Annenberg
Public Policy Centre.
Ashton, J.R. and S. Donnan (1981). Suicide by burning as an epidemic phenomenon: An
analysis of 82 deaths and inquests in England and Wales in 1978–79. Psychological
Medicine, 11(4), 735–39.
Barraclough, B., D. Shepherd and C. Jennings (1977). Do newspaper reports of
coroners inquests incite people to commit suicide? British Journal of Psychiatry,
131(5), 528–32.
BBC (2003). Values, standards and principles: BBC Producers’ Guidelines. London: British
Broadcasting Corporation.
Beautrais, A.L., L.J. Horwood and D.M. Fergusson (2004). Knowledge and attitudes
about suicide in 25-year-olds. Australian and New Zealand Journal of Psychiatry,
38(4), 260–65.
Berkowitz, L. (1984). Some effects of thoughts on the anti and prosocial influences
of media events: A cognitive neoassociationistic analysis. Psychological Bulletin,
95(3), 410–27.

187
Farah Kidwai

Bhattacharya, S. (2003). Media coverage boosts ‘charcoal burning’ suicides. British


Medical Journal, 326(7400), 498.
Björkqvist, K. (1985). Violent films, levels of anxiety, and aggression. 3rd European Isra
Conference, 3–7 September, Parma, Italy.
Boyd-Barrett, O. and P. Braham (1987). Media, knowledge & power. London: Routledge.
Cheng, A.T.A., K. Hawton, T.H.H. Chen, A.M.F. Yen, C.Y. Chen, L.C. Chen et al. (2007).
The influence of media coverage of a celebrity suicide on subsequent suicide at-
tempts. Journal of Clinical Psychiatry, 68(6), 862–66.
Cheng, A.T.A., K. Hawton, C.T.C. Lee and T.H.H. Chen (2007). The influence of media
reporting of the suicide of a celebrity on suicide rates: A population-based study.
International Journal of Epidemiology, 36(6), 1229–34.
Davidson, L.E. and M.S. Gould (1989). Contagion as a risk factor for youth suicide.
In Alcohol, Drug Abuse, and Mental Health Administration. Report of the
Secretary’s Task Force on Youth Suicide, Risk factors for youth suicide (pp. 88–109).
Washington, DC: US Department of Health and Human Services, Public Health
Service.
Etzendorfer, E., G. Sonneck and S. Nagel-Kuess (1992). Newspaper reports and suicide.
The New England Journal of Medicine, 327(7), 502–03.
Ferguson, T.J. and B.G. Rule (1983). An attributional perspective on anger and ag-
gression. In R. Geen and E. Donnerstein (Eds), Aggression: Theoretical and Empirical
Reviews (pp. 41–74). New York: Academic Press.
Gerbner, G. and L. Gross (1976). Living with television: the violence profile. Journal
of Communication, 26(2), 172–99.
Goldstein, B. (Producer) and A. Leyland (Director) (2006). The making of a martyr
[Film]. Isreal: a2b Film Productions.
Gould, M. (2001). Suicide and the media. In H. Hendin and J. Mann (Eds), Suicide
Prevention: Clinical and Scientific Aspects (pp. 200–24). New York, NY: Academy
of Science.
Gundlach, J. and S. Stack (1990). The impact of hyper media coverage on suicide:
New York City, 1910–1920. Social Science Quarterly, 71(3), 619–27.
Hawton, K. and P. Coulter (2003). Suicide and the Media: Pitfalls and Prevention. Report
on a Seminar organised by the Reuters Foundation Programme at Green College
and the Oxford University Centre for Suicide Research (CSR).
Holding, T.A. (1975). Suicide and “The Befrienders”. British Medical Journal, 3(5986),
751–53.
Hong Kong Journalists Association. (2002). Hong Kong Journalists Association guidelines
on coverage of suicides. Hong Kong.
Huston, A.C. and J.C. Wright (1983). Children’s processing of television: The infor-
mative functions of formal features. In J. Bryant and D.R. Anderson (Eds),
Children’s Understanding of TV: Research on Attention and Comprehension
(pp. 37–68). New York: Academic Press.
Linz, D., E. Donnerstein and S. Penrod (1984). The effects of long-term exposure to
filmed violence against women. Journal of Communication, Reprinted in Media in
Society: Readings in Mass Communication, 34(3), 130–47.

188
Reporting Suicide

Phillips, D.P., K. Lesyna and D.J. Paight (1992). Suicide and media. In R.W. Maris,
A.L. Berman, and J.T. Maltsberger (Eds), Assessment and prediction of suicide
(pp. 499–519). New York: Guilford.
Phillips, D.P. and L.L. Cartensen (1986). Clustering to teenage suicides after television
news stories about suicide. The New England Journal of Medicine, 315(11), 685–89.
Romer, D., P.E. Jamieson and K.H. Jamieson (2006). Are news reports of suicide
contagious? A stringent test in six U.S. cities. Journal of Communication, 56(2),
253–70.
Rule, B.G. and T.J. Ferguson (1984). An overview of the relations among attribution,
moral evaluation, anger and aggression. In A. Mummendey (Ed.), Social psychology
of aggression: From individual behavior towards social interaction (pp. 41–74). Berlin:
Springer-Verlag.
Schmidtke, A. and H. Hafner (1988). The Werther effect after television films: new
evidence for an old hypothesis. Psychological Medicine, 18(3), 665–76.
Scott, T.A. (2008). Army tries to combat soldier suicide, “Beyond the front,” Washington
Post, Wednesday, 8 October 2008. Retrieved 8 Oct 2008 from www.washingtonpost.
com/wp-dyn/.../AR2008100702780.html
Stack, S. (1987). Celebrities and suicide: A taxonomy and analysis, 1948–1983. American
Sociological Review, 52(3), 401–12.
Stack, S. (2000a). Suicide: A 15-year review of the sociological literature. Part II:
Modernization and social integration perspectives. Suicide and Life-Threatening
Behavior, 30(2), 145–62.
Stack, S. (2000b). Media impacts on suicide: A quantitative review of 293 findings. Social
Science Quarterly, 81(44), 957–71.
Stack, S. (2003). Media coverage as a risk factor in suicide. Journal of Epidemiology and
Community Health, 57(4), 238–40.
Wasserman, I.M. (1984). Imitation and suicide: A re-examination of the weather effect.
American Social Review, 49, 427–36.
Weimann, G. and G. Fishman (1995). Reconstructing suicide: Reporting suicide in the
Israeli press. Journal of Mass Communication, 72(3), 551–58.
World Health Organization. (2000). Preventing Suicides: A Resource for Media
Professionals. Geneva: WHO, Department of Mental Health.
Yip, P.S., K.W. Fu, K.C.T. Yang, B.Y.T. Ip, C.L.W. Chan, E.Y.H. Chen et al. (2006).
The effect of a celebrity suicide on suicide rates in Hong Kong. Journal of Affective
Disorders, 93(1), 245–52.

189
Farah Kidwai

190
SecƟon II

Assessment: People-at-Risk
192
10

Suicide: Its Assessment and PredicƟon


Pٮ㫃 M绫Êփ—«ùƒù

M ulti-faceted problem of suicidal behaviour has been the subject of


extensive studies by many researchers (Brown et al., 2000; Goldstein
et al., 1991). It has resulted in development of a host of different models,
scales of measurement to track and understand the factors responsible for
the act of suicide. Although significant risk factors have been identified and
models have emerged, no major breakthrough is yet evident to prevent
the suicidal behaviour and to help the vulnerable person to cope with sui-
cidal intent and adverse life situations. Perhaps, the root of suicidality is
embedded deep in human psyche.
This chapter on assessment starts with an introduction and clinical and
empirical approaches to assessment and prediction of suicide, which will
be followed by the details of the assessment tools and discussion regarding
the same with a conclusive note on role of assessment in prevention and
prediction.
In this chapter an attempt has been made to provide a systematic ex-
amination of the measures that assess suicidal ideation and behaviour
across the age, from childhood to older adulthood. The assessment tools
encompass: (a) suicide ideation and behaviour; (b) cognitive and affective
factors underlying suicidal behaviour; (c) lethality of suicide attempts;
(d) attitudes and knowledge concerning suicide; and (e) brief screening
measures of suicidality. Although some measures do not directly assess
suicidal behaviour, the variables closely associated with suicide have been

193
Pritha Mukhopadhyay

given due consideration. Some of these factors are potentially modifiable


with appropriate timely interventions. The present review mostly includes
suicide assessment instruments with adequate reliability and validity.
A few tests, which appear to be in their developing phases, have been
included as it might be helpful to the interested researchers. Measures
concerned with psychopathology have not been prioritised as it is the part
of clinical history-taking. Many of the measures of personality have not
been considered as these are not of great utility in comprehensive suicide
risk evaluations (Johnson et al., 1999). Minnesota Multiphasic Personality
Inventory (MMPI), Thematic Apperception Test (TAT) and Rorschach
Inkblot Test have not been found to be predictors of suicidality (Beck et al.,
1979). Since the factors not identified previously may prove significant in a
given person, suicide risk assessment should be based on clinical judgement
duly considering the clinical details (Jacobs, 2003). It is deemed desirable
that thorough psychiatric examination, identification of risk and protective
factors, a distinction between the modifiable and stable risk factors are to
be incorporated in case conceptualisation—a prerequisite condition of as-
sessment of suicidal behaviour (Meichenbaum, 2005).
Heterogeneity in definition of suicidal behaviour is an additional
problem in developing suicide measures and evaluating them from the
same platform. The definitions provided by O’Carroll and colleagues
(1996) have been considered by the reviewers (Brown, 2008; Goldston,
2003) as useful ones in understanding the domain of suicidality.

CLINICAL AND EMPIRICAL APPROACHES


TO ASSESSMENT AND PREDICTION

Clinical evaluation is an important part of assessment in suicidal behaviour.


Clinical assessment includes a detailed interview that elicits information
regarding the person’s life experience, temperament and character and
adaptive need structure to enable clinicians to understand the diathesis
underlying suicidal behaviour. This would give direction to empirical ap-
proach of assessment, which is concerned with the identification of
persons at risk based on response of the person on standardised tools.
The assessment requires inclusion of both the clinical and empirical ap-
proaches together to reach a final decision. Since the assessment tool for

194
Suicide, Assessment and Prediction

suicidal behaviour is inferential in nature, decision may not go beyond


subjectivity. As a matter of fact, even prediction loses its accuracy when it
is not anchored on scientific theoretical framework and clinical insight. It
may well be the reason to adopt clinical approach to comprehend suicidal
behaviour. The development of the tool, Collaborative Assessment and
Management of Suicidality (CAMS), has helped clinicians to improve
their methods to identify, assess and conceptualise treatment plans and to
view suicide-prone individuals in an empathic, matter-of-fact and non-
judgemental fashion (Jobes et al., 2007). Agreement between clinical and
empirical evaluation should be achieved for an in-depth understanding
of a case that may add significance to the concept of typology and varied
nomenclature in suicidal behaviour. Information regarding previous
suicidal history, personality constellation, sensitivity to life experiences
and vulnerability along with diagnosis of any psychiatric illness, including
Axis I and Axis II disorders which could be a comorbid condition with
suicidality, are essential in predicting suicidal intent. A comprehensive
approach for understanding a case would include precise clinical interview,
self-report in conjunction with objective behavioural measures.
However, diagnostic diversity and non-specificity of life stressors along
with the reports of stressful life events not being a significant predictor
(Dogra et al., 2008) suggest the difficulty in prediction of suicidal be-
haviour. Even a significant biological marker, like low 5-HIAA, has not
been reported to serve as a single good predictor of suicidality (Brown
et al., 1992).
However, review of the current literature (Berk et al., 2007; Dervic
et al., 2007; Dhar and Basu, 2006; Haaga et al., 1991; Sil and Basu, 2007;
Wilson et al., 2007) suggests the existence of two distinct groups of patients,
irrespective of diagnostic category, namely (a) those who are characterised
by depression, hopelessness, suicidal ideation and pre-meditation and
(b) those with impulsivity and comparatively lower level of depression
or hopelessness without any contemplation. This indicates the utility of
assessment of psychic constellation through different psychological tools.

VARIOUS MEASURES OF ASSESSMENT

Suicidal ideation is the thought serving the means to one’s death.


It involves the attitude towards suicide, the intent and the plan to

195
Pritha Mukhopadhyay

attempt it. The suicide intent may vary in seriousness depending on


the degree of suicide intent and specificity of suicide plans. It may be
manifested from transient thoughts with respect to worthlessness of life
and death wish, to permanent recurring plans for killing oneself and pre-
occupation with self-destruction that reflects depressive mood and
hopelessness, where suicide is seen as a mode for coping with such a mood.
Suicide ideation may be chronic or acute in nature.
The measures of suicidal behaviour can be divided broadly in the
domains of cognition, affect and behaviour. The various psychological
tools used to measure suicidality and associated factors along with its
psychometric properties are summarised in the Appendix to this chapter.

EvaluaƟon of the Scales

It is quite evident from the table (see Appendix) that a host of measures
of suicide-related behaviours are available for assessment of cognitive, af-
fective and behavioural aspects underlying suicidal behaviour in adults
and in children. Most of these measures have adequate internal reliability
and concurrent validity. However, it is evident that no single instru-
ment and single-time assessment can yield an accurate estimate of the
intention and prediction of suicidal behaviour. In terms of prediction, a
multigating procedure is deemed essential to overcome the limitation of a
single test since suicidal behaviour is the function of interplay of multiple
factors that makes its prediction difficult. The same risk factors, when
accompanied by a protective factor could have a different implication
than when it is not so. Poor predictive validity of the assessment tool
may also be ascribed to the rarity of the incidence of suicide (Maris et al.,
1992). Perhaps, the assessment would be effective and helpful in suicide
prevention and prediction when suicidal behaviour is a cry for help akin
to the psyche of deliberate self-harmers and individuals with parasuicidal
intent than those with intent of terminating their lives.
Questionnaires administered by clinicians to a patient, in a face-to-face
situation, in a conducive environment, could enhance the efficacy of the
tool, eliciting more genuine response of the patient instead of depending
more on self-reporting measures. The research report (Joiner et al., 1999)
reveals that the intensity of suicide intent and planning is less on self-report
measures than when it is conducted by a clinician.

196
Suicide, Assessment and Prediction

Although different scales of suicide ideation measure current intensity


of patient’s specific attitude, behaviours and plan to commit suicide on
the day of the interview, Scale of Suicide Ideation-Worst (SSI-W; Beck
et al., 1997)—an interviewer-administered rating scale for suicide
ideation—yields more accurate estimate of suicide risk. This may be
attributable to the better elicitation of response with the ambience of
empathy associated with interviewing. Also the demand of the test is
to make the respondent recall the approximate date and circumstances
when the patient was experiencing most intense desire to commit suicide.
The assessment of mental state immediately prior to attempt is definitely
useful and enhances knowledge for understanding and prediction of sui-
cidal behaviour, but this retrospective reporting is subject to retrieval
falsification unless collected in an emergency set-up just after the incident.
However, it has been reported (Beck et al., 1997) that 16 and above score
on SSI-W has 14 times more likelihood to predict suicide commitment.
Similarly it has six times more likelihood of suicide commitment in case
of 2 or more scores on Scale for Suicide Ideation (SSI; Beck et al., 1997).
Sequential Emotion and Event Form for Suicidal Adolescents (SEESA;
Negron et al., 1997) also makes an attempt to give a detailed cognitive,
affective and behaviour status at the time of suicidal episode.
A very few tests, like SSI, SSI-W and Adult Suicidal Ideation Question-
naire (ASIQ), claim to have predictive validity, though most of the tests
lack predictability. However, the randomised clinical trials have shown
the sensitivity of the tests like SSI, Modified Scale for Suicide Ideation
(MSSI), Suicide Probability Scale (SPS), Suicide Behaviours Questionnaire
(SBQ) and Self Monitoring Suicide Ideation Scale (SMSI) to the changes
in suicidal thinking and level of depression and hopelessness which have
immense clinical implication and protective value for any subsequent
suicidal incidents. Negative correlation between SBQ and Reasons for
Living Inventory (RFL) reveals SBQ’s sensitivity to one’s appraisal for
reasons to live also. Efficacy of brief screening tools, like Paykel suicide items
and suicide ideation screening questionnaire, might have been enhanced
because of adopting interview technique. However, the psychometric
properties of the screening tool may require improvement.
On the other hand, the tests of suicidal behaviour are subject to cri-
ticism for its poor sensitivity and specificity. It may be attributable to the
inferential nature of these tests. Beck and his co-workers (1990) reported
the possibility of false positive and false negative responses on the Beck

197
Pritha Mukhopadhyay

Hopelessness Scale (BHS) which may lead to inaccurate prediction,


though they claim 9 and above score on BHS has a good predictability for
suicidal behaviour. But it is also true that even one with his best intent to
answer honestly, may rate himself at the lower end of the response category,
subjectively perceiving his intent as not to be so highly scored.
Although suicide probability (Cull and Gill, 1988) and suicide status
scales (Jobes et al., 1997) measure the probability of suicide risk considering
the factors, like negative self perception, negative mood state or external
pressure, they do not give due consideration to the protective factors,
though suicidal status and probability are outcomes of the interplay of risk
factors and associated protective factors as well. Consideration of modi-
fiable risk and protective factors could help to assess the probability of
risk over some significant period of time.
Even though there exists a good number of scales to assess suicidal be-
haviour and new assessment tools are in the offing, but the assessments-
outcome of different scales are not always comparable as they are either
standardised on different populations or are developed based on different
theoretical constructs. Rather, application of the same tool across the
population, culture and different settings could develop huge sets of
database and would be helpful to bring out clarity in conceptualising the
aetiology of suicidal behaviour (Brown, 2008). Moreover, most of the tests
are standardised on the population of one’s own country and have not
been always duly tested in other populations. Although suicidal scales
are available for children and adolescents (Goldston, 2003), tools addressing
the need of old age is also relatively scarce (Brown, 2008).
A battery of tests including the evaluation of direct factors related to
suicidal behaviour and indirect potential factors that contribute to it is
essential for understanding the phenomenon in totality. The indirect tests
assess the personality constellation, cognitive and affective appraisals of
life events that underlie suicidality having the potency to ignite suicidal
desire that culminates in suicidal act. Assessment of depression, on Beck
Depression Inventory (BDI) or Hamilton Rating Scale for Depression
(HRSD), hopelessness on BHS, or use of Psychache scale that reveals the
penetrating value of distress, Psychache Need Questionnaire that indicates
the gap between one’s wish and wish fulfilment, Reasons For Living
denoting the worth of living cognition of the person, reasons for attempting
suicide as assessed on RFL along with sensitivity to life events as measured
by Life Event Scales and assessment of ego integrity on Ego Function Test
(EFA) test could be considered as a battery for assessment of suicidal

198
Suicide, Assessment and Prediction

disposition in a person. Consequently, the underlying determinants of


personality constellation, like family cohesion, interpersonal relationship,
attachment with significant persons, become the essential part of suicidal
assessment.
Several measures, such as SMSI (Clum and Curtin, 1993) or Linehan’s
diary card for monitoring suicide ideation and self-harm behaviours
(Linehan, 1993), specific items of Beck Depression Inventory-II (BDI-II;
Beck et al., 1996), may be applied repeatedly during intervention.
Clinicians may provide with adequate intervention to prevent suicide
attempt in patients who regularly attend psychotherapy sessions and
complete the scales, such as BDI-II and BHS, prior to each visit that
indicates hopelessness or suicidal tendency in the patient (Ellis and
Newman, 1998). However, it may be suggested that if the test developers
aim to correlate the assessment tool of suicidal behaviour with the measures
of a few other relevant domains of affect and cognition that under-
lies suicidal behaviour, the assessment procedure could be made simpler
and more precise.

CONCLUSION

Consideration of personality constellation factors, namely depression,


hopelessness and psychache, on the one hand, and impulsivity, aggression
and emotional instability on the other, which are empirically distinct
but highly correlated constructs, may yield a more integrated view for
understanding and predicting suicide. Perhaps prediction becomes more
difficult due to unique psychic constellation of each individual. Any dis-
crepancy between objective and subjective criteria favours subjective
criteria to decide suicide risk. This necessitates an idiographic approach
to investigate total individual psyche. Moreover, the prediction of suicide
potential is not equivalent to prediction of event of suicide and the
term prediction loses its significance and may be replaced with the term
‘estimation of risk’ (Motto, 1992). Furthermore, the aim of any assessment
is prevention and treatment. The information regarding the intensity of
suicide ideation and intent, severity of underlying psychopathology and
availability of protective factors are the determinants not only in choosing
treatment modality but also in taking decision for indoor or outdoor
treatment. The risk assessment, if suggestive, may cause immediate forceful

199
Pritha Mukhopadhyay

admission of the patient in hospital—thus maximising the benefit of


assessment outcome.
From prevention perspective, with increased number of people in the
community to come under high risk group, formulation of preventive
programme at the community level is also becoming essential. Though
imposing restriction onto the accessibility to the method employed in
making attempts on suicide seems effective (Gelder et al., 1988), its
wide application may prove impracticable. In preventive programme,
assessment of vulnerable and resilience factors of each age group, and
nature of environmental setting is required as these could be the target to
make any intervention programme effective.
An important preventive effort may be made, from government as
well as non-government levels, for the high-risk adolescent population
by organising school-based programmes, wherein the teachers, parents,
other mentors and students are psychoeducated about the warning signals
of suicide and learn when to refer someone who seems to be in danger
(Kalafat and Elias, 1994). A study by Mishara and Daigle as mentioned
by Gould and colleagues (2003) further suggests the necessity of develop-
ment of centres for crisis intervention and online help to provide effective
support for the help-seekers. However, although school-based programmes
have been acknowledged to improve attitude (Kalafat and Elias, 1994) and
help-seeking behaviour (Ciffone, 1993), its efficacy has been questioned
including the report of no benefits (Shaffer et al., 1991) and even showing
its detrimental effects (Shaffer et al., 1991) on the vulnerable persons. As
a result, skill training programme has been devised to immunise suicide-
prone persons against suicidality (Gould et al., 2003). Moreover, youth
at risk usually do not attend the programme though the person with
suicide intent usually visit doctors prior to the occurrence of the event
(Maris et al., 1992), indicating their desire to seek help in a very personal
and one-to-one setting.
Although efficacy of primary prevention programme has been proved
in Gotland project (Gelder et al., 1988), the suicide rate had come at the
baseline level three years after the termination of the project, indicating the
necessity of continuation of the programme. Community organisations
may come forward with continuous programme and operate in the seeking
mode for those who are troubled, and function as paraprofessionals
who help bridge the gap between mental health professionals and the
sufferers.

200
APPENDIX

Details of Assessment Tools with their Psychometric Properties


(Abbreviations used in the tables CV = Concurrent Validity; PV = Predictive Validity; CC = Chronbach
Coefficient; IR = Interrater Reliability)

Suicide Ideation Scales


Scales and authors Constructs Standardisation details
The Scale for Suicidal ideation Number of items: 21
Suicide Ideation Frequency of attempt Interviewer administered
(SSI; Beck et al., 1979) Degree of intent to kill Sample studied: Adult psychiatric inpatients and outpatients
with two major dimensions Validity:
of preparation and motivation CV = Significantly associated with BDI and HDRS
PV = Presence of suicidal ideation provides an independent
estimates of the risk for suicide
Reliability:
CC: r = .84 to .89 IR r = .83 to .98
Scale for Suicide Retrospective information Number of items: 19
Ideation-Worst (SSI-W; of patient’s most intense desire to Interviewer administered
Beck et al., 1997) commit suicide Sample studied: Adult psychiatric inpatients and outpatients
Two important dimensions: Validity:
preparation CV = Associated with measures of suicide ideation including
and motivation SSI, BDI and HDRS
(Appendix Continued)
(Appendix Continued)

Suicide Ideation Scales


Scales and authors Constructs Standardisation details
PV = Patients with higher score for high risk category were 14
times more likely to commit suicide than patients who scored
in the low risk category
Reliability:
CC : r = .88 IR Reported High
Beck Scale for Desire for death Number of items: 21
Suicide Ideation Preparation for suicide Self-report version of SIS
(BSSI; Beck and Steer, 1991) Actual suicide desire Sample studied: Adolescents and adult psychiatric inpatients
and outpatients
Validity:
CV = Highly correlated with SSI .90 to .94
Moderately correlated with BDI suicide item, .58 to .69
Moderately correlated with BDI, .64 to .75 and BHS, .53 to .62
Reliability:
CC : r = .87 to .97
Test-retest reliability r = .54
Modified Scale for Suicidal desire Number of items: 18
Suicide Ideation Preparation for attempt Semi-structured interview
(MSSI; Miller et al., 1986) Perceived capability of Sample studied: Adult psychiatric inpatients and outpatients
making an attempt Validity:
CV = Moderately high correlation with SSI (r = .74)
Moderate correlation with suicide item from BDI
(r = .60)
Significant correlation with total BDI (r = .34)
Zung Depression Scale (ZDI) (r = .45) and BHS (r = .46)
Reliability:
CC: r = .87 to .94
Item total correlation, .41 to .83
Test-retest reliability (r = .65)
Self Monitoring Intensity and duration of Number of items: 3
Suicide Ideation ideation and level of control Self-report measure for daily assessment
Scale (SMSI; Clum and Curtin, 1993) in making a suicide attempt Sample studied: Chronically and severely suicidal college
students (18–24 years)
Validity:
CV = Moderately correlated with SSI and MSSI ranging from
.46 to .56
Significantly correlated with BHS and ZDI
Suicide Probability Suicide ideation Number of items: 36
Scale (SPS; Cull and Gill, 1988) Hopelessness Self-reporting scale
Positive outlook Sample studied: Non-clinical adolescents and adults
Interpersonal closeness Validity:
Hostility CV = With following scales of MMPI Positively correlated
Angry impulsivity with depression (r = .44 to .73)
Psychopathic deviate (r = .48 to .63)
Paranoia (r = .47 to .61)
Schizophrenia (r = .56 to .68)
Suicide thread scale developed for the
MMPI (r = .67 to .71)
Reliability:
CC: r = .93; IR for subscale r = .62 to .89
Test-retest reliability r = .92
Positive and Negative Suicide Positive and negative thoughts Number of items: 20
Ideation Inventory (PANSI; related to suicide attempts Self-report measure
Osman et al., 1998) Sample studied: Undergraduate college
students (mean age = 20 years)
(Appendix Continued)
(Appendix Continued)

Suicide Ideation Scales


Scales and authors Constructs Standardisation details
Validity:
CV = Moderately and negatively correlated with SPS
(r = –.47); SBQ (r = –.21 to –.45)
Moderately and positively correlated with
SPS (r = .59) and with SBQ (r = .39 to .61)
Reliability:
IR: r = .80 to .93
Adult Suicidal Suicide ideation and behaviour Number of items: 25 items
Ideation Questionnaire in adults Self-report measure
(ASIQ; Reynolds, 1991a, 1991b) Sample studied: Undergraduate college students (mean age 20
years) and clinical population
Validity:
CV = Highly correlated with HRSD (r = .77); significant
correlation with measures of depression (r = .60); anxiety
(r = .41) and the history of prior suicide attempts (r = .36)
Validity in community sample:
Significantly correlated with measures of depression (r = .60);
hopelessness (r = .53); anxiety (r = .38); low self-esteem
(r = .48); history of prior suicide attempt (r = .33)
PV = Baseline ASIQ scores significantly predicated suicide
attempt in 3 months follow-up study in a sample of
psychiatric inpatients who had previously attempted suicide
Reliability:
CC: r = .96 to .98
Test-retest reliability: r = .95
Suicide Ideation Severity or intensity of Number of items: 10-items
Scale (SIS; Rudd, 1989) suicidal ideation Self-report scale
Sample studied: College students (16–30 years)
Validity:
CV = Moderately correlated with centre for epidemiologic
studies depression scale (r = .55) and the BHS (r = .49)
Reliability:
CC: r = .86
Item–Total Correlation r = .45 to .74
Reasons for Living Beliefs and expectations for not Number of items: 48
Inventory (RFL; committing suicide Self-report measure
Linehan et al., 1983) subscales include: Sample studied: General population (mean age, 36 years)
Survival and coping Validity:
Beliefs CV = Following subscales of survival and coping,
Responsibility to family responsibility to family, child-related concerns and moral
Child-related concerns objections—inversely related to measures of suicide ideation
Fear of suicide (r = –.13 to –.53) and suicide probability (r = –.28 to –.67).
Fear of social disapproval Survival and coping subscale negatively correlated with the
Fear of oral objections Beck Depression Inventory (r = –.68), the Beck
Hopelessness Scale (r = – .71) and the Suicide Intent Scale (r
= –.42)
Moderately and negatively correlated with the Scale for
Suicide Ideation (–.64) and the Beck Hopelessness Scale
(–.63) in a sample of college students
Reliability:
CC: r = .72 to .92 for each subscale
RFL total scale: r = .89
(Appendix Continued)
(Appendix Continued)

Suicide Ideation Scales


Scales and authors Constructs Standardisation details
Reasons for Motivation for suicide Number of items: 14
Attempting Suicide Two scales: Self-reporting questionnaire
Questionnaire Internal perturbation Sample studied: Clinical and non-clinical adult population
(RASQ; Holden and McLeod, 2000) Extrapunitive/manipulative Validity:
motivation PV = Internal perturbations correlate .41 with previous
suicide history
Reliability
CC:
Internal perturbations scale: r = .71 to .87
Extrapunitive/manipulative motivations
scale: r = .80 to .86
Child Suicide Suicidal behaviours in Comprehensive semi-structured interviews
Potential Scale youths scales assessing Sample studied: Child and adolescent psychiatric (6–12 years)
(CSPS; Pfeffer et al., 1979) ego functioning and inpatients and outpatients and corresponding non-clinical
ego defence mechanisms population
Validity:
CV:
Perception of death as temporary and use of introjection on
CSPS consistently associated with suicidality among child
psychiatric inpatients
PV = CSPS ratings of suicidal behaviour among adolescent
psychiatric inpatients predict CSPS rating of suicidal
behaviour one year later
With Suicidal Ideation Questionnaire-Junior High School
Version (SIQ-Jr) ratings of suicidality
6 to 8 months later; child psychiatric inpatients and non-
clinical controls Childrens’ three and six times more scores
on suicidal ideation and suicide attempts respectively on
the Spectrum of Suicidal Behaviour likely to make suicide
attempts over the 6- to 8-year follow-up
IR
Concept of Death Scales: r = .89 to .90
Ego Mechanism Scales: r = .65 to 1.0
Ego Defense Scales: r = .83 to 1.0
Suicide Status Scales
Scales and authors Constructs Standardisation details
Suicide Status Form Psychological pain Number of items: 12
(SSF; Jobes et al., 1997) External pressures Agitation Six self-reporting and six clinician administering items
Hopelessness Sample studied: Non-suicidal (18–26 years) and suicidal
Low self regard college students (17–55 years)
Overall risk of suicide Validity:
CV: Moderately and negatively correlated with Linehan
Reasons for Living Scale (r = –.42)
Test-retest reliability: r = .35 to .69
Sequential Emotion Emotional, cognitive and Two parallel semi-structured questionnaires
and Event Form for Suicidal behavioural staus of suicidal Sample studied: Adolescent suicide ideators and attemptors
Adolescents (SEESA; Negron et al., adolescents before, during (mean age 12 years)
1997) and after the suicidal
episode
(Appendix Continued)
(Appendix Continued)

Suicide Status Scales


Scales and authors Constructs Standardisation details
Suicide Intent Scale Seriousness of the intent Number of items: 15
(SIS; Beck, Schuyler and to commit suicide among Interviewer administered
Herman, 1974; as cited in attempted suicide patients Sample studied: Admitted suicidal patient after attempt
Brown, 2008) (18–63 years)
Validity:
CV = Moderate correlations with measures of depression
(r = .17 to .62) and with measures of hopelessness (r = .31 to
.41), highly correlated with self-report version of SIS (r = .87);
relates to lethality of suicide attempts (r = .38)
PV = SIS total scale did not predict completed suicide
Reliability:
CC: r = .95; IR: r = .81 to .95; lethality of intent: r = 90;
planning: r = .74
Domain of Affect
Scales and authors Constructs Standardisation details
Psychache Needs Psychological needs: Number of items: 35
Questionnaire (PNQ; Seven subscales: Achievement, Self-reporting questionnaire
Munchua, 2003) affiliation, autonomy, Sample studied: Adult non-clinical sample
counteraction, order, Validity:
avoidance and succorance Differentiate between suicide attempters and non-attempters
in a university undergraduate sample
Reliability:
CC: r = .82
Separation Anxiety Child’s reaction to separation Items: Semi-projective instrument with 20 drawings
Test (SAT; Hansburg, 1980) and loss Sample studied: Adolescents (12–16 years)
Validity:
Demonstrated by identifying seven different factors
Reliability:
Reported satisfactory split-half and test-retest reliability
State–Trait Anger State Anger, Trait Anger, Angry Number of items: 44
Expression Temperament, Angry Reaction, Self-reporting questionnaire
Inventory (Speilberger, 1988) Anger turned inwards, Anger Sample studied: Adolescents and adults
turned outwards, Anger Control, Validity:
Anger Expression CV-r = Trait Anger: Buss-Durkee Hostility
Inventory: Male (M), .71; Female (F), .66;
MMPI Hostility Scale: M, .59; F, .43;
MMPI Overt Hostility Scale: M, .32; F, .37; Eysenck
Personality Questionnaire (EPQ) Neuroticism: M, .50; F, .49;
Psychoticism: M, .21; F, .20; Lie Scale: M, –.20; F, –.25
State Anger: with (EPQ) Neuroticism: M, .43; F, .27;
Psychoticism: M, .26; F, .27.
Reliability:
CC: r = .73 to .85
Domain of Suicide Behaviour
Scales and Authors Constructs Standardisation Details
Suicide Behaviours Questionnaire Suicide thoughts and behaviour Number of items: 4 items (abbreviated)
(SBQ; Linehan, 1981) Self-reporting questionnaire
Sample studied: Female psychiatric outpatients (mean age 32
years) and college students
Validity:
(Appendix Continued)
(Appendix Continued)

Domain of Suicide Behaviour


Scales and authors Constructs Standardisation details
CV = Significantly correlated with SSI r = .69; negatively
related with Linehan RLI r = –.34. Items measuring self-harm
on SBQ and diagnostic interview for borderline personality
disorder (BPD)—moderately to highly correlated
(r = .61 to .93)
Reliability:
CC: r = .75 (clinical sample) r = .80 (non-clinical sample)
Test-retest reliability: r = .95
Suicidal Behaviours Five behavioural domains: past Number of items: 34
Questionnaire (Revised) (SBQ-14; suicidal ideation, future suicidal Self-reporting questionnaire
Linehan, 1996) ideation, past suicide threats; future Sample studied: Non-clinical adult population
suicide attempts and likelihood of Validity:
dying in a future suicide attempt CV = Positively correlated with item of SSI and SCI (r = .36
to .51)
Positively correlated with SSI, BDI and BHS (r = .55 to .62)
Negatively correlated with Linehan RLI r = .46)
Reliability:
CC: r = .73 to .92
Impulsivity Control Impulsivity Number of items: 15
Scale (ICS; Plutchik et al., 1989) Self-reporting scale
Sample studied: Adolescents and adults
Validity:
Correlates with other measures of suicide and violent risk
Reliability:
IR: r = .77
Risk and Methods of Lethality
Scales and authors Constructs Standardisation details
Lethality Scales Medical lethality of a suicide Interviewer-administered scales
(LS; Beck et al., 1975) attempts Sample studied: Admitted patient with suicide attempt (mean
Eight separate scales age 29 years)
according to the method of attempt Validity:
CV = Correlation between suicide intent and medical lethality
was found to be low (r = .19); highly correlated with SIS
(r = .73); moderately correlated with RRR measure (r = .60)
Reliability:
IR: r = .80
Escape Potential Scale (EPS; Two aspects (composite score): Number of items: 2
Baumeister, 1990) Escapability Controllability Sample studied: Parasuiciders and non-clinical adults
Reliability:
CC: r = .63
Defeat Scale Degree to which most stressful r Number of items: 4
(O’Connor and Leenaars, 2003) ecent life event led to feelings Sample studied: Parasuiciders and non-clinical adults
of defeat, rejection, loss and failure Reliability:
CC : r = .86
Rescue Scale (Sherbourne Availability of social support to Number of items: 18
and Stewart, 1991) patients with chronic conditions Multi-dimensional instrument
Sample studied: Parasuiciders and non-clinical adults
Reliability:
CC: r = .83 to .93
Risk-Rescue Rating Scale (RRR; Lethality and intent of a suicide Number of items: 10
Weissman and Worden, 1972) attempt Interviewer-administered scale
Sample studied: Admitted suicide attempters (10–60 years or
older)
Validity:
(Appendix Continued)
(Appendix Continued)
Risk and Methods of Lethality
Scales and authors Constructs Standardisation details
CV = Moderately correlated with Beck’s Lethality Scale
(r = .60); positively associated with high scores on SIS
(r = .38)
Reliability:
IR: Kappa = .67; IR: Kappa = .59
Self-Inflicted Injury Focuses on the assessment of injury Number of items: 7
Severity Form (SIISF; Potter et al., lethality. There are seven methods of Interviewer-administered scale
1998) injury category rated according to the Sample studied: Admitted patient with self-inflicted wound
degree of lethality (13–34 years)
Validity:
CV: High rate of agreement with RRR (Kappa = .88)
Reliability:
IR: Kappa = .94; IR: on ‘near fatality’ ratings r = .93
Domain of Cognition
Scales and authors Constructs Standardisation details
Firestone Assessment of Self- Self-destructive thoughts; Number of items: 84
Destructive Thoughts (FAST; self-defeating composite; addictions Self-report questionnaire
Firestone and Firestone, 1996) composite, self-annihilating Sample studied: Adult psychiatric inpatients and outpatients
composite suicide intent composite and on-clinical college students
Validity:
CV = The suicide intent composite subscale was highly
correlated with suicide ideation subscale of SPS (r = .85) and
BSSI (r = .81)
Reliability:
CC: r = .84 to .97
Internal consistency: r = .76 to .91
Test-retest reliability: r = .63 to .94
Depression and Hopelessness
Scales and authors Constructs Standardisation details
Beck Depression Severity of depression Number of items: 21
Inventory (BDI; Beck Self-report inventory
and Steer, 1987) Sample studied: Adolescents and adults
Validity:
CV: Average correlation of the BDI with clinical ratings of
psychiatric patients (r = .72) and non-psychiatric patients
(.60); with Hamilton Rating Scale for Depression (.73), for
psychiatric patients and non-psychiatric patients (.74)
PV: 2 or higher score indicates 6.9 times (95 percent, CI:
3.7–12.6) more likelihood to commit suicide
Reliability:
CC: r = .73 to .95
Hamilton Depression Depressive symptom severity Number of items: 21
Rating Scale (HDRS; Interviewer-administered rating scale
Hamilton, 1960) Sample studied: Affective disorder of depressive type
Validity:
CV = Highly correlated with ASIQ, SSI and suicide item of BDI
PV = Patients who scored 2 or higher on HRSD suicide item
were 4.9 times (95 percent, CI: 2.7–9.0) more likely
to commit suicide than patients who scored less than 2
Reliability:
IR : r = .92
Test-retest reliability: r = .64
(Appendix Continued)
(Appendix Continued)

Depression and Hopelessness


Scales and authors Constructs Standardisation details
Reynolds Adolescent Current level of an adolescent’s Number of items: 30
Depression Scale-2 (RADS-2; depressive symptomatology Self-reporting scale
Reynolds,1987, 1988) Dysphoric mood, anhedonia/ Sample studied: Adolescents with depressive symptomatology
negative affect, negative self- Validity:
evaluation and somatic CV = Correlation with BDI-II (r = .84), correlation with
complaints Hamilton interview on retesting was .76
Reliability:
CC: r = .91
Test-retest reliability: r = .87
Children’s Depression 27 sets of items yields a total score Number of items: 27 sets of item
Inventory (CDI; Kovacs, 1982) and five subscale scores on: mood, Self-report inventory
interpersonal problems, Sample studied: Children and adolescents (7 to 17 years)
ineffectiveness, anhedonia and Validity:
negative self-esteem CV: Correlation of CDI suicidal items and remaining total
scores in non-clinical group, r = .45; psychiatric sample,
r = .52; newly diagnosed diabetic youths, r = .22; second
sample of school children, r = .49; sexually abused youth,
r = .27.
PV: Moderately predictive of suicidal second sample of school
children, r = .49; sexually abused youth, r = .27.
PV: Moderately predictive of suicidal ideation one year later,
r = .39
Reliability:
Test-retest reliability: 50 percent of the youths endorsed initial
score at the second testing 6 to 9 weeks later
Children’s Depression To diagnose severity of 17 Brief rating scale based on a semi-structured scale
Rating Scale (CDRS-R; symptoms of depression Interviewer administered
Poznanski et al., 1984) Sample studied: Child (or an adult informant who knows the
child well)
Validity:
Good concordance with diagnosis of depression
Corroborates with recovery
Correlates with other tests of depression severity
Reliability:
IR: r = .86
Beck Hopelessness Extent of negative expectancies Number of items: 20 true–false statements
Scale (BHS; Beck et al., 1974) about the future Sample studied: Psychiatric inpatients and outpatients with
suicide ideation and attempt
Validity:
CV = Correlations with clinical ratings of hopelessness,
r = .66 to .74
Correlation coefficients between the BHS and the BDI
pessimism item, r = .42 to .64 in clinical samples
PV = Scores of 9 and above have accurate prediction of
eventual suicide completion in outpatients
Multi-dimensional Hope in chronically ill patients Sample studied: Inpatients and in non-clinical population—
Hope Scale (Raleigh minimum age, 13 years reported
and Boehm, 1994) Validity:
CV = Significant negative correlation with the BHS,
r = –.45
Reliability:
CC: r = .95
Test-retest reliability: r = .82
(Appendix Continued)
(Appendix Continued)
Depression and Hopelessness
Scales and authors Constructs Standardisation details
State Hope Scale Present status in hope and Number of items: 6
(Snyder et al., 1996) described as ‘Goal scale for Self administered
the present’ Sample studied: Non-clinical adults
Validity:
CV: with Trait Hope Scale, r = .79
With self-esteem, .45 to .75; with negative affect of Positive
and Negative Affect Scale (PANAS), –.37 to –.50
Reliability:
CC: Agency subscale, .79 to .95
CC: Pathway subscale, .59 to .93
Test-retest: .82 to .93
Adult Dispositional (Trait) Hope as disposition, described Number of items: 12
Hope Scale (Snyder et al., 1991) as ‘the future scale’ Self-reporting scale
Sample studied: Non-clinical college students and individual
in psychological treatment
Validity:
CV: with optimism (test of Life-Orientation) .50 to .60 with
hopelessness scale of Becket, r = –.51, and BDI –.42
Discriminant validity with Self-Conscious Scale
Reliability:
CC: .74 to .84.
Test-retest: r = .80 for individual in psychological treatment
Children’s Hope Dispositional index of hope Number of items: 6
Scale (Snyder et al., 1997) Combination of agentic and Self-reporting scale
pathways thinking towards goals, i.e., Sample studied: Children
perceived capability to produce Validity:
routes to those goals CV: r = .38, reported adequate discriminant, and incremental
validity
Reliability:
CC: r = .72 to .86
Item-remainder coefficients = .27 to .68
Test-retest reliability: r = .71
Hopelessness Scale for Children Hopelessness Number of items: 17
(HPLS; Kazdin et al., 1986) (Modification of the BHS) Self-reporting scale
Sample studied: Adolescents and psychiatric patients
Validity:
CV = Positive correlation (r = .71) with a five-item
questionnaire used by Beck to validate the BHS
PV: With severity of depressive symptoms, depressive
diagnoses, poor self-esteem, poor self-rated social skills,
anxiety and perfectionism in girls, difficult temperament,
lower estimated intellectual functioning, suicidal ideation,
suicidal tendencies and suicide attempts
Moderate stability (r = .57) in HPLS scores over a six-week
period of time among child psychiatric inpatients
Reliability:
Internally consistent (r = .89) among adolescent psychiatric
inpatients
Multidimensional Self-oriented Number of items: 45
Perfectionism Scale Other-oriented Self-reported scale
(MPS; Hewitt and Flett, 1991) Socially prescribed dimensions Sample studied: Clinical and non-clinical adult population
Validity:
Item-to-subscale total correlations:
Self-oriented items: r = .51 and .73
Other-oriented items: .43 and .64
Socially prescribed items: .45 and .71
Reliability:
(Appendix Continued)
(Appendix Continued)

Depression and Hopelessness


Scales and authors Constructs Standardisation details
CC:
Self-oriented perfectionism, r = .86
Other-oriented perfectionism, r = .82
Socially prescribed perfectionism, r = .87
Child and Adolescent Self-oriented and socially Number of items: 22
Perfectionism Scale prescribed-oriented perfectionism Self-reporting scale
(CAPS; Flett et al., 1992) Sample studied: Adolescent (11–19 years)
Reliability:
Test-retest reliability (one week), r = .80;
CC: r = .85
Extended Attributional Style 3 dimensions: Items: 12 hypothetical negative events
Questionnaire (EASQ; Joiner and Internal–external Self-rating scale
Metalsky,1999) Stable–unstable Sample studied: University undergraduates
Global–specific Reliability:
CC: r = .86 (achievement events); .82
(interpersonal events)
Means–End Problem-Solving Measure social problem skill by Validity:
Procedure (Platt and Spivack, 1975) generating social scenarios and Parasuiciders discriminated from normal,
asking for possible solutions and generating fewer and less relevant solutions
obstacles
Self Injury Implicit Association Test Reaction time measure of implicit Items: Series of images of self-injury—related or neutral
(SI-IAT; Nock and Banaji, 2007) association between self-injury and Sample studied: Suicide ideators, suicide attempters and non-
oneself to detect suicide ideators suicidal adolescents (12–19 years)
and attempters Validity:
PV = 74–77 percent accuracy in discriminating randomly
selected suicidal individual from non-suicidal ones
Domain of Personality
Scales and authors Constructs Standardisation details
Ego Function Assessment Scale- Separate dimensions for each Number of items: 120
Modified (EFA-M; Basu et al., 1996) of the 12 ego functions Self-reporting scale
(adapted from Bellack, 1989) Reality testing, judgement, sense of Sample studied: College students (20–23 years)
reality, drive control, object relation, Validity:
thought process, adaptive regression Internal consistency validity between each item and
in service of ego (ARISE) defensive corresponding subscale (r = .32 to .84); CV = Psychoticism
function, stimulus barrier, scale, .20 to .36 (excepting ARISE scale)
autonomic function, synthetic/ Neuroticism scale, .30 to .55 (excepting ARISE scale)
integrative function, sense of Reliability:
mastery and competence Reliability coefficient for each subscale
Chronbach, .50 to .79; Split half, .52 to .79
Life Style Index Ego defence mechanisms of Number of items: 97
(LSI; Plutchik et al., 1979) compensation, denial displacement, Self-reporting questionnaire
intellectualisation, projection, Sample studied: Non-clinical college students
reaction formation, regression, Reliability:
repression Internal consistency for each of the defence mechanisms
obtained from a sample of inpatients and college students and
test-retest reliability are respectively as follows:
Displacement (.69, .62, .76) Intellectualisation (.58, .30, .61)
Projection (.86, .75, .75)
Reaction formation (.73, .63, .76)
Regression (.65, .56, .38) Repression (.55, .38, .48)
Compensation (.59, .43, .61) Denial (.54, .52, .55)
(Appendix Continued)
(Appendix Continued)

Domain of Personality
Scales and authors Constructs Standardisation details
Rorschach Suicide Suicide proneness Response from 10 cards
Index Constellation Affective variables (vista responses, Sample studied: Clinical and non-clinical adolescents
(RSIC; Exner, 2002) colour-shading Validity: Four of six of the features on this index selected 64
blends, colour-dominated responses percent of suicidal subjects
and morbid content), cognitive PV: Estimated S-CON score of 7 or more predictor of
distortion (inaccurately perceived near-lethal suicide attempts
human movement responses [M-]
and special scores)
Epigenetic Assessment Describes 10 levels of personality Rating of narrativised speech sample
Rating System (EARS; Wilson, organisation Sample studied: Adolescents (12–16 years)
Passik and Kuras, 1989; as cited in Validity:
Feldman and Wilson, 1997) Validated demonstrating factorial independence of
psychological dimension and arousal at various levels of
personality dimensions across the task conditions
Reliability:
IR: r = .80
Domain of Interpersonal Relationship
Scales and Authors Constructs Standardisation Details
Family Adaptability and Cohesion Family relationships: Number of items: 20
Evaluation Scales (FACES IV; Olson, evaluate communication Self-report survey
Gorall and Tiesel, 2004; as cited in styles, family interactions Sample studied: Adolescents and adults (12 years onwards)
Olson and Gorall, 2006) and flexibility Validity:
r = .91 to .93
Reliability:
CC: Disengaged = .87,
Enmeshed = .77, Rigid = .83, Chaotic = .85, Cohesion = .89,
Balanced Flexibility = .80
Parental Bonding Instrument Two aspects: Number of items: 25
(PBI; Parker et al., 1979) Care Self-report questionnaire
Protection Sample studied: Adolescents (12–18 years)
Validity:
Convergent validity with family, corroborative witnesses
and twin studies and studies using independent raters on
construct of care and protection
Reliability:
Internal consistency (split half)
Care: r = .88; Protection: r = .74
Life Event Scales
Scales and authors Constructs Standardisation details
Impact of Events Scale Intrusion Number of items: 15 items
(IES; Horowitz et al., 1979) Avoidance Self-reporting scale
Degree to which respondents have Sample studied: Adult parasuicider and non-clinical control
experienced intrusive and avoidant Reliability:
thoughts relating to a specified event CC: Intrusion, r = .83; Avoidance, r = .72
that occurred in the last 6 months
Life Experience Relatively frequently Number of items: 47 (general), 10 items (students)
Survey (LES; occurring life events Self-reporting scale
Sarason et al., 1978) Reported adequate validity
Reliability:
6-week test-retest: r = .63
(Appendix Continued)
Domain of Attitude
Scales and authors Constructs Standardisation details
Multi Attitude Attraction to life Number of items: 30
Suicide Tendency Repulsion to life Sample studied: Adolescents, non-clinical high school
Scale for Adolescents Attraction to death students, psychiatric inpatients and outpatients
(Orbach et al., 1991) Repulsion to death Validity:
CV: r = with Israelian Index of suicide potential
Attraction to life, .66; repulsion to life, .64
Attraction to death, .48; repulsion to death, .28
Reliability:
Internal consistency
Attraction to life, .83; repulsion to life, .76
Attraction to death, .76; repulsion to death, .83
Total scale = .92
Brief Screening Measures
Scales and authors Constructs Standardisation details
Paykel Suicide Items Questions with increasing levels of Number of items: 5
(PSI; Paykel et al., 1974) intent to assess suicidality during Interviewer administered
the past week, month, year or Samples studied:
lifetime. On residents of Adults between 18 and 60 years and above
psychiatric catchments area Validity:
CV = with any suicidal feelings during the past year,
psychiatric symptoms with social isolation, somatic
complaints and had a greater proportion of two or more
negative life events in the past year than non-suicidal controls
Suicidal subjects with hospital admission for emotional
problems or for taking tranquilizers in the past year
Suicidal Ideation Screening Disturbance in sleep and mood, Number of items: 4
Questionnaire (SIS-Q; guilt and hopelessness during Samples studied:
Cooper-Patrick et al., 1994) the past year Patients receiving care in general medical setting (18–70
years)
Interviewer administered
Validity:
CV: Correctly identified 84 percent of general medical
patients with suicide ideation
Pritha Mukhopadhyay

REFERENCES

Basu, J., M. Banerjee and P. Mukhopadhyay (1996). Applicability of the ego function
assessment scale on college population. Indian Journal of Clinical Psychology,
23(1), 40–46.
Baumeister, R.F. (1990). Suicide as escape from self. Psychological Review, 97(1),
90–113.
Beck, A.T., G. Brown, R.J. Berchick, B.L. Stewert and R.A. Steer (1990). Relationship
between hopelessness and ultimate suicide: A replication with psychiatric out-
patients. American Journal of Psychiatry, 147(2), 190–95.
Beck, A.T. and R.A. Steer (1987). Manual for Beck Depression Inventory. San Antonio,
TX: Psychological Corporation.
Beck, A.T. and R.A. Steer (1991). Manual for the Beck Scale for Suicide Ideation.
San Antonio, TX: Psychological Corporation.
Beck, A.T., R. Beck and M. Kovacs (1975). Classification of suicidal behaviors:
I. Quantifying intent and medical lethality. American Journal of Psychiatry,
132(3), 285–87.
Beck, A.T., G.K. Brown and R.A. Steer (1997). Psychometric characteristics of the scale
for suicide ideation with psychiatric outpatients. Behavior Research and Therapy,
35(11), 1039–46.
Beck, A.T., M. Kovacs and A. Weissman (1979). Assessment of suicidal intention: The
scale for suicide ideation. Journal of Consulting and Clinical Psychology, 47(2),
343–52.
Beck, A.T., R.A. Steer and G.K. Brown (1996). Manual for the Beck Depression
Inventory-II”. San Antonio, TX: Psychological Corporation.
Beck, A.T., A. Weissman, D. Lester and L. Trexler (1974). The measurement of
pessimism: The hopelessness scale. Journal of Consulting and Clinical Psychology,
42(6), 861–65.
Bellack, K. (1989). The ego function assessment: A manual. New York: Wiley.
Berk, M.S., E. Jeglic, G.K. Boown, G.R. Henriques and A.T. Beck (2007). Characteristics
of recent suicide attempters with and without borderline personality disorder.
Archives of Suicide Research, 11(1), 91–104.
Brown G.K. (2008). A review of suicide assessment measures for intervention research
with adults and older adults. National Institute of Mental Health. Retreived
7 September 2008 from: http://www.nimh.nih.gov/suicideresearch/adultsuicide.
pdf, p. 10.
Brown, G.K., A.T. Beck, R.A. Steer and J.R. Grisham (2000). Risk factors for suicide
in psychiatric outpatients: a twenty year prospective study. Journal of Consulting
and Clinical Psychology, 68(3), 371–77.
brown, G.L., M.I. Linnoila and F.K. Goodwin (1992). Impulsivity, aggression, and
associated affect: Relationship to self destructive behavior and suicide. In R.W.

224
Suicide, Assessment and Prediction

Maris, A.L. Berman, J.T. Maltsberger and R.L. Yufit (Eds), Assessment and
Prediction of Suicide (pp. 589–606). New York: The Guilford Press.
Ciffone, J. (1993). Suicide prevention: A classroom presentation to adolescents. Social
Work, 38(2), 197–203.
Clum, G.A. and L. Curtin (1993). Validity and reactivity of a system of self-monitoring
suicide ideation. Journal of Psychopathology and Behavioral Assessment, 15(4),
375–85.
Cooper-Patrik, L., R.M. Crum and D.E. Ford (1994). Identifying suicidal ideation in
general medical patients. Journal of the American Medical Association, 272(22),
1757–62.
Cull, J.G. and W.S. Gill (1988). Suicide probability scale manual. Los Angeles: Western
Psychological Services.
Dervic, K., M.F. Grunebaum, A.K. Burke, J.J. Mann and M.A. Oquendo (2007). Cluster-
C personality disorder in major depressive episodes: The relationship between
hostility and suicidal behaviour. Archives of Suicide Research, 11(1), 83–90.
Dhar, S. and S. Basu (2006). A comparative study of number of life events, presumptive
stress and different ego functions of students with low and high suicidal risks.
Indian Journal of Clinical Psychology, 33(2), 159–64.
Dogra, A.K., S. Basu and S. Das (2008). The roles of personality stressful life events
meaning in life, reasons for living on suicidal ideation: A study on college students
SIS. Journal of Projective Psychology and Mental Health, 15(1), 52–57.
Ellis, T.E. and C.F. Newman (1998). Choosing to Live: How to Defeat Suicide through
Cognitive Therapy. Oakland, CA: New Harbinger.
Exner, J. (2002). Rorschach: A Comprehensive System: Basic Foundations and Principles
of Interpretation. Chicheter: John Wiley and Sons.
Feldman, M. and A. Wilson (1997). Adolescent suicidality in urban minorities and its
relationship to conduct disorders, depression and separation anxiety. Journal of
American Academy of Child and Adolescent Psychiatry, 36(1), 75–84.
Firestone, R.W. and L.A. Firestone (1996). Firestone assessment of self-destructive
thoughts. San Antonio, TX: Psychological Corporation.
Flett, G.L., P.L. Hewitt, D.J. Boucher, L.A. Davidson and Y. Munro (1992). The
Child-Adolescent Perfectionism Scale: Development, validation, and association
with adjustment. Department of Psychology Reports, York University, Toronto,
No. 203.
Gelder, M., D. Gath, R. Mayyou and P. Cowen (1988). Oxford Textbook of Psychiatry
(3rd Ed). London: Oxford University Press.
Goldstein, R.B., D.W. Black, A. Nasrallah and G. Winokur (1991). The Prediction
of suicide: Sensitivity, specificity and predictive value of a multivariate model
applied to suicide among 1906 patients with affective disorders, Archives of General
Psychiatry, 48(5), 418–22.
Goldston, D.B. (2003). Measuring suicidal behavior and risk in children and adolescents.
Washington, DC: American Psychological Association.

225
Pritha Mukhopadhyay

Gould, M.S., T. Greenberg, D.M. Velting and D. Shaffer (2003). Youth suicide risk
and preventive interventions: A review of the past 10 years. Journal of American
Academy of Child and Adolescent Psychiatry, 42(4), 386–405.
Haaga, D.A., M.J. Dyck and D. Ernst (1991). Empirical status of cognitive theory of
depression. Psychological Bulletin, 110(2), 215–36.
Hamilton, M. (1960). A rating scale for depression. Journal of Neurology, Neurosurgery
and Psychiatry, 23, 56–62.
Hansburg, H.G. (1980). Adolescent Separation Anxiety Test. Melbourne: Krieger
Publishing.
Hewitt, P.L. and G.L. Flett (1991). Perfectionism in the self and social contexts: Con-
ceptualization, assessment, and association with psychopathology. Journal of
Personality and Social Psychology, 60(3), 456–70.
Holden, R.R. and L.D. McLeod (2000). The structure of the Reasons for Attempting
Suicide Questionnaire (RASQ) in a nonclinical adult population. Personality and
Individual Differences, 29(4), 621–28.
Horowitz, M., N.J. Wilner and W. Alvarez (1979). Impact of events scale: A measure
of subjective stress. Psychosomatic Medicine, 41(3), 209–18.
Jacobs, D. (2003). Suicide assessment. University of Michigan Depression Center
ColloquiumSeries, Presented at University of Michigan.
Jobes, D.A., M.M. Moore and S.S. O’Connor (2007). Working with suicidal clients using
the Collaborative Assessment and Management of Suicidality (CAMS). Journal of
Mental Health Counseling, 29(4), 283–300.
Jobes, D.A., A.M. Jacoby, P. Cimbolic and L.A.T. Hustead (1997). Assessment and
treatment of suicidal clients in a university counseling center. Journal of Counseling
Psychology, 44(4), 368–77.
Johnson, W.B., R. Lall, B. Bongar and M.D. Nordlund (1999). The role of objective
personality inventories in suicide risk assessment: An evaluation and proposal.
Suicide and Life-Threatening Behavior, 29(2), 165–85.
Joiner, T.E. and G.I. Metalsky (1999). Factorial construct validity of the extended attri-
butional style questionnaire. Cognitive Therapy and Research, 23(1), 105–13.
Joiner, T.E., M.D. Rudd and M.H. Rajab (1999). Agreement between self-and
clinician-rated suicidal symptoms in a clinical sample of young adults: Explaining
discrepancies. Journal of Consulting and Clinical Psychology, 67(2), 171–76.
Kalafat, J. and N. Elias (1994). An evaluation of a school-based suicide awareness
intervention. Suicide and Life-Threatening Behavior, 24(3), 224–33.
Kazdin, A., A. Rodgers and D. Colbus (1986). The hopelessness scale for children:
Psychometric characteristics and concurrent validity. Journal of Consulting and
Clinical Psychology, 54(2), 241–45.
Kovacs, M. (1982). Children’s depression inventory. Pittsburgh: Western Psychiatric
Institute and Clinic.
Linehan, M.M., J.L. Goodstein, S.L. Nielsen and J.A. Chiles (1983). Reasons for staying
alive when you are thinking of killing yourself: The reasons for living inventory.
Journal of Consulting and Clinical Psychology, 51(2), 276–86.

226
Suicide, Assessment and Prediction

Linehan, M.M. (1981). Suicidal behaviors questionnaire. Unpublished inventory,


University of Washington, Seattle, Washington.
Linehan, M.M. (1993). Cognitive-behavioral treatment of Borderline Personality Disorder.
New York: Guilford.
Linehan, M.M. (1996). Suicidal Behaviors Questionnaire (SBQ). Unpublished
manuscript, Department of Psychology, University of Washington, Seattle, WA.
Maris, R.W., A.L. Berman and J.T. Maltsberger (1992). Summary and conclusions:
What have we learned about suicide assessment and prediction? In R.W. Maris,
A.L. Berman, J.T. Maltsberger and R.L. Yufit (Eds), Assessment and Prediction of
Suicide (pp. 640–72). New York: The Guilford Press.
Meichenbaum, D. (2005). 35 years of working with suicidal patients: lessons learned.
Canadian Psychologist, 46(2), 64–72.
Miller, I.W., W.H. Norman, S.B. Bishop and M.G. Dow (1986). The modified scale
for suicide ideation: Reliability and validity. Journal of Consulting and Clinical
Psychology, 54(5), 724–25.
Motto, J.A. (1992). An integrated approach to estimating suicide risk. In R.W. Maris,
A.L. Berman, J.T. Maltsberger and R.L. Yufit (Eds), Assessment and Prediction of
Suicide (pp. 625–39). New York, The Guilford press.
Munchua, M.M. (2003). The underlying need structure of psychache and its role in
the statistical prediction of suicidal manifestations. Unpublished Master’s Thesis,
Queen’s University, Kingston, Ontario, Canada.
Negron, R., J. Piacentini, G. Flemming, M. Davies and D. Shaffer (1997). Microanalysis
of adolescent suicide attempters and ideators during the acute suicidal
episode. Journal of American Academy of Child and Adolescent Psychiatry, 36(11),
1512–19.
Nock, M.K. and M.R. Banaji (2007). Prediction of suicide ideation and attempts among
adolescents using a brief performance based test. Journal of Consulting and Clinical
Psychology, 75(5), 707–15.
O’Carroll, P.W., A.L. Berman, R.W. Maris, E.K. Moscicki, B.L. Tanney and
M.M. Silverman (1996). Beyond the Tower of Babel: A nomenclature for
suicidology. Suicide and Life-Threatening Behavior, 26(3), 237–52.
O’Connor, R.C. and A.A. Leenaars (2003). A thematic comparison of suicide notes
drawn from Northern Ireland and the United States. Current Psychology, 22(4),
339–47.
Olson, D.H. and D.M. Gorall (2006). FACES IV & the Circumplex Model. Retrieved
10 October 2008 from http://www.facesiv.com/pdf/3.innovations.pdf
Orbach, I., I. Milstein, D. Har-Even, A. Apter, S. Tiano and A. Elizur (1991).
Multi attitude suicide tendency scale for adolescents. Psychological Assessment,
3(3), 398–404.
Osman, A., P.M. Gutierrez, B.A. Kopper, F.X. Barrios and C.E. Chiros (1998). The
positive and negative suicide ideation inventory: Development and validation.
Psychological Reports, 82(3, part 1), 783–93.

227
Pritha Mukhopadhyay

Parker, G., H. Tupling and I.B. Brown (1979). A parental bonding instrument. British
Journal of Medical Psychiatry, 52, 1–20.
Paykel, E.S., J.K. Myers, J.J. Lindenthal and J. Tanner (1974). Suicidal feelings in the
general population: A prevalence study. British Journal of Psychiatry, 124(5),
460–69.
Pfeffer, C.R., H.R. Conte, R. Plutchik and I. Jerret (1979). Suicidal behaviour in latency
age children: An empirical study. Journal of American Academy of Child Psychiatry,
18(4), 679–92.
Platt, J.J. and G. Spivack (1975). Unidimensionality of the Means-Ends Problem-Solving
(MEPS) procedure. Journal of Clinical Psychology, 31(1), 15–16.
Plutchik R., H. Kellermen and H.R. Conte (1979). A structural theory of ego defense
and emotions. In C.E. Izard (Ed.), Emotions in Personality and Psychopathology
(pp. 229–57). New York: Plenum.
Plutchik, R., H.M. van Praag, H.R. Conte and S. Picard (1989). Correlates of suicide
and violence risk: The suicide risk measure. Comprehensive Psychiatry, 30(4),
296–302.
Potter, L.B., M. Kresnow, K.E. Powell, P.W. O’Carroll, R.K. Lee, R.F. Frankowski et al.
(1998). Identification of nearly fatal suicide attempts: Self-inflicted injury severity
form. Suicide and Life-Threatening Behavior, 28(2), 174–86.
Poznanski, E.O., J.A. Grossman, Y. Buchsbaum, M. Banegas, L. Freeman and R. Gibbons
(1984). Preliminary study of the reliability and validity of the children’s depression
rating scale. Journal of American Academy of Child Psychiatry, 23(2), 191–97.
Raleigh, E.H. and S. Boehm (1994). Development of multidimensional hope scale.
Nursing Measures, 2(2), 155–67.
Reynolds, W.M. (1987). Suicidal Ideation Questionnaire. Odessa, FL: Psychological
Assessment Resources.
Reynolds, W.M. (1988). Reynolds Adolescent Depression Scale-2. California: Western
Psychological Services.
Reynolds, W.M. (1991a). Psychometric characteristics of the adult suicidal ideation
questionnaire in college students. Journal of Personality Assessment, 56(2),
289–307.
Reynolds, W.M. (1991b). Adult Suicide Ideation Questionnaire: Professional manual.
Odessa, FL: Psychological Assessment Resources.
Rudd, M.D. (1989). The prevalence of suicidal ideation among college students. Suicide
and Life-Threatening Behavior, 19(2), 173–83.
Sarason, I.G., J.H. Johnson and J.M. Siegel (1978). Assessing the impact of life changes:
Development of the life experiences survey. Journal of Consulting and Clinical
Psychology, 46(5), 932–46.
Shaffer, D., A. Garland, V. Vieland, M.M. Underwood and C. Busner (1991). The
impact of curriculum–based suicide prevention programme for teenagers. Journal
of American Academy of Child and Adolescent Psychiatry, 30(4), 588–96.

228
Suicide, Assessment and Prediction

Sherbourne, C.D. and A.L. Stewart (1991). Rescue scale. The Medical Outcomes Study’s
(MOS) measure of social support. The MOS Social Support Survey. Social Science
and Medicine, 32(6), 705–14.
Sil, M. and S. Basu (2007). A study of hope, hopelessness, reasons for living and suicidal
ideation in college students. Indian Journal of Clinical Psychology, 34(1), 76–82.
Snyder, C.R., B. Hoza, W.E. Pelham, M. Rapoff, L. Ware, M. Danovsky et al. (1997).
The development and validation of the children’s hope scale. Journal of Pediatric
Psychology, 22(3), 399–421.
Snyder, C.R., C. Harris, J.R. Anderson, S.A. Holleran, L.M. Irving, S.T. Sigmon
et al. (1991). The will and the ways: Development and validation of an individual
differences measure of hope. Journal of Personality and Social Psychology, 60(4),
570–85.
Snyder, C.R., S.C. Sympson, F.C. Ybasco, T.F. Borders, M.A. Babyak, Higgins et al.
(1996). Development and validation of the state hope scale. Journal of Personality
and Social Psychology, 70(2), 321–35.
Speilberger, C.D. (1988). State-Trait Anger-Expression Inventory. Odessa, Florida:
Psychological Assessment Resources.
Weissman, A.D. and J.W. Worden (1972). Risk-rescue rating in suicide assessment.
Archives of General Psychiatry, 26(6), 553–60.
Wilson, S.T., B. Stanley, M.A. Oquendo, P. Goldberg, G. Zalsman and J.J. Mann (2007).
Comparing impulsiveness, hostility, and depression in borderline personality
disorder and bipolar II disorder. Journal of Clinical Psychiatry, 68(10), 1533–39.

229
11

Substance Use and Suicidal Behaviour


N SH M S , A S H U SH K

W ith suicide accounting for over 1 million deaths worldwide every


year, suicidal behaviour is a pressing global health and social
concern (WHO, 2001). Such behaviour commonly occurs as a result of inter-
actions between multiple predisposing factors and current environmental
factors that influence its development. These may be neurobiological,
genetic, medical, psychological, social, economic or cultural (Beautrais
et al., 1996; Norstrom et al., 1995).
Particularly influential in suicidal behaviour is the presence of psy-
chiatric disorders. The majority of individuals who commit suicide have
a diagnosable mental disorder and, indeed, suicidal behaviour is more fre-
quent among psychiatric patients (Henriksson et al., 1993; Mann, 2002).
In their meta-analysis of 3,275 suicides across the world, Arsenault-
Lapierre, Kim and Turecki (2004) reported that ‘on average, 87.3 % of
the subjects who committed suicide had a mental disorder’. Affective,
substance-related, personality and psychotic disorders accounted for the
majority of these diagnoses.
Numerous studies have reported a significant association between sub-
stance abuse and suicidal behaviour, especially in young people (Fowler
et al., 1986; Neeleman and Farrel, 1997). In fact, substance abuse itself
is viewed as a form of self-destructive behaviour by some researchers.
Meninger (1938), for example, described substance abuse as ‘chronic

230
Substance Use and Suicidal Behaviour

suicide’, suggesting that substance abuse is, perhaps, at least partially moti-
vated by suicidal impulses, either consciously or unconsciously.
The San Diego Suicide Study by Rich and colleagues (1986) was one of
the first investigations to report an association between substance abuse and
suicide. This study also suggested that this association is stronger among
individuals aged under 30 as compared to those aged 30 and above. Ac-
cording to a 2004 World Health Organization study, substance-related
disorders are involved in 17 percent of completed suicides (Bertolote
et al., 2004). Molnar, Berkman and Buka (2001), in their assessment of
data from the US National Comorbidity Survey, concluded that alcohol
and drug abuse are associated with a suicide risk 6.2 times greater than
the average risk.
Several studies of suicide among adolescents and young adults in
the United States have consistently reported high rates of substance use
(60–70 percent; Brent et al., 1988; Shafii et al., 1985), although similar
studies in Europe have reported relatively lower rates of substance use
associated with suicide (30–47 percent; Appleby et al., 1999; Marttunen
et al., 1991). Asian research exploring the association between suicidal
behaviour and substance use is unfortunately still scarce.
The risk of suicidal behaviour is associated not only with substance
abuse, but also with substance use and dependence. Borges, Walters and
Kessler (2000) found that current substance use, even in the absence of
abuse or dependence is a significant risk factor for unplanned suicides
among individuals who have suicidal ideation.

THE CONTINUUM OF SUICIDAL BEHAVIOUR

Suicidal behaviour spans a spectrum that, in addition to completed suicide,


also encompasses suicidal ideas and thoughts of varying intensity and spe-
cificity, and other acts of deliberate self-harm of varying intent and
lethality. The tripartite classification system suggested by Beck, Resnik and
Lettieri (1973) categorises these behaviours as suicidal ideation, suicide
attempt and completed suicide.
Completed suicide is defined as self-inflicted death, and suicide attempt
as self-destructive act with a non-fatal outcome, both accompanied
by evidence (explicit or implicit) that the person intended to die

231
Nishi Misra et al.

(American Psychiatric Association, 2003). At the mildest end, suicidal


ideation is a preoccupation with intrusive thoughts of ending one’s own
life which includes suicide threats, direct expressions of the wish to die,
and indirect indicators of suicide planning (Harter et al., 1992). The
three forms of suicidal behaviour—completed suicide, suicide attempt
and suicidal ideation—can, thus, be understood as a continuum of self-
harming behaviours.

TYPES OF STUDIES

Studies exploring the association between substance use and suicide can
be broadly classified into three categories based on their research design
(Hillman et al., 2000):

1. Psychological autopsy studies—which involves the retrospective


psychiatric assessment of the deceased by means of a proxy-based
interview process with best informants on the deceased (Arsenault-
Lapierre et al., 2004).
2. Case behaviour (cases) and a representative sample of those who
do not (controls), on the risk factors of interest (Breslow and Day,
1980).
3. Longitudinal studies—which involve repeated observations of the
same sample for suicidal behaviour, with assessments of exposure
to the risk factors, at regular intervals (Beautrais, 2003).

THE SUBSTANCES OF ABUSE

Substance use has generally been found to increase the likelihood of a


suicide attempt, but the specific association between the different types
of substances of abuse (with the exception of alcohol) and suicidal be-
haviour has not been clearly established (Borges et al., 2000). In many
cases of suicidal behaviour abuse of drugs and alcohol are co-occurring,
which makes it difficult to distinguish the contributions of each separately
(Harris and Barraclough, 1997). Research discussing the role of alcohol

232
Substance Use and Suicidal Behaviour

in elevating suicide risk is considerably more extensive and robust than


similar research on the role of other drugs of abuse in suicidal behaviour.
Felts, Chenier and Barnes (1992), in their study on adolescents in North
Carolina, found that the use of cocaine/crack was more closely associated
with a self-reported incidence of attempted suicide than was the use of
alcohol, marijuana or needle drugs.
Analysis of data on 7,227 suicides in 2003 (Serpi et al., 2005) indicated
that among suicide victims who tested positive for substances, 33.3
percent tested positive for alcohol, 16.4 percent for opiates, 9.4 percent for
cocaine, 7.7 percent for marijuana and 3.9 percent for amphetamines.

Alcohol

The association between alcohol use and suicidal behaviour, both in


alcohol dependent and non-alcohol dependent populations, has long been
recognised and documented (Wilcox et al., 2004). According to a meta-
analysis of mortality studies by Inskip, Harris and Barraclough (1998), the
lifetime risk of suicide is 7 percent for alcohol dependence.
Several studies across the world have consistently reported a high pre-
valence of alcohol use disorders among people who committed suicide
(20–56 percent; Caces and Harford, 1998; Pirkola et al., 2000; Vijayakumar
and Rajkumar, 1999).
Alcohol dependence has also been found to be associated with at-
tempted suicide and suicidal ideation (Borges et al., 2000). An increased
risk of suicide attempts in alcoholism has been reported, ranging from 17
(Schuckit, 1986) to 29 percent (Whitters et al., 1985).
Certain other variables have been found to be associated with suicidal
behaviour among alcoholics. In many studies, suicidal behaviour has been
more strongly associated with alcohol use among males than among
females (Ohberg et al., 1996; Pirkola et al., 2000). On the other hand,
other studies have reported that the risk for suicide among those with
alcohol use disorders is higher for females (e.g., Roy et al., 1990). Other
demographic variables associated with suicide attempts among alcoholics
are younger age, lower socio-economic status and family history of alcohol
abuse (Roy et al., 1990).
With respect to per capita alcohol consumption, Stack (2000) reported
that the greater the alcohol consumption, the higher is the suicide rate,

233
Nishi Misra et al.

although the strength of this association varied considerably across coun-


tries. Positive and significant associations between per capita alcohol
consumption and male suicide rates have been found in a number of
countries, for example, Denmark (Skog, 1993), France (Norstrom, 1995),
Norway (Rossow, 1993) and Portugal (Skog et al., 1995). This relationship
between alcohol consumption and suicide rate has also been evidenced in
the former Union of Soviet Socialist Republics (USSR), where, as a part
of an anti-alcohol campaign, the prices of alcohol were increased and its
availability was reduced. As a result, the consumption of alcohol reduced
considerably, as did the suicide rate (Wasserman et al., 1998).
The severity of alcoholism has been shown to be associated with sui-
cidal behaviour (Conner and Duberstein, 2004). For example, Kockott
and Feuerlein (1968, as cited in Lester and Beck, 1975) reported that
individuals suffering from alcohol use disorder, who have a history of
delirium tremens, have a higher incidence of suicidal behaviour than those
who have not experienced delirium tremens. It has also been found that
suicidal behaviour is more common in the later stages of alcoholism than
in the earlier stages (Robins and Murphy, 1965).
Alcohol intoxication, even without abuse or dependence, has been
shown to predict suicidal behaviour. Autopsy studies have found alcohol
to be present in 20–50 percent of persons who commit suicide (Ohberg
et al., 1996). Suokas and Lonnqvist (1995) reported that 62 percent of
suicide attempters had consumed alcohol prior to or at the time of their
attempt. In fact, it has been claimed that immediate alcohol intoxication
presents a greater risk for suicidal behaviour than the habitual consumption
pattern (Borges and Rosovsky, 1996).

Cannabis/Marijuana

In a case-control study comparing 302 consecutive hospital admissions of


medically serious suicide attempts with 1,028 randomly selected control
participants, Beautrais, Joyce and Mulder (1999) reported that 16.2 percent
of the attempters met the DSM-III-R criteria for cannabis abuse/
dependence at the time of the attempt, compared to only 1.9 percent of
comparison subjects. However, when this association was controlled for
socio-demographic factors and psychiatric comorbidity, it fell short of
significance. More recently, however, regular or excessive cannabis use

234
Substance Use and Suicidal Behaviour

has been shown to be significantly associated with suicidal behaviour even


after adjusting for socio-demographic factors and comorbidity (Kung
et al., 2003).
This association between cannabis use and suicidal behaviour is found
to be more common among young people. In two recent studies of high
school students in France, on the basis of hierarchical multiple regression
analysis, Chabrol and colleagues reported that cannabis was a significant
predictor of suicidal behaviour including suicidal ideation (Chabrol,
Chauchard and Girabet, 2008; Chabrol, Mabila and Chauchard, 2008).
A study of 1,265 children for over a 21-year period found that cannabis
use, particularly heavy or regular use, was associated with later increases
in suicidal thoughts and suicide attempts (Fergusson et al., 2002). Heavy
cannabis use has also been associated with a relative risk for suicide four
times that of non-users in a prospective study of Swedish conscripts
(Andreasson and Allebeck, 1990). Lesser/casual use, however, was not
associated with an increased risk. A cross-sectional study of twin pairs
discordant for lifetime cannabis dependence found that individuals who
were cannabis dependent had odds of suicidal ideation and suicide at-
tempt that were 2.5 to 2.9 times higher than those of their non-cannabis
dependent co-twin (Lynskey et al. 2004).

Cocaine

Patients with cocaine dependence have been shown to have a greater risk
of suicidal behaviour. For example, Marzuk et al. (1992) found that one
out of every five suicide victims in New York aged 21–30 tested positive for
cocaine. A significantly greater risk of attempting suicide among cocaine
abusers was also reported by the US Epidemiologic Catchment Area survey
of 13,673 participants (Petronis et al., 1990).
More recently, Darke and Kaye (2004), in their study of injecting and
non-injecting cocaine users, found that 31 percent had attempted suicide,
and 18 percent had done so on more than one occasion. Moreover, the
injecting cocaine users were significantly more likely to have attempted
suicide and used more violent methods than non-injecting cocaine users.
The use of cocaine has also been found to be associated with suicidal
ideation. In a sample of 777 patients referred for evaluation of chemical

235
Nishi Misra et al.

dependency in a psychiatric emergency service, Garlow, Purselle and


D’Orio (2003) reported that 43.7 percent of the patients with only a
cocaine use disorder expressed suicidal ideation, compared to 38 percent
of those with both cocaine and alcohol use disorders, 24.3 percent with
only an alcohol use disorder. Some of the characteristics of the cocaine-
dependent patients who had attempted suicide are: a family history of
suicidal behaviour, more childhood trauma, higher personality scores
for introversion, neuroticism and hostility, more comorbidity with other
substance dependence, and psychiatric and physical disorders (Roy, 2001).

Heroin and Other Opiates

Suicide has been implicated as a frequent cause of death among heroin


users. There are a number of reports indicating elevated rates of suicide
and suicidal behaviour among opiate users (Oyefeso et al., 1999). The
proportion of deaths among heroin users attributed to suicide has been
found to range from 3 (O’Doherty and Farrington, 1997) to 35 percent
(Engstrom et al., 1991). In their meta-analysis of the literature on suicide
risk, Harris and Barraclough (1997) reported that death due to suicide
among heroin users occurs at 14 times the rate of matched peers. A
relatively recent study undertaken to determine the lifetime and recent
histories of attempted suicide among entrants to treatment for heroin
dependence reported that a lifetime history of attempted suicide was
reported by 34 percent of subjects (Darke et al., 2004).

THE PRECIPITATING FACTORS

The relationship between substance use and suicidal behaviour can be


conceptualised from a biological, psychological and a socio-economic
perspective. From a biological perspective there is evidence that alcohol
influences the serotonin neurons in the brain stem and causes a reduction
of serotonin transporter function in the prefrontal cortex (Malone, 1999).
On the other hand, disorders of central serotonergic neurotransmission,
as reflected by low levels of 5-HIAA, have been associated with suicidal
behaviour (Mann et al., 1999). A deficient serotogenic system has also been

236
Substance Use and Suicidal Behaviour

implicated as a predisposing factor to personality traits like impulsivity


and aggression (Coccaro 1989; Linnoila et al., 1993), both of which are
strongly associated with the use of substances such as alcohol, cocaine,
cannabis and ecstasy (e.g., Donovan et al., 1998; Guy et al., 1994).

Impulsivity

Individuals with more impulsive traits are known to be at greater risk


for suicidal behaviour (Horesh et al, 1999). Studies of non-fatal suicidal
behaviour have found 50 percent of attempters saying later that they did
not think of it for more than one hour beforehand (Williams, 1997).
At the other end of this relationship, impulsivity has been identified
both as a risk factor and predictor of substance use and abuse (Guy et al.,
1994). Studies exploring the association between substance use and
impulsivity have reported two different aspects of the association—higher
levels of impulsivity lead to acquisition of substance abuse and misuse, and
substances of abuse increase impulsivity (Perry and Carroll, 2008). It must
be noted, however, that while some studies have associated impulsivity
with the use of drugs (Butler and Montgomery, 2004; Donovan et al.,
1998), others have failed to find this association (McCann et al., 1994;
Ricaurte et al., 1990). Further research is, therefore, required to better
understand the role of impulsivity in the relationship between substance
use and suicidal behaviour.

Aggression

The alcohol–aggression relationship has been extensively researched in


adolescents (Dembo et al., 1997) as well as adults (Giancola et al., 2002;
Hoaken and Pihl, 2000).
With respect to suicidal behaviour, aggressiveness has been shown to be
a major underlying process facilitating such behaviour. Several studies have
identified aggression as a factor that increases the risk of suicidal behaviour
(Korn, et al., 1992; Plutchik and van Praag, 1989). It, therefore, seems
plausible that the use of substances may intensify aggressive behaviour,
which, in turn, may lead to suicidal behaviour.

237
Nishi Misra et al.

Deficient Cognitive Flexibility

Several researchers have studied the effect of alcohol and other substances
of abuse on the cognitive performance of individuals. Indeed, one of the
cognitive functions most commonly found to be negatively correlated to
substance use is cognitive flexibility. For example, Zorko and colleagues
(2004) compared the performance of alcohol dependents with controls on
a battery of neuropsychological tests and found that the alcoholics ex-
hibited more impairment in cognitive flexibility than did the controls,
amphetamine and heroin users (Ornstein et al., 2000).
In their review of factors for suicide attempts, Rudd and Joiner (1998)
listed cognitive rigidity as one of the factors that may increase the risk for
suicidal behaviour. Lack of cognitive flexibility can, therefore, be seen as
a possible link between substance use and suicidal behaviour, although
further research is necessary to prove the causality.

Deficient Problem-solving

Several studies comparing individuals displaying suicidal behaviour


and controls have found significant differences in their performance on
problem-solving tasks. Patsiokas, Clum and Luscomb (1979) compared a
group of 49 male suicide attempters with 48 controls and found that the
suicide attempters had greater difficulty in solving the problems.
It is also a well established fact that the use of substances leads to
impairment in problem-solving. Studies on cognitive impairments in
alcoholics have largely pointed towards problem-solving as a significantly
correlated factor (e.g., Zorko et al., 2004).

Psychosocial Stressors/Adverse Life Events

Psychosocial stressors are commonly found in suicidal behaviour reported


by clinical studies, with interpersonal conflicts (usually family or marital
discord), break-up of a significant relationship, and financial problems
being the most frequently reported (Weissman, 1974). In a retrospective
study of adolescent suicides, Marttunen, Aro and Lönnqvist (1992) reported
a high level of psychosocial stress in the year preceding the suicide.

238
Substance Use and Suicidal Behaviour

Incidences of current separation, divorce and current employment have


been reported to be higher among individuals with a history of suicide
attempts (Preuss et al., 2002).

COMORBIDITY OF SUBSTANCE USE, PSYCHIATRIC


DISORDERS AND SUICIDAL BEHAVIOUR

Comorbidity of psychiatric illnesses and substance misuse leads to an in-


crease in suicide attempts (Moselhy and Conlon, 2001). Alcohol abuse and
use of illegal drugs was reported in 69.7 percent of suicide attempters.
Comorbid depression and substance abuse is the most frequent category
in suicide attempters. Substance abuse is major comorbidity in bipolar
patients. Although rates decrease in older age groups, substance abuse is
still present in the elderly. According to the National Comorbidity Survey
(Kessler et al., 1997), the lifetime co-occurrence of psychiatric disorders
with alcohol abuse and dependence in men was 35.8 percent for anxiety
disorder (generalised anxiety disorder, panic, phobia), 28.1 percent for
mood disorder (major depression, dysthymia, mania), 29.5 percent for
drug dependence disorder and 16.9 percent for antisocial personality
disorder. Among women with alcohol dependence, 60.7 percent had an
anxiety disorder, 34.7 percent drug dependence disorder and 7.8 percent
antisocial personality disorder.
Various types of interactions are possible in case of comorbidity. The
first is a causal relationship in which the psychiatric disorder results in
alcohol and substance use disorder. On experiencing symptoms of psy-
chopathology, the individual resorts to greater consumption of the sub-
stance thus resulting in substance use disorder. Second, a psychiatric
disorder or symptoms may occur as a consequence of alcohol use, and
may persist after abstinence. Third, a person’s problems with alcohol and
psychiatric symptoms may develop simultaneously. Some of the risk
factors associated with one disorder may also be associated with the other.
In the fourth possibility, neither disorder may cause the other, but the
two become meaningfully linked over time. Last, it is also possible that
the disorders simply co-occur and co-exist with no meaningful inter-
relationship (Rosenthal and Westreich, 1999).

239
Nishi Misra et al.

Personality Disorder

Addiction was historically thought to be a symptom of personality disorder.


With the advent of DSM-III, substance use disorders were defined as a
clinical syndrome and separated from personality pathology. Since then,
psychiatric comorbidity in substance abusers is an area of research. Only
a few studies have however assessed the influence of Axis II disorders on
suicide attempts in substance use disorder (Koller et al., 2002; Ravandal
and Vaglum, 1999). Studies have found borderline personality disorder
(Darke et al., 2004; Ravandal and Vaglum, 1999) and antisocial personality
disorder (Koller et al., 2002; Roy et al., 1990) to be associated with suicide
attempts in this clinical group. The median rates for personality disorders
across different studies range from a low of 44 percent for alcohol de-
pendent individuals to a high of 79 percent for opioid dependent
patients, Cluster B personality disorders—antisocial, borderline and less
frequently narcissistic and histrionic personality disorders—are the most
prevalent.
In treated substance abusers, both Cluster C (avoidant and dependent
and less frequently obsessive compulsive disorder) and Cluster A (paranoid
and less frequently schizoid and shizotypal disorder) are most common.
Rounsaville et al. (1998) found Cluster B disorders to be most prevalent
(61 percent), followed by Cluster C (34 percent) and Cluster A
(22 percent). Antisocial (46 percent), borderline (30 percent) and avoidant
(20 percent) disorders were identified as the most common specific
personality disorders.

Affective Disorder

In a national epidemiological survey conducted in the United States, Grant


et al. (2005) reported that among the 9.35 percent of substance abusers,
19.67 percent had at least one independent mood disorder during the
same 12-month period. Nearly one-third of patients with major depressive
disorder also have substance use disorders, and it has been associated with
a greater risk of suicides and greater social and personal impairment (Davis
et al., 2008; Kessler et al., 1999).
Depressed alcoholics have significantly higher suicidality than subjects
with either depression or alcohol dependence. Alcohol dependence and

240
Substance Use and Suicidal Behaviour

depression act additively to produce greater suicide risk among subjects


with both disorders (Cornelius et al., 1996). Periods of major depression
during sustained abstinence are a risk factor for increased number of
suicide attempts over the lifetime. Major depression with onset before the
onset of substance dependence predicts severity of suicidal intent (Hasin
et al., 2002). Adults aged 18 or older who reported binge alcohol consumption
were more likely to report past year major depressive episode (MDE) than
their counterparts who had not engaged in binge drinking (8.7 versus
7.3 percent). Adults with past year MDE and past month binge alcohol use
reported more past year suicidal thoughts and suicide attempts than those
with MDE who did not binge drink. Rates of past year suicidal thoughts
and suicide attempts were also higher among adults with past year MDE
who had used illicit drugs during the past month than adults with past
year MDE who had not used illicit drugs (Substance Abuse and Mental
Health Services Administration, 2003).
Tondo et al. (1999) after a thorough review of literature reported
association of alcohol abuse with major affective disorders and of some
substance (polyabuse, alcohol, heroin, cocaine and even tobacco but not
marijuana or hallucinogens) with suicidal behaviour. The study reported
a tendency for bipolar I, mainly non-mixed patients, to have a relatively
high risk of substance abuse and low risk of suicide attempts, proving
that depressive or dysphoric (bipolar II, non-bipolar and bipolar I) are
more lethal. An epidemiological study (Kessler et al., 1997) reported an
association between alcoholism and mood disorders. A total of 60.7 percent
of people with bipolar I disorder had a lifetime diagnosis of a substance
use disorder; and 40.7 percent had a drug abuse diagnosis. Mania and
alcohol use disorders are 6.2 times more likely to occur than would be
expected by chance and they co-occur more often than do alcoholism and
unipolar depression.
Suicides in alcoholics are largely dependent on the co-occurrence of
a depressive episode. In a Veterans Administration Study, Lehmann,
McCormick and Mc Cracken (1995) found that 77 percent of suicide
completers who were diagnosed with substance abuse had a diagnosis of
affective disorder (39 percent). Of patients who completed suicide, 5 percent
had a comorbid psychosis and substance abuse; 67 percent of people with
Post Traumatic Stress Disorder (PTSD) who completed suicide had a
comorbid disorder, usually an affective disorder or substance abuse.

241
Nishi Misra et al.

Substance use disorders comorbid with other psychiatric disorders,


socio-demographic disadvantage and adverse childhood experiences are
associated with increased suicidal risk (Beautrais et al., 1996). More men
(49 percent) than women with bipolar disorder (29 percent) met the cri-
teria for lifetime alcoholism. Bipolar disorder is found to be frequently
comorbid with substance use disorders but it is difficult to diagnose among
patients with active substance use (Sloan et al., 2000).
When bipolar I disorder is compared with bipolar II disorder,
Chengappa, Levine, Gershon and Kupfer (2000) found 57.8 percent sub-
jects with bipolar I disorder to be dependent on one or more substances
or alcohol, 28.2 percent dependent on two substances or alcohol and
11.3 percent abused three or more substances or alcohol. Alcohol was the
only common drug among either bipolar I or bipolar II subjects. Bipolar
I subjects had higher rates of these comorbid conditions than bipolar II
subjects. Gender differences in lifetime prevalence of alcohol abuse and
bipolar illness in the same study revealed that more men (49 percent) than
women (29 percent) with bipolar disorder met the criteria for lifetime
alcoholism. However, the risk for having alcoholism was greater for women
with bipolar disorder.

Eating Disorders

Bulimia nervosa, binge eating disorder and alcohol/drug abuse are fre-
quently comorbid. Around 20–40 percent of females with bulimia also
report a history of problems with alcohol or drug abuse (Beary et al., 1986;
Hall et al., 1989). Amongst bulimic adolescents, substance use was found
to be related to an increased likeliness of attempted suicide, stealing and
sexual intercourse (Wiederman and Pryor, 1996). Stock, Goldberg, Corbett
and Katzman (2002) concluded that adolescents with restrictive eating
disorders used significantly less alcohol, tobacco and cannabis than the
general adolescent population. They also found lower use of substances in
adolescents with binging and purging symptoms. Personality traits, such
as impulsivity, were discovered as a common factor between bulimia and
substance abuse (Wiederman and Pryor, 1996). Guilt was regarded as one
of the emotions associated with both eating and alcohol abuse (Frank,
1991; Potter-Efron, 1989) and high rates of social anxiety (Striegel-Moore
et al., 1993).

242
Substance Use and Suicidal Behaviour

It is proposed that substance abuse provides relief from anxiety, de-


pression and other psychosocial problems, and an addictive personality
predisposes individuals to both eating disorders and alcohol abuse.
Individuals who develop an addiction to one substance may develop
psychological and behavioural patterns that leave them vulnerable to
develop addictions to other substances (Brisman and Siegel, 1984).
Eating disorder patients are already at an increased risk for morbidity
and mortality, so alcohol and drug use pose additional dangers for these
patients.

Anxiety Disorder

Goodwin and colleagues (2002) found prevalence of comorbid anxiety


disorders, which is an undetected and untreated problem, among patients
with severe affective disorders and substance use comorbidity. Thus, the
most frequently diagnosed comorbid Axis I conditions are anxiety and
mood disorders, while the most frequently observed Axis II disorders are
Cluster B-borderline personality disorder (BPD) and antisocial personality
disorder, followed by Cluster C, avoidant personality disorder (PD), passive
aggressive PD and obsessive compulsive PD; and then Cluster A; schizoid
PD. Polysubstance dependent subjects were more likely to be diagnosed
with anxiety disorders or bipolar disorder than were those who were
not polysubstance dependent or were dependent only on alcohol.
Polysubstance dependent men were at highest risk of Axis II disorders
(Salloum and Thase, 2000). Generalised anxiety disorder (GAD) serves
as a protective factor in suicide attempts in substance abusers. This may
be due to the fact that GAD is related to avoidant behaviour and greater
impulse control.

PREVENTIVE MANAGEMENT OF SUBSTANCE


USE AND SUICIDAL BEHAVIOUR

Individuals suffering from substance use disorders, including alcoholism,


are at increased risk for suicide; about 15–16 percent of which is higher
than the general population (Murphy, 2000). Presence of depression,

243
Nishi Misra et al.

impulsive behaviour and social and financial problems increase the risk.
Evidence suggests that comorbid psychiatric disorders are associated with
poorer alcohol treatment retention and outcomes (Kranzler et al., 1996).
Patients with personality disorders have high treatment dropout and
relapse rates (Nace et al., 1986). Since comorbidity of substance use
with other mental disorders is associated with additional risks and poor
outcomes, it is essential that patients with substance use disorder should
be screened for other mental health disorders.
Before referring to a drug and alcohol treatment plan, it is important
to consider whether the patient should be dealt with on an inpatient or
outpatient basis. Full risk assessment for suicidal attempts in such a person
and mental status examination are essential. Intervention is required as
early as possible in the initial course of the illness because as the illness
progresses, there is a need for more intense psychological and medical
treatment as well as group and residential therapeutic community care.
Managing in the ‘initial phase’ requires that the safety of the suicidal
person first needs to be looked into. In the absence of a comorbid mental
disorder, further contact may be unnecessary, although the opportunity
for further follow-up should be left open. For those who are profoundly
suicidal with a severe mental disorder, hospitalisation may be necessary. In
the ‘subsequent phase’, the person may be allowed to deal with his or her
current interpersonal difficulties by involving significant other people. In
the ‘later phase’, the person needs to be encouraged to use his or her coping
skills so that he or she is able to adapt to any future crises on his or her
own. If there are signs and symptoms of a mental disorder, antidepressant,
anti-anxiety or anti-psychotic medication may be provided to the suicidal
individual. It is also imperative to be aware of the potential risk of suicide
with such drugs because of the toxicity of the anti-depressant used.
Individuals with comorbid disorders present a confusing array of
symptoms. Social workers should therefore do continuous monitoring of
symptoms that may emerge later in the treatment process. If symptoms
disappear quickly during periods of abstinence from alcohol or substance
use, they are not suggestive of a comorbid psychiatric disorder. Re-
evaluation of symptoms after a period of abstinence is needed. In case the
symptoms were not present earlier, chances are there for the presence of
an independent psychiatric disorder. Monitoring of phenomenology, time,
course and aetiology is therefore needed (Rosenthal and Westreich, 1999).
Questions that need to be asked are: Does the consumption of substance

244
Substance Use and Suicidal Behaviour

provide relief or exacerbation of symptoms? Whether the amelioration


of psychiatric symptoms has an effect on pattern of substance use? Social
workers need to become acquainted with the symptoms of psychiatric
disorders and have knowledge of the effects of intoxication and withdrawal
of alcohol and other substances. They should be skilled in asking questions
about the drinking or substance abuse intake pattern.
Preventive management of substance use and suicide can be dealt at
three levels: primary, secondary and tertiary. Considering the prevention
of substance abuse suicidal cases at the primary prevention level requires
effective treatment of mood and other psychiatric disorders (Sher et al.,
2001) and modification of social, economic and biological conditions, like
reduction of poverty, violence, divorce rates and promotion of healthy
lifestyle (Maris, 2002). People with comorbid psychiatric disorders are
difficult to manage, relapse more often and adhere poorly to treatment
regimen (Sitharthan et al., 1999).Treating comorbid depression, which
accompanies in most of these cases is another viable option (Mason and
Kocsis, 1991).
At the secondary level, decreasing the likelihood of a suicide attempt
in the high-risk group is essential. It includes diagnosis and treatment of
existing psychiatric illnesses, assessment of suicide risk and the reduction
in access to lethal means (Mann, 2002). Useful questions in evaluation of
psychological risk are: How has the patient reacted in response to stress
in the past and how effective are his coping strategies? Has the patient
attempted suicide in the past? If so, how frequently and under what
circumstances? What are the similarities in the current circumstances?
Questioning about current depression and suicidal thoughts in addition to
one’s pattern of drinking and substance use are helpful. Providing support
at the home front and removal of any lethal means is another preventive
measure that needs to be adopted by family members (Brent et al., 1987).
Training for recognition of at-risk behaviour is necessary. Not only the
health workers but the family members and other people who come in
contact with the client, like the teachers and police officers, need training
in recognition of such behaviour. The American Association of Suicidology
(Rudd et al., 2006) has developed a mnemonic (is path warm?) to help
identify key warning signs for suicide. They are:

Ideation: talking about or threatening to kill oneself


Substance Abuse: increased substance abuse

245
Nishi Misra et al.

Purposelessness, Anxiety: anxiety, agitation or changes in sleep pattern


Trapped: feeling like there is no way out
Hopelessness, Withdrawal: from friends, family and society
Anger, Recklessness, Mood changes

Tertiary prevention is aimed at diminishing the effects of suicide


attempts through clinical treatment and rehabilitation. Education of
mental health professionals and assessment of family members who are
influenced by suicide attempt is needed.
Some additional points that need consideration in preventive
management of substance use and suicidal behaviour are as follows:

1. Implementing technically sound suicide prevention programmes at


both the state and national levels and improvement and expansion
of surveillance systems are needed. Ongoing collection of data at
national and state level drawing attention to the magnitude of pro-
blem and evaluating the impact of suicide prevention strategies is
needed.
2. Change in alcohol policies, like increasing the minimum legal
drinking age (MLDA) from 18 to 21 years is needed as an increase
in number of suicides in 18 to 20 years has been reported compared
to a 21 year MLDA.
3. Treatment should focus on prevention of comorbidity and pro-
vision of rehabilitation. A methodical screening and assessment
can ease the diagnostic challenge of distinguishing symptoms of
comorbid disorders from manifestations of substance intoxication
and withdrawal. Treatment should first focus on the medication
that is effective for treating co-occurring substance use.
4. The stigma of mental illness and substance abuse prevents people
from seeking help fearing prejudicial and discriminating behaviour.
The media can play a significant role in influencing public attitudes,
particularly if it is extensive, prominent, sensationalist and/or
explicitly describes the method of suicide. Individuals who belong
to the vulnerable group engage in imitative behaviour on reading
or watching coverage on suicide. The media can act as a preventive
means by educating the public about suicide, and by encouraging
those at risk of suicide to seek help. Changing media representation
of suicidal behaviour can help a lot in reducing suicide rates.

246
Substance Use and Suicidal Behaviour

5. Health messages that provide young people with information, life


skills and values are essential.
6. Better enforcement of regulations and strict penalties for the sale
and supply of tobacco and alcohol to minors is needed.
7. More research is needed to improve the interface of addiction
treatment with suicide interventions. Incorporating suicide
prevention into drug abuse treatment and looking especially at
the effects of intoxication and withdrawal on suicidal behaviours
are other areas of concern.

BARRIERS TO PREVENTIVE MANAGEMENT OF SUBSTANCE


USE AND SUICIDE

Some of the barriers to effective preventive management of substance use


and suicide are: poor integration of primary health and mental health
services, difficulties in making an appropriate diagnosis, marginalisation
and stigmatisation of clients and families, failure to involve families in
treatment plan and lack of training skills and commitment of staff. These
barriers, if overcome would result in better management and prevention
of this widespread problem and its ill effects.

CONCLUSION AND FURTHER RESEARCH

The focus of this chapter has been on the evidence for a link between
substance use and suicidal behaviour, as well as on the possible factors
that may mediate this relationship and suggestions on suicide prevention
and intervention specifically for substance abusers. While the relationship
between substance use and suicidal behaviour has been established
beyond doubt, there still remains ambiguity about several aspects of this
relationship. Methodological limitations in studies have resulted in a
lack of consistency among them, which makes it difficult to generalise
the findings. For instance, studies do not always distinguish between the
different forms of suicidal behaviour or between the levels of severity
of substance use. Studies are also often limited by the fact that alcohol

247
Nishi Misra et al.

and drugs are frequently both used together, which makes it difficult to
disentangle their independent contribution. And finally, most such studies,
relying on psychological autopsies and self-reports, may be limited by
the personally sensitive nature of the issue of suicide as well as substance
use; information may be withheld because of the illicit nature of these
behaviours or to avoid social undesirability.
Some areas of exploration for future research could be to study the
variables that link the two behaviours either as mediating causal factors,
or as moderating factors that influence the relationship between substance
use and suicidal behaviour. It may be interesting to study the interplay
between these different biological, personality, cognitive and psychosocial
variables, along with demographic variables, in determining the risk for
suicide. Does the risk of suicidal behaviour increase because of such as-
sociations? Other research questions that must be addressed are whether
the different substances of abuse are associated with differential risk for
suicidal behaviour and whether these associations are causal or merely
correlational.

REFERENCES

American Psychiatric Association (2003). Practice guideline for the assessment and
treatment of patients with suicidal behaviors. American Journal of Psychiatry, 160
(supplement 11), 1–60.
Andreasson, S. and P. Allebeck (1990). Cannabis and mortality among young men: A
longitudinal study of Swedish conscripts. Scandinavian Journal of Social Medicine,
18(1), 9–15.
Appleby, L., J. Cooper, T. Amos and B. Faragher (1999). Psychological autopsy study of
suicides by people aged under 35. British Journal of Psychiatry, 175(2), 168–74.
Arsenault-Lapierre, G., C. Kim and G. Turecki (2004). Psychiatric diagnoses in 3275
suicides: a metaanalysis. BMC Psychiatry, 4, 37.
Beary, M.D., J.H. Lacey and J. Merry (1986). Alcoholism and eating disorders in women
of fertile age. British Journal of Addiction, 81(5), 685–89.
Beautrais, A.L. (2003). Methodological issues in suicide research: Application of case
control and cohorts designs in the study of suicidal behaviours. In L. Vijaykumar
(Ed.), Suicide Prevention: Meeting the Challenge Together (pp. 68–78). Hyderabad,
India: Orient Longman.
Beautrais, A.L., P.R. Joyce and R.T. Mulder (1999). Cannabis use and serious suicide
attempts. Addiction, 94(8), 1155–64.

248
Substance Use and Suicidal Behaviour

Beautrais, A.L., P.R. Joyce, R.T. Mulder, D.M. Fergusson, B.J. Deavoll and
S.K. Nightingdale (1996). Prevalence and comorbidity of mental disorders in
persons making serious suicide attempts: A case-controlled study. American Journal
of Psychiatry, 153(8), 1009–14.
Beck, A.T., H.L.P. Resnik and D. Lettieri (Eds). (1973). Measurement of suicidal
behaviors. New York: Charles Press.
Bertolote, J.M., A. Fleischmann, D. De Leo and D. Wasserman (2004). Psychiatric
diagnoses and suicide: Revisiting the evidence. Crisis, 25(4), 147–55.
Birckmayer, J. and D. Hemenway (1999). Minimum-age drinking laws and youth
suicide, 1970–1990. American Journal of Public Health, 89(9), 1365–68.
Borges, G. and H. Rosovsky (1996). Suicide attempts and alcohol consumption in an
emergency room sample. Journal of Studies on Alcohol, 57(5), 543–48.
Borges, G., E.E. Walters and R.C. Kessler (2000). Associations of substance use,
abuse, and dependence with subsequent suicidal behaviour. American Journal of
Epidemiology, 151(8), 781–89.
Brent, D., J. Perper and C. J. Allman (1987). Alcohol, firearms, and suicide among
youth: Temporal trends in Allegheny County, Pennsylvenia, 1960 to 1983. Journal
of the American Medical Association, 257(24), 3369–72.
Brent, D., J. Perper, C. Goldstein and D. Kolko (1988). Risk factors for adolescent
suicide: A comparison of adolescent suicide victims with suicidal inpatients.
Archives of General Psychiatry, 45(6), 581–88.
Breslow, N.E. and N.E. Day (1980). Statistical methods in cancer research (Vol. 1).
The analysis of case control studies. Lyon: International Agency for Research
on Cancer.
Brisman, J. and M. Siegel (1984). Bulimia and alcoholism: Two sides of the same coin?
Journal of Substance Abuse Treatment, 1(2), 113–18.
Butler, G.K.L. and A.M.J. Montgomery (2004). Impulsivity, risk taking and recreational
‘ecstasy’ (MDMA) use. Drug and Alcohol Dependence, 76(1), 55–62.
Caces, F.E. and T. Harford (1998). Time series analysis of alcohol consumption and
suicide mortality in the United States, 1934–1987. Journal of Studies on Alcohol,
59(4), 455–61.
Chabrol, H., E. Chauchard and J. Girabet (2008). Cannabis use and suicidal behaviours
in high-school students. Addictive Behaviours, 33(1), 152–55.
Chabrol, H., J.D. Mabila and E. Chauchard (2008). Influence of cannabis use on suicidal
ideations among 491 high-school students. L’Encephale, 34(3), 270–73.
Chengappa, K.N., J. Levine, S. Gershon and D.J. Kupfer (2000). Lifetime prevalence
of substance or alcohol abuse and dependence among subjects with bipolar I and
II disorders in a voluntary registry. Bipolar Disorder, 2(3), 191–95.
Coccaro, E.F. (1989). Central serotonin and impulsive aggression. British Journal of
Psychiatry, 155(Supplement 8), 52–62.
Conner, K.R. and P.R. Duberstein (2004). Predisposing and precipitating factors for
suicide among alcoholics: Empirical review and conceptual integration. Alcoholism
Clinical and Experimental Research, 28(Supplement 5), 6S–17S.

249
Nishi Misra et al.

Cornelius, J.R., I.M. Salloum, N.L. Day, M.E. Thase and J. J. Mann (1996). Patterns of
suicidality and alcohol use in alcoholics with major depression. Alcoholism: Clinical
and Experimental Research, 20(8), 1451–55.
Darke, S. and S. Kaye (2004). Attempted suicide among injecting and noninjecting
cocaine users in Sydney, Australia. Journal of Urban Health, 81(3), 505–15.
Darke, S., J. Ross, M. Lynskey and M. Teesson (2004). Attempted suicide among
entrants to three treatment modalities for heroin dependence in the Australian
Treatment Outcome Study (ATOS): Prevalence and risk factors. Drug and Alcohol
Dependence, 73(1), 1–10.
Davis, L., A. Uezato, J.M. Newell and E. Frazier (2008). Major depression and co-morbid
substance use disorders. Current Opinion in Psychiatry, 21(1), 14–18.
Dembo, R., K. Pacheco, J. Schmeidler, L. Fisher and S. Cooper (1997). Drug use and
delinquent behavior among high risk youths. Journal of Child and Adolescent
Substance Abuse, 6(2), 1–25.
Donovan, J.M., S. Soldz, H.F. Kelley and W.E. Penk (1998). Four addictions: The MMPI
and discriminant function analysis. Journal of Addictive Diseases, 17(2), 41–55.
Engstrom, A., C. Adamsson, P. Allebeck and U. Rydberg (1991). Mortality in patients
with substance abuse: A follow-up in Stockholm County, 1973–1984. International
Journal of the Addictions, 26(1), 91–106.
Felts, W.M., T. Chenier and R. Barnes (1992). Drug use and suicide ideation and
behavior among North Carolina public school students. American Journal of
Public Health, 82(6), 870–72.
Fergusson, D.M., L.J. Horwood and N. Swain-Campbell (2002). Cannabis use and
psychosocial adjustment in adolescence and young adulthood. Addiction, 97(9),
1123–35.
Fowler, R.C., C.L. Rich and D. Young (1986). San Diego suicide study: II. Substance
abuse in young cases. Archives of General Psychiatry, 43(10), 962–65.
Frank, E.S. (1991). Shame and guilt in eating disorders. American Journal of
Orthopsychiatry, 61(2), 303–06.
Garlow, S.J., D. Purselle and B. D’Orio (2003). Cocaine use disorders and suicidal
ideation. Drug and Alcohol Dependence, 70(1), 101–04.
Giancola, P.R., E.L. Helton, A.B. Osborne, M.K. Terry, A.M. Fuss and J.A. Westerfield
(2002). The effects of alcohol and provocation on aggressive behavior in men and
women. Journal of Studies on Alcohol, 63(1), 64–73.
Goodwin, R.D., D.A. Stayner, M.J. Chinman, P. Wu, J.K. Tebes and L. Davidson (2002).
The relationship between anxiety and substance use disorders among individuals
with severe affective disorders. Comprehensive Psychiatry, 43(4), 245–52.
Grant, B.F., F.S. Stimson, D.S. Hasin, D.A. Dawson, S.P. Chou, W.J. Ruan et al. (2005).
Prevalence, correlations and comorbidity of bipolar I disorder and Axis I and II
disorder: Research from National Epidemiological Survey on alcoholism and
related conditions. Journal of Clinical Psychiatry, 66(10), 1205–15.
Guy, S.M., G.M. Smith and P.M. Bentler (1994). Consequences of adolescent drug
use and personality factors on adult drug use. Journal of Drug Education, 24(2),
109–32.

250
Substance Use and Suicidal Behaviour

Hall, R.C., T.P. Beresford, B. Wooley, L. Tice and A.K. Hall (1989). Covert drug abuse
in patients with eating disorders. Psychological Medicine, 7(4), 247–55.
Harris, E.C. and B. Barraclough (1997). Suicide as an outcome for mental disorders.
British Journal of Psychiatry, 170(3), 205–28.
Harter, S., C. Marold and N. Whitesell (1992). Model of psychosocial risk factors leading
to suicidal ideation in young adolescents. Development and Psychopathology,
4(1), 167–88.
Hasin, D., X. Liu, E. Nunes, S. McCloud, S. Samet and J. Endicott (2002). Effects of
major depression on remission and relapse of substance dependence. Archives of
General Psychiatry, 59(4), 375–380.
Henriksson, M.M., H.M. Aro, M.J. Marttunen, M.E. Heikkinen, E.T. Isometsa, K.I.
Kuoppasalmi et al. (1993). Mental disorders and comorbidity in suicide. American
Journal of Psychiatry, 150, 935–40.
Hillman, S.D., S.R. Silburn, A. Green and S.R. Zubrick (2000). Youth suicide in Western
Australia involving cannabis and other drugs. Perth: Western Australian Drug
Abuse Strategy Office.
Hoaken, P.N.S. and R.O. Pihl (2000). The effects of alcohol intoxication on aggressive
responses in men and women. Alcohol and Alcoholism, 35(5), 471–77.
Horesh, N., D. Gothelf, H. Ofek, T. Weizman and A. Apter (1999). Impulsivity as a
correlate of suicidal behaviour in adolescent psychiatric inpatients. Crisis, 20(1),
8–14.
Inskip, H.M., E.C. Harris and B. Barraclough (1998). Lifetime risk of suicide for
affective disorder, alcoholism and schizophrenia. British Journal of Psychiatry,
172(1), 35–37.
Kessler, R.C., G. Borges and E.E. Walters (1999). Prevalence of and risk factors for
lifetime suicide attempts in the National Comorbidity Survey. Archives of General
Psychiatry, 56(7), 617–26.
Kessler, R.C., R.M. Crum, L.A. Warner, C.B. Nelson, J. Schulenberg and J.C. Anthony
(1997). Lifetime co-occurrence of DSM-III-R alcohol abuse and dependence with
other psychiatric disorders in the national comorbidity survey. Archives of General
Psychiatry, 54(4), 313–21.
Koller, G., U.W. Preuss, M. Bottlender, K. Wenzel and M. Soyka (2002). Impulsivity
and aggression as predictors of suicide attempts in alcoholics. European Archives
of Psychiatry and Clinical Neuroscience, 252(4), 155–60.
Korn, M.L., A.J. Botsis, M. Kotler, R. Plutchik, H.R. Conte, G. Finkelstein et al. (1992).
The suicide and aggression survey: A semistrcutured instrument for the measure-
ment of suicidality and aggression. Comprehensive Psychiatry, 33(6), 359–65.
Kranzler, H.R., F.K. Del Boca and B.J. Rounsaville (1996). Comorbid psychiatric
diagnosis predicts three-year outcomes in alcoholics: A post-treatment natural
study. Journal of Studies on Alcohol, 57(6), 619–26.
Kung, H.C., J.L. Pearson and X. Liu (2003). Risk factors for male and female suicide
decedents ages 15–64 in the United States. Results from the 1993 National
Mortality Followback Survey. Social Psychiatry and Psychiatric Epidemiology,
38(8), 419–26.

251
Nishi Misra et al.

Lehmann, L., R.A. McCormick and L. Mc Cracken (1995). Suicidal behaviour among
patients in VA Health Care System. Psychiatric Services, 46(10), 1069–71.
Lester, D. and A.T. Beck (1975). Attempted suicide in alcoholics and drug addicts.
Journal of Studies on Alcohol, 36(1), 162–64.
Linnoila, M., M. Virkkunen, T. George and D. Higley (1993). Impulse control disorders.
International Clinical Psychopharmacology, 8(Supplement 1), 53–56.
Lynskey, M.T., A.L. Glowinski, A.A. Todorov, K.K. Bucholz, P.A.F. Madden, E.C. Nelson
et al. (2004). Major depressive disorder, suicidal ideation, and suicide attempt
in twins discordant for cannabis dependence and early-onset cannabis use.
Archives of General Psychiatry, 51, 1026–32.
Malone, K.M. (1999). Is there a biology of suicide? Irish Journal of Psychological
Medicine, 16(4), 121–22.
Mann, J.J. (2002). A current perspective of suicide and attempted suicide. Annals of
Internal Medicine, 136(4), 302–11.
Mann, J.J., M. Oquendo, M.D. Underwood and V. Arango (1999). The neurobiology
of suicide risk: A review for the clinician. Journal of Clinical Psychiatry, 60
(Supplement 2), 7–11.
Maris, R.W. (2002). Suicide. Lancet, 360(9329), 319–26.
Mason, B.J. and J.H. Kocsis (1991). Desipramine treatment of alcoholism.
Psychopharmacological Bulletin, 27(2), 155–61.
Marttunen, M., H. Aro, M. Henriksson and J. Loennqvist (1991). Mental disorders in
adolescent suicide: DSM-III—R Axes I and II diagnoses in suicides among 13- to
19-year-olds in Finland. Archives of General Psychiatry, 48(9), 834–38.
Marttunen, M.I., H.M. Aro and J.K. Lönnqvist (1992). Adolescent suicide: Endpoint
of long-term difficulties. Journal of the American Academy of Child and Adolescent
Psychiatry, 31(4), 649–54.
Marzuk, P.M., K. Tardiff, A.C. Leon, M. Stajic, E.B. Morgan and J.J. Mann (1992).
Prevalence of cocaine use among residents of New York City who committed suicide
during a one-year period. American Journal of Psychiatry, 149(3), 371–75.
McCann, U.D., A. Ridenour, Y. Shaham and G.A. Ricaurte (1994). Serotonin
neurotoxicity after (±)-3, 4-methylenedioxymethamphetamine (MDMA;
“Ecstasy”): A controlled study in humans. Neuropsychopharmacology, 10(2),
129–38.
Meninger, K. (1938). Man against himself. New York: Harcourt, Brace & World.
Molnar, B.E., L.F. Berkman and S.L. Buka (2001). Psychopathology, childhood sexual
abuse and other childhood adversities: Relative links to subsequent suicidal
behavior in the U.S. Psychological Medicine, 31(6), 965–77.
Moselhy, H.F and W. Conlon (2001). Natural history of affective disorders: Comorbidity
as a predictor of suicide attempts. International Journal of Psychiatry in Clinical
Practice, 5(3), 203–06.
Murphy, G.E. (2000). Psychiatric aspects of suicidal behaviour: Substance abuse. In
K. Hawton and K. Van Heeringen (Eds), International handbook of suicide and
attempted suicide (pp. 135–46). Chichester: John Wiley and Sons.

252
Substance Use and Suicidal Behaviour

Murphy, G.E., R.D. Wetzel, E. Robins and L. Mcevoy (1992). Multiple risk factors
predict suicide in alcoholism. Archives of General Psychiatry, 49(6), 459–63.
Nace, E.P., J.J. Saxon, Jr. and N. Shore (1986). Borderline personality disorder and
acoholism treatment: A one-year follow-up study. Journal of Studies on Alcohol,
47(3), 196–200.
Neeleman, J. and M. Farrell (1997). Suicide and substance misuse. British Journal of
Psychiatry, 171(4), 303–04.
Norstrom, T. (1995). Alcohol and suicide: A comparative analysis of France and
Sweden. Addiction, 90(11), 1463–69.
Norstrom, P., D. Shalling and M. Asberg (1995). Temperamental vulnerability in
attempted suicide. Acta Psychiatrica Scandinavica, 92(2), 155–60.
O’Doherty, M. and A. Farrington (1997). Estimating local opioid addict mortality.
Addiction Research, 4(4), 321–27.
Ohberg, A., E. Vuori, I. Ojanpera and J. Lonnqvist (1996). Alcohol and drugs in suicides.
British Journal of Psychiatry, 169(1), 75–80.
Ornstein, T.J., J.L. Iddon, A.M. Baldacchino, B.J. Sahakian, M. London, B.J. Everitt
et al. (2000). Profiles of cognitive dysfunction in chronic amphetamine and heroin
abusers. Neuropsychopharmacology, 23, 113–26.
Oyefeso, A., H. Ghodse, C. Clancy and J.M. Corkery (1999). Suicide among drug addicts
in the UK. British Journal of Psychiatry, 175(3), 277–82.
Patsiokas, A.T., G.A. Clum and R.L. Luscomb (1979). Cognitive characteristics of suicide
attempters. Journal of Consulting and Clinical Psychology, 47(3), 478–84.
Perry, J.L. and M.E. Carroll (2008). The role of impulsive behavior in drug abuse.
Psychopharmacology, 200(1), 1–26.
Petronis, K., J. Samuels, E. Moscicki and J. Anthony (1990). An epidemiologic
investigation of potential risk factors for suicide attempts. Social Psychiatry and
Psychiatric Epidemiology, 25(4), 193–99.
Pirkola, S.P., E.T. Isometsa, M.E. Heikkinen and J.K. Lönnqvist (2000). Suicides of
alcohol misusers and non-misusers in a nationwide population. Alcohol and
Alcoholism, 35(1), 70–75.
Plutchik, R. and H. van Praag (1989). The measurement of suicidality, aggressivity
and impulsivity. Progress in Neuropsychopharmacology and Biological Psychiatry,
13(Supplement), 23–34.
Potter-Efron, R.T. (1989). Guilt, shame and alcoholism: Treatment issues in clinical
practice. New York: Haworth Press.
Preuss, U.W., M.A. Schuckit, T.L. Smith, G.P. Danko, K. Buckman, L. Bierut et al.
(2002). Comparison of 3190 alcohol-dependent individuals with and without
suicide attempts. Alcoholism, Clinical and Experimental Research, 26, 471–77.
Ravandal, E. and P. Vaglum (1999). Overdoses and suicide attempts: different relations
to psychopathology and substance abuse? A 5-year prospective study of drug
abusers. European Addiction Research, 5(2), 63–70.
Ricaurte, G.A., K.T. Finnegan, I. Irwin and J.W. Langston (1990). Aminergic metabolites
in cerebrospinal fluid of humans previously exposed to MDMA: Preliminary
observations. Annals of the New York Academy of Sciences, 600(1), 699–710.

253
Nishi Misra et al.

Rich, C.L., D. Young and R.C. Fowler (1986). San Diego suicide study I: Young vs old
subjects. Archives of General Psychiatry, 43(6), 577–82.
Robins, E. and G. Murphy (1965). The physician’s role in the prevention of suicide. In
L. Yochelson (Ed.), Symposium on suicide (pp. 84–91). Washington, DC: George
Washington University Press.
Rosenthal, R.N. and L. Westreich (1999). Treatment of persons with dual diagnoses
of substance use disorder and other psychological problems. In B.S. Mc Crady
and E.E. Epstein (Eds), Addictions: A Comprehensive Guidebook (pp. 439–76).
New York: Oxford University Press.
Rossow, I. (1993). Suicide, alcohol and divorce: aspects of gender and family
integration. Addiction, 88(12), 1659–65.
Rounsaville, B.J., H.R. Kranzler, S. Ball, H. Tennen, J. Poling and E. Triffleman (1998).
Personality disorder in substance abusers: Relationship to substance use. Journal
of Nervous and Mental Diseases, 186(2), 87–95.
Roy, A. (2001). Characteristics of cocaine-dependent patients who attempt suicide.
American Journal of Psychiatry, 158(8), 1215–19.
Roy, A., D. Lamparski, J. DeJong, V. Moore and M. Linnoila (1990). Characteristics of
alcoholics who attempt suicide. American Journal of Psychiatry, 147(6), 761–65.
Rudd, M.D., A.L. Berman, T.E . Joiner, Jr., M.K. Nock, M.M. Silverman, M. Mandrusiak
et al. (2006). Warning signs for suicide: Theory, research and clinical applications.
Suicide and Life-Threatening Behaviour, 36, 255–62.
Rudd, M.D. and T. Joiner (1998). The assessment, management, and treatment of
suicidality: Toward clinically informed and balanced standards of care. Clinical
Psychology Review, 5(2), 135–50.
Salloum, I.M. and M.E. Thase (2000). Impact of substance abuse on the course and
treatment of bipolar disorder. Bipolar Disorder, 2(3, Part 2), 269–80.
Schuckit, M.A. (1986). Primary men alcoholics with histories of suicide attempts.
Journal of Studies on Alcohol, 47(1), 78–81.
Serpi, T.L., B. Wiersma, H. Hackman, L. Ortega, B.J. Jacquemin, K.S. Weintraub et al.
(2005). Homicide and suicide rates-National Violent Death Reporting System, six
states, 2003. Morbidity and Mortality Weekly Report, 54, 377–80.
Shafii, M., S. Carrigan, J.R. Whittinghill and A. Derrick (1985). Psychological autopsy
of completed suicide in children and adolescents. American Journal of Psychiatry,
42(9), 1061–1106.
Sher, L., M.A. Oquendo and J. J. Mann (2001). Risk of suicide in mood disorders.
Clinical Neuroscientific Research, 1(5), 337–44.
Sitharthan, T., S. Singh, P. Kranitis, J. Currie, P. Freeman, G. Murugesan et al. (1999).
Integrated drug and alcohol intervention: Development of an opportunistic
intervention program to reduce alcohol and other substance use among psychiatric
patients. Australian and New Zealand Journal of Psychiatry, 33, 676–83.
Skog, Ole-Jorgen. (1993). Alcohol and Suicide in Denmark 1911–1924—Experiences
from a Natural Experiment. Addiction, 88(9), 1189–93.
Skog, Ole-Jorgen, Z. Teixeira, J. Barrias and R. Moreira (1995). Alcohol and suicide—
portuguese experience. Addiction, 90(8), 1053–61.

254
Substance Use and Suicidal Behaviour

Sloan, K.L., B. Kivlahan and A.J. Saxon (2000). Detecting bipolar disorders among
treatment seeking substance abusers. American Journal of Drug and Alcohol Abuse,
26(1), 13–23.
Stack, S. (2000). Suicide: A 15-year review of the sociological literature. Part II:
modernization and social integration perspectives. Suicide and Life-Threatening
Behavior, 30(2), 163–76.
Stock, S.L., E. Goldberg, S. Corbett and D.K. Katzman (2002). Substance use in female
adolescents with eating disorders. Journal of Adolescent Health, 31(2), 176–82.
Striegel-Moore, R.H., L.R. Silberstein and J. Rodin (1993). The social self in bulimia
nervosa: public self-consciousness, social anxiety and perceived fraudulence.
Journal of Abnormal Psychology, 102(2), 297–303.
Substance Abuse and Mental Health Services Administration (2003). The NHSDA
report: Substance use and the risk of suicide among youths. Retrieved 14 November
2008 from http://oas.samhsa.gov /2k2/suicide/suicide.htm
Suokas, J. and J. Lonnqvist (1995). Suicide attempts in which alcohol is involved: A
special group in general hospital emergency rooms. Acta Psychiatrica Scandinavica,
91(1), 36–40.
Tondo, L., R.J. Baldessarini, J. Hennen, G.P. Minna, P. Salis, L. Scamonatti et al. (1999).
Suicide attempts in major affective disorder patients with comorbid substance
use disorders. Journal of Clinical Psychiatry, 60, 63–69.
Vijayakumar, L. and S. Rajkumar (1999). Are risk factors for suicide universal? A case-
control study in India. Acta Psychiatrica Scandinavica, 99(6), 407–11.
Wasserman, D., A. Varnik and G. Eklund (1998). Female suicides and alcohol con-
sumption during perestroika in the former USSR. Acta Psychiatrica Scandinavica
supplement, 394, 26–33.
Weissman, M.M. (1974). The epidemiology of suicide attempts, 1960 to 1971. Archives
of General Psychiatry, 30(6), 737–46.
Whitters, A.C., R.J. Cadoret and R.B. Widmer (1985). Factors associated with suicide
attempts in alcohol abusers. Journal of Affective Disorders, 9(1), 19–23.
Wiederman, M. W. and T. Pryor (1996). Substance use in impulsive behaviours among
adolescents with eating disorders. Addictive Behaviour, 21(2), 269–72.
Wilcox, H.C., K.R. Conner and E.D. Caine (2004). Association of alcohol and drug
use disorders and completed suicide: An empirical review of cohort studies. Drug
and Alcohol Dependence, 76 (Supplement), 11–19.
Williams, J.M.G. (1997). Cry of pain: Understanding suicide and self harm.
Harmondsworth, Middlesex, England: Penguin Books.
World Health Organization. (2001). World Health Organization: Suicide prevention.
Retrieved 26 October 2005 from http://www.who.int/mental_health/prevention/
suicide/suicideprevent/en/
Zorko, M., A. Marusic, Z. Cebasek-Travnik and V. Bucik (2004). The frontal lobe
hypothesis: Impairment of executive cognitive functions in chronic alcohol in-
patients. Psychiatria Danubina, 16(1/2), 21–28.

255
12

Suicide Risk in Bipolar Disorder


Maurizio Pompili, Marco Innamorati,
Enrica De Simoni, Ilaria Falcone,
Gaspare Palmieri, Laura Sapienza
and Roberto Tatarelli

B ipolar disorders (BD) are of particular public health significance as they


are prevalent, severe and disabling. A review of the literature, published
between 1988 and 2002, indicated that the lifetime prevalence rate of DSM-
III or DSM-III-R diagnosed bipolar I (BD-I) and bipolar II (BD-II) in the
population of the United States, the Netherlands and Hungary was 0.8–3.0,
0.2–2.0, and 4.4–15.8, respectively (Rihmer and Angst, 2005). Miklowitz
and Johnson (2006), using data from the National Comorbidity Survey
replication, estimated a lifetime prevalence rate of 3.9 percent for BD-I and
BD-II (Kessler et al., 2005). However, prevalence of BDs much depends
on the different criteria for diagnosis used in the different studies;
including subsyndromal forms and using a BD spectrum dramatically
increases the prevalence rates (Akiskal, 1996; Faravelli et al., 2006; Judd and
Akiskal, 2003; Pompili et al., 2006; Rihmer and Angst, 2005). Research
indicates that when subthreshold hypomania (i.e., subthreshold BD-II) is
also considered, the combined lifetime prevalence of BD-I and BD-II dis-
orders is more than 5 percent (Rihmer and Angst, 2005). Women and men
are equally likely to develop BD-I, although women report more episodes
of depression than men and, correspondingly, are more likely to meet or
resemble the criteria for BD-II (Leibenluft, 1997; Schneck et al., 2004).

256
Suicide Risk in Bipolar Disorder

The National Comorbidity Survey replication study reported that the


median age of onset for BD was 25 years. Approximately 25 percent of
patients had onset by age 17. The onset of BD appears to be occurring at
a younger age with successive birth cohorts (Rice et al., 1987; Ryan et al.,
1992; Wickramaratne et al., 1989).
BD is often associated with elevated risks of premature mortality
(Muller-Oerlinghausen et al., 2002), adverse outcomes of medical
disorders, accidents and complications from comorbid substance use
disorders. However, the major source of premature mortality in BD
patients is suicide (Ahrens et al., 1995; Harris and Barraclough, 1997;
Rihmer, 2005; Tondo and Baldessarini, 2005; Tondo et al., 2003). The risk
of suicide is approximately 400–1400/100,000/year (about 0.9 percent/
year), or 25–90 times higher than the general population rate of 0.015
percent/year (Baldessarini, Pompili and Tondo, 2006a; Baldessarini et al.,
2006; Tondo and Baldessarini, 2005; Tondo et al., 2003). Suicide accounts
for 15–20 percent of deaths among BD patients (Baldessarini, Pompili
and Tondo, 2006a; Baldessarini et al., 2006; Goodwin and Jamison, 2007;
Tondo and Baldessarini, 2005). In a 40-year follow-up of 406 BD-I and
BD-II patients admitted to the University Psychiatric Hospital of Zurich
between 1959 and 1963, 11 percent of the patients committed suicide
(Angst et al., 2005). As many as 50 percent of people with BD attempt
suicide during their lifetime (Jamison, 2000).
According to Akiskal (2007), the higher prevalence of suicide risk and
acts among BD patients is due to the fact that BD-II accounts for 30–58
percent of all major depressions in psychiatric practice (versus a 5 percent
prevalence of BD-II spectrum in the general population). Mixed states
are very common in the BD-II spectrum even though the term mixed
refers to mania and not hypomania. The cyclothymic temperament is
the most prevalent temperament in BD-II. Rapid mood shifts are the
hallmark of BD-II, and BD-II suicides are the most lethal—they use the
most aggressive methods.

Bipolar depression

We combined data from four studies (Isometsa et al., 1994; Rouillon


et al., 1991; Tondo et al., 1998; Valtonen et al., 2005) and found that most

257
Maurizio Pompili et al.

suicides and suicide attempts (78–89 percent) occurred during times of


major depression (pure or mixed depression), about 0–7 percent during
the euthymic phase and 11–20 percent during the dysphoric (mixed)
mania phase.
The most robust short-term predictors of completed suicide are
a strong wish to die, suicidal ideation or plans, the communication
of suicide intent during depression and having few reasons for living
(Hawton and van Heeringen, 2000; Isometsa et al., 1994; Oquendo
et al., 2000). BD depressives with a history of suicide attempts have a
more severe symptomatology in general, report more hopelessness, self-
blame, guilt and current suicidal ideation, and have more aggressive/
impulsive personality traits than non-suicidal BD depressives (Bulik
et al., 1990; Fagiolini et al., 2004; Leverich et al., 2003; Lopez et al., 2001;
Oquendo et al., 2000, 2004).
Recently, it has also been demonstrated that mixed depressive episodes
(major depression plus three or more hypomanic symptoms, which
is twice as common in BD-II depression than in unipolar depression
[Benazzi, 2006] or agitated depression), that are greatly overlapping
(almost identical) conditions (Akiskal and Benazzi, 2003; Akiskal et al.,
2005; Benazzi et al., 2004; Maj et al., 2003), also increase the risk of all
forms of suicidal behaviour in both unipolar and BD patients (Akiskal
et al., 2005). Patients with mood disorders who exhibit volatile and erratic
moods associated with dysphoria and agitation or who present the classic
mixed states are at particularly high suicide risk.
Akiskal et al. (2005) investigated the clinical and familial characteristics
of patients with agitated and non-agitated unipolar depression. They found
that depressive mixed states, distractibility, racing/crowded thoughts,
talkativeness, weight loss, suicidal ideation and a familiy history of BD-II
in first degree relatives were significantly more common in agitated than
in non-agitated patients with unipolar depressive disorders, supporting
the bipolar nature of agitated unipolar depression. Among the depressive
mixed symptoms, they found an association between suicidal ideation,
psychomotor activation and racing thoughts. A forward stepwise logistic
regression found that agitated depression was predicted by depressive
mixed states, talkativeness and suicidal ideation as the independent,
significant, positive predictors.
Suicidal behaviour in BD patients is not restricted exclusively to
depressive episodes. Out of the 576 consecutively admitted BD-I manic

258
Suicide Risk in Bipolar Disorder

patients, 51 (9 percent) had suicidal ideation at admission, including


13 patients (3 percent) who attempted suicide during the index episode
(Sato et al., 2004). However, if a distinction is made between dysphoric
(mixed) and pure mania, it has been found that 26–55 percent of patients
with mixed (dysphoric) mania had current suicidal ideation, in contrast
to 2–7 percent of patients with pure mania (Dilsaver et al., 1994; Sato et
al., 2004; Strakowski et al., 1996). A detailed analysis of the psychatric
history of 31 consecutive suicide victims with BD-I showed that, in the
vast majority of cases, the suicide occurred during a major depressive
episode, and only in a few cases (11 percent) did it occur during a mixed
(dysphoric) mania (Isometsa et al., 1994).
Fiedorowicz et al. (2008) studied suicidality in participants with major
affective disorders in the National Institute of Mental Health (NIMH)
Collaborative Depression Study (CDS) who were followed prospectively
for up to 25 years. A total of 909 participants meeting prospective diag-
nostic criteria for major depressive and BD were followed through 4,204
mood cycles. Suicidal behaviour was defined as suicide attempts or com-
pletions. Mixed-effects, grouped-time survival analysis assessed the risk
of suicidal behaviour and the differential effects of risk factors for suicidal
behaviour by polarity. In addition to polarity, the main effects of age,
gender, hopelessness, marital status, prior suicide attempts and active
substance abuse were modelled, with mood cycle as the unit of analysis.
These authors found that, after controlling for age of onset, there were
no differences in prior suicide attempts by polarity although bipolar
participants had more prior severe attempts. During follow-up, 40 cycles
ended in suicide and 384 cycles contained at least one suicide attempt.
Age, hopelessness and active substance abuse, but not polarity, predicted
suicidal behaviour. They concluded that the effects of risk factors did not
differ by polarity.
Valtonen and colleagues (2005) investigated the incidence of suicide
attempts in different phases of BD as a part of a naturalistic, prospective,
18-month study representing psychiatric inpatients and outpatients
with DSM-IV BD. These authors found that, compared to the other
phases of the illness, the incidence of suicide attempts was 37-fold higher
(95 percent confidence interval [CI] for relative risk [RR]: 11.8–120.3)
during combined mixed and depressive mixed states, and 18-fold higher
(95 percent CI: 6.5–50.8) during major depressive phases. In Cox’s
proportional hazards regression models, combined mixed (mixed or

259
Maurizio Pompili et al.

depressive mixed) or major depressive phases and prior suicide attempts


independently predicted suicide attempts. No other factor significantly
modified the risks related to these time-varying risk factors; their
population-attributable fraction was 86 percent. These data indicate that
the incidence of suicide attempts varies remarkably between illness phases,
with mixed and depressive phases involving the highest risk by time. Time
spent in high-risk illness phases is likely the major determinant of overall
risk for suicide attempts among BD patients.

Substance abuse

BD is often comorbid with substance abuse disorder (Baldessarini,


Pompili and Tondo, 2006a; 2006b). In the Epidemiologic Catchment Area
(ECA) study of substance abuse, Regier et al. (1990) found high lifetime
prevalence rates of substance abuse or dependence in individuals with
BD. These authors estimated that up to 56 percent of the subjects with
BD had substance abuse or dependence and 44 percent had some type
of alcohol-related diagnosis. People with dual disorders are best served
when common aetiology, risk factors and treatments are assumed for the
combined syndrome. Individuals with dual disorders frequently have their
psychiatric symptoms misunderstood in substance abuse programmes or
their substance abuse problems ignored at mental health centres. Substance
abuse or dependence is associated with a higher suicide risk (Harris and
Barraclough, 1997; Pompili, Lester et al., 2007).
The comorbidity between two disorders both of which are associated
with increased suicide risk is often the mixture that precipitates the
suicide. Substance abuse mediates increased suicide risk by determining
poorer outcome. Some of the factors that predispose to poorer outcome
are listed in Table 12.1.
Substance abuse may also result in a number of impairments that
contribute to suicide risk such as worse cognitive performance, impairment
of working memory, verbal learning and memory deficits, and lack of
appropriate orientation engaging in problem-solving (Pompili, Lester
et al., 2007).
Akiskal et al. (2003) reported that, compared to non-cyclothymic BD-II
patients, cyclothymic BD-II patients reported significantly more lifetime

260
Suicide Risk in Bipolar Disorder

Table 12.1â•… Factors Contributing to Increased Suicide Risk in Cases of Comorbid


Substance Abuse Disorder and Bipolar Disorder

 Earlier onset
 Several relapses
 Treatment non-adherence and more side effects from treatment
 Poor response to medication
 More hospitalisations
 Increased risk for violence
 Increased medical costs
 Dysphoric-irritable mixed states
Source: Authors.

suicide attempts (49 percent versus 38 percent) and had more current
hospitalisations for suicidal risk (61 percent versus 50 percent). During
a 2–4 year prospective follow-up of 80 juvenile inpatients with a current
major depressive episode, having a cyclothymic-sensitive temperament
at baseline, significantly predicted not only the bipolar outcome but also
suicidal behaviour during the follow-up. Among these young patients
81 percent of those with a cyclothymic-sensitive temperament had at least
one episode of suicidal ideation or attempt versus 36 percent of subjects
without such a temperament (Kochman et al., 2005). Investigating the
affective temperament profiles of 150 non-violent suicide attempters
(121 of them with a current major depressive episode) and 717 healthy
controls, research (Rihmer et al., 2007) indicated that, compared to con-
trols, suicide attempters scored significantly higher on four of the five
affective temperaments (depressive, cyclothymic, irritable and anx-
ious temperaments). On the other hand, no significant difference between
the suicide attempters and controls was found for the hyperthymic
temperament.
Maser et al. (2002) investigated temperament in completed suicides
and attempted suicides. They found that attempters and completers
shared core characteristics: previous attempts, impulsivity, substance
abuse and psychic turmoil within a cycling/mixed BD. The temperament
traits of impulsivity and assertiveness were the best prospective predictors
of completed suicides beyond 12 months with a sensitivity level of
74 percent and specificity level of 82 percent.
Pompili, Rihmer et al. (2008) investigated 150 psychiatric inpatients
with BD-I, BD-II, Major Depressive Disorder (MDD) and psychotic
disorders for temperament, personality traits and suicide risk. The patients

261
Maurizio Pompili et al.

were administered the Temperament Evaluation of Memphis, Pisa,


Paris and San Diego autoquestionnaire-Rome (TEMPS-A-Rome), the
Minnesota Multiphasic Personality Inventory, 2nd edition (MMPI-2) and
the Beck Hopelessness Scale (BHS) and evaluated for suicide risk through
the critical items of the Mini International Neuropsychiatric Interview
(MINI). Irritable temperament and social introversion were the strongest
predictors of suicide risk, and the hyperthymic temperament appeared to
be a protective factor both for hopelessness and suicide risk.
Previous research has led to the hypothesis that patients with white
matter hyperintensities in the brain may be at higher risk for suicide
because of possible disruption of neuroanatomic pathways (Taylor
et al., 2001). Mood regulation depends on the complex system composed
of the prefrontal cortex, amygdala-hippocampus complex, thalamus,
and basal ganglia and the extensive connections between these areas
(Soares and Mann, 1997). Lesions in one specific part or disruption
of interconnections can result in malfunctions in other areas. Mood
regulation abnormalities could confer a biological vulnerability which, in
combination with environmental stressors, results in suicidal behaviour.
The aetiology of white matter hyperintensities with respect to suicidality
and mood disorders could be hypoxic-ischemic insults during birth which
are especially common in premature infants. Perinatal white matter lesions
represent damage of association-commissural and projection fibres and
may lead to disturbances in the organisation and use of neural systems
(Judas et al., 2005; Peterson, 2003). Pompili and colleagues (Pompili,
Ehrlich et al., 2007, Pompili, Innamorati et al., 2008) conducted two
studies on inpatients with MDD or BD. The authors found an increased
prevalence of white matter hyperintensities in adults with major affective
disorders and a history of suicide attempt, compared to similar patients
without such a history. In the most recent study (Pompili, Innamorati
et al., 2008), the presence of periventricular white matter hyperintensity
was robustly associated with suicidal behaviours even after controlling
for age (odds ratio: 8.08).

Suicide attempts

Attempted suicide, particularly in the case of BD, is one of the most


powerful clinical predictors of repeated attempts (Oquendo et al., 2004;

262
Suicide Risk in Bipolar Disorder

Slama et al., 2004) and of completed suicide (Goodwin and Jamison, 2007;
Hawton and van Heeringen, 2000; Isometsa et al., 1994). About one-third
of BD patients have a lifetime history of one or more suicide attempts
(Table 12.2), and up to 56 percent of suicides with BD have made at least
one prior suicide attempt (Goodwin and Jamison, 2007; Isometsa et al.,
1994; Romero et al., 2007).

Table 12.2â•… A Check-list of Risk Factors for Suicidal Behaviour in Bipolar Disorder

Risk factors Protective factors


 Early in the course of illness  Good family and social support
 Younger age  Pregnancy
 Depressive, mixed dysphoric-irritable states  Post-partum period
 Caucasian ethnicity  Children at home
 Unmarried  Holding strong religious belief
 Previous depression  Adherence to pharmacotherapy
 Previous dysphoric-agitated states  Participation in psychoeducation
programmes
 Hopelessness
 Prior suicide attempts
 Substance or alcohol abuse
 Impulsivity
 Poor compliance with treatment
 Stressful life events
 Childhood sexual/physical abuse
 Current suicidal ideation
 Limited access to support or clinical services
 Family history of suicide
Source: Authors.

The general population ratio of suicide attempts to completed suicides


varies with such factors as age, sex, ethnicity, comorbid conditions and
the accuracy of case identification, especially for suicide attempts of varied
severity and potential lethality (Tondo and Baldessarini, 2005). Given these
caveats, this ratio among BD patients may be as low as 3 (Carlson et al.,
1974), and averages about 5 (Tondo et al., 2003). This ratio is much lower
than that for the general population, in which the ratio of attempts/suicides
is typically 10, as high as 25 in the United States, and averages about 16
internationally (Tondo et al., 2003; WHO, 2003).
Baldessarini, Pompili and Tondo (2006a) recently reviewed 60
published studies of the risk of suicide attempts among more than 70,000

263
Maurizio Pompili et al.

persons with mood disorders. The annual risk of suicide attempts was
approximately 4.2 percent per year among patients with BD, suggesting
that the great majority of such patients may make at least one attempt
in their lifetime. The risk of an attempt was not significantly higher among
those with BD as compared with those with depressive or unspecified
mood disorders. In patients with major affective disorders, attempted
suicide is the most powerful predictor of future completed suicide (Cheng
et al., 2000; Gibb et al., 2005; King et al., 2001; Rihmer, 2005; Suominen
et al., 2004). Considering the 10 studies (including a total of 3,187 patients)
in which unipolar, BD-I and BD-II patients were analysed separately, it
has been found that the lifetime rates of suicide attempts in unipolar, BD-I
and BD-II patients were 13, 26 and 33 percent, respectively (Rihmer, 2005).
Community-based epidemiological studies from the United States (Chen
and Dilsaver, 1996; Kessler et al., 1999) and from Hungary (Szadoczky
et al., 2000) have shown that the lifetime rate of prior suicide attempts
was 1.5 to 2.5 higher in bipolar than in unipolar patients.
Grunebaum et al. (2006) studied patients with BD for the presence
or absence at baseline evaluation of a history of suicide attempt. The
regression analysis showed that a history of suicide attempts in BD
patients was associated with recent suicidal ideation, more psychiatric
hospitalisations, lifetime aggressive traits and an earlier age at onset of a
first mood episode.
Investigating the frequency of current suicidal ideation in 605 unipolar
major depressives, 103 bipolar II and 81 bipolar I depressives a recent study
from Italy found that 16.5 percent of patients were actually suicidal, and
the bipolar/unipolar risk of suicidality was 2.2 (Tondo et al., 2008).

Treatments for suicide in BD patients

Lithium and Mood Stabilizers

In a large population-based sample of more than 20,000 persons treated


for BD, Goodwin et al. (2003) found that risk of attempted or completed
suicide was 1.5 to 3 times higher during periods of treatment with divalproex
than during periods of treatment with lithium. This difference in risk

264
Suicide Risk in Bipolar Disorder

was consistent across all outcome measures (suicide, a suicide attempt


resulting in hospitalisation, and a suicide attempt treated in the emergency
room—ER) and across the two study sites. Results for carbamazepine were
qualitatively similar to those for divalproex but (reflecting the smaller
sample size) much less precise. A recent, comprehensive meta-analysis of
studies of lithium adds additional support to the impression that lithium
has major beneficial effects against both suicide and suicide attempts and
that these effects are found consistently across almost all trials reported over
the past three decades, including trials involving randomisation and
double-blind assessments (Baldessarini et al., 2006). This larger analysis
considered a total of 45 reports involving 53,472 patients with BD or more
broadly defined manic-depressive disorder (including unipolar recurrent
major depressive and schizoaffective disorder) treated and evaluated for an
average exposure of nearly 348,000 patient-years with or without lithium.
Risks for both completed suicide and suicide attempts were reduced by
nearly fivefold, or 80 percent.
The US Food and Drug Administration (FDA, 2008) analysed reports
of suicidality (suicidal behaviour or ideation) from placebo-controlled
clinical studies of 11 drugs used to treat epilepsy as well as psychiatric
disorders and other conditions. These drugs are commonly referred to as
antiepileptic drugs. In the FDA’s analysis, patients receiving antiepileptic
drugs had approximately twice the risk of suicidal behaviour or ideation
(0.43 percent) as compared to patients receiving a placebo (0.22 percent).
The increased risk of suicidal behaviour and suicidal ideation was observed
as early as one week after starting the antiepileptic drug and continued
for 24 weeks. The results were generally consistent among the 11 drugs.
Patients who were treated for epilepsy, psychiatric disorders and other
conditions were all at increased risk for suicidality when compared to a
placebo, and there did not appear to be a specific demographic subgroup
of patients to which the increased risk could be attributed. The relative risk
for suicidality was higher in the patients with epilepsy compared to patients
who were given one of the drugs for psychiatric or other conditions.╯
A growing number of anticonvulsants have demonstrated antimanic
efficacy and are occasionally studied for possible long-term mood-
stabilising effects in patients with BD. Most of these agents remain
largely unexamined for possible beneficial effects on suicidal behaviour.
Yerevanian et al. (2007b) studied 405 veterans with BD followed for a
mean of 3 years treated with lithium, divalproex and carbamazepine

265
Maurizio Pompili et al.

monotherapies. These authors found that the three anticonvulsants


showed similar benefits in protecting bipolar patients from non-lethal
suicidal behaviour when careful analysis of clinical data was done to con-
firm medication adherence/non-adherence. The authors observed that
there was a 16-fold increase in the non-lethal suicidal event rate after dis-
continuation compared with mood-stabiliser monotherapy.

Antidepressants

Despite controversies in the involvement of antidepressant agents in the


management of bipolar disorders, during acute depressive phases these
agents can offer a valid therapeutic option (Henry and Demotes-Mainard,
2006), especially for BD-II (Amsterdam and Brunswick, 2003). To some
extent, the lack of demonstrated effectiveness of antidepressant treatment
in reducing suicide risk in some studies may reflect continuing low rates
of closely supervised antidepressant treatment, particularly in young men,
and inadequate dosing and duration of sustained treatment for many
depressed patients (Baldessarini, 2006), as well as misdiagnosis of BD as
unipolar depression and using antidepressant monotherapy in patients
with unrecognised underlying bipolarity which is the one of the major
sources of treatment-resistance and ultimately the worsening of depression
(Inoue et al., 2006; O’Donovan et al., 2008; Sharma et al., 2005).
Alternatively, suicidal behaviour may require more than depressed
mood, and the relevant factors (possibly including elements of anger,
aggression and impulsivity) may not be ameliorated by antidepressant
treatment (Baldessarini et al., 2005). Moreover, in some vulnerable
patients, mixed dysphoric-agitated states in BD can be induced by anti-
depressant monotherapy (unprotected by mood stabilisers or atypical
antipsychotics/benzodiazepines). This may not be recognised clinically
and thus not accurately differentiated from worsening depression
(Akiskal and Mallya, 1987; Inoue et al., 2006; O’Donovan et al., 2008).
Such states, as well as other adverse behavioural responses to antidepressant
treatment (such as insomnia, restlessness, irritability, agitation and
mixed states), in patients with mood disorder may well increase the risk
of aggressive-impulsive acts, including perhaps suicidal behaviour in
some adults and children, and such responses may be particularly likely
among BD patients in depressive or mixed states (Baldessarini et al., 2005;
O’Donovan et al., 2008).
266
Suicide Risk in Bipolar Disorder

Antipsychotics

Among the various antipsychotics, clozapine has been reported as an


antisuicidal agent in several studies (Hennen and Baldessarini, 2005),
including an important randomised trial comparing clozapine with
olanzapine (Meltzer et al., 2003). Despite insufficient proof of reduction
in mortality and limitation only to the treatment of patients with chronic
psychotic disorders (schizophrenia or schizoaffective disorders), this
drug is the first treatment of any kind to receive regulatory approval by
the FDA for reducing the risk of suicidal behaviour, a precedent-setting
development. Clozapine treatment involves, like lithium prophylaxis,
close medical supervision and regular blood testing to minimise the risk
of potentially lethal side effects. Such procedures increase the levels of
patient’s interaction with medical personnel which has been reported as
a contributor factor for preventing suicide.
A recent study by Yerevanian et al. (2007a) involving 405 bipolar
patients demonstrated a higher risk of suicidal behaviour associated with
antipsychotic treatment, compared with mood-stabiliser monotherapy.
There was a nearly 10-fold greater risk of non-lethal suicidal behaviour during
antipsychotic monotherapy compared with mood-stabiliser monotherapy,
with intermediate risk during mood stabiliser + antipsychotic treatment
periods. It is therefore suggested that clinicians who add antipsychotics to
mood stabilisers to treat breakthrough (hypo)mania in BD patients should
keep their patients on this supplementary medication for as short a time
as possible since antipsychotics (particularly in longer use than needed)
may increase the switch from (hypo)mania into depressive/mixed states
and, therefore, increase the risk of suicidal behaviour.

Electroconvulsive Therapy (ECT)

A study by Prudic and Sackeim (1999) found that ECT responders and
non-responders showed a large decrease in scores on the suicide item of
the Hamilton Rating Scale for Depression, and this decrease was greater
than the average improvement on other items. This would confirm that
some of the therapeutics available for psychiatric disorders may not have
a real impact on symptoms but can have an independent effect on suicide
risk. ECT is reported as being the most effective and rapid treatment for

267
Maurizio Pompili et al.

emerging or ongoing suicidality in severe depressive illness and has been


considered to be the treatment of choice in emergency situations involving
high suicide risk (‘Consensus conference. Electroconvulsive therapy’, 1985;
Weiner, 2000). However, the effectiveness of ECT for sustained suicide
prevention has not been proven and requires further study, including in
patients with BD (‘Practice guideline for the assessment and treatment
of patients with suicidal behaviors’, 2003).

Psychosocial Interventions

Recently, several effective psychosocial interventions for BD have been


developed (including psychoeducation, cognitive-behavioural therapy,
and interpersonal and social rhythm therapy) (Fountoulakis et al., 2008),
primarily for noncompliant, drug-intolerant and drug-nonresponsive
patients. These strategies may be effective alone, but are used mostly in
combination with mood stabilisers (Bauer, 2001; Rucci et al., 2002).
Since they are designed for relapse/recurrence prevention, these inte-
rventions might have efficacy for suicide prevention as well. However,
at present, there is only one published study concerning the effect on
suicidality of intensive psychosocial treatment that specifically targeted
suicidality in BD. Rucci et al. (2002) investigated the lifetime rates of suicide
attempts among 175 BD-I patients during a 2-year period of intensive
pharmacotherapy (lithium, valproate, carbamazepine) and one of two
adjunctive psychosocial interventions (either interpersonal and social
rhythm therapy, which is a psychotherapy specific for BD, or intensive
clinical management, which is a nonspecific psychosocial intervention).
In addition to receiving mood stabilisers, patients were randomly assigned
to treatment with one of the two psychosocial interventions. Forty-seven
patients did not complete the acute phase, and 20 failed to complete the
2-year long maintenance phase. Before entry into the study, the rate of
suicide attempts per 100 patient-months was 1.05. However, during the
acute and the 2-year maintenance phase, the rates were 0.31 (a 3.4-fold
reduction) and 0.06 (a 17.5-fold reduction), respectively (both differences
were significant). Patients receiving psychotherapy specific for BD showed
3.7-fold lower frequency of suicide attempts during the maintenance
phase than those receiving nonspecific intensive clinical management
(0.07 versus 0.26 attempt per 100 patient-months, respectively,

268
Suicide Risk in Bipolar Disorder

a significant difference). Before the patients entered the protocol,


50 patients had made 72 suicide attempts, but none of them attempted
suicide during the acute and maintenance phase of the study. The com-
bined rate of suicide attempts per 100 patent-months during the acute
and maintenance phase was 0.17, which is a six-fold reduction when com-
pared with the pre-treatment period. The findings suggest that disorder-
specific psychotherapy might be more effective than nonspecific clinical
management in augmenting the effect of long-term treatment with mood
stabilisers in BD-I patients.

Conclusions

This comprehensive update supports the emerging conclusion that BD is


a highly prevalent, often severe, sometimes disabling, and potentially fatal
illness. It is associated with a very high risk of suicide, especially early in the
illness when sustained clinical interventions, and even the diagnosis, may
not have been established, and suicidal risk continues over many years,
eventually accounting for perhaps 15 percent of deaths. The epidemiology
of suicide risk in BD, specifically, is methodologically limited and has often
been confounded by a lack of separation of BD-I and BD-II from each
other, or from recurrent major unipolar depressive illnesses.
The depressive and dysphoric-agitated mixed phases of BD, particularly
following repeated episodes of severe depression, appear to be especially
dangerous and life threatening, as well as being particularly challenging to
diagnose and treat effectively and safely (Marneros et al., 2004). BD also
is associated with very high rates of comorbid substance-use and anxiety
disorders (Krishnan, 2005), as well as impulsivity, lack of insight and poor
treatment-adherence, all further complicating the effective clinical care of
such patients and probably adding to their suicidal risk.
Rucci et al. (2002) found a consistent reduction in suicide attempt risk
associated with a combination of pharmacotherapy (mostly lithium) and
either a highly-structured psychosocial treatment designed specifically
for individuals with BD-I or a more general supportive intervention
offered in the context of a research clinic environment geared towards
optimal care of patients with this condition. This reduction was more pro-
nounced during the maintenance phase than during the acute phase.

269
Maurizio Pompili et al.

Preliminary evidence suggests that intensive clinical management


or psychotherapy adds to the effect of lithium or other appropriate
pharmacotherapy.
Short-term interventions that are widely accepted empirically for
managing acute suicidality include close clinical supervision, rapid
hospitalisation and use of ECT. However, there is little evidence that
various types of clinical interventions, including specific mood-altering
medical treatments and widely accepted psychosocial therapies, have
long-term effectiveness in reducing suicidal risk in BD patients. Lithium
maintenance treatment appears to be a unique exception in having
considerable research support for a sustained reduction in the suicidal
risk in persons with BD and possibly other forms of major affective
illnesses, including demonstrated reductions in mortality. Anticonvulsant
and atypical antipsychotic agents, as well as the less toxic modern anti-
depressants, all require specific research assessment of their long-term
ability to limit premature mortality from all causes in BD and other major
affective disorders, and specifically to reduce suicidal risk.
Interventions such as social skills training, vocational rehabilitation and
supportive employment are, therefore, very important in the prevention
of suicide in patients. Broadly speaking, these kinds of therapies focus
on working out daily problems rather than achieving psychological
insight. It has become increasingly clear that supportive, reality-oriented
therapies are generally of great value in the treatment of patients with
BD. In particular, supportive psychotherapy aims at offering the patient
the opportunity to meet with the therapist and discuss the difficulties en-
countered in daily activities. Patients are, therefore, encouraged to discuss
concerns about medications and side effects as well as social isolation,
money, stigma, etc. The therapist has an active role as he or she gives sug-
gestions and shares good and bad periods empathically. The nature of these
treatments and their availability vary greatly from place to place.

References

Ahrens, B., B. Muller-Oerlinghausen, M. Schou, T. Wolf, M. Alda, E. Grof et al. (1995).


Excess cardiovascular and suicide mortality of affective disorders may be reduced
by lithium prophylaxis. Journal of Affective Disorders, 33(2), 67–75.

270
Suicide Risk in Bipolar Disorder

Akiskal, H.S. (1996). The prevalent clinical spectrum of bipolar disorders: Beyond
DSM-IV. Journal of Clinical Psychopharmacology, 16(2 Suppl. 1), 4S–14S.
Akiskal, H.S. (2007). Targeting suicide prevention to modifiable risk factors: Has bipolar
II been overlooked? Acta Psychiatrica Scandinavica, 116(6), 395–402.
Akiskal, H.S. and F. Benazzi (2003). Delineating depressive mixed states. Their
therapeutic significance. Clinical Approaches in Bipolar Disorders, 2(2), 41–47.
Akiskal, H.S., F. Benazzi, G. Perugi and Z. Rihmer (2005). Agitated “unipolar”
depression re-conceptualized as a depressive mixed state: implications for the
antidepressant-suicide controversy. Journal of Affective Disorders, 85(3), 245–58.
Akiskal, H.S., E.G. Hantouche and J.F.Allilaire (2003). Bipolar II with and without
cyclothymic temperament: “Dark” and “sunny” expressions of soft bipolarity.
Journal of Affective Disorders, 73(1–2), 49–57.
Akiskal, H.S. and G. Mallya (1987). Criteria for the “soft” bipolar spectrum: treatment
implications. Psychopharmacology Bulletin, 23(1), 68–73.
Amsterdam, J.D. and D.J. Brunswick (2003). Antidepressant monotherapy for bipolar
type II major depression. Bipolar Disorders, 5(6), 388–95.
Angst, J., F. Angst, R. Gerber-Werder and A. Gamma (2005). Suicide in 406 mood-
disorder patients with and without long-term medication: A 40 to 44 years’ follow-
up. Archives of Suicide Research, 9(3), 279–300.
Baldessarini, R.J. (2006). Drugs and the treatment of psychiatric disorders: Anti-
depressant and antianxiety agents. In J.G. Hardman, L.E. Limbird and
A.G. Gilman (Eds), Goodman and Gilman’s the pharmacological basis of
therapeutics (11th ed.). New York: McGraw-Hill.
Baldessarini, R.J., M. Pompili and L. Tondo (2006a). Bipolar disorder. In R.I. Simon
and R.E. Hales (Eds), American psychiatric textbook of suicide assessment and
management (pp. 277–99). Washington, DC: American Psychiatric Publishing.
Baldessarini, R.J., M. Pompili and L. Tondo (2006b). Suicide in bipolar disorder: Risks
and management. CNS Spectrums, 11(6), 465–71.
Baldessarini, R.J., M. Pompili, L. Tondo, E. Tsapakis, F. Soldani, G.L. Faedda et
al. (2005). Antidepressants and suicidal behavior: Are we hurting or helping?
Clinical Neuropsychiatry, 2(1), 73–75.
Baldessarini, R.J., L. Tondo, P. Davis, M. Pompili, F.K. Goodwin and J. Hennen (2006).
Decreased risk of suicides and attempts during long-term lithium treatment:
A meta-analytic review. Bipolar Disorders, 8(5 Pt 2), 625–39.
Bauer, M.S. (2001). An evidence-based review of psychosocial treatments for bipolar
disorder. Psychopharmacology Bulletin, 35(3), 109–34.
Benazzi, F. (2006). Mood patterns and classification in bipolar disorder. Current Opinion
in Psychiatry, 19(1), 1–8.
Benazzi, F., A. Koukopoulos and H.S. Akiskal (2004). Toward a validation of a new
definition of agitated depression as a bipolar mixed state (mixed depression).
European Psychiatry, 19(2), 85–90.
Bulik, C.M., L.L. Carpenter, D.J. Kupfer and E. Frank (1990). Features associated with
suicide attempts in recurrent major depression. Journal of Affective Disorders,
18(1), 29–37.

271
Maurizio Pompili et al.

Carlson, G.A., J. Kotin, Y.B. Davenport and M. Adland (1974). Follow-up of 53 bipolar
manic-depressive patients. British Journal of Psychiatry, 124(579), 134–39.
Chen, Y.W. and S.C. Dilsaver (1996). Lifetime rates of suicide attempts among subjects
with bipolar and unipolar disorders relative to subjects with other axis I disorders.
Biological Psychiatry, 39(10), 896–99.
Cheng, A.T., T.H. Chen, C.C. Chen and R. Jenkins (2000). Psychosocial and psychiatric
risk factors for suicide. Case-control psychological autopsy study. British Journal
of Psychiatry, 177(4), 360–65.
Consensus conference. Electroconvulsive therapy (1985). Journal of the American
Medical Association, 254(15), 2103–08.
Dilsaver, S.C., Y.W. Chen, A.C. Swann, A.M. Shoaib and K.J. Krajewski (1994).
Suicidality in patients with pure and depressive mania. American Journal of
Psychiatry, 151(9), 1312–15.
Fagiolini, A., D.J. Kupfer, P. Rucci, J.A. Scott, D.M. Novick and E. Frank (2004).
Suicide attempts and ideation in patients with bipolar I disorder. Journal of Clinical
Psychiatry, 65(4), 509–14.
Faravelli, C., S. Rosi, M. Alessandra Scarpato, L. Lampronti, S.G. Amedei and N. Rana
(2006). Threshold and subthreshold bipolar disorders in the Sesto Fiorentino
Study. Journal of Affective Disorders, 94(1–3), 111–19.
Fiedorowicz, J.G., A.C. Leon, M.B. Keller, D.A. Solomon, J.P. Rice and W.H. Coryell
(2008). Do risk factors for suicidal behavior differ by affective disorder polarity?
Psychological Medicine, 30(5), 1–9.
Fountoulakis, K.N., X. Gonda, M. Siamouli and Z. Rihmer (2008). Psychotherapeutic
intervention and suicide risk reduction in bipolar disorder: A review of the
evidence. Journal of Affective Disorders, 113(1), 21–29.
Gibb, S.J., A.L. Beautrais and D.M. Fergusson (2005). Mortality and further suicidal
behaviour after an index suicide attempt: A 10-year study. Australian and
New Zealand Journal of Psychiatry, 39(1–2), 95–100.
Goodwin, F.K., B. Fireman, G.E. Simon, E.M. Hunkeler, J. Lee and D. Revicki (2003).
Suicide risk in bipolar disorder during treatment with lithium and divalproex.
Journal of the American Medical Association, 290(11), 1467–73.
Goodwin, F.K. and K.R. Jamison (Eds). (2007). Manic-depressive illness (2nd ed.).
New York: Oxford University Press.
Grunebaum, M.F., S.R. Ramsay, H.C. Galfalvy, S.P. Ellis, A.K. Burke, L. Sher et al.
(2006). Correlates of suicide attempt history in bipolar disorder: A stress-diathesis
perspective. Bipolar Disorders, 8(5 Pt 2), 551–57.
Harris, E.C. and B. Barraclough (1997). Suicide as an outcome for mental disorders.
A meta-analysis. British Journal of Psychiatry, 170(3), 205–28.
Hawton, K. and K. van Heeringen (Eds). (2000). The international handbook of suicide
and attempted suicide. Chichester: John Wiley and Sons.
Hennen, J. and R.J. Baldessarini (2005). Suicidal risk during treatment with clozapine:
A meta-analysis. Schizophrenia Research, 73(2–3), 139–45.

272
Suicide Risk in Bipolar Disorder

Henry, C. and J. Demotes-Mainard (2006). SSRIs, suicide and violent behavior: Is


there a need for a better definition of the depressive state? Current Drug Safety,
1(1), 59–62.
Inoue, T., S. Nakagawa, Y. Kitaichi, T. Izumi, T. Tanaka, T. Masui et al. (2006).
Long-term outcome of antidepressant-refractory depression: The relevance of
unrecognized bipolarity. Journal of Affective Disorders, 95(1–3), 61–67.
Isometsa, E.T., M.M. Henriksson, H.M. Aro and J.K. Lonnqvist (1994). Suicide in
bipolar disorder in Finland. American Journal of Psychiatry, 151(7), 1020–24.
Jamison, K.R. (2000). Suicide and bipolar disorder. Journal of Clinical Psychiatry, 61
(Suppl 9), 47–51.
Judas, M., M. Rados, N. Jovanov-Milosevic, P. Hrabac, R. Stern-Padovan and I.
Kostovic (2005). Structural, immunocytochemical, and mr imaging properties
of periventricular crossroads of growing cortical pathways in preterm infants.
American Journal of Neuroradiology, 26(10), 2671–84.
Judd, L.L. and H.S. Akiskal (2003). The prevalence and disability of bipolar spectrum
disorders in the US population: Re-analysis of the ECA database taking into account
subthreshold cases. Journal of Affective Disorders, 73(1–2), 123–31.
Kessler, R.C., G. Borges and E.E. Walters (1999). Prevalence of and risk factors for
lifetime suicide attempts in the National Comorbidity Survey. Archives of General
Psychiatry, 56(7), 617–26.
Kessler, R.C., W.T. Chiu, O. Demler, K.R. Merikangas and E.E. Walters (2005).
Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the
National Comorbidity Survey Replication. Archives of General Psychiatry, 62(6),
617–27.
King, E.A., D.S. Baldwin, J.M. Sinclair, N.G. Baker, M.J. Campbell and C. Thompson
(2001). The Wessex Recent In-Patient Suicide Study, 1. Case-control study of
234 recently discharged psychiatric patient suicides. British Journal of Psychiatry,
178(6), 531–36.
Kochman, F.J., E.G. Hantouche, P. Ferrari, S. Lancrenon, D. Bayart and H.S. Akiskal
(2005). Cyclothymic temperament as a prospective predictor of bipolarity and
suicidality in children and adolescents with major depressive disorder. Journal of
Affective Disorders, 85(1–2), 181–89.
Krishnan, K.R. (2005). Psychiatric and medical comorbidities of bipolar disorder.
Psychosomatic Medicine, 67(1), 1–8.
Leibenluft, E. (1997). Issues in the treatment of women with bipolar illness. Journal of
Clinical Psychiatry, 58 (Suppl 15), 5–11.
Leverich, G.S., L.L. Altshuler, M.A. Frye, T. Suppes, P.E. Keck, Jr., S.L. McElroy
et al. (2003). Factors associated with suicide attempts in 648 patients with bipolar
disorder in the Stanley Foundation Bipolar Network. Journal of Clinical Psychiatry,
64(5), 506–15.
Lopez, P., F. Mosquera, J. de Leon, M. Gutierrez, J. Ezcurra, F. Ramirez et al. (2001).
Suicide attempts in bipolar patients. Journal of Clinical Psychiatry, 62(12),
963–66.

273
Maurizio Pompili et al.

Maj, M., R. Pirozzi, L. Magliano and L. Bartoli (2003). Agitated depression in bipolar
I disorder: Prevalence, phenomenology, and outcome. American Journal of
Psychiatry, 160(12), 2134–40.
Marneros, A., S. Rottig, A. Wenzel, R. Bloink and P. Brieger (2004). Affective
and schizoaffective mixed states. European Archives of Psychiatry and Clinical
Neuroscience, 254(2), 76–81.
Maser, J.D., H.S. Akiskal, P. Schettler, W. Scheftner, T. Mueller, J. Endicott et al. (2002).
Can temperament identify affectively ill patients who engage in lethal or near-
lethal suicidal behavior? A 14-year prospective study. Suicide & Life-Threatening
Behavior, 32(1), 10–32.
Meltzer, H.Y., L. Alphs, A.I. Green, A.C. Altamura, R. Anand, A. Bertoldi et al. (2003).
Clozapine treatment for suicidality in schizophrenia: International Suicide
Prevention Trial (InterSePT). Archives of General Psychiatry, 60(1), 82–91.
Miklowitz, D.J. and S.L. Johnson (2006). The psychopathology and treatment of bipolar
disorder. Annual Review of Clinical Psychology, 2, 199–235.
Muller-Oerlinghausen, B., A. Berghofer and M. Bauer (2002). Bipolar disorder. Lancet,
359(9302), 241–47.
O’Donovan, C., J.S. Garnham, T. Hajek and M. Alda (2008). Antidepressant
monotherapy in pre-bipolar depression; Predictive value and inherent risk. Journal
of Affective Disorders, 107(1–3), 293–98.
Oquendo, M.A., C. Waternaux, B. Brodsky, B. Parsons, G.L. Haas, K.M. Malone et
al. (2000). Suicidal behavior in bipolar mood disorder: clinical characteristics of
attempters and nonattempters. Journal of Affective Disorders, 59(2), 107–17.
Oquendo, M.A., H. Galfalvy, S. Russo, S.P. Ellis, M.F. Grunebaum, A. Burke et al. (2004).
Prospective study of clinical predictors of suicidal acts after a major depressive
episode in patients with major depressive disorder or bipolar disorder. American
Journal of Psychiatry, 161(8), 1433–41.
Peterson, B.S. (2003). Brain imaging studies of the anatomical and functional
consequences of preterm birth for human brain development. Annals of the
New York Academy of Sciences, 1008, 219–37.
Pompili, M., S. Ehrlich, E. De Pisa, J.J. Mann, M. Innamorati, A. Cittadini et al. (2007).
White matter hyperintensities and their associations with suicidality in patients
with major affective disorders. European Archives of Psychiatry and Clinical
Neuroscience, 257(8), 494–99.
Pompili, M., P. Girardi, R. Tatarelli and D. Lester (2006). Subthreshold bipolar traits
and suicide risk among undergraduates. Psychological Reports, 98(2), 417–18.
Pompili, M., M. Innamorati, J.J. Mann, M.A. Oquendo, D. Lester, A. Del Casale
et al. (2008). Periventricular white matter hyperintensities as predictors of
suicide attempts in bipolar disorders and unipolar depression. Progress in Neuro-
Psychopharmacology & Biological Psychiatry, 32(6), 1501–07.
Pompili, M., D. Lester, P. Girardi and R. Tatarelli (2007). High suicide risk after the
development of cognitive and working memory deficits caused by cannabis, cocaine
and ecstasy use. Substance Abuse, 28(1), 25–30.

274
Suicide Risk in Bipolar Disorder

Pompili, M., Z. Rihmer, H.S. Akiskal, M. Innamorati, P. Iliceto, K.K. Akiskal et al.
(2008). Temperament and personality dimensions in suicidal and nonsuicidal
psychiatric inpatients. Psychopathology, 41(5), 313–21.
Practice guideline for the assessment and treatment of patients with suicidal behaviors.
(2003). American Journal of Psychiatry, 160(Suppl 11), 1–60.
Prudic, J. and H.A. Sackeim (1999). Electroconvulsive therapy and suicide risk. Journal
of Clinical Psychiatry, 60 (Suppl 2), 104–10.
Regier, D.A., M.E. Farmer, D.S. Rae, B.Z. Locke, S.J. Keith, L.L. Judd et al. (1990).
Comorbidity of mental disorders with alcohol and other drug abuse. Results from
the Epidemiologic Catchment Area (ECA) Study. Journal of the American Medical
Association, 264(19), 2511–18.
Rice, J., T. Reich, N.C. Andreasen, J. Endicott, M. Van Eerdewegh, R. Fishman et al.
(1987). The familial transmission of bipolar illness. Archives of General Psychiatry,
44(5), 441–47.
Rihmer, Z. (2005). Prediction and prevention of suicide in bipolar disorders. Clinical
Neuropsychiatry, 2(1), 48–54.
Rihmer, Z. and J. Angst (2005). Epidemiology of bipolar disorder. In S. Kasper and
R. M. A. Hirscfeld (Eds), Handbook of bipolar disorder (pp. 21–35). New York:
Taylor and Francis.
Rihmer, A., S. Rózsa, Z. Rihmer, X. Gonda, K.K. Akiskal and H.S. Akiskal (2007).
Affective temperament-types and suicidal behaviour. European Psychiatry, 22
(Suppl 1), S244.
Romero, S., F. Colom, A.M. Iosif, N. Cruz, I. Pacchiarotti, J. Sanchez-Moreno et al.
(2007). Relevance of family history of suicide in the long-term outcome of bipolar
disorders. Journal of Clinical Psychiatry, 68(10), 1517–21.
Rouillon, F., D. Serrurier, H.D. Miller and M.J. Gerard (1991). Prophylactic efficacy
of maprotiline on unipolar depression relapse. Journal of Clinical Psychiatry, 52
(10), 423–31.
Rucci, P., E. Frank, B. Kostelnik, A. Fagiolini, A.G. Mallinger, H.A. Swartz et al.
(2002). Suicide attempts in patients with bipolar I disorder during acute and
maintenance phases of intensive treatment with pharmacotherapy and adjunctive
psychotherapy. American Journal of Psychiatry, 159(7), 1160–64.
Ryan, N.D., D.E. Williamson, S. Iyengar, H. Orvaschel, T. Reich, R.E. Dahl et al. (1992).
A secular increase in child and adolescent onset affective disorder. Journal of the
American Academy of Child and Adolescent Psychiatry, 31(4), 600–05.
Sato, T., R. Bottlender, A. Tanabe and H.J. Moller (2004). Cincinnati criteria for mixed
mania and suicidality in patients with acute mania. Comprehensive Psychiatry,
45(1), 62–69.
Schneck, C.D., D.J. Miklowitz, J.R. Calabrese, M.H. Allen, M.R. Thomas, S.R.
Wisniewski et al. (2004). Phenomenology of rapid-cycling bipolar disorder: Data
from the first 500 participants in the Systematic Treatment Enhancement Program.
American Journal of Psychiatry, 161(10), 1902–08.

275
Maurizio Pompili et al.

Sharma, V., M. Khan and A. Smith (2005). A closer look at treatment resistant
depression: Is it due to a bipolar diathesis? Journal of Affective Disorders, 84(2–3),
251–57.
Slama, F., F. Bellivier, C. Henry, A. Rousseva, B. Etain, F. Rouillon et al. (2004). Bipolar
patients with suicidal behavior: Toward the identification of a clinical subgroup.
Journal of Clinical Psychiatry, 65(8), 1035–39.
Soares, J.C. and J.J. Mann (1997). The anatomy of mood disorders—Review of structural
neuroimaging studies. Biological Psychiatry, 41(1), 86–106.
Strakowski, S.M., S.L. McElroy, P.E. Keck, Jr. and S.A. West (1996). Suicidality among
patients with mixed and manic bipolar disorder. American Journal of Psychiatry,
153(5), 674–76.
Suominen, K., E. Isometsa, J. Suokas, J. Haukka, K. Achte and J. Lonnqvist (2004).
Completed suicide after a suicide attempt: A 37-year follow-up study. American
Journal of Psychiatry, 161(3), 562–63.
Szadoczky, E., J. Vitrai, Z. Rihmer and J. Furedi (2000). Suicide attempts in the
Hungarian adult population: Their relation with DIS/DSM-III-R affective and
anxiety disorders. European Psychiatry, 15(6), 343–47.
Taylor, W.D., M.E. Payne, K.R. Krishnan, H.R. Wagner, J.M. Provenzale, D.C. Steffens
et al. (2001). Evidence of white matter tract disruption in MRI hyperintensities.
Biological Psychiatry, 50(3), 179–83.
Tondo, L. and R.J. Baldessarini (2005). Suicidal risk in bipolar disorder. Clinical
Neuropsychiatry, 2(1), 55–65.
Tondo, L., R.J. Baldessarini, J. Hennen and G. Floris (1998). Lithium maintenance
treatment of depression and mania in bipolar I and bipolar II disorders. American
Journal of Psychiatry, 155(5), 638–45.
Tondo, L., G. Isacsson and R. Baldessarini (2003). Suicidal behaviour in bipolar disorder:
Risk and prevention. CNS Drugs, 17(7), 491–511.
Tondo, L., B. Lepri and R.J. Baldessarini (2008). Suicidal status during antidepressant
treatment in 789 Sardinian patients with major affective disorder. Acta Psychiatrica
Scandinavica, 118(2), 106–15.
U.S. Food and Drug Administration [FDA]. (2008). Information for healthcare
professionals suicidality and antiepileptic drugs. Retrieved 08/12, 2008, from
http://www.fda.gov/Cder/Drug/InfoSheets/HCP/antiepilepticsHCP.htm
Valtonen, H., K. Suominen, O. Mantere, S. Leppamaki, P. Arvilommi and E. T. Isometsa
(2005). Suicidal ideation and attempts in Bipolar I and II Disorders. Journal of
Clinical Psychiatry, 66(11), 1456–62.
Weiner, R.D. (Ed.). (2000). Practice of electroconvulsive therapy: Recommendations for
treatment, training, and privileging. A task force report of the American Psychiatric
Association (2nd ed.). Washington: American Psychiatric Association.
Wickramaratne, P.J., M.M. Weissman, P.J. Leaf and T.R. Holford (1989). Age, period
and cohort effects on the risk of major depression: Results from five United States
communities. Journal of Cinical Epidemiology, 42(4), 333–43.

276
Suicide Risk in Bipolar Disorder

World Health Organization (WHO). (2003). International suicide rates. Retrieved


1 November 2005, from www.who.int/mental_health/Topic_Suicide/suicide_rates.
html
Yerevanian, B.I., R.J. Koek and J. Mintz (2007a). Bipolar pharmacotherapy and
suicidal behavior Part 3: Impact of antipsychotics. Journal of Affective Disorders,
103(1–3), 23–28.
Yerevanian, B.I., R.J. Koek and J. Mintz (2007b). Bipolar pharmacotherapy and suicidal
behavior. Part I: Lithium, divalproex and carbamazepine. Journal of Affective
Disorders, 103(1–3), 5–11.

277
13

Depression and Suicide


EòA S‘«A½½›Ù Aė MAĥٛ— Wʽ¥›ÙݗÊÙ¥

W ithout any doubt, major depressive disorder is one of the most


common psychiatric disorders worldwide. Furthermore, compared
to other psychiatric disorders it is associated with the highest suicide risk.
The lifetime prevalence for mood disorders in the newer World Mental
Health Surveys (Kessler et al., 2007; Wittchen et al., 1999) is currently
between 3.3 and 21.4 percent. The impact of depression on life quality is
comparable with that of severe physical diseases. The comorbidity quota
with other psychiatric disorders, for example, addiction, anxiety disorders,
somatoform disorders, as well as physical diseases, is high. At present it
is known that, on the one hand, it is depressive mood within psychiatric
and physical diseases, on the other hand, it is major depression which
are the primary affective disorders that lead to suicide rather than other
psychiatric disorders (Bertolote et al., 2004; Harris and Barraclough,
1997, 1998; Schneider, 2003; Wolfersdorf, 2002, 2008). In this context,
it becomes understandable that the current WHO/EU-health care policy
points out the importance of depression prevention as well as prevention
of suicidal behaviour and declares them as major topics of European and
worldwide health policy.
Depressive mood, feelings of worthlessness and insufficiency, guilt and
hopelessness, lack of perspective, agonising states of uneasiness as well
as mental and emotional pressure make patients suffering from major

278
Depression and Suicide

depression a high-risk group for suicidal behaviour (Haenel and Pöldinger,


1986; Hole, 1973; Sainsbury, 1986; Wolfersdorf et al., 1992, 2002).
Suicidal tendency is defined as the sum of all ways of thinking and
behaving that include the ambition to die or the acceptance of death due
to risky behaviour. To this also belong active suicidal behaviour and the
forbearance of activities to keep oneself alive (Wolfersdorf, 2002, 2008).
Suicidal tendency therefore can be discussed as a human way of thinking
and behaving but not as an illness itself.
Griesinger (1845) described suicidal behaviour as ‘… not always the
symptom or result of a psychiatric disorder but also due to a tiredness of
life dependent on external circumstances’. Therewith he already made
an essential distinction between suicidal tendencies as an expression of
a life crisis, for example, reactive-depressive stress disorder on the one
hand and suicidal tendency in the context of a psychiatric disorder on
the other hand.

MENTAL DISORDERS AND SUICIDE: AN OVERVIEW

Psychiatric disorders are: next to acute psychosocial stresses and strains,


former suicide attempts and consequently fewer inhibitions concerning
suicide, as well as physical diseases accompanied by heavy pains and
suffering, the most important risk factor for suicidal behaviour. According
to Harris and Barraclough (1998), suicide mortality in patients suffering
from major depression is increased by a factor of 21, with politoxicomania
13, with bipolar affective disorders 12 and with dysthymia by a factor of
approximately 12 compared to the general population.
Bertolote and colleagues (2004) recently pointed out the important
role of psychiatric disorders for suicidal tendencies. In the general
population, affective disorders clearly dominate with a prevalence of
44 percent, followed by addiction with a prevalence of 19 percent and
schizophrenia with 7.5 percent. By contrast, with inpatients, affective dis-
orders account for 21 percent, followed by schizophrenia with 20 percent.
This represents the current knowledge that in the context of inpatients,
young schizophrenic men are at a similar risk to commit suicide in a
psychotherapeutic setting as patients suffering from major depressive
disorder (Morgan and Stanton, 1997; Roy, 1984; Wolfersdorf, 1989).

279
Eva Schaller and Manfred Wolfersdorf

In Table 13.1, all psychological autopsy studies with the focus on affective
disorders of which we are aware are listed. Both the large percentage of
men as well as the numerous occurrences of affective disorders, which are
between 30 and 100 percent in suicidal tendencies and behaviour, clearly
appear in all studies.
Harris and Barraclough (1997) found a standard mortality rate (SMR) of
1,209 for psychiatric diseases, 2,035 for major depression, 1,505 for bipolar
affective disorders and 1,210 for dysthymia. According to these figures,
major depression seems to bear the highest suicide mortality rate among all
affective disorders. Within a group of 479 people who died from suicide in
southern Germany, Wolfersdorf, Faust and Hölzer (1992) found 36 percent
suffering from the beginning of a depressive episode, another 27 percent
at the point where the depressive episode was fading away, and within
about one-third of all suicides, typical depressive symptomatology could
be diagnosed. Kung, Pearson and Lin (2003) reported on the results of the
National Mortality Followback Survey in the United States. In a group
of men aged between 45 and 64 years and in women of all age groups,
they found depressive symptomatology significantly higher in suicides
compared to the control group. Cheng (1995) examined suicides, looking
for presuicidal risk factors. They identified the following risk factors: loss
events, suicidal behaviour in first degree relative, major depressive episode
to International Classification of Diseases 10 (ICD-10) within 87.1 percent,
emotionally instable personality disorder within 61.9 percent and sub-
stance dependencies within 27.6 percent (comparison of 113 suicides
with 226 controls). Also in the European Study on the Epidemiology of
Mental Disorders (ESEMED), Bernal and colleagues (2007) investigated in
total 21,425 people from Belgium, France, Germany, Italy, the Netherlands
and Spain for suicidal tendency in the quest for risk factors. Suicide
attempts were found to be most common among patients suffering from
general anxiety disorders (12 percent), followed by alcohol dependence
(12 percent), major depressive disorder (8 percent) and dysthymia
(10 percent), as well as PTSD (11 percent).
Hall and Platt (1999) found major depressive episodes in 43 percent of
analysed suicide attempts, followed by adjustment disorders with anxiety
and depression (15 percent), anxiety disorders (10 percent) as well as
schizophrenia (2 percent). Mann, Waternaux, Haas and Malone (1999)

280
Table 13.1 Psychiatric Disorders, Especially Depressive Disorders and Suicide in a
Community-based Study Using Psychological Autopsy

Number of Affective Major depressive Bipolar disorder Dysthymia Schizophrenia


Authors/Year/ Country suicides Male (%) disorders (%) episode (%) (%) (%) (%)
Robins et al. (1959) 134 77 n.a. 45 n.a. n.a. 2
Dorpat and Ripley (1960) 114 68 n.a. 30 n.a. n.a. 12
Barraclough and Pelles (1975) 100 53 n.a. 70 n.a. n.a. n.a.
Beskow (1979) 271 100 n.a. 45 n.a. n.a. n.a.
Chynoweth et al. (1980) 135 63 n.a. 55 n.a. n.a. 4
Mitterauer (1981) 145 n.a. n.a. n.a. n.a. n.a. 18
Rich et al. (1986) 283 71 n.a. 47 n.a. n.a. 6
Arato et al. (1988) 200 64 58 34 24 2 9
Shaffi et al. (1988) 21 n.a. 52 n.a. n.a. n.a. n.a.
Runeson (1989) 58 72 43 36 5 2 16
Asgard (1990) 104 0 60 35 1 20 5
Petronis et al. (1990) 40 30 60 53 7 n.a. n.a.
Marttunen et al. (1991) 53 83 52 23 n.a. 4 6
Apter et al. (1993) 43 n.a. n.a. n.a. n.a. n.a. n.a.
Wolfersdorf et al. (1992) 454 72 66 66 n.a. n.a. 7,5
Brent et al. (1992) 67 n.a. 39 n.a. n.a. n.a. n.a.
Henriksson et al. (1993) 229 75 n.a. 59 n.a. n.a. n.a.
Lesage et al. (1994) 75 100 53 40 n.a. n.a. 9
Cheng (1995) 116 62 87 87 0 n.a. 7
Rhimer et al. (1995) 115 77 50 16 22 n.a. 6
Conwell et al. (1996) 141 80 47 28 1 n.a. 16
Shaffer et al. (1996) 120 n.a. 57 n.a. n.a. n.a. n.a.
Heilä et al. (1987) 1,397 n.a. n.a. n.a. n.a. 11 7
(Table 13.1 Continued)
(Table 13.1 Continued)

Number of Affective Major depressive Bipolar disorder Dysthymia Schizophrenia


Authors/Year/ Country suicides Male (%) disorders (%) episode (%) (%) (%) (%)
Foster et al. (1997) 117 79 36 32 n.a. n.a. 11
Appleby et al. (1999) 84 81 23 n.a. n.a. n.a. 19
Brent et al. (1999) 140 85 48 n.a. n.a. n.a. 0
Vijayakumar and Rajkumar (1999) 100 n.a. 30 n.a. n.a. n.a. n.a.
Phillips et al. (2002) 519 48 66 36 n.a. 30 n.a.
Coryell and Young (2005) 33 45 100 80 20 n.a. n.a.
Bertolote et al. (2004)(meta-analyse) 19,716 n.a. 30 n.a. n.a. n.a. 14
Schneider et al. (2001) 163 64 37 18 5 5 14
Agoub et al. (2006) 51(suicidal n.a. 45 24 n.a. 21 n.a.
ideation)
McGirr et al. (2007) 156 81 156 156 0 0 0
Bernal et al. (2007) 686/159 48 45/29 34/13 n.a. 11/16 n.a.
(suicidal
ideation)
Depression and Suicide

compared 184 people after suicide attempts to a control group and found
no difference concerning the severity of depression or psychosis in an
assessment by others, while in the self-evaluation scale (Beck’s Depression
Inventory) patients reported significantly higher scores compared to the
matched controls; they also reported higher scores in aggression scales,
hostility scales and impulsiveness scales among patients with suicide
attempts.
In summary, it can be said that among psychiatric diseases, suicide
risk is clearly increased, which makes them an important risk factor for
suicidal behaviour.

DEPRESSION AND SUICIDE

Depression and Suicidal Behaviour: Epidemiological Data

In all psychological autopsy studies, major depressive disorders are the


most frequent psychiatric disorders and hence form the high risk group
for suicidal behaviour, while among psychiatric inpatients, it is above all
young psychotic patients (Wolfersdorf, 2000). According to psychological
autopsy studies, 40–60 percent of all people who died by committing
suicide suffered from a primary depressive episode at the time of suicide.
Lifetime suicide mortality in depression (all grades of severity) is estimated
at 3–4 percent (Blair-West et al., 1997; Wolfersdorf, 2002, 2008), suicide
mortality in major depression is currently between 12 and 15 percent
(Guze and Robins, 1970; Miles, 1977; Wolfersdorf, 2007).
In a more than 40-year long follow-up study with 406 patients suffering
from affective disorders, Angst and colleagues (2005) found a standard
mortality ratio (SMR) of 26.4 for the unipolar and 11.7 for the bipolar
affective disorders, both significant with p < .05.
Schneider, Philipp and Müller (2001) observed 280 patients with a
major depressive episode for more than 5 years and found 16 suicides
(5.7 percent). Coryell and Young (2005) observed 785 patients with
Research Diagnostic Criteria (RDC) major depressive disorder form 1976
to 1990 and found 33 (4.2 percent) suicides.

283
Eva Schaller and Manfred Wolfersdorf

Risk Factors for Suicide in Depression: Psychopathology

The last two decades of suicide research were marked by the attempt to
describe different risk factors for suicide to reduce the likelihood of future
suicidal behaviour. Various current studies provide evidence of specific
psychopathological phenomena and their meaning for acute short-term or
long-term suicide prevention. By means of discriminate function analysis,
Pokorny (1983) describes the following predictors: diagnoses of depression
and schizophrenia, a prehistory of suicide attempts, sleeplessness and
feelings of guilt. However, the predictive usefulness was low: only 35 of
67 suicides could be identified. Goldstein and colleagues (1991)
investigated 1,906 patients with affective disorders and found the usual
risk factors as number of former suicide attempts, occurrence of the
suicide ideas per admission, frequency of bipolar disorders, male gender or
outcome by discharge. But this model also did not allow a reliable identi-
fication of those who passed away by suicide later.
Researchers (Steiner et al., 1992, 1993) explored the interaction between
hopelessness, measured with Beck’s Hopelessness Scale and suicidal be-
haviour in the course of depressive illnesses within the scope of a one-year
follow-up. The motive for doing so was Beck’s notice (Beck et al., 1985,
1990) that raised scores in the hopelessness scale are linked to a raised
suicide risk. From a total of 62 depressed patients, two passed away by sui-
cide in the year of follow-up and eight committed suicide attempts, so that
a total of 16 percent of the whole group showed suicidal actions in the
first year after inpatient therapy at a special depression unit. This reveals
a clear trend, although not statistically significant.
McGirr and colleagues (2007) compared depressive suicides to de-
pressive controls without a suicide attempt (Table 13.2): significant dif-
ferences were found in the suicides with regard to loss of appetite, sleeping
disturbances, feelings of worthlessness and guilt, as well as suicide ideas
and desires to die.
According to Bernal and colleagues (2007), the risk factors for suicidal
behaviour can be summarised as follows: female gender, younger age,
divorced or widowed; also the existence of a psychiatric diagnosis like major
depressive disorder, dysthymia, general anxiety disorder, post-traumatic
stress disorder or alcohol dependency. Schneider et al. (2001) investigated
psychopathological predictors which can be assigned to 16 suicides of
280 depressive patients. During a five-year follow-up they found highly

284
Depression and Suicide

Table 13.2 Major Depressive Disorder and Suicide: Comparison of


Depressive Suicide versus Depressive Non-suicide Controls

Symptoms Suicides (%) Controls (%) OR (95% CI)


Depressed mood 95.4 91.9 1.668
Anhedonia/apathy 84.9 84.7 .863
Weight or appetite loss 85.3 68.2 2.564 (p < .05)
Weight or appetite gain 14.3 19.7 .834
Insomnia 60.3 2.371 (p < .05)
Hypersomnia 33.3 .531 (p < .05)
Psychomotoric disturbance 61.0 49.4 1.583
Fatigue 95.2 .229 (p < .05)
Feelings of worthlessness/guilt 69.8 53.2 2.389 (p < .01)
Concentration or indecisiveness 61.5 75.9 .493 (p < .05)
Thoughts of death or suicidal ideation 96.1 67.5 12.585 (p < .001)
Psychotic features 6.9 2.5 1.222
Source: Adapted from McGirr et al. (2007).

significant mood congruent preoccupations and delusions (p = .008)


(suicides 88 percent, non-suicides 54 percent), hypochondrical pre-
occupations and delusions (p = .002, suicides 37 percent, non-suicides
8 percent), severe hopelessness (p = .022; suicides 68.8 percent, non-
suicides 39.7 percent), delusions of reference (p = .06), initial insomnia
(p = .02) and recurrence (p = .01). Keller and Wolfersdorf (1993) found an
increased suicide risk for the existence of hopelessness; Wolfersdorf and
Vogel (1987) described ‘delusional depression’ as a risk factor.
In their long-term follow-up between 1976 and 1990, Coryell and Young
(2005) identified that suicides were significantly more often in inpatients
(p = .05), had significantly more former suicide attempts (p = .018), re-
ported significantly more frequent hopelessness (p = .02) and in total scored
higher on measurement of suicidal tendencies (p = .002). Rhimer, Rutz
and Pihlgren (1995) tested suicide attempts in the forefront of completed
suicide; within 48 percent of 25 suicides of depressive patients they found
significantly (p < .05) more suicide attempts than with non-depressive sui-
cides (23 percent). Also, Isometsä and colleagues (1994), Barraclough
and Pelles (1975), Pokorny (1983) and Fawcett and colleagues (1987) de-
scribed the impact of former suicide attempts several times. In their
Finnish study, Sokero and colleagues (2003) examined 269 patients with
DSM-IV diagnosis of major depression concerning suicidal tendencies, and
found suicide ideas during the current period of depression in 58 percent
of all patients, and actual suicide attempts by 15 percent of patients.

285
Eva Schaller and Manfred Wolfersdorf

As significant risk factors for suicide ideas, they found hopelessness, alcohol
abuse, poor social integration, lack of social support; as risk factors for
suicide attempts they saw major depression, existing alcohol abuse,
younger age and bad social adaptation.
The study by Vuorilehto and his colleagues (2006) showed that 32 percent
of 137 patients with DSM-IV major depressive disorder have already had
suicide ideas, 17 percent reported suicide attempts in the prehistory;
lifetime suicide mortality was predicted by psychiatric anamnesis as well
as the existence of a comorbid personality disorder; suicide attempts were
predicted most reliably by severity of depressive episode. Wolfersdorf
(1989) found in a multiple prediction analysis of depressed patients
who later died from suicide and depressed patients without suicidal be-
haviour, a common prediction value with Multiple Preanalysis (MUP)
Lambda = 0.31, that is, the combination of hopelessness, depressive
delusion and psychomotoric inhibition resulted in an allocation security
of 31 percent for the identification of suicides. If one suicide attempt in the
prehistory is added to this, allocation security protection rises to 35 percent.
The existence of hopelessness, depressive delusion and a lack of psy-
chomotoric inhibition allow the identification of about one-third of all
deaths by suicide.
To summarise, the comparison of non-suicidal and suicidal depressed
inpatients shows significant differences: suicidal depressed patients show
significantly more frequent sleeping disturbances, in particular, problems
in getting to sleep, suicide attempts before stationary admission, stationary
admission due to suicidal tendency, developmental disturbances in early
childhood and youth as well as suicide attempts among relatives (Metzger
and Wolfersdorf, 1988; Modestin and Knopp, 1988).
If one looks at the data of the Weissenauer follow-up study (Steiner
et al., 1993; Wolfersdorf, Steiner et al., 1990) concerning the scores in
different self-judgement scales, as well as assessments by others for ac-
quisition of depression, hopelessness or physical discomfort, non-suicidal
depressive patients do not differ from the ones reporting suicide ideas,
death wishes or former suicide attempts. Anxiety disorders and reduced
level of self-esteem are exhibited by suicidal depressed patients at time of
stationary admission; suicidal depressed patients with suicide ideas de-
scribe themselves as suffering from significantly more hopelessness, more
anxious and also diminished self-esteem. Suicidal depressed patients with
suicide attempts in their prehistory suffer from significantly reduced

286
Depression and Suicide

self-esteem. Consequently, lack of self-esteem, hopelessness and anxiety


distinguish suicidal depressives from non-suicidal depressives, besides the
other risk factors mentioned earlier.
Besides, well-known psychosocial risk factors such as living alone, being
unmarried, divorced or widowed are to be considered especially among
men, but so are unemployment or migration background. Religion on
the other hand works as a significant protective factor.
An interesting approach has been time-related clinical predictors of
suicide, an approach in particular used by Fawcett and his colleagues (1990)
as well as Coryell and Young (2005). In a group of 154 suicides, McGirr,
Renaud, Séguin, Alda and Turecki (2008) found that 74.7 percent com-
mited suicide during their primary depressive episode, 18.8 percent during
the second and 6.5 percent during further episodes. At the same time,
alcohol abuse as well as anxiety was reported during the index episode.
Coryell and Young (2005) stated that in suicides that happen one year after
a depressive episode, significantly more suicide intentions are reported
than earlier occurring suicides (see Table 13.3).

Table 13.3 Suicidal versus Non-suicidal Depressed Inpatients of the Weissenau


Depression Treatment Unit Therapist´s Assessment at Admission
(Significantly Discriminating Variables of a Patient´s Questionnaire)

Non-suicidal Suicidal
(n = 67) (n = 66)
Variables of the questionnaire (n) (%) (n) (%) Chi
Paranoid ideas 8 13 1 2 ∗
Insomnia, especially early insomnia 42 63 52 81 ∗
Suicide attempt prior to admission 16 24 28 42 ∗
Admission because of suicide intent 5 8 33 55 ∗∗
Developmental disturbances in early childhood 8 12 18 29 ∗
Suicide attempts among relatives 1 2 7 11 ∗

Source: Wolfersdorf et al. (1990).


Note: ∗ p ≤ .05; ∗∗ p ≤ .01.

Fawcett and his colleagues (1990) distinguished predictors for suicide in


the first year after stationary treatment related to alcohol abuse, anhedonia,
anxiety, concentration loss and sleeping disturbances, and later occurring
suicides which distinguished themselves above all by hopelessness and
suicide ideas. In addition, significant psychosocial factors for suicidal
behaviour were male gender, being unmarried, separated or widowed and

287
Eva Schaller and Manfred Wolfersdorf

having had previous suicide ideas or suicide intentions. Other authors


(Beck et al., 1985; Keller and Wolfersdorf, 1993; Schneider et al., 2001;
Wolfersdorf and Niehus, 1993; Wolfersdorf et al., 1987; Wolfersdorf, Hole
et al., 1990) found hopelessness and delusional symptomatology during
the index-episode significantly more often with suicide deaths.

Suicide and Bipolar Affective Disorder

Bipolar affective disorders are considered as a high risk group for suicide.
In their cumulative 17-year follow-up in Scotland, Sharma and Marker
(1994) found a suicide rate of 16 percent in all deaths and calculated
a 23 times higher suicide risk for those with bipolar illness than in the
general population. In Finland, 46 of 1,297 (3 percent) suicide cases within
one year were identified as having bipolar illness (Isometsä et al., 1994;
Isometsä and Lonnqvist, 1997). According to our own overview (Table 13.1),
there is a range of 1–22 percent, nevertheless, the majority is clearly less
than 10 percent. Lönnqvist (2000) reports on psychological autopsy
studies according to DSM-III criteria with data of 1–5 percent. Within suicides,
the depressive episodes dominate at the time of suicide with most suicides
being men. Because most of the suicides take place in a depressive or mixed
manic-depressive state, most of the risk factors also apply to the depressive
suicides with basic bipolar affective disorder (Dilsaver et al., 1994; Lester,
1993; Wolfersdorf et al., 2005).

The Suicidal Depressed Patient: Clinical Picture

Against the background of the description of discerning factors between


suicides and non-suicides made earlier, as well as the inclusion of
clinical experience, the clinical, psychopathological picture of a suicide-
threatened patient is outlined in Table 13.4. Thereby, cognitive symptoms,
psychomotoric symptoms, physical symptoms and also the behaviour of
the patient are distinguished.
The typical suicide-threatened depressive patient whose suicidal actions
happen quite close to acute psychopathology is depressive for the first
time or has had several depressive episodes. He suffers predominantly
from feelings of worthlessness, thoughts of guilt and self-depreciation,

288
Depression and Suicide

Table 13.4 Symptoms Significantly Differentiating Suicides and Non-suicides in


Studies with Depressive Disorder Patients

z Severity of depression
z Thoughts of death, suicide ideations, threatening suicide
z Thoughts of worthlessness and guilt
z Thoughts of helplessness and hopelessness
z Somatic symptoms especially insomnia, fatigue and anhedonia , as well as loss of
appetite and weight
z Disturbances of concentration or decisiveness
z Psychotic features (delusions, hallucinations, paranoid ideas)
z Comorbidity with drug or alcohol abuse/dependency, anxiety disorders, personality
disorders
z Tendency to impulsivity and aggression
Source: Authors’ compilation from various sources (e.g., Wolfersdorf, 1995, 2000, 2007,
2008; Wolfersdorf, Steiner et al., 1990; Wolfersdorf et al. 1992; Wolfersdorf and
Niehus, 1993; Wolfersdorf and Vogel, 1987).

but also altruistic ideas, for example, that it would be better for others if he
was no longer there (Metzger and Wolfersdorf, 1988; Steiner et al., 1993;
Wolfersdorf, 1995, 2007; Wolfersdorf and Niehus, 1993; Wolfersdorf and
Vogel, 1987; Wolfersdorf, Hole et al., 1990).
Lacking self-esteem, feelings of hopelessness, increasing tendency
towards favouring a suicidal solution, as well as in particular cases a
delusional disturbed perception, are typical for the depressive patient with
a heightened suicide risk. Moreover, there are feelings of restlessness and
agitation, sleeping disturbances as well as retreatment and loss of social
contacts at the behavioural level (Table 13.5).
It becomes clear that, in this case, it concerns the description of a heavily
depressed person who shows a raised suicidal risk. If one look, however,
at long-term predictors for raised suicide risk, it is not the acute clinical
picture, but the aforementioned hopelessness with regard to the expected
course of disease which is most important. A greater closeness to suicidal
tendencies and not least insufficient or missing continuity of treatment
are also crucial factors.

Suicide Prevention in Depressed Patients

Primary preventive considerations predominantly deal with the health care


politics field and extend from awareness programmes to de-stigmatisation,
289
Eva Schaller and Manfred Wolfersdorf

Table 13.5 The Depressed Patient with Suicide Risk—Clinical


Psychopathological Picture in the Presuicidal Situation

Psychomotoric Psychosomatic
Cognitive symptoms symptoms symptoms Behaviour
Ideas of worthlessness Inner restlessness Sleep disturbances, Loss of contacts
Ideas of guilt and self- Agitation insomnia and relations with
depreciation others, retreat
Altruistic ideas (the world Hostility
would be better without Cry for help
oneself) Verbal signs of
Lack of self-esteem ambivalence
Feelings of helplessness and
hopelessness
Narrowness of thinking
Depressive delusions
Source: Authors’ compilation from various sources (e.g., Wolfersdorf, 1995, 2007;
Wolfersdorf, Steiner et al., 1990; Wolfersdorf and Niehus 1993; Wolfersdorf and
Vogel, 1987).

whereas secondary preventive measures contain above all early detection,


as well as outpatient, inpatient and post-stationary treatment. However,
tertiary preventive approaches deal in particular with questions of relapse
prophylaxis, treatment of comorbidity groups, addiction prophylaxis, pain
treatment and also old people’s care.
In suicide prevention the main features in establishing a good and
reliable therapeutic relationship are diagnostic clarification of suicidal
tendency and of the psychosocial crisis or psychopathology; caring
therapeutic measures, including high frequency protective care; therapy
of the basic illness by crisis intervention, adequate psychopharmacotherapy
and psychotherapeutic intervention. Today, general depression treat-
ment encompasses the pillars: psychotherapy, biological measures
of treatment, sociotherapeutic-psychoeducative and occupational
therapeutic elements, as well as self-help programmes and cooperation
with relatives. In the treatment of acute suicidal tendency within major
depressive disorder, however, alongside high frequency caring therapeutic
conversation and sedating-anxiolytic psychopharmacotherapy, a crucial
aspect is safeguarding measures to come through the phase of acute sui-
cidal tendency. According to relapse-prophylactic aspects, long-term
psychopharmacotherapy and psychotherapy, as well as the comprehension
of relatives are important suicide-preventive factors.

290
Depression and Suicide

REFERENCES

Agoub, M., D. Moussaoui and N. Kadri (2006). Assessment of suicidality in a Moroccan


metropolitan area. Journal of Affective Disorders, 90(2), 223–26.
Angst, J., F. Angst, R. Gerber-Werder and A. Gamma (2005). Suicide in 406 mood-
disorder patients with and without long-term medication: A 40 to 44 years’
follow-up. Archives of Suicide Research, 9(3), 279–300.
Appelby, L., J. Shaw, T. Amos, R. McDonnell, K. Kiernan, S. Davies et al. (1999).
Suicide within 12 month of contact with mental health services: National clinical
survey. British Medical Journal, 318(7193), 1235–39.
Apter, A., A. Bleich, R.A. King, S. Kron, A. Fluch, M. Kotler et al. (1993). Death without
warning? A clinical post-mortem study of suicide in 43 Israeli adolescent males.
Archives of General Psychiatry, 50(2), 138–42.
Arato, M., E. Demeter, Z. Rihmer and I.E. Somogy (1988). Retrospective psychiatric
assessment of 200 suicide in Budapest. Acta Psychiatrica Scandinavica, 77(4),
454–56.
Asgard, U. (1990). A psychiatric study of suicide among urban Swedish women. Acta
Psychiatrica Scandinavica, 82(2), 115–24.
Barraclough, B.M. and D.J. Pelles (1975). Depression followed by suicide: A comparison
of depressed suicide with living depressive. Psychological Medicine, 5(1), 55–61.
Beck, A.T., G. Brown, R.J. Preschick, B.L. Steward and R.A. Steer (1990). Relationship
between hopelessness and ultimate suicide: A replication study with psychiatric
outpatients. American Journal of Psychiatry, 147(11), 559–63.
Beck, A.T., R.A. Steer, M. Kovacs and B. Garrison (1985). Hopelessness and eventual
suicide: A ten-year prospective study of patients hospitalized with suicidal audition.
American Journal of Psychiatry, 142(5), 559–63.
Bernal, M., J.M. Haro, S. Bernert, T. Brugha, R. de Graaf, R. Bruffaerts et al. (2007).
Risk factors for suicidality in Europe: Results from the ESEMED study. Journal of
Affective Disorders, 101(1), 27–34.
Bertolote, J.M., A. Fleischmann, D. De Leo and D. Wasserman (2004). Psychiatric
diagnosis and suicide: Revisiting the evidence. Crisis, 25(4), 147–55.
Beskow, J. (1979). Suicide and mental disorder in Swedish men. Acta Psychiatrica
Scandinavica, 277(supplement), 1–138.
Blair-West, G.W., G.W. Mellsop and M.L. Eyeson-Annan (1997). Down-rating lifetime
suicide risk in major depression. Acta Psychiatrica Scandinavica, 95(3), 259–63.
Brent, D.A., M. Baugher, J. Bridge, T. Chen and L. Chiappett (1999). Age-and sex-
related risk factors for adolescent suicide. Journal of American Academy of Child
and Adolescent Psychiatry, 38(12), 1497–1505.
Brent, D.A., J.A. Perper, G. Moritz, C. Allmen, A. Friend, C. Roth et al. (1992).
Psychiatric risk factors for adolescent suicide: a case-control study. The Journal of
the American Academy of Child and Adolescent Psychiatry, 32(3), 521–29.

291
Eva Schaller and Manfred Wolfersdorf

Cheng, A.T. (1995). Mental illness and suicide. A case-control study in East Taiwan.
Archives of General Psychiatry, 52(7), 594–603.
Chynoweth, R., J.I. Tonge and J. Amstrong (1980). Suicide in Brisbane: A retrospective
psychosocial study. Australian and New Zealand Journal of Psychiatry, 14(1),
37–45.
Conwell, Y., P.R. Duberstein, C. Cox, J.H. Herrman, N.T. Forbes and E.D. Cain
(1996). Relationship of age and axis I diagnoses in victims of completed suicide:
A psychological autopsy study. American Journal of Psychiatry, 153(8), 1001–08.
Coryell, W. and E.A. Young (2005). Clinical Predictors of suicide in Primary Major
Depressive Disorder. Journal of Clinical Psychiatry, 66(4), 412–17.
Dilsaver, S.C., Y.W. Cheng and A.C. Swann (1994). Suicidality inpatients with pure and
depressive mania. American Journal of Psychiatry, 151(9), 1312–15.
Dorpat, T.L. and H.S. Ripley (1960). A study of suicide in the Seattle area. Comprehensive
Psychiatry, 1, 349–59.
Fawcett, J., W. Scheftner, D. Clark, D. Hedeker, R. Gibbons and W. Coryell (1987).
Clinical predictors of suicide inpatients with major affective disorders: A control
prospective study. American Journal of Psychiatry, 144(1), 35–40.
Fawcett, J., W. Scheftner, L. Fogg, D. Clark, M.A. Young, D. Hedeker et al. (1990).
Time-related predictors of suicide in major affective disorders. American Journal
of Psychiatry, 147(9), 1189–94.
Foster, T., K. Gillesbie and R. Mc Clelland (1997). Mental disorders and suicide in
northern Ireland. British Journal of Psychiatry, 170(5), 447–52.
Goldstein, R.B., D.W. Black, A. Nasrallah and G. Winokur (1991). The prediction of
suicide. Archives of General Psychiatry, 48(5), 418–22.
Griesinger, W. (1845). Melancholy accompanied by impulses to destroy. In
W. Griesinger (Ed.), The pathology and therapy of psychological illnesses, for doctors
and students (pp. 191–207). Stuttgart: Krabbe.
Guze, S.B. and E. Robins (1970). Suicide and primary affective disorder. British Journal
of Psychiatry, 117(539), 437–38.
Haenel, T. and W. Pöldinger (1986). Recognition and assessment of suicidal tendency.
In K. Kisker, H. Lauter, J.E. Mayer, C. Müller and E. Strömgren (Eds), Psychiatrie
der Gegenwart 2 (pp. 107–32). Berlin: Springer.
Hall, R.C.W. and D.E. Platt (1999). Suicide risk assessment: A review of risk factors
for suicide in 100 patients who made severe suicide attempts. Psychosomatic,
40(1), 18–27.
Harris, C.E. and B.M. Barraclough (1997). Suicide as an outcome for mental disorder.
British Journal of Psychiatry, 170(3), 205–08.
Harris, C.E. and B.M. Barraclough (1998). Excess mortality of mental disorder. British
Journal of Psychiatry, 173(1), 11–53.
Heilä, H., E.T. Isometsä, M.M. Henriksson, M.E. Keikkinen, M.J. Marttunen and
J. Lönnqvist (1987). Suicide and schizophrenia: A nationwide psychological autopsy
study on age-and sex-specific clinical characteristics of 92 suicide victims with
schizophrenia. American Journal of Psychiatry, 154(9), 1235–42.

292
Depression and Suicide

Henriksson, M.M., H.M. Aro, M.J. Marttunen, M.E. Heikkinen, E.T. Isometsä,
K. Kuoppasalme et al. (1993). Mental disorders and comorbidity in suicide.
American Journal of Psychiatry, 150(6), 935–40.
Hole, G. (1973). Suicidal tendency and loss of self-worth in depressed people. Psy-
chotherapy and Medical Psychology, 23, 233–38.
Isometsä, E.T. and J.K. Lönnqvist (1997). Suicide in mood disorders. In J.A. Botsis,
C.R. Soldatos and C.N. Stefanis (Eds), Suicide: Biopsychosocial approaches
(pp. 33–47). Amsterdam: Elsevier.
Isometsä, E.T., M.M. Henriksson, H.M. Aro, M.E. Heikkinen, K.I. Kuoppasalmi and
J.K. Lönnqvist (1994). Suicide in major depression. American Journal of Psychiatry,
151(4), 530–36.
Keller, F. and M. Wolfersdorf (1993). Hopelessness and the tendency to commit suicide
in the course of depressive disorders. Crisis, 14(4), 173–77.
Kessler, R.C., M. Angermeyer and J.C. Antony (2007). Lifetime prevalence and age- of –
onset distributions of mental disorders in the World Health Organization’s World
Mental Health Survey Initiative. World Psychiatry, 6(3), 168–76.
Kung, A.C., J.H. Pearson and X. Lin (2003). Risk factors for male and female suicide
decendents age 15–64 in the United States: Results from the 1993 National
Mortality Follow back Survey. Social Psychiatry and Psychiatric Epidemiology,
38(8), 419–26.
Lesage, A.D., R. Boyer, F. Grunberg, C. Vanier, R. Morissette, C. Menard-Buteau et al.
(1994). Suicide and mental disorders: A case-control study of young men. American
Journal of Psychiatry, 151(7), 1063–68.
Lester, D. (1993). Suicidal behaviour in bipolar and unipolar affective disorders:
A meta-analysis. Journal of Affective Disorders, 27(2), 117–21.
Lönnqvist, J.K. (2000). Psychiatric aspects of suicidal behaviour: In K. Hawton and
K. van Heeringen (Eds), The international handbook of suicide and attempted suicide
(pp. 107–20). New York: Wiley & Sons.
Mann, J.J., C. Waternaux, G.L. Haas and K.M. Malone (1999). Toward a clinical
model of suicidal behavior in psychiatric patients. American Journal of Psychiatry,
156(2), 181–89.
Marttunen, M.J., H.M. Aro, M.M. Henriksson and J.K. Lönnqvist (1991). Mental
disorders in adolescent suicide. DSM-III-R axes I and II diagnoses in suicides
among 13- to 19-years-olds in Finland. Archives of General Psychiatry, 48(9),
834–39.
McGirr, A., J. Renaud, M. Séguin, M. Alda and G. Turecki (2008). Course of major
depressive disorder and suicide outcome: A psychological autopsy study. Journal
of Clinical Psychiatry, 69(6), 966–70.
McGirr, A., J. Renaud, M. Séguin, M. Alda, C. Benkelfat, A. Lesage et al. (2007). An
examination of DSM-IV depressive symptoms and risk for suicide completion
in major depressive disorder: A psychological autopsy study. Journal of Affective
Disorders, 97(1–3), 203–09.

293
Eva Schaller and Manfred Wolfersdorf

Metzger, R. and M. Wolfersdorf (1988). Suicides among patients treated in a world


specializing in affective disorders. In H.J. Möller, A. Schmidtke and R. Welz (Eds),
Current issues of suicidology (pp. 1–108). Berlin: Springer.
Miles, C. (1977). Conditions predisposing to suicide: A review. Journal of Nervous and
Mental Disease, 164(4), 231–46.
Mitterauer, B. (1981). Können Selbstmorde in einem Psychiatrischen Krankenhaus
verhindert werden? Psychiatrische Praxis, 8, 25–30.
Modestin, J. and W. Knopp (1988). Study on suicide in depressed inpatients. Journal
of Affective Disorders, 15(22), 157–62.
Morgan, H.G. and R. Stanton (1997). Suicide among psychiatric inpatients in a changing
clinical scene. British Journal of Psychiatry, 171(6), 561–63.
Petronis, K.R., J.F. Samuels, E.K. Moscicki and J.C. Anthony (1990). An epidemiologic
investigation of potential risk factors for suicide attempts. Social Psychiatry and
Psychiatric Epidemiology, 25(4), 193–99.
Phillips, M.R., G. Yang, Y. Zhang, L. Wang, H. Ji and M. Zhou (2002). Risk factors for
suicide in China: A nationl case-control psychological autopsy study. The Lancet,
360 (9347), 1728–36.
Pokorny, A.D. (1983). Prediction for suicide in psychiatric patients. Report of a
prospective study. Archives of General Psychiatry, 40(3), 249–57.
Rhimer, Z., W. Rutz and H. Pihlgren (1995). Depression and suicide on Gotland. An
intensive study of all suicides before and after a depression-training programme
for general practitioners. Journal of Affective Disorders, 35(4), 147–52.
Rich, C.L., D. Young and R.C. Fowler (1986). San Diego Suicide Study I: Young versus
old subjected. Archives of General Psychiatry, 43, 577–82.
Robins, E., S. Gassner, J. Kayes, R.H. Wilkinson and G.E. Murphy (1959). The com-
munication of suicidal intent: A study of 134 consecutive cases of successful
(completed) suicide. American Journal of Psychiatry, 115(8), 724–33.
Roy, A. (1984). Suicide in recurrent affective disorder patients. Canadian Journal of
Psychiatry, 29(44), 319–22.
Runeson, B. (1989). Mental disorder in youth suicide. Acta Psychiatrica Scandinavica,
79, 490–97.
Sainsbury, P. (1986). Depression, suicide, and suicide prevention. In A. Roy (Ed.),
Suicide (pp. 73–88). London: Williams & Wilkins.
Schneider, B. (2003). Risk factors for suicide. Regensburg: Roderer.
Schneider, B., M. Philipp and M.J. Müller (2001). Psychopathological predictors of
suicide in patients with major depression during a 5-year follow-up. European
Journal of Psychiatry, 16(5), 283–88.
Shaffer, D., M.S. Gould, P. Fisher, P. Trautman, D. Moreau, M. Kleinman et al. (1996).
Psychiatric diagnosis in child and adolescent suicide. Archives of General Psychiatry,
53(4), 339–48.
Shaffi, M., J. Steltz-Lenarsky, A.M. Derrick, C. Beckner and J.R. Whittinghill (1988).
Comorbidity of mental disorders in the post-mortem diagnosis of completed
suicide in children and adolescents. Journal of Affective Disorders, 15(3), 227–33.

294
Depression and Suicide

Sharma, R. and H.R. Marker (1994). Mortality in affective disorder. Journal of Affective
Disorders, 31(22), 91–96.
Sokero, T.P., T.K. Melartin, H.J. Rytsälä, U.S. Leskelä, Lestelä- P.S. Mielonen and
E.T. Isometsä (2003). Suicidal ideation and attempts among psychiatric patients
with major depressive disorder. Journal of Clinical Psychiatry, 64(9), 1094–100.
Steiner, B., M. Wolfersdorf, F. Keller and G. Hole (1993). The relationship between
hopelessness and a tendency to commit suicide in the course of depressive
disorders. In K. Böhme, R. Freytag, C. Wächtler and H. Wedler (Eds), Suicidal
behavior. The state of the art (pp. 769–74). Regenburg: Roderer.
Steiner, B., M. Wolfersdorf and F. Keller (1992). The course of disease in delusional
depressed patients. Fundamenta Psychiatrica, 6, 31–36.
Vijayakumar, L. and S. Rajkumar (1999). Are risk factors for suicide universal? A case-
control study in India. Acta Psychiatrica Scandinavica, 99(6), 407–11.
Vuorilehto, M., T. Melartin and E. Isometsä (2006). Suicidal behaviour among primary-
care patients with depressive disorders. Psychological Medicine, 36(2), 203–10.
Wittchen, H.U., U. Müller, H. Pfister, S. Winter and B. Schmidtkunz (1999). Affective,
somatoform and anxiety disorders in Germany – First results from the nationwide
survey. Gesundheitswesen, 61, 216–22.
Wolfersdorf, M. (1989). Suicide in psychiatric inpatients (pp. 1–297). Regensburg:
Roderer.
Wolfersdorf, M. (1995). Depression and suicidal behaviour: Psychopathological
differences between suicidal and non-suicidal depressive patients. Archives of
Suicide Research, 1(4), 273–88.
Wolfersdorf, M. (2000). The suicidal patient in hospital (pp. 1–222). Stuttgart:
Wissenschaftliche Verlagsgesellschaft.
Wolfersdorf, M. (2002). Depression and suicide: New clinical aspects and research
results. In G. Laux (Ed.), Depression 2000 (pp. 163–81). Berlin: Springer.
Wolfersdorf, M. (2007). Depression and suicide prevention. In O. Elmer (Ed.),
Psychotherapie affektiver Störungen (pp. 71–89). Tübingen: DGVT-Verlag.
Wolfersdorf, M. (2008). Depression and Suicide. Bundesgesundheitsblatt –
Gesundheitsforschung – Gesundheitsschutz, 51(4), 1–8.
Wolfersdorf, M. and E. Niehus (1993). Depressive inpatients and suicidal behaviour:
A comparison of suicidal and non-suicidal depressives. Schweizer Archiv für
Neurologie und Psychiatrie, 144(6), 575–83.
Wolfersdorf, M. and R. Vogel (1987). Types of hospital suicide and depression: Some
selected results of a study on suicides committed during psychiatric hospitalization.
Crisis, 8(1), 37–48.
Wolfersdorf, M., B. Lehle and L. Adler (2005). Bipolar affective disorders and suicide
during psychiatric in-patients treatment. Archives of Suicide Research, 9(3),
261–66.
Wolfersdorf, M., B. Steiner, F. Keller, M. Hautzinger and G. Hole (1990). Depression
and suicide. Is there a difference between suicidal and non-suicidal depressed
inpatients? European Journal of Psychiatry, 4(2), 235–52.

295
Eva Schaller and Manfred Wolfersdorf

Wolfersdorf, M., F. Keller and B. Steiner (1987). Delusional depression and suicide.
Acta Psychiatrica Scandinavica, 76(4), 359–61.
Wolfersdorf, M., F. Neher, C. Franke and C. Mauerer (2002). Schizophrenia and
affective psychoses. In T. Bronisch, P. Götze, A. Schmidtke, M. Wolfersdorf (Eds),
Suicidality (pp. 175–201). Schattauer, Stuttgart.
Wolfersdorf, M., G. Hole, B. Steiner and F. Keller (1990). Suicide risk in suicidal versus
nonsuicidal depressed inpatients. Crisis, 11(2), 85–97.
Wolfersdorf, M., V. Faust and R. Hölzer (1992). Suicide in the region of Ravensburg/
Oberschwaben. Results from a study of 508 suicides on the basis of police
documents. Schweizer Archiv für Neurologie und Psychiatrie, 143(6), 485–95.

296
14

The Suicidal Soldier


LƒRÝ M›«½çà ƒÄ— Lƒã«ƒ NR禫ƒÃ

E pidemiology and empirical research in the fields of mental health, the


social sciences and biology have contributed to the increasing recognition
of the multi-dimensional and overlapping risk factors of suicidal behaviour.
It is common to include the spectrum of suicidal ideation, via deliberate
self-harm with either expressed or inferred suicidal intent or suicide when
using the term ‘suicidal behaviour’. Completed and attempted suicide share
many common aspects: depressive and other psychiatric disorders, psy-
chological aspects such as poorer problem-solving skills and higher levels
of impulsitivity, social factors such as adversities during childhood and
alcohol habits, and biological factors such as increased age. However, dif-
ferences between completed suicides and attempted suicides also exist,
such as increased vulnerability of the male gender for completed suicide
and the female gender for attempted suicide.
Understanding thus gained has been used to design national prevention
strategies in the early 1990s, especially in developed nations witnessing strong
increases in their suicide rates, such as Norway and Finland. Although
some promising treatment approaches have been identified (Hawton et al.,
2000), our knowledge of effective treatments of suicidal individuals is quite
limited. Furthermore, suicide prevention directed at high-risk groups such
as psychiatric patients, even if proven effective in preventing suicide among
patients, is likely to have very limited impact in non-clinical populations.

297
Lars Mehlum and Latha Nrugham

Collective societies such as the armed forces is a typical non-clinical arena


requiring specialised suicide prevention strategies, adapted to specific
challenges and needs.
A collective society is one wherein the individual, more or less willingly,
takes on a collective identity or, in other words, accepts himself as a part
of a collective identity, with the interdependency of the parts determining
the roles and functions of the parts. The goal of the collective is considered
more important than personal goals which are subservient to the collective’s
goals. Although, this subservience of personal goals can, by itself be stressful
and threatening to personal integrity, the collective is also expected to take
care of the individual’s needs at all levels. Similarly, a threat to the collective
is also supposed to elicit an appropriately defensive response from the in-
dividuals. This kind of specific and complex organisation, whether formal
or informal, is the hallmark of collective societies. Therefore, suicidal
phenomena within a collective society, like the defence forces, needs to
be dealt with differently than the rest of society.
A study of US Air Force, Army, Marine Corps and Navy recruits from
1980 through 2004 revealed that overall suicide rates among military
recruits were lower than those of comparably aged US civilians (Scoville
et al., 2007). This is the usual finding when suicide in the military is
compared with other segments of the population of the same gender and
age (Hytten and Weisaeth, 1989; Mahon et al., 2005; Wasserman, 1992),
including long-term (20 years) mortality of suicide attempters (Mehlum,
1994). This is probably due to what is commonly known as ‘healthy
worker effect’ and is seen as the selection effect of the ruling out of many
psychiatric conditions while being admitted to the armed forces and con-
tinuing health monitoring. Although US army personnel with service in
Iraq had higher suicide rates than the average US army rates, it was lower
than the civilian male population rate (Nelson, 2004). However, there are
exceptions to this pattern in military subgroups and in certain countries.
For example, a significantly increased standard mortality rate (SMR) of
1.4 was found among former peacekeepers from Norway, which only dis-
appeared when adjusted for marital status, indicating the vulnerability of
single military personnel (Thoresen et al., 2003). This study also found
significant increases in suicide by firearms, as early and recent studies
of suicides among active military duty personnel in the United States
have, indicating the continuing importance of gun control as a prevention
strategy regardless of geography (Helmkamp, 1995; Scoville et al., 2007;

298
The Suicidal Soldier

Thoresen et al., 2003). The vulnerability of male as compared to female


military recruits is the same as in the rest of society (Scoville et al., 2007).
Military personnel in the age range of 17 to 24 years have been found to
be the most vulnerable, accounting for 48 percent of the suicides and the
highest age group-specific rate (Helmkamp, 1995).
It is also known that while suicide is the eighth leading cause of death
in the United States, it is the third leading cause of death in its military
population (Sentell et al., 1997). Suicide is also the leading cause of sudden
violent deaths (deaths by unintentional injuries, suicide or homicide)
(Scoville et al., 2004) which account for three-fourths of active duty
military deaths (Helmkamp and Kennedy, 1996; Powell et al., 2000).
Therefore, effective management of suicidal behaviour in the military
has an important role to play in decreasing the overall death rate among
military personnel.
In this chapter, we will try to describe the emergence, maintenance
and prevention of suicidal behaviour in military settings by examining
environmental and individual factors, using published research and our
own experiences with suicide and suicide prevention within Norwegian
Armed Forces and in other nations. This examination will have three
parts: risk factors, protective factors and assessment, wherein each part is
composed of specific topics. Assessment of risk and protective factors will
be dealt with in the environmental and individual contexts. The individuals
we refer to are basic trainees, officers and veterans. We continuously use
the male gender while referring to the individual as the bulk of soldiers
are male. However, a section is also devoted to the nature of risks faced by
the female soldier, as female soldiers are now increasing in numbers.
Unless specifically mentioned as suicide attempts or suicidal ideation,
we use available information on suicide in the military setting throughout
this chapter. A summary of this information and knowledge will also in-
clude suggestions for effective prevention measures.

SUICIDAL BEHAVIOUR IN MILITARY SETTINGS

In this section, we examine the various environmental and individual


factors which contribute to suicidal behaviour among military
personnel, whether as independent/solitary factors or as interacting/

299
Lars Mehlum and Latha Nrugham

combining factors. For most military personnel, the armed forces serve
as a total institution, comprising not only of a job and working hours,
but most areas of life (Mehlum and Schwebs, 2001). As such, we discuss
risk and protective factors of suicide in the military as an environment in
which many soldiers spend most of their time, round the clock.

RISK FACTORS

SelecƟon and Training Processes

Due to the traditionally stringent requirements of military selection pro-


cesses which uses several steps to ensure that the healthiest candidates
are selected, persons with severe mental health problems or other vulner-
abilities which might impair their functioning have little chance of getting
into the military. The healthy young adults who are thus selected are
then trained to become healthier and provided with support to maintain
their physical, mental and social health. This means that the suicide rates
within the military have often remained lower than comparable civilian
populations, regardless of time and place, as stated earlier.
However, exposure to certain conditions and environments may pro-
vide elevated risk levels to both basic trainees, active duty personnel and
veterans resulting in increases in mental health problems and in suicidal
phenomena such as suicidal ideation and suicide attempts.

Rapid Changes in Social IntegraƟon

Although a high level of social integration is a protective factor of suicidal


behaviour, rapid decreases in levels of social integration may result in
higher suicide rates (Durkheim, 1897) as was seen among Norwegian
youth in the 1980s (Mehlum et al., 1999). Rapid transitions in the im-
mediate social environment imply a high level of uncertainty and un-
predictability to which the individual must adjust. In environments
where such rapid transitions are frequent and the individual has limited
capacities to adjust or is not able to adjust can result in a perceived loss

300
The Suicidal Soldier

of social support, loss of face, shame and dishonour. In certain cases, this
may result in the development of a feeling of being trapped without any
escape routes other than suicide as a viable alternative. Such helplessness
is taken up in the job demand-control model of suicide (Karasek and
Theorell, 1990) and has been empirically tested among male Japanese
workers in a multicentre study with supportive findings (Tsutsumi et al.,
2007). The authors conclude that job redesign aimed at increased worker
control could be a worthwhile strategy in preventing, or at least reducing
the risk of death by suicide.
The military lifestyle with frequent moves and relocations implies
constant changes in the social environment which may be perceived as en-
riching but which may also make it difficult for usual social relationships
to be developed or maintained and for family structures to survive. Among
male peacekeepers, the suicidal vulnerability of being single has been
documented (Thoresen and Mehlum, 2006; Thoresen et al., 2003). This
vulnerability may be extended to those military personnel who are in
uncertain and unpredictable work environments.

Loss of Individualism and Conformity Pressure

The stamp of the collective can be so strong as to suppress individualism to


an extent that the person experiences a loss of individualism. In a military
setting, the uniformity expected and required may disturb the balance
between interdependence and autonomy. What might be experienced
as group cohesion and social support providing safety and insurance to
most might be experienced as eradication of the individual personality
by those who are vulnerable. This threat of eradication of individualism
may by itself, if large and intense enough, or in combination with several
other factors such as loss of social support and loss of meaning, lead to the
consideration of suicide as a viable option (Mehlum, 1992). Not only are
military personnel required to adhere to the uniformity requirements of
their work setting, but they are also required to conform to a prescribed set
of norms and codes during times when they are not at work. The inability
of those who desire to be different yet are afraid of facing the conse-
quences of non-conformity may build up as internal tension, which over
a period of time, may also lead to considering suicide as a real option.

301
Lars Mehlum and Latha Nrugham

Military Lifestyle

The earlier mentioned frequent residential moves and changes in job


profiles so typical to the military lifestyle with their potential for increased
stress in the individual, family and social life may lead to detrimental effects
on families such as interpersonal conflicts, alcohol abuse, divorce, anger
outbursts and violence which may not or may be self-directed. The easy
availability of alcohol and the military traditions of drinking have been
reported to be risk factors of suicidal behaviour (Mehlum et al., 2006;
Rossow and Amundsen, 1997; Thoresen and Mehlum, 2004).

TraumaƟc Stress Exposure

The exposure of military personnel to high rates of traumatic events during


combat and peacekeeping operations is logically expected and scientifically
documented (Mehlum and Weisaeth, 2002; Prigerson, et al., 2001, 2002).
Exactly how and specifically how much such exposure contributes to
later mental health issues has been extensively investigated, with much
evidence for a positive association (Mehlum and Weisaeth, 2002; Sareen
et al., 2007, 2008; Southwick et al., 1995) but also with some negative
findings (Horn et al., 2006; Unwin et al., 1999; Wong et al., 2001). The
differences in these findings may be reflections of real differences between
the contexts of the samples studied or methodological differences as is
possible between sample compositions, questionnaire and interviews used
as assessment tools, and cross-sectional and longitudinal designs. These
methodological differences are often highlighted as the rationale for new
studies using uniform or standard assessment methodology (Sareen et al.,
2007; Thoresen et al., 2003; Unwin et al., 1999).
Recently published reports, although cross-sectional and retrospective
in nature, overcome many of the limitations of the earlier studies and
reveals that deployment to combat operations and witnessing atrocities
were associated with increased prevalence of mental disorders, including
suicidal ideation and perceived need for care (Sareen et al., 2007, 2008).
Further, their findings also reveal that after adjusting for the effects of
exposure to combat and witnessing atrocities, deployment to peacekeeping
operations was not associated with an increased prevalence of mental
disorders (Sareen et al., 2007).

302
The Suicidal Soldier

The latest published report on the subject investigated Norwegian male


peacekeepers seven years after redeployment (Thoresen and Mehlum,
2008). In this sample, 6 percent of the veterans and 17 percent of the
subset of prematurely repatriated personnel reported suicidal ideation.
Service stress exposure level, even after adjustment for background factors,
repatriation status, negative life events, social support, alcohol con-
sumption, marital and occupational status, predicted suicidal ideation. The
findings indicated that post-traumatic stress disorder (PTSD) symptoms
and general mental health problems combined to mediate the relationship
between service stress exposure and suicidal ideation.
In a fine-grained analysis on how negative events could be related to
suicidal crises among predominantly young male US army personnel it
was found that negative events were related to the intensity of suicidal
crises among single attempters along with non-attempters but not among
multiple attempters, while negative events were related to the duration of
suicidal crises among multiple attempters but not among single attempters
and non-attempters (Joiner and Rudd, 2000). Although the study was based
on correlational data which eliminates causal inferences and demonstrated
relatively small effect sizes, the findings revealed that persons highly
vulnerable to suicide and a higher risk of suicide (multiple attempters)
reported persistent states of distress and lack of responsiveness to external
events. In other words, the authors state that negative events and severity
of suicidality were relatively independent for multiple attempters in this
sample, replicating a similar relationship between stress and depression,
explained by a sensitisation of previous suicide-related experiences which
make the suicidal behaviours more accessible to the individual, making it
easier for a suicidal crisis to be triggered, which is likely to be more severe
than the earlier ones. It is useful to note here that an in-depth study of
suicides among Norwegian soldiers over an eight-year period led the
authors to conclude that the suicides were frequently precipitated by acute
crises and the distance from home to the duty post had an important
role as a possible causal factor (Hytten and Weisaeth, 1989).

Loss of Meaning

This is one of the comparatively lesser researched dimensions of suicidal


phenomena in military settings. An early study of the concept named

303
Lars Mehlum and Latha Nrugham

‘sense of coherence’ (Antonovsky, 1993) and its relationship to suicidal


ideation among basic military trainees found that those reporting suicidal
ideation had lower mean scores of a sense of coherence, indicating that
when environmental stress may more easily initiate a suicidal process in
individuals with a low sense of coherence (Mehlum, 1998). This finding has
been replicated in other military samples (Giotakos, 2003). A later study
investigated peacekeepers from Norway nearly seven years after service
and found that perceived lack of meaningfulness with respect to military
mission significantly and multivariately predicted a higher level of post-
traumatic stress symptom scores (Mehlum and Weisaeth, 2002).
Loss of meaning is the term used to describe the concept of ‘existential
vacuum’ espoused by Frankl (1967), which he believed could arise when an
individual experienced a sense of loss of meaning in life, a sense of complete
emptiness and an absence of a purpose for continuing to live, wherein
suicide may appear as a plausible alternative to the distress generated by
such experiences of loss. This loss of meaning may be related to the distress
or mental pain considered unbearable by the individual experiencing it
and which has been termed as ‘psychache’ by Shneidman (1993). A sense of
inner rupture, being damaged and shattered, losing parts of the self, losing
self-coherence, losing the ability to function fully and losing connectedness
with others have been found to be at the core of mental pain (Bolger,
1999). Empirical evidence of the relationship between suicidal behaviour
and a loss of meaning has been published (Harlow et al., 1986; Lester and
Badro, 1992; Mehlum, 1998; Orbach, Mikulincer, Gilboa-Schechtman and
Sirota, 2003; Orbach, Mikulincer, Sirota and Gilboa-Schechtman, 2003;
Petrie and Brook, 1992).

Loss of Social Support

This is also one of the lesser researched dimensions of suicidal phenomena


in the military. However, Durkheim’s theory (1897), with and without its
criticisms, provides theoretical pillars for explanations about how social
support could be related to suicide rates. Durkheim postulates that both,
a too high and too low level of social integration, can result in higher
suicide rates. This perception of the presence and absence of social support
by the individual and its acceptance or rejection is the level and grade of

304
The Suicidal Soldier

connectedness experienced by the individual to the group he either belongs


or desires to belong to. Social support could also mean the presence of
those relationships that a person desires to have and has.
The suicidal impact of the disruption of such relationships or the
unavailability of circumstances to form age-appropriate relationships, such
as romantic relationships, due to the unique nature of military settings
has also been documented (Thoresen and Mehlum, 2006; Thoresen
et al., 2003). A report of greater social isolation was one of the variables
associated with higher suicidal intent among emergency department
patients (Haw and Hawton, 2008). It is important to underline here that it
is the perception of the individual of the social support that has been found
to be associated with suicidal phenomena, not the objective assessment
of social networks or social supports available to the individual. This
perception of loss of social support by the individual at risk may not be
easy to detect among military personnel as in civilian settings due to the
regimentation of daily life and the increased levels of group life in military
settings. Remote postings with little contact with other group members or
possibilities for a normal social life may be situations resulting in loss of
support. In other cases, social exclusion of vulnerable or randomly selected
individuals is seen and may result in them being bullied, which in reality
represents negative group processes.

Downsizing and Involuntary RepatriaƟon

While employment has its own set of stresses that an employee has to cope
with, unemployment comes with an additional set of stresses which the
employee may not be able to cope with, such as loss of face and fellowship,
loss of rank and income, loss of role and function, loss of residence and
opportunities to lead a family life, amongst other losses such as special
privileges, perks and social status.
During unstable economic times, downsizing travels under different
names such as budget adjustments, staff re-organisation, workforce opti-
misation and streamlining before moving on to more clear ones such as
free-time, pay-cuts, reduction in force, voluntary retirements, lay-offs,
attrition and downsizing. Military work settings are placed in the larger
contexts of their own societies and cannot remain untouched by the

305
Lars Mehlum and Latha Nrugham

stability or instability of these societies. With war becoming more of a


remote threat than an actual event, maintaining a military with personnel
not in combat can drain national budgets. During prolonged periods of
peace, a large number of ready-for-combat personnel may appear to be
redundant and disaster tasks may not be as frequent as to require the
maintenance of a relatively large military force on a continuous basis.
Thus, staff-on-assignment may appear to be an attractive proposition,
leading to military personnel having to involuntarily leave the work and
life they prepared for and chose for themselves. With personnel available
for hire on task basis, the long-term commitments of the employer also
lessen and when financial leashes pull the budget, ex-servicemen may be
disillusioned ex-heroes experiencing the pain and distress of dishonour for
reasons beyond their control and not directly attributable to themselves.
Unemployment rates have been well-documented as correlates of higher
suicide rates in civilian settings (Agerbo et al., 2007; Blakely et al., 2003;
Kposowa, 2001). Among ex-military personnel, the additional stresses of
seeking employment in civilian settings and not being trained to work or
live in such settings may be present.
Involuntary repatriation also led to re-adaptation problems in a
prospective study of among UN military observers in South Lebanon
(Mehlum et al., 2006). The authors found that these veterans reported
significantly higher levels of war zone stressors than peacekeepers, and
significantly more post-traumatic stress (PTS) symptoms at follow-up
(four years later), higher alcohol consumption levels during service
and at follow-up, and more social adaptation problems to their lives at
home after service. Exposure during the mission and problems with social
adaptation after homecoming predicted PTS symptoms at follow-up in
this sample. A similar finding on repatriation was also reported among
Norwegian male veterans of peacekeeping operations (Thoresen et al.,
2006). Among Vietnam veterans, longitudinal models estimated causal
relationships among PTSD, drug dependence and suicidality over a 25-
year period (Price et al., 2004). Their results revealed evidence of strong
continuity of PTSD, drug dependence and suicidality over time. The causal
role of drug dependence on PTSD and suicidality was limited to young
adulthood while evidence was stronger for self-medication during later
adulthood in this sample.

306
The Suicidal Soldier

Psychopathology

The variable ‘mental disorders’ had the strongest association with suicide
in a review of psychological autopsy studies (Cavanagh et al., 2003).
Depressive syndromes and alcohol abuse/dependence were the two most
prevalent disorders in a psychological autopsy of all suicides in a calendar
year in Finland (Lonnqvist et al., 1995). Among Axis II disorders, borderline
personality disorders have shown excessive mortality by suicide in several
studies (Black et al., 2004). Among depressed patients, panic attacks,
severe psychic anxiety, diminished concentration, global insomnia,
moderate alcohol abuse and anhedonia were associated with suicide within
one year, and three other symptoms: severe hopelessness, suicidal ideation
and history of previous suicide attempts were associated with suicide oc-
curring after one year (Fawcett et al., 1990). Imperative hallucinations,
hopelessness, impulsivity and aggression are psychopathological pheno-
mena considered sufficient to explain an outcome of suicide directly,
without diagnoses (Maris et al., 2000). Five constructs have been consistently
associated with completed suicide: impulsivity/aggression, depression,
anxiety, hopelessness and self-consciousness/social disengagement
(Conner et al., 2001). Among severely suicidal predominantly young adult
males in the US army, hopelessness, depressive symptoms and suicidal
ideation was found to be a single syndrome (Shahar et al., 2006). This
finding may not be due to the gender differential of this sample as a similar
association was found in a cross-sectional study of university students who
were predominantly female (Chioqueta and Stiles, 2005).
A greater risk of suicidal acts was found in the combination of major
depression, PTSD and a Cluster B personality disorder than major de-
pression alone among outpatients (Oquendo et al., 2005). However, as
most psychiatric patients do not attempt suicide and as most suicide
attempters do not suicide, a psychiatric disorder and a previous history
of suicide attempt are both to be considered as necessary but insufficient
conditions for suicide (Mann et al., 1999). These authors found that
suicide attempters reported higher scores on subjective depression and
suicidal ideation, greater rates of lifetime aggression and impulsivity, higher
frequencies of comorbid borderline personality disorder, smoking, past
substance use disorder or alcoholism, family history of suicidal acts, head
injury and childhood abuse history with fewer reasons for living. On the

307
Lars Mehlum and Latha Nrugham

basis of these findings, the authors proposed a stress-diathesis model in


which the risk for suicidal acts is determined not merely by a psychiatric
illness (the stressor) but also by a diathesis.
Whether findings from clinical studies in non-military samples such
as these may directly apply to military populations is unclear. The stress-
diathesis model is, however, probably relevant to apply when trying to
understand suicidal behaviour in a high stress environment such as is the
case often in the military. For example, studies exploring the risk of suicide
among peacekeepers found that although peacekeeping per se does not
increase overall suicide risk, military lifestyles may strain interpersonal
relationships, encourage alcohol abuse and contribute to psychiatric
illness and suicide in a minority of vulnerable individuals irrespective of
peacekeeping assignment (Thoresen and Mehlum, 2004; Thoresen et al.,
2003; Wong et al., 2001).

Easy Access to Firearms

The best documented environmental risk factor of suicide in military


settings is easy access to firearms (Desjeux et al., 2004; Helmkamp, 1995,
1996; Hytten and Weisaeth, 1989; Mahon et al., 2005; Marttunen et al., 1997;
Scoville et al., 2007; Thoresen and Mehlum, 2006; Thoresen et al., 2003).
Among other occupations with easy access to lethal means are doctors
and nurses, who die of overdose (Agerbo et al., 2007). When the range
of alternatives is limited, the ‘ease of access’ (attainability) and the ‘readiness
for use’ (availability) probably are most likely to define the choice of a
particular method for suicide, for example jumping from tall buildings
in New York, insecticides in Sri Lanka, drowning in Norway, firearms
in the United States and Norway (Maris et al., 2000; Thoresen and
Mehlum, 2006). Firearms have been documented to be the preferred means
of suicide across studies of suicide among military personnel (Helmkamp,
1995; Scoville et al., 2007; Thoresen et al., 2003).
The preference for firearms is in keeping with the choice of male suicides
for irreversible means of death, involving negligible or nil probability of
rescue or survival. Such use of firearms also indicates the negative im-
plications of imparting knowledge of and training in the use of lethal
means. The overwhelming majority of suicidal gunshot wounds result

308
The Suicidal Soldier

from single shots, most of which were aimed at the head, with multiple
wounds in roughly 1 percent of suicides in the general population (Maris
et al., 2000). Among regular duty Irish military personnel, suicides that
took place on duty occurred predominantly when personnel were alone
shortly after duty commencement in the morning (Mahon et al., 2005).

Females in the Military

It was when the impact of combat trauma on women was explored that
sexual harassment, assault and trauma among both genders were revealed
(Fontana et al., 1997, 2000; Himmelfarb et al., 2006; Kang et al., 2005; Suris
and Lind, 2008; Yaeger et al., 2006). Most of these studies simultaneously
reported on the positive relationship between PTS symptoms/disorder and
unwanted sexual harassment/contact, which was found to be the same for
both genders. However, none of these studies studied the link between
this specific vulnerability and suicidal phenomena.

PROTECTIVE FACTORS

Leadership

During the implementation of the suicide prevention strategy for the


Norwegian Armed Forces, the following statement was made by the first
author of this chapter when he took on the task of selling the strategy
into the different levels of leadership throughout the organisation: ‘Good
military leadership is good primary suicide prevention in the military.’ The
suicide prevention programme in the Norwegian Armed Forces represents
a comprehensive strategy involving a range of initiatives and measures, and
may be regarded as the military sector’s contribution to the Norwegian
National Strategy for Suicide Prevention launched in 1994 (Norwegian
Board of Health, 1995).
The impact of leadership on suicide prevention has been documented
earlier (Mehlum and Schwebs, 2001). In a military unit, there is a danger
that a suicide may lead to erosion of essential qualities such as group

309
Lars Mehlum and Latha Nrugham

cohesion, motivation and confidence in the leadership. Hence, the attitudes


of the leaders to suicide prevention should not only be guided by the
general and self-evident ethical obligation to prevent tragic deaths, but also
by the need to protect the unit and counteract destructive group processes.
This is one of the reasons why the Norwegian Armed Forces regard their
leaders as a natural backbone of their suicide prevention programme. The
competence of the leaders in handling personal relations, building groups
out of individuals, their multi-faceted roles versus their men, and the
sheer number of leaders and their widespread distribution throughout the
organisation—these are factors that we should not forget when planning
for prevention of suicide in an organisation such as the military.
Leaders have great influence over the daily lives of their men. They
set standards and values, give orders, control outcomes and provide in-
formation, support and care. They certainly have a lot more possibilities
than medical doctors to counteract the negative tendencies in individuals
or groups of military personnel. Good military leadership will inherently
represent good primary suicide prevention. In addition to this, it is es-
sential that leaders possess basic knowledge about the high-risk signs of
suicide. Their task is, however, not to give any form of treatment; this is
the responsibility of medical personnel to whom persons at risk should
be immediately referred.
The strong demands put on the individual to adapt to the organisation
and the group will usually give rise to growth and maturation in the
young man and woman. But in some this is a highly stressful situation
causing adjustment reactions or formation of psychiatric symptoms in
some way or the other. The process of group formation may sometimes
result in destructive group processes. There is always the possibility of
‘scape-goating’ or alienation or bullying of individuals characterised by a
minority language, colour, religious belief or sexual orientation. Leaders
have a responsibility of counteracting such negative consequences of
socialisation and group formation. Fortunately, there are some potentially
protective resources in the military that should be utilised for preventive
purposes. First, we should mention the clear and visible signs of authority
and rules. Second, the clear expectation that everybody must take part
in essential activities and perform their best and that leaders will control
the quality. Third, the systematic emphasis put on the small group as the
forum for acquiring social and military skills and fostering group cohesion.

310
The Suicidal Soldier

These are factors that may protect against alienation, symptom formation
and suicidal behaviour.
The armed forces deal with and educate large groups of young people
under crucial years of their development. Most leaders see it as an obvious
responsibility for the military organisation to fulfil its mission in a way
that enhances growth and maturation in the individual and reduces the
risk of developing emotional problems or personal crises. Hence, the sui-
cide prevention strategy has been very well received throughout the or-
ganisation. One success factor has probably been to anchor the suicide
prevention strategy in the line of command and to involve a wide range of
different professional groups. Medical personnel played an important role
giving professional advice, teaching, evaluating and conducting research
and clinical work. But the responsibility of suicide prevention has remained
with commanders where it rightfully belongs.
In Norway, the Chief of Personnel at the National Military Headquarters
took the initiative to form and lead a Council for Suicide Prevention with
the mandate to survey the different elements of the suicide preventive
strategy, see to it that they are really implemented and to ensure that the
different parts of the organisation collaborated as good as possible in order
to reach the common goals. This council has served as a very important
forum for collaboration, to discuss and to decide upon new initiative,
to evaluate results, to counteract inefficient use of resources and to keep
suicide prevention on the agenda.
An increasing number of nations are currently establishing national
strategies for suicide prevention according to recommendations from the
World Health Organization. The armed forces of every nation should take
part in these collaborative efforts to face one of the most serious challenges
to the public health of the world today.

Training OpportuniƟes

Experiences from the implementation of suicide prevention in the


Norwegian Armed Forces also reveal that the training period is a time-
window of opportunities. This has also been documented before (Mehlum
and Schwebs, 2001). Studies within the Norwegian Armed Forces have
shown that the majority of soldiers who attempt suicide display more or

311
Lars Mehlum and Latha Nrugham

less specific prodromal signs and symptoms of suicide (Mehlum, 1990,


1992; see Rudd et al., 2006 and Van Orden et al., 2006 for a list of warning
signs and myths). However, these signals have often been inadequately
perceived and misunderstood by peers, leaders and medical personnel.
It is reasonable to believe that these problems may be due to a lack of
knowledge in the military personnel about the warning signs of suicide. In
order to increase both, the awareness in military personnel about suicide
and the knowledge about suicide prevention in health care personnel and
other specialised groups within the military, a wide range of education
measures were undertaken. Such an educational approach actually implies
a tremendous task in a large and complex organisation like the military.
The first group to be targeted was the privates for whom a specific two-
hour information package was developed. This package, consisting of a
video, a lecture and an outline for a group discussion, focuses on basic
knowledge about the high risk signs of suicide and where to get help
during an emotional crisis. The package is administered during the first
couple of weeks in the conscript’s basic training period. Suicide and suicide
prevention are now included as topics in leadership lectures at officers’
schools, military and staff academies. Finally, medical personnel and other
specialised personnel groups such as chaplains and welfare personnel,
being key groups in suicide preventive work, have been updated through
their specific continued education courses.

Clarity of Roles and Tasks

Due to the highly organised and detailed nature of the responses required
from the military, be they combat operations, peacekeeping operations or
disaster responses; clarity of roles and tasks of each person is of paramount
importance to keep the chain working and the links functioning. In
order that one’s fellow solider, subordinate and superior can function
as expected, clear expectations and performances are required at each
level. This clarity removes much of the ambiguity that could jeopardise
the organisation and also provides each person with a role and function,
instilling in them a sense of purpose by being specifically useful to the
team and the larger organisation, which is again placed within the larger
society.

312
The Suicidal Soldier

Common Goals and MoƟvaƟon

As an individual is not expected to function alone or in isolation but as a


team member, he necessarily shares the motivation of the team and its goal.
This is crucial to the team’s functioning and well-being. In life-threatening
missions/tasks, this sharing is life-saving and its absence can lead to
negligence or at the worst, to him becoming a traitor and jeopardising
not only the success of the mission/task, but also the lives of his fellow
soldiers, subordinates and superiors, if not more. The trust required in
being part of such a team ensures that the team works together to bring
the sharing of motivation and goals to the appropriate level. This pulling
together of the team can function as a protective factor for a suicidal solider
in ways similar to ways in which parents of young children refrain from
carrying out a suicidal plan, stopped by their own sense of responsibility
towards the care of their children or unable to bear the thought of letting
their children down.

Common Value Systems

A member of team having a clear understanding of his own role and


function, performing in alignment with the team’s goal and motivation
and sharing the team’s value systems is the ideal team member desired
by each one of us, regardless of the setting, whether civilian or military.
The degree to which the team is able to balance individual identity with
the team identity indicates its capacity to perform and deliver results as
expected. This would also mean stepping in with support for each other
as and when required, which often leads to strong emotional bonds and
interpersonal relationships, as the team is not allowed to be vulnerable
despite individual vulnerabilities. Such opportunities present themselves
often in military settings and are utilised as such.

Group Cohesion and a Sense of Coherence

Such a group as described here would logically have a high level of cohesion.
As the cohesion is task based and limited to time, place and situation,

313
Lars Mehlum and Latha Nrugham

it can also be very specific in nature, allowing a high level of flexibility


once its role and function are over, with the possibility of coming together
again as per requirement. This flexibility and openness removes from
the unit the suffocation which could otherwise be reasonably expected.
It has been found that a well-functioning social network is a protective
factor against suicidal phenomena in late adolescence (Ystgaard, 1997;
Ystgaard et al., 1999). This purposeful connectedness to the team leads
to a sense of coherence, at the individual and unit level, which is reflected
back to the organisational level. This sense of coherence in the individual
which implies a tendency for the individual to regard life as having some
purpose, to believe that difficulties can be overcome and that the world is
understandable was found to be associated with a lower level of current
suicidal ideation (Mehlum, 1998).

Social Support and Welfare Systems

Mehlum and Schwebs (2001) report on how the welfare service of the
armed forces can give valuable contributions in reducing some of the in-
herent stress of the military lifestyle. Individually tailored services can be
offered in order to solve social or financial problems. Even more import-
ant are group-oriented measures for improvement of the social milieu;
sports, culture, education or hobby activities. For basic trainees, the welfare
services represent a very important resource easing their social integration
and reducing the stress of adaptation.
Mental and suicidal crises may arise at times, in places or in situations
where professional help or support from comrades, colleagues or leaders
is unavailable. Furthermore, even if they need it desperately, many regular
employees will feel it problematic to take the first step to seek help from
medical personnel who belong to the same military system as themselves
in fear of losing prestige or career opportunities. In order to lower the
threshold for help-seeking and to reach as many persons as possible
regardless of time of day or place, the Norwegian Armed Forces established
a crisis telephone service in 1994. This is called the ‘Green Line’ and is
open 24 hours a day. Its use is free of charge, even if long-distance and it
is possible to call anonymously. An evaluation performed after the first
three years of operation showed that this service has become well known

314
The Suicidal Soldier

in the essential target groups and that they experienced it as a valuable


service that they would want to use if ever in need of it. It is estimated that
each year about 2–3 percent of all Norwegian military conscripts use this
telephone service at least once.
Although interpersonal support was not directly assessed in a study of
the catharsis effect of a suicidal crisis among predominantly young males
attending a US army medical centre, decreased suicidality was found, but
due to the increased severity of symptoms of the attempters, single and
multiple as compared to ideators, the authors chose to interpret the de-
creased suicidality as the gradual action of interpersonal support rather
than as emotional catharsis (Walker et al., 2001).
Perception of increased social support was a predictor of lower
levels of suicidal ideation, independent of the severity of depression
and hopelessness, minimised by the level of life satisfaction and level of
exhibited self-esteem in university students (Chioqueta and Stiles, 2007).

Medical Support

The experiences of the Norwegian Armed Forces have been documented


on this factor too (Mehlum and Schwebs, 2001). Although the Norwegian
Armed Forces have found it necessary to drastically expand the scope
of suicide prevention beyond the medical service and to extend the re-
sponsibility for the implementation of its various aspects to wider groups
of personnel, medical personnel remain an indispensable resource. A
crucial function is to diagnose and treat persons in acute suicidal crises.
This is done both by the general military physicians and by personnel
working in the psychiatric services in the armed forces. In some cases, the
suicidal crisis is evoked by some stressor or problem that can be rapidly
alleviated and the soldier can continue his service. Some soldiers have more
severe mental problems or personality disorders that make it necessary
to separate them from further service. Frequently soldiers find it difficult
to seek medical personnel for emotional, social or existential problems. To
seek the military chaplain is an alternative to many of these and many are
those soldiers whose suicidal crises have been uncovered by the chaplain.
In most cases, these soldiers will be referred to medical personnel.

315
Lars Mehlum and Latha Nrugham

ASSESSMENT AND MANAGEMENT

As specific tools to assess the performance of an organisation or a unit


with reference to its suicide prevention do not exist; routine check-ups
of the organisation’s member individual’s proneness towards suicidal
phenomena can work as a proxy indicators of the way in which the or-
ganisation is able to prevent and manage suicidal behaviour. However,
it is important to mention here that a systematic review of suicide
prevention strategies around the world reported on the study of five pre-
ventive strategies: education and awareness for the general public and for
professionals; screening tools for at-risk individuals; treatment of psy-
chiatric disorders; restricting access to lethal means; and responsible media
reporting of suicide (Mann et al., 2005). The authors concluded that
physician education in depression recognition and treatment, and restrict-
ing access to lethal means where the method was common were found to
prevent suicide and thus reduce suicide rates while the other interventions
needed more evidence of efficacy. They also mention, as part of means
restriction, that restrictions on access to alcohol coincided with decreases in
overall suicide rates in the former Union of Soviet Socialist Republics and
Iceland. A specific recommendation made to reduce suicide attempts by
firearms in the military was to implement stringent ammunition control
procedures and placing weapons in racks immediately after coming off the
line (Scoville et al., 2004). At the individual level, several questionnaires
and clinical interviews with excellent psychometric properties to assess
psycho-social and psychiatric dimensions are available, as seen in the
studies referred to earlier.
Management of suicidal behaviour is a difficult clinical task and only a
handful of empirically-tested clinical approaches with theoretical bases are
available (Brown et al., 2005; Cipriani et al., 2005; Jobes, 2006; Jobes et al.,
2005; Linehan, 1993; Rudd et al., 2001). Of particular interest to the armed
forces is the Collaborative Assessment and Management of Suicidality
(CAMS) tested by Jobes and associates with outpatients from the US airforce.
The CAMS assessment uses both qualitative and quantitative responses,
assessing psychological pain, stress, agitation, hopelessness, self-hate and
overall suicide risk, with a heavy emphasis on the clinician’s collaboration

316
The Suicidal Soldier

with the patient. In CAMS, the patient’s perspective is treated as the assess-
ment gold standard, modification of clinician behaviours is also an
objective and suicidality itself is the primary focus of care instead of
treating suicidality as a symptom of a psychiatric disorder. It is also
pertinent to note here a study on discrepancies between clinician and
self-ratings of suicidal symptoms in a predominantly young male sample
from attendees of a major US army medical centre, which found that
clinicians were more likely to assess the patient as high in suicidality than
the patients themselves approximately in 50 percent of the assessments
(Joiner et al., 1999). This finding need not necessarily be due to the gender
or age skewness of this sample as a similar discrepancy between experts
and clinicians was revealed in a later study of case vignettes and authors
of both the studies concluded that this discrepancy had much to do with
the better-safe-than-sorry attitude of the clinicians which is protective of
themselves and the patients (Wagner et al., 2002).

SUMMARY

The prevention of suicidal phenomena in the military presents unique chal-


lenges and opportunities. The organised nature of the military provides
an opportunity for preventive intervention and post-event manage-
ment, and evaluation while the knowledge and training imparted along
with the unsupervised access given to lethal means makes it into an imple-
mentation challenge. Similarly, the predominance of the male gender and
the stamp of the collective on the individual, the fellowship, the group
cohesion, the preparation for combat and disaster responses make suicide
prevention here simultaneously a challenge and an opportunity. The
abstract disconnection between the collective and the individual may lead
to a sense of entrapment if the individual is unable to move out of the
collective which requires such an adherence from him. We see that some
of those very factors which were discussed earlier in the section on risk
factors were now discussed as protective factors. This is an indicator of
the complexity of the processes involved in the making and unmaking of
a suicidal soldier.

317
Lars Mehlum and Latha Nrugham

REFERENCES

Agerbo, E., D. Gunnell, J.P. Bonde, P.B. Mortensen and M. Nordentoft (2007). Suicide
and occupation: The impact of socio-economic, demographic and psychiatric
differences. Psychological Medicine, 37(8), 1131–40.
Antonovsky, A. (1993). The structure and properties of the sense of coherence scale.
Social Science and Medicine, 36(6), 725–33.
Black, D.W., N. Blum, B. Pfohl and N. Hale (2004). Suicidal behaviour in borderline
personality disorder: Prevalence, risk factors, prediction, and prevention. Journal
of Personality Disorders, 18(3), 226–39.
Blakely, T.A., S.C. Collings and J. Atkinson (2003). Unemployment and suicide.
Evidence for a causal association? Journal of Epidemiology and Community Health,
57(8), 594–600.
Bolger, E.A. (1999). Grounded theory analysis of emotional pain. Psychotherapy
Research, 9(3), 342–62.
Brown, G.K., T. Ten Have, G.R. Henriques, S.X. Xie, J.E. Hollander and A.T. Beck
(2005). Cognitive therapy for the prevention of suicide attempts: A randomized
controlled trial. Journal of the American Medical Association, 294(5), 563–70.
Cavanagh, J.T., A.J. Carson, M. Sharpe and S.M. Lawrie (2003). Psychological autopsy
studies of suicide: A systematic review. Psychological Medicine, 33(3), 395–405.
Erratum in: Psychological Medicine, 33(5), 947.
Chioqueta, A.P. and T.C. Stiles (2005). Personality traits and the development
of depression, hopelessness and suicide ideation. Personality and Individual
Differences, 38(6), 1283–91.
Chioqueta, A.P. and T.C. Stiles (2007). The relationship between psychological buffers,
hopelessness, and suicidal ideation: Identification of protective factors. Crisis,
28(2), 67–73.
Cipriani, A., H. Pretty, K. Hawton and J.R. Geddes (2005). Lithium in the prevention
of suicidal behavior and all-cause mortality in patients with mood disorders: A
systematic review of randomized trials. American Journal of Psychiatry, 162(10),
1805–19.
Conner, K.R., P.R. Duberstein, Y. Conwell, L. Seidlitz and E.D. Caine (2001). Psy-
chological vulnerability to completed suicide: A review of empirical studies. Suicide
and Life-Threatening Behaviour, 31(4), 367–85.
Desjeux, G., J. Labarère, L. Galoisy-Guibal and R. Ecochard (2004). Suicide in the French
armed forces. European Journal of Epidemiology, 19(9), 823–29.
Durkheim, E. (1897). Le suicide. Paris: Alcan.
Fawcett, J., W.A. Scheftner, L. Fogg and D.C. Clark (1990). Time-related predictors
of suicide in major affective disorder. American Journal of Psychiatry, 147(9),
1189–94.
Fontana, A., B. Litz and R. Rosenheck (2000). Impact of combat and sexual harassment on
the severity of posttraumatic stress disorder among men and women peacekeepers
in Somalia. Journal of Nervous and Mental Disorder, 188(3), 163–69.

318
The Suicidal Soldier

Fontana, A., L.S. Schwartz and R. Rosenheck (1997). Posttraumatic stress disorder
among female Vietnam veterans: a causal model of aetiology. American Journal of
Public Health, 87(2), 169–75.
Frankl, V.E. (1967). Psychotherapy and Existentialism: Selected Papers on Logotherapy.
New York: Simon and Schuster.
Giotakos, O. (2003). Suicidal ideation, substance use and sense of coherence in Greek
male conscripts. Military Medicine, 168(6), 447–50.
Harlow, L.L., M.D. Newcomb and P.M. Bentler (1986). Depression, self-derogation,
substance use and suicide ideation: Lack of purpose in life as a mediational factor.
Journal of Clinical Psychology, 42(1), 5–21.
Haw, C. and K. Hawton (2008). Life problems and deliberate self-harm: Associations
with gender, age, suicidal intent and psychiatric and personality disorder. Journal
of Affective Disorders, 109(1–2), 139–48.
Hawton, K., E. Townsend, E. Arensman, D. Gunnell, P. Hazell, A. House et al. (2000).
Psychosocial versus pharmacological treatments for deliberate self harm. Cochrane
Database of Systematic Reviews, 2, CD001764.
Helmkamp, J.C. (1995). Suicides in the military: 1980–1992. Military Medicine,
160(2), 45–50.
Helmkamp, J.C. (1996). Occupation and suicide among males in the US Armed Forces.
Annals of Epidemiology, 6(1), 83–88.
Helmkamp, J.C. and R.D. Kennedy (1996). Causes of death among U.S. military
personnel: A 14-year summary, 1980–1993. Military Medicine, 161(6), 311–17.
Himmelfarb, N., D. Yaeger and J. Mintz (2006). Posttraumatic stress disorder in female
veterans with military and civilian sexual trauma. Journal of Traumatic Stress,
19(6), 837–46.
Horn, O., L. Hull, M. Jones, D. Murphy, T. Browne, N.T. Fear et al. (2006). Is there an
Iraq war syndrome? Comparison of the health of UK service personnel after the
Gulf and Iraq wars. The Lancet, 367(9524), 1742–46.
Hytten, K. and L. Weisaeth (1989). Suicide among soldiers and young men in the Nordic
countries 1977–1984. Acta Psychiatrica Scandinavia, 79(3), 224–28.
Jobes, D.A. (2006). Managing suicidal risk: A collaborative approach. New York: The
Guilford Press.
Jobes, D.A., A. Wong, S. A.K. Conrad, J.F. Drozd and T. Neal-Walden (2005). The
collaborative assessment and management of suicidality versus treatment as
usual: a retrospective study with suicidal outpatients. Suicide and Life-Threatening
Behavior, 35(5), 483–97.
Joiner, T.E. Jr. and M.D. Rudd (2000). Intensity and duration of suicidal crises vary as a
function of previous suicide attempts and negative life events. Journal of Consulting
and Clinical Psychology, 68(5), 909–16.
Joiner, T.E. Jr., D.M. Rudd and M.H. Rajab (1999). Agreement between self- and
clinician-rated suicidal symptoms in a clinical sample of young adults: explaining
discrepancies. Journal of Consulting and Clinical Psychology, 67(2), 171–76.

319
Lars Mehlum and Latha Nrugham

Kang, H., N. Dalager, C. Mahan and E. Ishii (2005). The role of sexual assault on the
risk of PTSD among Gulf War veterans. Annals of Epidemiology, 15(3), 191–95.
Karasek, R. and T. Theorell (1990). Healthy work: Stress, productivity, and the
reconstruction of working life. New York: Basic Books.
Kposowa, A.J. (2001). Unemployment and suicide: A cohort analysis of social factors
predicting suicide in the US National Longitudinal Mortality Study. Psychological
Medicine, 31(1), 127–38.
Lester, D. and S. Badro (1992). Depression, suicidal preoccupation and purpose of life
in a subclinical population. Personality and Individual Differences, 13(1), 75–76.
Linehan, M.M. (1993). Skills training manual for treating borderline personality disorder.
New York: The Guilford Press.
Lonnqvist, J.K., M.M. Henriksson, E.T. Isometsa, M.J. Marttunen, M.E. Heikkinen,
H.M. Aro et al. (1995). Mental disorders and suicide prevention. Psychiatry and
Clinical Neurosciences, 49(Supplement 1), S111–16.
Mahon, M.J., J.P. Tobin, D.A. Cusack, C. Kelleher and K.M. Malone (2005). Suicide
among regular-duty military personnel: A retrospective case-control study of
occupation-specific risk factors for workplace suicide. American Journal of
Psychiatry, 162(9), 1688–96.
Mann, J.J., C. Waternaux, G.L. Haas and K.M. Malone (1999). Towards a clinical
model of suicidal behaviour in psychiatric patients. American Journal of Psychiatry,
156(2), 181–89.
Mann, J.J., A. Apter, J. Bertolote, A. Beautrais, D. Currier, A. Haas et al. (2005).
Suicide prevention strategies: A systematic review. Journal of the American Medical
Association, 294(16), 2064–74.
Maris, R.W., A.L. Berman and A.M. Silverman (2000). Comprehensive textbook of
suicidology. New York: The Guilford Press.
Marttunen, M., M. Henriksson, S. Pelkonen, M. Schroderus and J. Lonnqvist (1997).
Suicide among military conscripts in Finland: A psychological autopsy study.
Military Medicine, 162(1), 14–18.
Mehlum, L. (1990). Attempted suicide in the armed forces: A retrospective study of
Norwegian conscripts. Military Medicine, 155(12), 596–600.
Mehlum, L. (1992). Prodromal signs and precipitating factors in attempted suicide.
Military Medicine, 157(11), 574–77.
Mehlum, L. (1994). Young male suicide attempters 20 years later: The suicide mortality
rate. Military Medicine, 159(2), 138–41.
Mehlum, L. (1998). Suicidal ideation and sense of coherence in male conscripts. Acta
Psychiatrica Scandinavia, 98(6), 487–92.
Mehlum, L., K. Hytten and F. Gjertsen (1999). Epidemiological trends of youth suicide
in Norway. Archives of Suicide Research, 5(3), 193–205.
Mehlum, L., B.O. Koldsland and M.E. Loeb (2006). Risk factors for long-term
posttraumatic stress reactions in unarmed UN military observers: A four-year
follow-up study. Journal of Nervous and Mental Disorders, 194(10), 800–04.

320
The Suicidal Soldier

Mehlum, L. and R. Schwebs (2001). Suicide prevention in the military: Recent


experiences from the Norwegian armed forces. International Review of the Armed
Forces Medical Services, 74(2), 71–74.
Mehlum, L. and L. Weisaeth (2002). Predictors of posttraumatic stress reactions in
Norwegian U.N. peacekeepers 7 years after service. Journal of Traumatic Stress,
15(1), 17–26.
Nelson, R. (2004). Suicide rates rise among soldiers in Iraq. The Lancet, 363 (9405),
300.
Norwegian Board of Health (1995). The National Plan for Suicide Prevention:
1994–1998. Oslo: Norwegian Board of Health.
Orbach, I., M. Mikulincer, E. Gilboa-Schechtman and P. Sirota (2003). Mental pain
and its relationship to suicidality and life meaning. Suicide and Life-Threatening
Behavior, 33(3), 231–41.
Orbach, I., M. Mikulincer, P. Sirota and E. Gilboa-Schechtman (2003). Mental pain: A
multidimensional operationalization and definition. Suicide and Life-Threatening
Behavior, 33(3), 219–30.
Oquendo, M., D.A. Brent, B. Birmaher, L. Greenhill, D. Kolko, B. Stanley et al. (2005).
Posttraumatic stress disorder comorbid with major depression: factors mediating
the association with suicidal behavior. American Journal of Psychiatry, 162(3),
560–66.
Petrie, K. and R. Brook (1992). Sense of coherence, self-esteem, depression and
hopelessness as correlates of reattempting suicide. British Journal of Clinical
Psychology, 31(part 3), 293–300.
Powell, K.E., L.A. Fingerhut, C.M. Branche and D.M. Perrotta (2000). Deaths due to
injury in the military. American Journal of Preventive Medicine, 18(Supplement 3),
26–32.
Price, R.K., N.K. Risk, A.H. Haden, C.E. Lewis and E.L. Spitznagel (2004). Post-
traumatic stress disorder, drug dependence, and suicidality among male Vietnam
veterans with a history of heavy drug use. Drug and Alcohol Dependence, 76
(Supplement), S31–43.
Prigerson, H.G., P.K. Maciejewski and R.A. Rosenheck (2001). Combat trauma: Trauma
with highest risk of delayed onset and unresolved posttraumatic stress disorder
symptoms, unemployment, and abuse among men. Journal of Nervous and Mental
Disorder, 189(2), 99–108.
Prigerson, H.G., P.K. Maciejewski and R.A. Rosenheck (2002). Population attributable
fractions of psychiatric disorders and behavioral outcomes associated with combat
exposure among US men. American Journal of Public Health, 92(1), 59–63.
Rossow, I. and A. Amundsen (1997). Alcohol abuse and mortality: A 40-year prospective
study of Norwegian conscripts. Social and Scientific Medicine, 44(2), 261–67.
Rudd, M.D., A.L. Berman, T.E. Joiner, Jr. and M.H. Rajab (2001). Treating suicidal
behaviour: An effective, time-limited approach. In D. Barlow (Series Ed.),
Treatment manuals for practitioners. New York: The Guilford Press.

321
Lars Mehlum and Latha Nrugham

Rudd, M.D., A.L. Berman, T.E. Joiner, Jr., M.K. Nock, M.M. Silverman, M. Mandrusiak
et al. (2006). Warning signs for suicide: theory, research, and clinical applications.
Suicide and Life-Threatening Behavior, 36(3), 255–62.
Sareen, J., S.L. Belik, T.O. Afifi, G.J. Asmundson, B.J. Cox and M.B. Stein (2008).
Canadian military personnel’s population attributable fractions of mental disorders
and mental health service use associated with combat and peacekeeping operations.
American Journal of Public Health, 98(12), 2191–98.
Sareen, J., B.J. Cox, T.O. Afifi, M.B. Stein, S.L. Belik, G. Meadows et al. (2007). Combat
and peacekeeping operations in relation to prevalence of mental disorders and
perceived need for mental health care: Findings from a large representative sample
of military personnel. Archives of General Psychiatry, 64(7), 843–52.
Scoville, S.L., J.W. Gardner and R.N. Potter (2004). Traumatic deaths during U.S.
Armed Forces basic training, 1977-2001. American Journal of Preventive Medicine,
26(3), 194–204.
Scoville, S.L., M.E. Gubata, R.N. Potter, M.J. White and L.A. Pearse (2007). Deaths
attributed to suicide among enlisted U.S. Armed Forces recruits, 1980–2004.
Military Medicine, 172(10), 1024–31.
Sentell, J.W., M. Lacroix, J.V. Sentell and K. Finstuen (1997). Predictive patterns of
suicidal behavior: The United States armed services versus the civilian population.
Military Medicine, 162(3), 168–71.
Shahar, G., L. Bareket, M.D. Rudd and T.E. Joiner, Jr. (2006). In severely suicidal young
adults, hopelessness, depressive symptoms, and suicidal ideation constitute a single
syndrome. Psychological Medicine, 36(7), 913–22.
Shneidman, E.S. (1993). Commentary: Suicide as psychache. Journal of Nervous and
Mental Disorders, 181(3), 145–47.
Southwick, S.M., C.A. Morgan, 3rd, A. Darnell, D. Bremner, A.L. Nicolaou, L.M. Nagy
et al. (1995). Trauma-related symptoms in veterans of operation desert storm: A
2-year follow-up. American Journal of Psychiatry, 152(8), 1150–55.
Suris, A. and L. Lind (2008). Military sexual trauma: a review of prevalence and
associated health consequences in veterans. Trauma and Violence Abuse, 9(4),
250–269.
Thoresen, S. and L. Mehlum (2004). Risk factors for fatal accidents and suicides in
peacekeepers: is there an overlap? Military Medicine, 169(12), 988–93.
Thoresen, S. and L. Mehlum (2006). Suicide in peacekeepers: risk factors for suicide
versus accidental death. Suicide and Life-Threatening Behavior, 36(4), 432–42.
Thoresen, S. and L. Mehlum (2008). Traumatic stress and suicidal ideation in Norwegian
male peacekeepers. Journal of Nervous and Mental Disorders, 196(11), 814–21.
Thoresen, S., L. Mehlum and B. Moller (2003). Suicide in peacekeepers—A cohort
study of mortality from suicide in 22,275 Norwegian veterans from international
peacekeeping operations. Social Psychiatry and Psychiatric Epidemiology, 38(11),
605–10.
Thoresen, S., L. Mehlum, E. Røysamb and A. Tønnessen (2006). Risk factors for
completed suicide in veterans of peacekeeping: Repatriation, negative life events,
and marital status. Archives of Suicide Research, 10(4), 353–63.

322
The Suicidal Soldier

Tsutsumi, A., K. Kayaba, T. Ojima, S. Ishikawa, N. Kawakami and Jichi Medical


School Cohort Study Group (2007). Low control at work and the risk of suicide
in Japanese men: a prospective cohort study. Psychotherapy and Psychosomatics,
76(3), 177–85.
Unwin, C., N. Blatchley, W. Coker, S. Ferry, M. Hotopf, L. Hull et al. (1999). Health of
UK servicemen who served in Persian Gulf War. The Lancet, 353(9148), 169–78.
Van Orden, K.A., T.E. Joiner, Jr., D. Hollar, M.D. Rudd, M. Mandrusiak and
M.M. Silverman (2006). A test of the effectiveness of a list of suicide warning signs
for the public. Suicide and Life-Threatening Behavior, 36(3), 272–87.
Wagner, B.M., S.A. Wong and D.A. Jobes (2002). Mental health professionals’
determinations of adolescent suicide attempts. Suicide and Life-Threatening
Behavior, 32(3), 284–300.
Walker, R.L., T.E. Joiner, Jr. and M.D. Rudd (2001). The course of post-crisis suicidal
symptoms: How and for whom is suicide “cathartic”? Suicide and Life-Threatening
Behavior, 31(2), 144–52.
Wasserman, I.M. (1992). Economy, work, occupation and suicide. In R.W. Maris,
A.L. Berman, J.T. Maltsenberger and R.I. Yufit (Eds), Assessment and prediction
of suicide (pp. 520–40). New York: The Guildford Press.
Wong, A., M. Escobar, A. Lesage, M. Loyer, C. Vanier and I. Sakinofsky (2001). Are
UN peacekeepers at risk for suicide? Suicide and Life-Threatening Behavior,
31(1), 103–12.
Yaeger, D., N. Himmelfarb, A. Cammack and J. Mintz (2006). DSM-IV diagnosed
posttraumatic stress disorder in women veterans with and without military sexual
trauma. Journal of General Internal Medicine, Supplement 3, S65–69.
Ystgaard, M. (1997). Life stress, social support and psychological distress in late
adolescence. Social Psychiatry and Psychiatric Epidemiology, 32(5), 277–83.
Ystgaard, M., K. Tambs and O.S. Dalgard (1999). Life stress, social support and
psychological distress in late adolescence: A longitudinal study. Social Psychiatry
and Psychiatric Epidemiology, 34(1), 12–19.

323
15

Suicidal IdeaƟon and Behaviour


among Asian Adolescents
AĦ›½ N¦ƒ-ÃƒÄ L›çĦ, Cƒã«ù Yç®-‘«® FÊĦ
ƒÄ— Cƒã«›Ù®Ä› A½›øƒÄ—Ùƒ M‘BÙ®—›-C«ƒÄ¦

S uicide is a major issue of concern among adolescents worldwide.


This chapter is an attempt to focus on trends related to suicide and
suicidal ideation among teenagers in Asia. Besides this overview, adequate
consideration will be given to different factors related to adolescent suicide
trends, encompassing psychological, environmental and socio-cultural
influences. Among these, Internet stands out as particularly important
for clinicians and researchers to consider in future studies. Although parts
of this chapter will highlight issues that are of particular concern for Asian
adolescents, some findings seem relatively universal across societies. Suicide
research on adolescents from across the globe will be reviewed here, but
with the goal of discussing its particular relevance to Asian cultures. The
chapter will conclude with review of some signs and proposed preventive
measures for fighting against adolescents’ suicide.

ADOLESCENT SUICIDE IN ASIA: PATTERNS AND TRENDS


Given the large population of Asia, suicidal deaths in Asia account for more
than 60 percent of the world’s total number of suicides (WHO Mortality

324
Adolescents’ Suicide in Asia

Database, 2008). Nevertheless, because most studies concerning suicide


have been carried out in Western cultures, mostly in the United States of
America, there are relatively less known facts about suicide in Asia (Beautrais,
2006). As the research concern about suicide deaths in Asia is increasing,
researchers have attempted specific profiles of suicide in Asian countries so
as to facilitate the development of suicide prevention programmes that can
fit the unique cultures and situations of various Asian nations. According to
recently updated statistics (Hendin et al., 2008), suicide is one of the leading
causes of death among adolescents but not among older people, although
the elderly are the age group with the highest suicide rate in most Asian
countries. In Hong Kong, Japan and the Republic of Korea, completed
suicide was the most common cause of death among young people
aged 15–24 years in 2006. For some other Asian countries such as
Thailand and Sri Lanka, young people actually had the highest suicide
rates across any age group, while China had high rates among both the
young and the old. Table 15.1 shows statistical data of suicide death rates
of adolescents (aged 15–24) in for some Asian countries from year 1990
to 2006.
The adolescent suicide rate in China stayed at a high rate of around
15 per 100,000 per year throughout the 1990s, but showed a decreasing
trend near the beginning of the 2000s, especially among young women
(as shown in Table 15.1). This downward tendency in this decade was
noted in Philips’s (2008) study on China’s changing rates and patterns in
suicide from 1987 to 2006. In contrast, rates in Japan rose overall from
seven in 1990 to 14.1 (per 100,000) in 2006, together with a rise in the
ranking of suicide as the leading cause of death among adolescents from
its previous ranking as third (WHO Mortality Database, 2008). Suicide
rates in Hong Kong, Korea, Thailand and Singapore have also undergone
fluctuations within the range of 6–12.2 per 100,000 in the past decade or
so (WHO Mortality Database, 2008). The relatively low suicide rate in
Taiwan has remained stable over time with small increases since the year
2000 (Department of Health of the Republic of China, 2008).
According to police records from India (National Crime Records
Bureau, 2006), the suicide rate for youths in 2006 was 10.5 (per 100,000)
and the rate for the total population in the country remained around the
same rate from 1996 to 2006. However, these government data may be
subject to considerable underestimation owing to inaccurate population
counts with registration systems varying in efficiency (especially in

325
Angel Nga-man Leung et al.

Table 15.1 Suicide Rates in Alternate Years among 15–24-year-old Adolescents per
100,000 of the PopulaƟon in Selected Asian Countries/Regions (1990–2006)

Year/Area China Hong Kong Japan Korea Singapore Thailand Taiwan


1990 M 10.7 7.4 9.2 10.7 13.1 12.9 5.0
F 22.2 6.9 4.7 5.5 9.2 9.0 3.5
T 16.2 7.1 7.0 8.1 10.8 11.0 4.3
1992 M 11.4 9.7 10.2 10.8 9.2 12.3 5.3
F 23.9 6.6 4.7 5.3 4.1 8.3 2.5
T 17.4 8.2 7.5 8.1 6.7 10.3 4.3
1994 M 10.0 9.5 12.0 11.0 10.8 7.8 6.1
F 20.1 8.7 5.1 5.9 8.8 4.3 2.6
T 14.8 9.1 8.6 8.5 9.8 6.1 4.4
1996 M 11.4 11.3 14.4 9.1 17.6 5.5
F 7.2 5.4 9.0 2.8 6.2 2.9
T 9.3 8.5 11.8 6.0 11.9 4.2
1998 M 5.4 13.9 16.8 15.1 10.2 18.1 5.1
F 8.6 7.5 7.4 9.4 7.2 5.6 3.3
T 6.9 10.7 12.2 12.3 8.7 11.9 4.2
2000 M 8.0 15.8 10.2 6.5 18.0 4.7
F 5.8 6.9 7.0 7.7 4.7 3.3
T 6.5 11.5 8.7 7.1 11.4 4.0
2002 M 13.2 14.7 9.7 6.9 13.8 7.7
F 8.8 6.3 7.1 9.5 3.8 3.6
T 12.6 10.7 8.5 8.2 8.9 5.9
2004 M 15.6 16.9 11.3 5.7 8.2
F 8.8 8.4 8.0 4.5 4.2
T 12.2 12.8 9.7 5.1 6.2
2006 M 9.2 11.6 18.2 9.7 6.2 9.2
F 4.5 5.2 9.7 8.8 1.3 4.5
T 6.9 9.3 8.8 9.3 3.8 6.9
Source: WHO Mortality Database (2008); Department of Health of the Republic
of China (2008).

rural areas). In addition, inaccuracy is likely because people are often un-
willing to bear the social and legal consequences associated with suicide
(Hendin et al., 2008). For example, the suicide rate in rural India reported
in a study using verbal autopsies (Joseph et al., 2003) was nine times that
of the official figure (Gururaj and Isaac, 2001).
Suicide trends and patterns for some other major countries in Asia with
insufficient official data (e.g., India) can perhaps be better understood by
referring to findings of local studies. Considering possibility of significant
differences in the standard of data collection and the verdict of suicide
death across countries, suicide rates should be interpreted with caution

326
Adolescents’ Suicide in Asia

in generalising these patterns in relation to regional differences and over


the time changes in suicidal behaviour (Cheng and Lee, 2000).
Among young people, suicide risk tends to increase with age in
adolescence (Pelkonen and Marttunen, 2003) across both male and female
populations (Shek et al., 2005). The major reason for the growing suicide
rates with age is the prevalence of psychopathology such as depression
and substance abuse in older adolescents (Granello and Granello, 2007).
Moreover, older adolescents are cognitively more able to make plans and
execute a lethal suicide attempt than are younger adolescents (Groholt
et al., 1998).
Although some previous studies have reported a higher number of
males than females committing suicide overall (e.g., Cheng and Lee, 2000),
the male-to-female ratio of suicides in Asian adolescents tends to be higher
than those of Western countries. Compared with the international mean
of 3.6 (males):1 (female) (Kelleher and Chambers, 2003), the average
gender ratios for Japan, Hong Kong, the Republic of Korea and Taiwan
were quite closer across the genders, that is within the range of 1.3–2.1 to 1
(see WHO Mortality Database [2008] and Table 15.1), from the year 2000
onwards. Indeed, the gender ratio was reverse in China. Chinese young
women were at higher risk for suicide than were Chinese young men.
This exceptional pattern has been attributed to the high rate of suicide
deaths for young females in the rural regions, which may be the result
of lower social status and marital problems for Chinese women (Zhang
et al., 2002). Yip’s (2001) studies reported a male/female suicide ratio of
0.8:1 and 0.5:1 in urban and rural areas, respectively, in the age group of
15- to 24-year-olds. However, newer studies (Yip and Liu, 2006; Yip et al.,
2005) have demonstrated a significant decrease in the rate of suicide
among young women in rural areas and a consequent increase in the
gender ratio for suicides in China. Authors of these studies further pointed
out that Chinese male-to-female ratios of suicide may come to resemble
those ratios of Western/developed countries, because of the expected
rising level of urbanisation in China and associated demographic changes
(e.g., the decline of the young female population in rural areas) and
cultural changes (e.g., in family and employment systems) (Yip and Liu,
2006; Yip et al., 2005).
Hendin and colleagues (2008) reviewed the most common methods
of suicide across Asia. In Japan, Pakistan and Thailand, most people
killed themselves by hanging. In Hong Kong and Singapore, jumping

327
Angel Nga-man Leung et al.

(especially from apartment buildings) was the preferred method of


suicide. In countries with large rural populations, such as India, China
and the Republic of Korea, pesticides poisoning has been widely used, and
importantly, the prevalence of this suicide method is regarded as one
major reason for the particularly high female suicide rate in these coun-
tries. Compared to young females in Western countries, for whom the
most frequently used means of suicide tends to be self-poisoning with
medication, a method that often has relatively low lethality, Asian women
in rural areas have more access to agricultural pesticides, which have
much higher toxicity (Philips, 2002). A study of 10- to 19-year-olds in a
rural region in southern India (Aaron et al., 2004) reported an alarmingly
high suicide rate of 148 per 100,000 for women, and 58 per 100,000 for
men. The easy availability of pesticides in these rural homes is perhaps
largely to blame for these high rates of suicide deaths, especially for young
females.
Along with the conventional methods, other newer methods of suicide
have emerged in individual regions in Asia as well. For example, the recent
surge in suicide attempts by hydrogen sulphide intoxication in Japan
and the widespread circulation of the method of producing the gas from
household detergents via Internet messages have been reported in local
and international news articles (Lewis, 2008). According to the report
issued by the National Police Agency of Japan, within the first five months
of 2008, nearly 520 people in Japan killed themselves by using hydrogen
sulphide gas, while there were only 30 such cases in the same period in
2007 (Kubota, 2008). Another example of novel suicide methods in Asia
is charcoal burning in Hong Kong (Yip, 2006).

SUICIDE RISK FACTORS IN ASIAN ADOLESCENTS

Risk factors for suicidal behaviour vary across ages and cultures (Guetzloe,
1989). The following discussion focuses on common risk factors cited in
the literature, particularly as they concern suicide in Asian adolescents.
The risk factors suggested here are categorised into: (a) psychological,
(b) environmental and (c) socio-cultural risk factors.
In terms of psychological risk factors, most studies of attempted
and completed suicide in adolescents indicate a high prevalence of

328
Adolescents’ Suicide in Asia

mental health disorders (Nrugham et al., 2008). Population-based


psychological autopsy studies on suicide in Asia (including Taiwan,
China, India and Singapore) show that 68–97 percent had diagnosable
psychiatric disorders (Vijayakumar, 2005). Depression is often the most
frequently cited diagnosis associated with suicide (Waldrop et al., 2007).
Although individual psychological factors may have been somewhat
under-researched in Asia as compared to the West (Vijayakumar, 2005),
large-sample studies have found that depressive symptoms significantly
distinguish suicide attempters from non-attempters equally in Eastern and
Western cultures (e.g., Stewart et al., 2006). Depression has been either
directly or indirectly implicated in the association between adverse events
and suicide attempts (Dube et al., 2001). Lee and colleagues (2006) have
showed that depression mediated the associations of academic- and family-
related risks and suicide ideation in a sample of Hong Kong adolescents.
Substance use is also a risk factor for adolescent suicide (Waldrop et al.,
2007). Studies indicate that many attempters have ingested alcohol and/or
illicit drugs at the time of their suicidal gestures (Hufford, 2001). Chemical
dependence has been found to be associated with suicide attempts at
higher rates in adolescents than adults (Holland and Griffin, 1984). Other
psychological risk factors for suicide include low self-esteem, hopelessness,
impulsivity and anxiety (Strauss et al., 2000; Wetzler et al., 1996). Of these,
hopelessness tends to be the most salient predictor of both—the new
onset of suicidal behaviour and the progression from general depression
to a high risk for suicide (Beautrais et al., 1999). Based on a 13-year
longitudinal study, Kuo, Gallo and Eaton (2004) revealed that hopelessness
is an independent risk factor for completed suicide, suicide attempts and
suicidal ideation, regardless of the presence of depression or other mental
health disorders. This pattern of association between hopelessness and
suicide behaviours and ideation tends to be similar in Western and Asian
countries (Zhang et al., 2002).
Family dynamics are related to suicide risk as well. Adolescents who
exhibit suicide behaviour have often experienced some form of family
dysfunction (Bridge et al., 2006). For example, those living in families
with high levels of conflict and other major stressors, are prone to
depression (Waldrop et al., 2007). This tends to be especially true for
those in traditional Chinese society, where great emphasis is placed on
close interpersonal ties, mutual dependence and harmony in the family

329
Angel Nga-man Leung et al.

(Harris and Molock, 2000). One example of family dynamics as a risk


factor is the range of children’s behavioural and emotional problems
including a higher risk for suicidal behaviour, resulting from parental
divorce in Chinese families (Dong et al., 2002). In addition, family physical
maltreatment is associated with greater risk for adolescents’ attempted
suicide and chronic suicidal behaviour in both Western (Santa Mina and
Gallop, 1998; Ystgaard et al., 2004) and Asian cultures (Lau et al., 2003).
Zeng and Le Tendre (1998) noted that, apart from family stresses,
schooling and examination expectations are major sources of stress
and problems related to suicide among students in Hong Kong, Japan,
South Korea and Taiwan, cultures with a notoriously strong emphasis
on educational attainment and exams. Owing to the strong impact
of Confucian tradition on the values of education and family life, for
example, Chinese children’s parents particularly emphasise the need for
academic success (Tse and Bagley, 2002). Children’s failures to succeed
academically are virtually universally attributed either to a lack of hard
work or disrespect for parents and teachers (Tse and Bagley, 2002).
Such interpretations may make unremarkable students more prone to
depression and hopelessness in Asia as compared to the West. Hong Kong
adolescents have rated academic concerns as one of the top stress-related
domains in their lives (Lee et al., 2006), and examination pressure itself
has been found contributing to adolescents’ mental health problems,
including depression and suicide intention (Lau et al., 1998).
Apart from family and academic concerns, the peer group is a third
major factor in adolescents’ psychosocial adjustment and, conversely,
suicide ideation. Peer victimisation and poor peer relationships are
strong predictors of adolescent suicidality (Granello and Granello, 2007).
Peer victimisation in both direct (i.e., physical or verbally bullying) and
relational (i.e., social exclusion, rumour-spreading) forms are associated
with suicidal thinking and behaviour (Baldry and Winkel, 2003), and this
is the case for both victims and bullies (Toros et al., 2004). Moreover,
suicidal individuals tend to have poor interpersonal relationships with
peers (King and Merchant, 2008). In a longitudinal study of 659 families
(Johnson et al., 2002), eight types of interpersonal difficulties were found
to account for suicide attempts, including difficulty in making friends,
frequent arguments or anger with peers, social isolation and lack of close
friends. Peer isolation is not merely a cause of loneliness and unhappiness,
such isolation also prevents further problem-solving because of a lack of

330
Adolescents’ Suicide in Asia

social resources (Johnson et al., 2002). In part, because of the power of peer
relationships, teenagers whose peers have either attempted or completed
suicide tend to be at greater risk for suicidal ideation themselves (Ho et al.,
2000). Although research on peer groups in relation to suicidal behaviour
in Asian adolescents remains sparse, links between peer difficulties and
depression are consistent across cultures (e.g., Lam et al., 2004).
The media is another additional factor in actual suicide cases across
societies. Considerable evidence for imitation effects from suicide reported
via newspaper and television has accumulated (Stack, 2000a). A likely
explanation for the media-related contagion effect is that vulnerable
individuals, who may have had some predispositions towards suicide
ideation but normally would not have carried out a suicidal attempt, may
be encouraged to act on their suicidal impulses because of media reports.
When newspaper stories about suicides are featured prominently, imitation
effect tends to be particularly strong. Moreover, such imitation effects are
particularly likely to affect young people (Stack, 2000b). Romer, Jamieson
and Jamieson (2006) found that increased numbers of television news
reports of suicides were associated with a significant increase in suicides
for those under the age of 25. Similarly, greater numbers of newspaper
reports on suicide were associated with suicide deaths across age groups
in a four-month study period. Recently in Hong Kong, deaths by suicide
increased substantially following the death of a well-known Hong Kong
pop singer who jumped from a high building (Yip et al., 2006). This
phenomenon again underscored the importance of extensive and dramatic
media coverage of suicides.
The influence of the Internet on adolescent suicide is a new concern for
researchers and the general public. Adolescents spend a large proportion
of their free time on computers and the Internet (Stanger and Gridina,
1999). There are worries about the shocking number of websites providing
detailed methods and means to commit suicide (Thompson, 1999). Alao
and Pohl (2006) summarised nine suicide attempts or completed suicides
related to the Internet. Five of the cases involved adolescents and young
adults from ages 16 to 25 years, while the rest ranged from 34 to 42 years
old. The adolescents and young adults had all gained information about
lethal means of committing suicide from the web. Some of them were
encouraged by others in chat rooms to commit suicide. Teenagers, especially
those lacking social support, may be particularly vulnerable to biased
opinions in those chat rooms (Alao and Pohl, 2006; Mehlum, 2000).

331
Angel Nga-man Leung et al.

Cases of Internet-related suicide pacts are also alarming. For instance, in


2004, there were two suicidal pacts involving seven young people, ranging
in age from 20 to 30 years, who committed suicide together as arranged
via the Internet (Harding, 2004). Another Internet suicide pact in 2006 in
Japan resulted in six deaths (Harding, 2004). Although Internet-related
suicide pacts are rare (indeed, suicide pacts are themselves a rare act), there
has been a sharp increase in the number of cases reported. For example,
in Japan, there were 34 Internet suicide pacts made in 2003, 55 in 2004
and 91 in 2005 (Harding, 2004). Apart from the use of online chat rooms
and forums, young people have increasingly tended to express suicidal
ideation in their blogs. The Samaritan Befrienders Hong Kong found more
than 3,000 blogs expressing emotional problems and suicidal ideations
(Yahoo News, 2008b). Furthermore, in 2008, at least five youngsters who
expressed suicidal thoughts or left death notes in their blogs subsequently
committed suicide (Yahoo News, 2008a). One of them left entries in his
blog culminating in a 24-day countdown to his suicide. He did actually
commit suicide at the end of that time (Yahoo News, 2008a). Thus, bloggers
sometimes communicate with one another about suicide, exchanging
suicidal methods or thoughts.
The influence of the Internet on adolescents in relation to suicide
likely depends on the nature of the Internet chat rooms and the group
components of the online chat rooms with which they become involved.
For instance, website users have reported gaining support, sympathy
and understanding from other online users about suicidal thoughts and
even actions as an online community (Baker and Fortune, 2008; Miller
and Gergen, 1998). Some believe that expressing their suicidal thoughts
or depression online is a reasonable way of coping and reducing their
self-harming behaviour (Baker and Fortune, 2008). The debate about
the influence of the Internet on adolescent suicide is ongoing. However,
research in this area is vital given the increasing prominence of the Internet
in modern society.

WARNING SIGNS FOR ADOLESCENT SUICIDE


ATTEMPTS OR BEHAVIOUR

Suicide seldom occurs without warning signs, and being able to identify
those common signs is important for future work with adolescents. Most of

332
Adolescents’ Suicide in Asia

such warning signals fall into one of the four major categories: behavioural,
emotional, physical and verbal (The Hong Kong Jockey Club, 2005).
Physical signs of suicidal ideation in adolescents may include sudden
changes in eating habits (e.g., loss of hunger or overeating) and altered
sleeping patterns (e.g., oversleeping or insomnia). Some chronic or un-
explained pains and illness can serve as signs as well. Suicidal adolescents
may also neglect their personal appearance or present a dramatic change
in personal appearance, particularly when the new style is thoroughly out
of character (Kalafat, 1990). Emotional signs of suicidal thoughts relate
to signs of apparent motivation or mood. For example, when young
people become suicidal, they may lose interest in friends, hobbies or
other activities formerly enjoyed; they sometimes also present a dramatic
change in personality, such as becoming withdrawn when they were
formerly extroverted (Caruso, n.d.). In addition, a sudden improvement
after a long period of emotional turmoil may actually signal an imminent
suicide (Granello and Granello, 2007); the outward feeling of calmness
may indicate that nothing matters anymore to them now since they have
already made a decision to carry out a suicide attempt. Other behaviours
correlated with suicide sometimes present themselves as suicide-related
gestures or attempts, including overdosing and self-cutting (Rudd et al.,
2006). Finally, some individuals may also make their suicidal ideation
clear, either by direct verbal warnings or through indirect and abstract
statements. Most victims of suicide had expressed their intentions through
some type of verbal ‘cry for help’ previously (Granello and Granello, 2007).
Such patterns are likely to be similar across cultures.

PREVENTION

Considering the variety of risk factors for adolescent suicide, in addition


to the fact that suicide itself is a multi-faceted phenomenon, there are a
number of potential preventive measures that can be considered in any
community. Early on, the Centre for Disease Control (CDC, 1992) sug-
gested that early screening programmes taking place at school showed
promising results. Such programmes are viewed as a cost-effective way to
locate high-risk students despite locating a fair number of ‘false positive’ at-
risk students as well (Shaffer and Craft, 1999). Alternatively, programmes

333
Angel Nga-man Leung et al.

aimed at skills training for adolescents can be considered more widely.


Workshops and training on improving stress and social management,
enhancing self-esteem, and increasing problem-solving skills are promising
for promoting the healthy development of adolescents (Pelkonen and
Marttunen, 2003). These kinds of training may be especially important for
preventing suicides in Asian adolescents, given that academic performance
is highly valued in most Asian countries (China, Hong Kong, Japan,
Korea) often to the extent of being detrimental of emotional and social
development (Zeng and Le Tendre, 1998).
It is important to provide opportunities for adolescents to understand
that they can realise their potential in areas other than academic success
(Chia, 1999). In addition, systematic training in problem-solving tech-
niques (D’Zurilla and Nezu, 1990; Orbach, 2003) could be taught in
groups to adolescents in an attempt to prevent suicidal thinking or actions.
Klingman and Hochdorf (1993) suggested that by changing automatic
negative thoughts and irrational thinking, negative self-beliefs could
become more positive and low self-esteem could be raised, eventually
resulting in a better self-image (Orbach, 2003).
Community gatekeepers, who are in close contact with adolescents,
including teachers, counselors, coaches and even physicians, police,
recreational staff, could be better trained to notice warning signs for sui-
cide or suicidal ideation in adolescents as well. Training can help to provide
these gatekeepers with the tools to identify at-risk teens, handle their
emotional episodes, and refer them to appropriate mental health services
(Gould and Kramer, 2001). Tierney (1994) showed that after attending
training programmes, the knowledge, attitudes and intervention skills
of community helpers increased significantly. Other community-based
prevention programmes, though not necessarily used in the adolescent
population, can also help to prevent suicides. For an example, Rihmer,
Rutz and Pihlgren (1995) trained physicians to identity depression. Such
training resulted in a significant reduction in suicide in the community.
In another multi-layered intervention programme aimed at reducing risk
factors and enhancing protective factors launched by the American Air
Force, a large drop of risks for suicide was observed (Knox et al., 2003).
In Hong Kong, a community-based prevention programme was
launched by the Hong Kong Jockey Club Centre for Suicide Research
and Prevention (CSRP, 2004), on Cheung Chau Island in 2002. This
island is infamous among Hong Kong people because of the number

334
Adolescents’ Suicide in Asia

of suicides using burning charcoal that have taken place there. Training
sessions for holiday flat owners, police and others interested in pre-
venting suicides were provided to educate them about warning signs of
suicide. Some practical strategies such as putting up posters with en-
couraging statements, suggesting that holiday flat owners not rent to
people who seemed depressed, and convincing storekeepers not to sell
medicines to teenagers were implemented. Perhaps correspondingly, there
was a marked decrease in the number of suicides in Cheung Chau, from
50 deaths in total in 1998–2002 to 11 in 2003–04 (CSRP, 2004). Thus, a
focus on public awareness may be helpful for reducing suicide attempts
and completions.
Attention to the media can clearly cut down on suicides as well.
After implementing media guidelines on reporting suicide cases, there
has been a significant drop in suicide rates (Etzersdorfer et al., 1992).
Thus, the CDC (1994) recommends that front-page coverage of suicidal
cases should be avoided. Also, detailed descriptions of the means of suicide,
as well as stories of the deceased, should not be reported. In addition,
referral resources involving mental health professionals should be included
along with any news about suicide in published materials (CDC, 1994).
Methods of suicides, very much a product of the region in which one
lives, can also be more carefully monitored and controlled in an effort to
cut down on suicides. For example, one of the most common means by
which adolescents commit suicides in Western countries (e.g., America)
is by using firearms (Gould and Kramer, 2001). Past studies have shown
that restrictions on access to firearms could reduce adolescent suicide
rates (Carrington and Moyer, 1994). However, in Asia, the most common
methods used for committing suicides are hanging, jumping from high-
rise buildings, ingesting pesticides and charcoal burning. Accordingly,
legislation aimed at preventing access to toxins such as pesticides should
be passed wherever possible. For example, in Sri Lanka, safer storage of
pesticides has reduced suicide rates (Hawton, 2008). Chia (1999) noted
that 75 percent of young suicide victims in Singapore jumped from
kitchens in their own home and, therefore, suggested that barriers should
be placed in kitchens. Window bars with locks could also be added. For
charcoal burning, popular in Hong Kong and nearby Macau, researchers
have suggested that crisis hotlines from suicide prevention centres could
be printed on the packages of charcoal to help those who plan to commit
suicide (Yip, 2006).

335
Angel Nga-man Leung et al.

For individuals, parents in particular, having the skills to talk with


suicidal people, especially adolescents, is particularly important. Many
adolescents faced with suicidal thoughts lack the appropriate coping
skills and knowledge to seek help. In 2006, the Suicide Crisis Intervention
Centre in Hong Kong (SCIC, 2006) suggested some simple tactics for
handling those who show some warning signs of suicidal behaviour. They
advise that others trust and accept that these individuals are really having
thoughts of suicide, listen carefully, and be patient in interacting with
suicidal adolescents who are expressing their feelings. Both encouraging
professional help and maintaining an ongoing relationship with these
adolescents are essential. In extreme cases, parents or other adults can even
consider removing lethal or toxic substances and locking up window bars
to reduce the means by which a suicide could be attempted as well.
The major treatments for suicidal attempters and ideators include
psychosocial and biological methods. Pelkonen and Marttunen (2003)
compared the efficacy of eight psychosocial treatments. They found
that cognitive-behavioural therapy (CBT), which aims at correcting the
cognition of adolescents, rather than correcting their depressive feelings and
behaviours, work well for most adolescents. Interpersonal psychotherapy
can also significantly reduce depression symptoms in adolescents (Rossello
and Bernal, 1999). Although research on the usefulness of selective
serotonin reuptake inhibitors (SSRIs) is still limited, there is evidence for
their effectiveness in treating major depressive disorders in children and
adolescents (Rossello and Bernal, 1999). Mood stabilisers are regarded
as a relatively good and effective first-line medication for teenagers
and children who are suffering from depression as well (Pelkonen and
Marttunen, 2003).

CONCLUSIONS

Suicide risk increases with age throughout adolescence. Adolescent


suicide is a complicated issue involving different risk factors, ranging
from psychological to environmental and socio-cultural. It is important
for the government, general public, teachers, parents and teenagers
themselves to be aware of the problem of adolescent suicide. A multi-
faceted approach to suicide prevention, ideally involving governments,

336
Adolescents’ Suicide in Asia

schools and the mass media, in addition to peers and families, will likely be
most effective in limiting suicides in Asia. Meanwhile, every one of us can
learn simple techniques for interacting with teenagers at-risk for suicide,
and a combination of strategies aimed at providing support and care to
adolescents around us, may be especially helpful at a local level.

REFERENCES

Aaron, R., A. Joseph, S. Abraham, J. Muliyil, K. George, J. Prasad et al. (2004). Suicides
in young people in rural southern India. The Lancet, 363(9415), 1117–18.
Alao, A. and E. Pohl (2006). Cybersuicide: review of the role of the internet on suicide.
Cyberpsychology & Behavior, 9(4), 489–93.
Baker, D. and S. Fortune (2008). Understanding self-harm and suicide websites: A
qualitative interview study of young adult website users. Crisis, 29(3), 118–22.
Baldry, A.C. and F.W. Winkel (2003). Direct and vicarious victimization at school and
at home as risk factors for suicidal cognition among Italian adolescents. Journal
of Adolescence, 26(6), 703–16.
Beautrais, A.L. (2006). Suicide in Asia. Crisis, 27(2), 55–57.
Beautrais, A.L., P.R. Joyce and R.T. Mulder (1999). Personality traits and cognitive
styles as risk factors for serious suicide attempts among young people. Suicide and
Life-Threatening Behavior, 29(1), 37–47.
Bridge, J.A., T.R. Goldstein and D.A. Brent (2006). Adolescent suicide and suicidal
behavior. Journal of Child Psychology and Psychiatry, 47(3–4), 372–94.
Carrington, P.J. and S. Moyer (1994). Gun control and suicide in Ontario. Journal of
Psychiatry, 151(4), 606–08.
Caruso, K. (n.d.) Suicide warning signs. Retrieved 17 November 2008 from http://www.
suicide.org/suicide-warning-signs.html
Centre for Disease Control [CDC]. (1992). Youth suicide prevention programmes: A
resource guide. Atlanta, GA: Atlanta Centre for Disease Control.
Centre for Disease Control [CDC]. (1994). Suicide contagion and the reporting of
suicide: Recommendations from a national workshop. Retrieved 1 November 2008
from http://www.cdc.gov/mmwr/preview/mmwrhtml/00031539.htm
Cheng, A.T.A. and C.S. Lee (2000). Suicide in Asia and the Far East. In K. Hawton and
K. Heeringen (Eds), The International Handbook of Suicide and Attempted Suicide
(pp. 24–48). New York: Wiley.
Chia, B.H. (1999). Too young to die. An Asian perspective on youth suicide. Singapore:
Times Books International.
Department of Health of the Republic of China (2008). Statistics on causes of death.
Retrieved 1 November 2008 from http://www.doh.gov.tw

337
Angel Nga-man Leung et al.

Dong, Q., Y. Wang and T.H. Ollendick (2002). Consequences of divorce on the
adjustment of children in China. Journal of Clinical Child and Adolescent Psychology,
31(1), 101–10.
Dube, S.R., R.F. Anda, V.J. Felitti, D.P. Chapman, D.F. Williamson and W.H. Giles
(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(24), 3089–96.
D’Zurilla, T.J. and A.M. Nezu (1990). Development and preliminary evaluation of the
social problems solving inventory. Psychological Assessment: A Journal of Consulting
and Clinical Psychology, 2(2), 156–63.
Etzersdorfer, E., G. Sonneck and S. Nagel-Kuess (1992). Newspaper reports and suicide.
New England Journal of Medicine, 327(7), 502–03.
Gould, M.S. and R.A. Kramer (2001). Youth suicide prevention. Suicide and Life-
Threatening Behavior, 31(Supplement), 6–31.
Granello, D.H. and P.F. Granello (2007). Suicide: An essential guide for helping
professionals and educators. Boston: Pearson/Allyn & Bacon.
Groholt, B., O. Ekeberh, L. Wichstrom and T. Haldorsen (1998). Suicide among children
and younger and older adolescents in Norway: A comparative study. Journal of the
American Academy of Child and Adolescent Psychiatry, 37(5), 473–81.
Guetzloe, E.C. (1989). Youth suicide: What the educator should know. Reston, VA:
Council for Exceptional Children.
Gururaj, G. and M.K. Isaac (2001). Epidemiology of suicides in Bangalore. Bangalore:
National Institute of Mental Health and Neuro Sciences. (Publication No 43.)
Harding, A. (2004). Japan’s internet ‘suicide clubs’. BBC News. Retrieved 1 November
2008 from http://news.bbc.co.uk/2/hi/programmes/newsnight/4071805.stm
Harris, T.L. and S.D. Molock (2000). Cultural orientation, family cohesion and family
support in suicide ideation and depression among African American college
students. Suicide and Life-Threatening Behavior, 30(4), 341–53.
Hawton, K. (2008). Safer storage of pesticides to prevent suicide: Results of a project in Sri
Lanka. Paper presented at the 3rd Asia Pacific Regional Conference of International
Association for Suicide Prevention, Hong Kong.
Hendin, H., L. Vijayakumar, J.M. Bertolote, H. Wang, M.R. Phillips and J. Pirkis (2008).
Epidemiology of suicide in Asia. In H. Hendin, M.R. Phillips, L. Vijayakumar, J.
Pirkis, H. Wang, P. Yip, D. Wasserman, J. Bertolote and A. Fleischmann (Eds),
Suicide and Suicide Prevention in Asia (pp. 7–18). Geneva: WHO Press.
Ho, T.P., P.W.L. Leung, S.F. Hung, C.C. Lee and C.P. Tang (2000). The mental health
of the peers of suicide completers and attempters. Journal of Child Psychology and
Psychiatry, 41(3), 301–08.
Holland, S. and A. Griffin (1984). Adolescent and adult drug treatment clients: Patterns
and consequences of use. Journal of Psychoactive Drugs, 16(1), 79–89.
Hong Kong Jockey Club Centre for Suicide Research and Prevention [CSRP] (2004).
Report 6: Community development suicide prevention programme on Cheung
Chau Island. Retrieved 15 November 2008 from http://csrp.hku.hk/files/589_
2772_489.pdf

338
Adolescents’ Suicide in Asia

Hong Kong Jockey Club Centre for Suicide Research and Prevention (2005). Student
suicide prevention practical tips & strategies. Retrieved 15 October 2008 from
http://csrp1.hku.hk/sss/intervention.html
Hufford, M.R. (2001). Alcohol and suicidal behavior. Clinical Psychology Review,
21(5), 797–811.
Johnson, J.G., P. Cohen, M.S. Gould, S. Kasen, J. Brown and J.S. Book (2002). Childhood
adversities, interpersonal difficulties, and risk for suicide attempts during
late adolescence and early adulthood. Archives of General Psychiatry, 59(8),
741–49.
Joseph, A., S. Abraham, J.P. Muliyil, K. George, J. Prasad, S. Minz et al. (2003).
Evaluation of suicide rates in rural India using verbal autopsies, 1994–99. British
Medical Journal, 326(7399), 1121–22.
Kalafat, J. (1990). Adolescent suicide and the implications for school response
programmes. School Counselor, 37(5), 359–69.
Kelleher, M.J. and D. Chambers (2003). Cross-cultural variation in child and adolescent
suicide. In Robert A. King and A. Apter (Eds), Suicide in Children and Adolescents
(pp. 170–97). Cambridge, United Kingdom: Cambridge University Press.
King, C.A. and C.R. Merchant (2008). Social and interpersonal factors relating to
adolescent suicidality: A review of the literature. Archives of Suicide Research,
12(3), 181–96.
Klingman, A. and Z. Hochdorf (1993). Coping with distress and self harm: the impact
of primary prevention programmes among adolescents. Journal of Adolescence,
16(2), 121–40.
Knox, K.L., D.A., Litts, G.W. Talcott, C. Feig and E. Caine (2003). Risk of suicide and
related adverse outcomes after exposure to a suicide prevention program in the
US Air Force: Cohort study. British Medical Journal, 327(7428), 1376–81.
Kubota, Y. (2008). Japan suicides near record high in 2007. Posted on 19 June 2008.
Retrieved 17 October 2008 from http://www.reuters.com/article/latestCrisis/
idUST1344
Kuo, W.H., J.J. Gallo and W.W. Eaton (2004). Hopelessness, depression, substance
disorder, and suicidality: A 13-year community-based study. Social Psychiatry &
Psychiatric Epidemiology, 39(6), 497–501.
Lam, T.H., S.M. Stewart, P.S.F. Yip, G.M. Leung, L.M. Ho, D.S.Y. Ho et al. (2004).
Suicidality and cultural values among Hong Kong adolescents, Social Science &
Medicine, 58(3), 487–98.
Lau, S., C.K.K. Siu and M.P.Y. Chik (1998). The self-concept development of Chinese
primary schoolchildren : A longitudinal study. Childhood, 5(1), 69–79.
Lau, T.F., K.K. Chan, K.W. Lam, Y.W. Choi and Y.C. Lai (2003). Psychological
correlates of physical abuse in Hong Kong Chinese adolescents. Child Abuse &
Neglect, 27(1), 63–75.
Lee, T.Y., B. Wong, W.Y. Chow and C. McBride-Chang (2006). Predictors of suicide
ideation and depression in Hong Kong adolescents: Perceptions of Academic and
Family Climates. Suicide and Life-Threatening Behavior, 36(1), 82–96.

339
Angel Nga-man Leung et al.

Lewis, L. (2008). Japan gripped by suicide epidemic. Posted on 19 June 2008. Times
Online. Retrieved 19 October 2008 from http://www.timesonline.co.uk/tol/news/
world/asia/article4170649.ece
Mehlum, L. (2000). The internet, suicide, and suicide prevention. Crisis, 21(4),
186–88.
Miller, J. and K. Gergen (1998). Life on the line: The therapeutic potentials of computer
mediated conversation. Journal of Marital and Family Therapy, 24(2), 189–202.
National Crime Records Bureau (2006). Accidental deaths and suicide in India.
New Delhi: Government of India.
Nrugham, L., B. Larsson and A.M. Sund (2008). Predictors of suicidal acts across
adolescence: Influences of familial, peer and individual factors. Journal of Affective
Disorders, 109(1–2), 35–45.
Orbach, I. (2003). Suicide prevention for adolescents. In R.A. King and A. Apter
(Eds), Suicide in Children and Adolescents (pp. 227–70). Cambridge: Cambridge
University Press.
Pelkonen, M. and M. Marttunen (2003). Child and adolescent suicide: Epidemiology,
risk factors, and approaches to prevention. Pediatric Drugs, 5(4), 243–65.
Philips, M.R. (2002). Risk factors for suicide in China: a national case-control
psychological autopsy study. Lancet, 360(9347), 1728–36.
Philips, M.R. (2008). China’s changing pattern of suicide, 1987–2006. Paper presented
at the 3rd Asia Pacific Regional Conference of International Association for Suicide
Prevention, Hong Kong.
Rihmer, Z., W. Rutz and H. Pihlgren (1995). Depression and suicide on Gotland. An
intensive study of all suicides before and after a depression-training programme
for general practitioners. Journal of Effective Disorders, 35(4), 147–52.
Romer, D., P.E. Jamieson and K.H. Jamieson (2006). Are news reports of suicide
contagious? A stringent test in six U.S. cities. Journal of Communication, 56(2),
253–70.
Rossello, J. and G. Bernal (1999). The efficacy of cognitive-behavioral and interpersonal
treatments for depression in Puetro Rican adolescents. Journal of Consulting Clinical
Psychology, 67(5), 734–45.
Rudd, M.D., A.L. Berman, T.E. Joiner, M.K. Nock, M.M. Silverman, M. Mandrusiak
et al. (2006). Warning signs for suicide: theory, research, and clinical applications.
Suicide and Life-Threatening Behavior, 36(2), 255–62.
Santa Mina, E.E. and R.M. Gallop (1998). Childhood sexual and physical abuse and
adult and adult self-harm and suicidal behavior: A literature review. Canadian
Journal of Psychiatry, 43(8), 793–800.
Shaffer, D. and L. Craft (1999). Methods of adolescent suicide prevention. Journal of
Clinical Psychiatry, 60 (Supplement), 70–74.
Shek, D.T.L., B.M. Lee and J.T.W. Chow (2005). Trends in adolescent suicide in Hong
Kong for the period of 1980 to 2003. The Scientific World Journal, 5(8), 702–23.
Stack, S. (2000a). Suicide: A 15-year review of the sociological literature. Part II:
Modernization and social integration perspectives. Suicide and Life-Threatening
Behavior, 30(2), 145–62.

340
Adolescents’ Suicide in Asia

Stack, S. (2000b). Media impacts on suicide: A quantitative review of 293 findings.


Social Science Quarterly, 81(4), 957–71.
Stanger, J.D. and N. Gridina (1999). Media in the home 1999: The fourth annual
survey of parents and children (Survey Series No. 5). Norwood, NJ: Philadelphia,
PA: Annenberg Public Policy Centre of the University of Pennsylvania. Retrieved
26 October 2006, from http://www.appcpenn.org/pubs.htm
Stewart, S.M., E. Felice, C. Claassen, B.D. Kennard, P.W. Lee and G.J. Emslie (2006).
Adolescent suicide attempters in Hong Kong and the United States. Social Science
and Medicine, 63(2), 296–306.
Strauss, J., B. Birmaher, J. Bridge, D. Axelson, L. Chiappetta, D. Brent et al. (2000).
Anxiety disorders in suicidal youth. Canadian Journal of Psychiatry, 45(8),
739–45.
Suicide Crisis Intervention Centre [SCIC]. (2006). Timely Rainfall in Desert. Hong Kong:
Suicide Crisis Intervention Centre, the Samaritan Befrienders Hong Kong.
Thompson, S. (1999). The Internet and its potential influence on suicide. Psychiatric
Bulletin, 23(8), 449–51.
Tierney, R.J. (1994). Suicide intervention training evaluation: a preliminary report.
Crisis, 15(2), 69–76.
Toros, F., N.G. Bilgin, T. Sasmaz, R. Bugdayci and H. Camdeviren (2004). Suicide
attempts and risk factors among children and adolescents. Yonsei Medical Journal,
45(3), 367–74.
Tse, W.L. and C. Bagley (2002). Suicidal behavior in Chinese youth: search for causes
and risk factors. In W.L. Tse and C. Bagley (Eds), Suicidal behavior, bereavement
and death education in Chinese adolescents: Hong Kong studies (pp. 1–24).
Aldershot: Ashgate Publishing Limited.
Vijayakumar, L. (2005). Suicide and mental disorders in Asia. International Review of
Psychiatry, 17(2), 109–14.
Waldrop, A.E., R.F. Hanson, H.S. Resnick, D.G. Kilpatrick, A.E. Naugle and B.E.
Saunders (2007). Risk factors for suicidal behavior among a national sample
of adolescents: Implications for prevention. Journal of Traumatic Stress, 20(5),
869–79.
Wetzler, S., G.M. Asnis, R.B. Hyman, C. Virtye, J. Zimmerman and J.H. Rathus (1996).
Characteristics of suicidality among adolescents. Suicide and Life-Threatening
Behavior, 26(1), 37–45.
WHO (World Health Organization) Mortality Database (2008). Number of registered
deaths. Retrieved 15 November 2008 from http://www.who.int/whosis/database/
mort/table1.cfm
Yahoo News (2008a). Desperate adolescents committed suicide after expressing suicidal
thoughts on Internet Blog but being ignored; five reported local Hong Kong cases
within 4 months, 11 May 2008. Retrieved 14 November 2008 from http://hk.news.
yahoo.com/article/080510/4/4eq0.html
Yahoo News (2008b). Suicide prevention by Samaritan Befrienders Hong Kong:
screening at-risk people from Internal Blogs and online counseling for

341
Angel Nga-man Leung et al.

identified suicidal groups. 3000 blogs were found with suicide-related emotional
disturbances, 29 June 2008. Retrieved 17 November 2008 from http://hk.news.
yahoo.com/article/080628/4/5kdk.html
Yip, P.S. (2001). An epidemiological profile of suicides in Beijing, China. Suicide and
Life-Threatening Behavior, 31(1), 62–70.
Yip, P.S. (2006). Suicide and Suicide Preventions in Hong Kong. Symposium presented
at Suicide Prevention ‘Making a Difference in Life Saving Work’. Retrieved from
http://csrp1.hku.hk/files/592_2820_525.pdf
Yip, P.S., K.W. Fu, K.C.T. Yang, B.Y.T. Ip, C.L.W. Chan, E.Y.H. Chen et al. (2006).
The effect of a celebrity suicide on suicide rates in Hong Kong. Journal of Affective
Disorders, 93(1), 245–52.
Yip, P.S. and K. Liu (2006). The ecological fallacy and the gender ratio of suicide in
China. British Journal of Psychiatry, 189(5), 465–66.
Yip, P.S., K.Y. Liu, J. Hu and X.M. Song (2005). Suicide rates in China during a decade
of rapid social changes. Social Psychiatry and Psychiatric Epidemiology, 40(10),
792–98.
Ystgaard, M., I. Hestetun, M. Loeb and L. Mehlum (2004). Is there a specific relationship
between childhood sexual and physical abuse and repeated suicidal behavior? Child
Abuse and Neglect, 28(8), 863–75.
Zeng, K. and G. Le Tendre (1998). Adolescent suicide and academic competition in
East Asia. Comparative Education Review, 42(4), 513–28.
Zhang, J., S. Jia, W. Wieczorek and C. Jiang (2002). An overview of suicide research in
China. Archives of Suicide Research, 6(2), 167–84.

342
About the Editors and Contributors

The Editors

Updesh Kumar, PhD, Scientist ‘F’ is Head of the Mental Health Division,
Defence Institute of Psychological Research (DIPR), Defence R&D Or-
ganisation (DRDO), Ministry of Defence, Government of India, Delhi.
Dr Kumar obtained his doctorate degree from Panjab University,
Chandigarh. He specialises in the area of suicidal behaviour and per-
sonality assessment. Starting with his doctoral thesis in the area of sui-
cidal behaviour, Dr Kumar, as principal investigator, has carried out
many major research projects related to suicidal behaviour. As Head,
Mental Health Division, he has psychologically analysed many suicide
incidents of defence personnel and followed the performance of cadets in
various training academies, namely National Defence Academy, Indian
Military Academy, Officers Training Academy, Air Force Academy, Air
Force Technical College and Naval Academy. With more than 18 years
of service as scientist, Dr Kumar has been the psychological assessor
(Psychologist) in various Services Selection Boards for eight years for the
selection of officers in Indian Armed Forces. Dr Kumar has developed
many psychological tests and assessment tools for the selection of officers
and recently the psychological test battery and manual for the selection of
Other Ranks (PBOR) in Indian Armed Forces. Dr Kumar has authored a
field manual on ‘Suicide and Fratricide: Dynamics and Management’ for
defence personnel, ‘Managing Emotions in Daily Life & at Work Place’ for
general population and ‘Overcoming Obsolescence & Becoming Creative
in R&D Environment’ for R&D organisations. Dr Kumar has conducted
many workshops and a National conference on the theme of psychology
related topics. Dr Kumar has edited a book, Recent Developments in

343
Suicidal Behaviour

Psychology, along with many other academic publications in the form of


journal articles and chapters in books.

Manas K. Mandal, PhD, Scientist ‘H’, Outstanding Scientist, is Director,


Defence Institute of Psychological Research (DIPR), Delhi. Dr Mandal
has obtained his Postgraduate and doctorate degrees from Calcutta Uni-
versity in 1979 and 1984, respectively. He has completed his post-doctoral
research programme at Delaware University (Fulbright Fellow), USA in
1986–87 and at Waterloo University (Shastri and Natural Sciences and
Engineering Research Council [NSERC] Fellow), Canada, in 1993–94.
Dr Mandal was a Professor of Psychology at IIT Kharagpur. He was also
a visiting professor at Kyushu University, Japan in 1997. During 2003 he
was a Fulbright Visiting Lecturer, Harvard University, USA. He has been
awarded various research fellowships and scientific awards at national
and international levels, such as International Scientific Exchange Award
(Canada), Fulbright Fellowship (USA), Shastri Fellowship (Canada),
Seymour Kety grant (USA), Career Award (India), University Gold Medal.
Recipient of four prestigious awards from the prime ministers of India,
Young Scientist Award (1986), Agni award for excellence in self-reliance
(2005), Scientist-of-the-Year Award (2006), Technology Spin-off award
for DIPR (2007), Dr Mandal has to his credit six books, and over 100
research papers/chapters in books (80 international and 25 national)
published in peer-reviewed journals/books. These researches are cited in
125 international journals and books with more than 500 citations.

The Contributors

Kathryn Kanzler Appolonio, Psy.D., is an active-duty USAF staff


psychologist for the Mental Health Clinic and the Clinical Health Psy-
chology Service at Lackland Air Force Base, Texas. She trains clinical
psychology interns in empirically-based treatments in her role as a core
faculty member in the Department of Psychology at Wilford Hall Medical
Center’s APA-accredited psychology internship programme. Dr Appolonio
received her doctorate in clinical psychology from La Salle University
and completed her internship at Wilford Hall Medical Center. Dr Appolonio’s

344
About the Editors and Contributors

clinical and research interests involve clinical health and primary care
psychology, as well as clinical suicidology.

Craig J. Bryan, Psy.D., is the Chief of Primary Care Psychology Services at


Kelly Family Medicine Clinic, and Suicide Prevention Program Manager
for Lackland Air Force Base, Texas. He is also a core faculty member in
the Department of Psychology at Wilford Hall Medical Center’s APA-
accredited psychology internship programme, where he trains clinicians
in the primary care behavioural health model. Dr Bryan received his
doctorate in clinical psychology from Baylor University, and completed
his clinical psychology internship at Wilford Hall Medical Center. Dr Bryan’s
primary areas of research include clinical suicidology and behavioural
health effectiveness research for integrated primary care clinics.

Erminia Colucci holds a Diploma in Education, and a first class honours


degree in Clinical Psychology from the University of Padua (Italy), Erminia
(Emy). She was trained as a researcher at The Australian Institute for
Suicide Research and Prevention (AISRAP), Griffith University, Brisbane.
In 2003 she won the Australian International postgraduate research
scholarships (University of Queensland International Postgraduate
Research Scholarship [UQIPRS] and International Postgraduate Research
Scholarship [IPRS]), which supported her PhD project ‘The Cultural
Meaning of Suicide: A Comparison between Italian, Indian and Australian
Students’ at The University of Queensland, where she was awarded a PhD
in Cultural Psychiatry. Her project was awarded the 2004 UQ Travel Award
and the 2005 Dr Helen Row–Zonta Memorial Prize. Since the last few
years she is a Research Fellow in the Centre for International Mental
Health (School of Population Health, The University of Melbourne),
where she is the Research Program Coordinator. She has also given
lectures in universities in India, Japan, Italy and Australia. Erminia has
presented papers on the cultural aspects of suicide, spirituality, gender
issues, qualitative research and arts-based research/prevention nationally
and internationally. She has also authored journal papers, book chapters
and other publications on these topics.

Enrica De Simoni, MD, is a resident in psychiatry at II Faculty of Medicine,


Sapienza University of Rome, Italy. She is interested in studying the role
of religion in mediating suicide risk. She recently carried out a number of

345
Suicidal Behaviour

review studies dealing with suicide risk related to sport activities, religion
and smoking.

Ilaria Falcone, MD, is a resident in psychiatry at II Faculty of Medicine,


Sapienza University of Rome, Italy. She is interested in studying the role
of impulsivity and aggression in mediating suicide risk. She has been
involved in the prevention of suicide in primary care. She carried out
studies dealing with suicide and pharmacological treatment.

Sandor Fekete, MD, PhD, is Professor and Head of the Department of


Psychiatry and Psychotherapy, University of Pecs, Hungary. He is a psy-
chiatrist, psychotherapist and neurologist as well as senior lecturer in
graduate/postgraduate education at the university, principal investigator
of collaborative studies on suicidal behaviour, and national focal point in
the WHO/EURO Network of Suicide Prevention and Research.

Cathy Yui-chi Fong obtained her Bachelor’s degree in psychology from


The Chinese University of Hong Kong. She was to begin her PhD studies
with a specialisation in educational psychology at the University of Hong
Kong in fall 2009. She has been involved in a number of research studies
on Chinese children’s reading development and impairment, including
co-authoring a chapter on this topic for a forthcoming book entitled The
Handbook of Chinese Psychology. Her interests include school readiness and
language and reading development, particularly in young children.

Marco Innamorati, Psy.D., is cognitive and behavioural psychotherapist


and Professor of Clinical Psychophysiology at Università Europea di Roma.
He collaborated at the validation for the Italian population of the Beck
Hopelessness Scale. He works for the prevention of suicide in high risk
groups such as youths and elderly.

Farah Kidwai holds a PhD in ‘Emotional Intelligence and Leadership’.


She specialises in the field of Applied Social Psychology and Personality
Assessment. She has a number of research studies to her credit and has
conducted several workshops in the area of interpersonal relations,
emotional intelligence and leadership in diverse organisations. A Scientist
‘E’ in the Defence R&D Organisation, she is the recipient of a number

346
About the Editors and Contributors

of scholarships and awards including Scientific Advisor’s Award for the


year 2001 by DRDO and Vice Chief of Army Commendation in 2008
by Indian Army. She was also recommended for the Award of Com-
monwealth Scholarship for Doctorate Degree in 1999 in Psychology by
the Government of India.

Angel Nga-man Leung obtained her Bachelor’s degree in Social Sciences


from the University of Hong Kong and her M.Phil. degree in psychology
from The Chinese University of Hong Kong. She is now a PhD candidate
in the Psychology Department of The Chinese University of Hong Kong
specialising in developmental psychology. Her major research interests are
parenting, adolescent depression and suicide ideation, and the influences
of online gaming on the social development of early adolescents.

Carmel McAuliffe is a Senior Researcher with the National Suicide Re-


search Foundation (NSRF) in Cork, Ireland. She is a psychology graduate
of University College Cork and has worked in suicide research since 1996
when she joined the NSRF as a Research Psychologist under the direction of
the late Dr Michael J. Kelleher. She has worked on several projects relating
to people who engage in deliberate self-harm including the WHO/EURO
Multicentre Study on Suicidal Behaviour and a randomised controlled
trial of Group Interpersonal Problem-Solving Skills Training for medically
treated deliberate self-harm patients. She has published in a number of
scientific peer-reviewed journals. She is currently working on a pilot
Suicide Support and Information System in the Cork region which aims
to better understand the causes of suicide and to improve the provision
of support to families bereaved by suicide.

Catherine Alexandra McBride-Chang is a Professor in the Psychology


Department of the Chinese University of Hong Kong specialising in de-
velopmental psychology. She has published journal articles on a variety
of topics, including adolescent suicide ideation, parenting, child abuse,
peer relations, creativity, and reading development and impairment. She
has also edited one book on Chinese children’s reading development and
authored a book entitled Children’s Literacy Development. She currently
serves as the Director of the Developmental Centre at The Chinese
University of Hong Kong.

347
Suicidal Behaviour

Lars Mehlum, MD, PhD, is the founding director of the National Centre
for Suicide Research and Prevention at the Institute of Psychiatry,
University of Oslo, Norway. He completed his medical training at the
University of Bergen in 1982, and his PhD at the University of Oslo in
1995. He is a board specialist of psychiatry in Norway and trained in psy-
chodynamic psychotherapy and dialectical behaviour therapy. He has
been a coordinator of the National Strategy for Suicide Prevention in
Norway since 1993, established a master’s degree on suicide prevention
at the University of Oslo, established a suicide preventive programme
in the Norwegian Armed Forces, headed a national suicide preventive
training programme for gate keepers, and co-founded the Norwegian
Association for Suicide Survivors. He has been actively supporting a
number of international suicide preventive initiatives throughout the years
and was president of the International Association for Suicide Prevention
(IASP) from 2003 to 2005. He has acted as an advisor for national suicide
preventive strategies in several countries. He is member of the Scientific
Advisory Council of the American Foundation for Suicide Prevention,
member of IASP, International Academy of Suicide Research and the
American Association of Suicidology. Apart from being the founding
editor of the journal Suicidologi published since 1996, he is also a member
of the editorial board of Suicide & Life-Threatening Behaviour and Archives
of Suicide Research. With his research group he focuses on the clinical
course of suicidal behaviour with respect to aetiological and prognostic
factors such as stressors and negative life events, major psychiatric illness
and the effectiveness of interventions, among them Dialectical Behaviour
Therapy. He has also conducted studies of the epidemiology of deliberate
self-harm and completed suicide in the general population and various
non-clinical populations. He has published several empirical papers in
scientific journals, book chapters and books. He has received several
national and international awards in recognition of his work.

Alec L. Miller, Psy.D., is Professor of Clinical Psychiatry and Behavioural


Sciences, Chief of Child and Adolescent Psychology, Director of the
Adolescent Depression and Suicide Program, Director of Mental
Health Services at P.S. 8 School-Based Mental Health Program, and
Associate Director of the Psychology Internship Training Program at
Montefiore Medical Center/Albert Einstein College of Medicine, Bronx,
NY. Dr Miller has published widely on adolescent suicide, Dialectical

348
About the Editors and Contributors

Behaviour Therapy (DBT), and borderline personality disorder, and has


trained thousands of mental health professionals in DBT. He is co-author
of Dialectical Behavior Therapy for Suicidal Adolescents.

Nishi Misra is Scientist ‘D’ at Defence Institute of Psychological Research,


Delhi. She obtained her postgraduation from Allahabad University
and M.Phil. in Medical & Social Psychology from Central Institute of
Psychiatry, Ranchi. She specialises in the area of Clinical Psychology and
has also served as an assessor in Services Selection Board, Allahabad. She
has worked extensively in the area of suicides and fratricides, job stress
and post-traumatic stress in Armed Forces. She is the co-author of the
popular field manual for the officers of the Indian Armed Forces titled
‘Suicide and Fratricide: Dynamics and Management’. She counsels DRDO
personnel as well. She has to her credit number of research projects and
publication’s.

Chad E. Morrow, Psy.D., is currently a captain and psychologist in the


United States Air Force. He is stationed at Maxwell AFB in Montgomery
where he serves as the Chief of the Mental Health Element, Alcohol and
Drug Abuse Prevention and Treatment (ADAPT) program manager,
Traumatic Stress Response program manager, Drug Demand Reduction
program manager and the Suicide Prevention program manager.
Dr Morrow received his doctorate in clinical psychology from La Salle
University, and completed his clinical psychology internship at Wilford
Hall Medical Center. Dr Morrow’s current research involves garnering
empirical support for managing risk/suicide in different settings, for
certain models of suicidal behavior, for the efficacy of integrated primary
care, and for the modification and use of empirically supported protocols
in primary care settings.

Swati Mukherjee is Scientist ‘C’ at Defence Institute of Psychological


Research, Delhi. She has obtained her M.Phil. degree from Delhi University.
She has been a gold medalist throughout her undergraduate and post-
graduate study in applied psychology. She is involved in many major
research projects of the institute. She has to her credit a few research
publications. She has edited a volume on ‘Recent Development in Psy-
chology’ and has published a manual on ‘Suicide & Fratricide’. She is cur-
rently working in the area of personality assessment for personnel selection.
Her areas of interests are Social Psychology and Peace Psychology.

349
Suicidal Behaviour

Pritha Mukhopadhyay is Professor in the Department of Psychology at


University of Calcutta, teaching both postgraduate and M.Phil. (Clinical
Psychology) courses and regularly renders services to the community as a
mental health professional. She has research experience of above 25 years in
clinical psychology from psychophysiological and psychoendocrinological
perspectives. She has published over 40 research papers in national and inter-
national journals and has authored four book chapters. She was awarded
Fulbright post doctoral fellowship and was fellow in Indian Council of
Medical Research (ICMR) projects, and has conducted UGC, All India Council
for Technical Education (AICTE) and Department of Science and Tech-
nology (DST) projects for research programmes, including intervention in
problem behaviour of children, using biofeedback in clinical population
and industrial personnel along with detection of neuropsychological
impairment in psychiatric disorders through neuroimaging and
neuropsychological techniques. Her recent research interest is on academic
readiness in children from Piagetian perspective and shared dysfunction in
Autistic sibship. She has supervised Doctoral thesis on neuropsychological
research in brain lesions and psychiatric illness, endophenotypes in
obsessive compulsive disorder and psychopathology in personality
disorders including psychosocial and neuropsychological approaches.

Latha Nrugham is currently working as Researcher with the National


Centre for Suicide Research and Prevention at the Institute of Psychiatry,
University of Oslo, Norway. She is also a Research Fellow at the Norwegian
University of Science and Technology (NTNU), pursuing her Doctorate
in Philosophy (PhD) in community medicine on attempted suicide
among Norwegian youth. She completed her M.Phil. in Psychiatric Social
Work from the National Institute of Mental Health and Neurosciences
(NIMHANS), India in 1994 and Master of Arts (MA) in Social Work from
the Tata Institute of Social Sciences (TISS), India in 1988 after finishing
her graduation in science from the University of Mumbai in 1986.
She has worked as Project Social Worker and Researcher with the field-
projects of TISS and University of Mumbai, and has been a Fellow of Child
Relief and You (CRY) and Consultant to Department for International
Development (DFID) in India. She has also worked as Clinical Social
Worker in hospital and outpatient settings in India and Norway. She has
published empirical papers in scientific journals and book chapters on
attempted suicide.

350
About the Editors and Contributors

Rory C. O’Connor is Professor of Psychology at University of Stirling and


a chartered health psychologist. He leads the Suicidal Behaviour Research
Group at Stirling which is the only dedicated suicide research group in
Scotland. He has an international reputation for his work on the risk and
protective factors associated with suicide and self-harm. Over the past
14 years he has co-authored some 70 peer-reviewed publications on the
psychological factors associated with suicide, self-harm and well-being.
He is the co-author of the book Understanding Suicidal Behaviour and is
one of the co-editors of the forthcoming book the International Handbook
of Suicide Prevention: Research, Policy and Practice (Wiley Blackwell). He is
a member of the American Association of Suicidology, the International
Academy for Suicide Research and the International Association for
Suicide Prevention. He is also the UK National Representative of the
International Association for Suicide Prevention (IASP).

Peter Osvath, MD, PhD, is an associate professor, psychiatrist and


psychotherapist. He presently works in the Department of Psychiatry and
Psychotherapy, University of Pecs, Hungary. He is also a senior lecturer
in graduate/postgraduate education at the university, and participant in
many collaborative studies on suicidal behaviour.

Gaspare Palmieri, MD, is a psychiatrist and psychotherapist at Casa di


Cura Villa Igea, Modena, Italy. He has been involved in studies dealing
with the Reasons for Living Inventory in various clinical and non-clinical
populations. He is engaged in research projects dealing with suicide risk
among psychiatric patients.

Maurizio Pompili, MD, psychiatrists and psychotherapist, is Professor of


Suicidology, II Faculty of Medicine, Sapienza University of Rome, Italy.
He is part of the community of the McLean Hospital—Harvard Medical
School, USA. He is a dedicated researcher of suicidology with more that
200 scientific publications in the field. He was the recipient of the American
Association of Suicidology’s 2008 Shneidman Award for outstanding early
career contributions to Suicidology. He is the National Representative to
the International Association for Suicide Prevention.

Amri Sabharwal is presently working as a Research Investigator at the


Defence Institute of Psychological Research, New Delhi. She completed

351
Suicidal Behaviour

her M.Sc. degree in Cognitive Neuropsychology from the University


of Essex, UK in 2006, with a dissertation on the dual process theory of
recognition memory. For her Bachelor’s degree from Osmania University
in 2005, she was awarded the gold medal for best overall performance in
psychology. She has been involved in defence-related research pertaining to
mental health, personality assessment and test development for personnel
selection, as well as other non-defence projects on cerebral lateralisation
and emotion recognition. Her area of research includes Cognitive Psy-
chology and Neuropsychology.

Laura Sapienza, MD, is a resident in psychiatry at II Faculty of Medicine,


Sapienza University of Rome, Italy. She is a member of the International
Association for Suicide Prevention. She is involved in suicide prevention
among psychiatric outpatients. She carried out researches dealing with
suicide risk among military personnel, psychiatric patients and non-
clinical populations.

Eva Schaller is psychologist and cognitive-behavioural orientated psy-


chological psychotherapist. At the State Mental Hospital Bayreuth,
Academic Hospital for Psychiatry of the University of Erlangen-Nuremberg,
she works at a unit for major depression, and also in outpatient care for
patients suffering form depressive and bipolar affective disorders. She
was born in Muenchberg, Bavaria, Germany. She studied psychology at
the University of Wuerzburg; after graduation she worked as a research
associate at the professorship for clinical psychology at the University of
Wuerzburg. Her focuses were anxiety disorders, especially specific phobias
and panic disorders. Her research topics are treatment of depressive
disorders, suicidal behaviour and outpatient care.

Roberto Tatarelli, MD, is Full Professor of Psychiatry and Chairman of


the Residency Training in Psychiatry at II Faculty of Medicine, Sapienza
University of Rome, Italy. He is the chairman of the Department of
Neuroscience at Sant’Andrea Hospital in Rome, Italy. He is a leading
expert on suicide and full member of the International Academy of Suicide
Research. He is author of more than 400 scientific publications.

Jitendra Kumar Trivedi is a Professor in the Department of Psychiatry,


Chhatrapati Shahuji Maharaj Medical University, Uttar Pradesh (formerly

352
About the Editors and Contributors

King George Medical University, Lucknow), India, since 1995. He has


taken the charge of Head, Department of Psychiatry recently. He has
done his medical graduation (MBBS) from the King George’s Medical
College, Lucknow University, Lucknow in 1973 and later MD in Psychiatry
from the same institute in 1977. He was appointed in the faculty in the
Department of Psychiatry, K.G. Medical College, Lucknow in 1978 and
since then is working at the same institution in various positions. He has
been a principal investigator for more than 25 multinational clinical trials
as well as ICMR and WHO sponsored projects. He has more than 200
publications in national and international journals as well as chapters in
books. He has made presentations on various aspects of mental health in
both national and international conferences. He has been the editor of
Indian Journal of Psychiatry (IJP) for six years and has been associated with
IJP for more than 18 years in various capacities. He was President of Indian
Psychiatric Society for the year 2004. He has been Zonal Representative
for Southern Asia—Zone-XVI of World Psychiatric Association from
year 2005 to 2008.

Kimberly A. Van Orden, MS, is a doctoral candidate in clinical psychology


at Florida State University and is completing her pre-doctoral internship
at Montefiore Medical Center/Albert Einstein College of Medicine, Bronx,
NY. Kim has a primary interest in suicide risk assessment and has co-
authored numerous papers on suicidal behaviour. She is also a co-author
of The Interpersonal Theory of Suicide: Guidance for Working with Suicidal
Clients. She received the American Association of Suicidology’s Student
Research Award, a Dissertation Research Award from the Melissa Institute
for Violence Prevention and Treatment, an APF/COGDOP Graduate
Research Scholarships, and a Scholar Award from P.E.O. International.

Sannidhya Varma is currently pursuing his Master’s degree in Psychiatry


from the Department of Psychiatry, CSM Medical College, Lucknow,
India. He has done his MBBS from Government Medical College and
Hospital, Chandigarh, and has been a meritorious student throughout.
He is academically and clinically oriented and has been active participant
in various zonal and national conferences. During his tenure of residency,
he has got training in the fields of general adult, child and adolescent,
geriatric, sex clinic and de-addiction psychiatry.

353
Suicidal Behaviour

Viktor Voros, MD, psychiatrist, Assistant Professor at the Department


of Psychiatry and Psychotherapy, University of Pecs, Hungary. Besides
clinical work, he is also a lecturer in graduate and postgraduate education
at the university, and participant in many collaborative studies on suicidal
behaviour.

Manfred Wolfersdorf, Professor, Dr med. Dr h. c., is psychiatrist and


psychodynamic orientated psychotherapist. He is medical director of
State Mental Hospital Bayreuth, Academic Hospital for Psychiatry of
the University of Erlangen-Nuremberg, head of the clinic for psychiatry,
psychotherapy and psychosomatic medicine and docent for ethics in
psychiatry at the University of Ulm and for psychiatry at the Stradins
University of Riga, Latvia. Professor Wolfersdorf was born in Amberg,
Bavaria, Germany, studied medicine at the University of Erlangen-
Nuremberg and got his medical education to become a psychiatrist at the
University of Ulm. In 1985 and 1987 he worked at the Institute on Self-
Destructive Behaviour and Suicide Prevention at the USC in Los Angeles,
USA. His research topics are depression treatment, severe depression and
suicidal behaviour.

354
Author Index

Author Index

Aaron, R., 328 Arato, M., 44, 281


Agerbo, E., 154, 306, 308 Arensman, E., 71, 73
Agoub, M., 282 Arensman et al., 2001, 66, 72
Ahrens, B., 257 Aro, H.M., 238
Akiskal, H.S., 32, 256–258, 260, 266 Arsenault-Lapierre, G., 232
Alao, A., 331 Asberg, M., 45, 154
Alda, M., 287 Asgard, U., 281
Alexopoulos, G.S., 161 Ashton, J.R., 176–177
Allan, S., 7 Averill, J.R., 101
Allebeck, P., 235
Allport, 1950, in Hill and Pargament, Baca-Garcia, E., 54, 154
2003, 124 Badro, S., 304
American Foundation for Suicide Bagley, C., 330
Prevention, 183 Baker, D., 332
American Psychiatric Association (APA), Baldessarini, R.J., 49–50, 257, 260, 263,
21, 27, 31–32, 232 265–267
Amsel, L., 58 Baldry, A.C., 330
Amsterdam, J.D., 266 Bale, E., 65
Amundsen, A., 302 Ballard, C.G., 112
Andreasson, S., 235 Banaji, M.R., 218
Angst, F., 283 Bancroft, J., 65
Angst, J., 256–257, 283 Bandura, A., 101
Anguelova, M., 51 Banki, C.M., 44
Antonovsky, A., 304 Barbe, R.P., 157
Apospori, E., 113 Barkey, K., 122
Appelby, L., 282 Barnes, R., 233
Applebaum, S., 65 Barraclough, B., 30, 32, 140, 154,
Appleby, L., 140, 231 176–177, 232–233, 236, 257, 260
Appolonio, K.K., 20 Barraclough, B.M., 278–281, 285
Apter, A., 92, 109, 281 Basu, J., 219
Arango, V., 45–46 Basu, S., 195

355
Suicidal Behaviour

Bauer, M.S., 267 Borowsky, I.W., 112


Baumeister, R.F., 6–7, 65, 211 Boyd-Barrett, O., 179
BBC, 179 Boyer, R., 34
Bear, D.M., 98 Braham, P., 179
Beary, M.D., 242 Braucht, G.N., 68
Beautrais, A.L., 154, 158, 182, 230, 232, Breault, K.D., 122
234, 242, 325, 329 Breed, W., 92
Beck, A.T., 12, 32, 194, 197, 199, 201–202, Brent, D.A., 50, 56, 154–155, 158, 160,
211, 213, 215, 231, 234, 284, 288 163, 231, 245, 281–282
Becker, E.S., 57 Breslow, N.E., 232
Bellack, K., 219 Brezo, J., 53–54
Bellivier, F., 52 Bridge, J.A., 160, 329
Ben, C.D., 47 Brisman, J., 243
Benazzi, F., 32, 258 Bristowe, E., 95
Berk, M.S., 195 Bronfenbrenner, V., 95
Berkman, L.F., 231 Brook, R., 304
Berkowitz, L., 100, 182 Brown, G.K., 34, 66, 73, 153–154, 156,
Bernal, G., 336 193, 316
Bernal, M., 280, 282, 284 Brown, G.L., 45, 195
Bernasco, W., 94 Brown, J., 29, 157
Bertolote, J.M., 231, 278–279, 282 Brown et al., 2000, 32
Beskow, J., 281 Brown G.K., 194, 198, 208
Beuhring, T., 112 Bruce, M.L., 161
Bhattacharya, S., 176 Brunswick, D.J., 266
Bhugra, D., 111, 119 Bryan, C.J., 20
Biddle, L., 141 Bryan, C.J., 21, 28, 34–37
Biegon, A., 48 Buchanan, A., 158
Bille-Brahe, U., 75 Bucik, V., 238
Birckmayer, J., 246 Buka, S.L., 231
Biswas, S., 111 Bulik, C.M., 258
Björkqvist, K., 183 Bunney, W., 44
Black, D.W., 284, 307 Burkhardt, M.A., 124
Blair-West, G.W., 283 Burstein, A.G., 21
Blakely, T.A., 306 Burt, V.K., 94
Block, J., 101 Busch, K.A., 156
Blumenthal, S.J., 161 Butler, G.K.L., 237
Boehm, S., 215 Buzan, R.D., 20
Boldt, M., 107, 115–116
Bolger, E.A., 304 Caces, F.E., 233
Bolton, C., 9 Canetto, S.S., 136–139
Bond, A., 65 Canli, T., 52
Bongar, B., 21, 26 Carlson, G.A., 263
Borges, G., 231–234 Carrington, P.J., 335

356
Author Index

Carriss, M.J., 68–69, 71 Colman, A.M., 96


Carroll, M.E., 237 Colucci, E., 107, 110–114, 116, 120–122,
Cartensen, L.L., 177 124–125, 127
Caruso, K., 333 Colucci, E.I., 127
Cavan, R., 99 Colucci (2008c), 128
Cavanagh, J.T., 157, 307 Conlon, W., 238
Cebasek-Travnik, Z., 238 Conner, K.R., 156–157, 234, 307
Centre for Disease Control (CDC), Consensus conference. Electroconvulsive
333, 335 therapy, 268
Chabrol, H., 235 Conwell, Y., 154, 156–159, 163, 281
Chagnon, F., 143 Cook, P.J., 155
Chambers, D., 327 Cooper, J.M., 96
Chang, E.C., 67 Cooper-Patrik, L., 223
Chauchard, E., 235 Corbett, S., 242
Chemtob, C.M., 20 Cordess, C., 97
Chen, R., 110 Cornelius, J.R., 241
Chen, T.H.H., 174, 177 Coryell, W., 282–283, 285, 287
Chen, Y.W., 264 Coulter, P., 175
Cheng, A.T., 264 Craft, L., 161, 333
Cheng, A.T. (the year is mismatching), Crane, C., 72
280–281 Crosby, A.E., 154
Cheng, A.T.A., 137, 174, 177, 327 Crouter, A., 95
Chengappa, K.N., 242 Crumbaugh, J.C., 124
Chenier, T., 233 Cull, J.G., 198, 203
Chia, B.H., 334–335 Currier, D., 3
Chiles, J., 65 Curtin, L., 199, 203
Chioqueta, A.P., 307, 315
Chithiramohan, R.N., 112 Dahlberg, L.L., 91
Chow, Y.W., 329 Dahm, P.F., 70
Chynoweth, R., 281 Daly, M., 97
Ciarrochi, J., 12 Darke, S., 235–236, 240
Ciffone, J., 200 Davidson, L.E., 174
Cipriani, A., 316 Davis, L., 240
Cittadini, A., 262 Day, N.E., 232
Clark, D.C., 34 Deakin, B., 47
Clum, G., 3 Dean, P.J., 7
Clum, G.A., 67–68, 70–71, 199, 203, 238 de Catanzaro, D., 159
Coccaro, E.F., 48, 237 Dein, S., 119
Cohen, A.R., 101 de Lara C., 51
Cohen, D., 95 De Leo, D., 107–108, 156
Cohen, Y., 109 De Luca, V., 51–54
Cohen-Sandler, R., 154 Dembo, R., 237
Collins, J.B., 95 Demotes-Mainard, J., 266

357
Suicidal Behaviour

Department of Health of the Republic Eronen, M., 94


of China, 325 Eshun, S., 112–113, 121
De Pisa, E., 262 Eskin, M., 110
Dervic, K., 121, 154, 157–158, 195 Etzendorfer, E., 173–174
De Simoni, E., 256 Etzersdorfer, E., 335
Desjeux, G., 308 Evans, J., 67
D’Haenan, H., 46 Everall, R.D., 115
Dhar, S., 195 Exline, J.J., 121
Dieserud, G., 67, 69, 72, 156 Exner, J., 220
Dilsaver, S.C., 259, 264, 288
Dogra, A.K., 195 Fagg, J., 65
Dollard, J., 100 Fagiolini, A., 258
Domino, G., 113 Falcone, I., 256
Dong, Q., 330 Faravelli, C., 256
Donnan, S., 176–177 Farberow, N.L., 115
Donnerstein, E., 183 Farrell, M., 230
Donovan, J.M., 237 Farrington, A., 236
Doob, L.W., 100 Faupel, C.E., 122
D’Orio, B., 236 Faust, V., 280
Dorpat, T.L., 281 Fawcett, J., 20, 34, 44, 285, 287, 307
Douglas, J. D., 115–116 Fekete, S., 136, 138, 141, 144–145,
Dube, S.R., 329 147–148
Duberstein, P.R., 154, 156, 158, 162, 234 Feldman, B.N., 21
DuRand, C.J., 94 Feldman, M., 220
Durkheim, as Boldt (1988), 115 Felts, W.M., 233
Durkheim, E., 99, 109, 127, 136, 300, 304 Ferguson, T.J., 182
Dusevic, N., 107 Fergusson, D.M., 182, 235
Dwivedi, Y., 53, 55 Ferracuti, F., 98
D’Zurilla, T., 66 Feshbach, S., 100
D’Zurilla, T.J., 66–67, 81, 334 Fetzer Institute, 124
Fiedorowicz, J.G., 259
Eaton, W.W., 329 Firestone, L.A., 212
Eddleston, M., 34 Firestone, R.W., 212
Egeland, J.A., 29 Fishman, G., 178
Eggert, L.L., 159 Flamenbaum, R., 7
Ehrlich, S., 262 Flett, G.L., 13, 156, 217–218
Elias, N., 200 Flitcraft, A., 95
Ellis, T.E., 199 Flouri, E., 158
Engel, G.L., 4 Folger, R., 101
Engstrom, A., 236 Fontana, A., 309
Erdõs, B.M., 144 Fortune, S., 332
Erikson, E., 145 Foster, T., 282
Erlangsen, A., 158 Foster, V.A., 20

358
Author Index

Fountoulakis, K.N., 267 Gooding, P., 72


Fowler, R.C., 230 Goodwin, F.K., 257, 263–264
Francis, L.J., 121 Goodwin, R.D., 243
Frank, E.S., 242 Gorall, D.M., 220
Frankl, V.E., 304 Gould, M., 173
Fredrickson, B.L., 16 Gould, M.S., 152, 157, 159, 161–162, 174,
Freedenthal, S., 21 200, 334–335
Freud, S., 91, 99 Granello, D.H., 327, 330, 333
Fu, Q., 32 Granello, P.F., 327, 330, 333
Fusé, T., 111 Grant, B.F., 240
Fuster, J.M., 45 Greenberg, T., 200
Greenberger, V., 48
Gallo, J.J., 329 Greening, L., 128
Gallop, R.M., 330 Gridina, N., 331
Gamma, A., 283 Griesinger, W., 279
Garbarino, J., 95 Griffin, A., 329
Garlow, S.J., 236 Groholt, B., 327
Gelder, M., 200 Gross, L., 178–179
Gellis, Z.D., 154 Grossoehme, D.H., 122
Genuis, M.L., 95 Grunebaum, M.F., 264
Gerber-Werder, R., 283 Guetzloe, E.C., 328
Gerbner, G., 178–179 Gundlach, J., 178
Gergen, K., 332 Gururaj, G., 326
Gershon, S., 242 Gutheil, T.G., 123
Giancola, P.R., 237 Gutierrez, P.M., 111–112
Gibb, S.J., 264 Guy, J.D., 21
Gibbons, R.D., 160 Guy, S.M., 237
Gil, A., 113 Guze, S.B., 283
Gil, A.G., 112–113
Gilbert, P., 7 Haaga, D.A., 195
Gilboa-Schechtman, E., 304 Haas, G.L., 280
Gill, W.S., 198, 203 Haenel, T., 279
Giotakos, O., 304 Hafner, H., 177
Girabet, J., 235 Hall, R.C., 242
Glinski, J., 158 Hall, R.C.W., 280
Goldberg, E., 242 Hall, S., 65
Goldfried, M.R., 66, 81 Hamilton, M., 213
Goldney, R.D., 148 Hammad, T.A., 160
Goldstein, B., 180 Hanau, M., 48
Goldstein, R.B., 193, 284 Handy, S., 112
Goldstein, R.C., 11 Hansburg, H.G., 209
Goldstein, T.R., 157 Harding, A., 332
Goldston, D.B., 194, 198 Harford, T., 233

359
Suicidal Behaviour

Harlow, L.L., 304 Hoaken, P.N.S., 237


Harmatz, M., 21 Hochdorf, Z., 334
Harris, C.E., 278–280 Holden, R.R., 7, 206
Harris, E.C., 30, 32, 140, 154, 232–233, Holding, T.A., 177
236, 257, 260 Hole, G., 279
Harris, T.L., 330 Holland, S., 329
Harriss, L., 65, 71, 155 Hölzer, R., 280
Harrow, M., 32 Home Office, 97
Harter, S., 232 Hong, C.J., 51
Hartless, G., 111, 113 Hong Kong Jockey Club Centre for
Hasin, D., 241 Suicide Research and Prevention
Haw, C., 71, 305 (CSRP), 333–335
Hawton, K., 65–66, 71–72, 136–140, 142, Hong Kong Journalists Association, 185
144, 156, 174–175, 177, 258, 263, Hood, R.W., 124
297, 305, 335 Horesh, N., 237
Hayes, J.A., 121 Horn, O., 302
Hayes, L.M., 154 Horner, S.D., 112
He, L., 51 Horowitz, M., 221
Heeringen, van K., 258, 263 Horwood, L.J., 182
Heide, K.M., 93 Houle, J., 143, 148
Heilä, H., 281 Houston, K., 71
Heisel, M.J., 156, 159, 163 Hoyer, G., 156
Heisel, M.J., 111, 156 Hsiung, S.C., 47
Helmkamp, J.C., 298–299, 308 Hufford, M.R., 329
Hempel, A.G., 96 Hughes, D., 154
Hemenway, D., 246 Hungarian Central Statistical Office, 146
Hendin, H., 325–327 Hunter, E.C., 12, 14
Hendin (1964, cited in Boldt, 1988), 115 Huston, A.C., 181
Hennen, J., 49–50, 267 Huth-Bocks, A.C., 156
Henriksson, M., 122 Hytten, K., 298, 303, 308
Henriksson, M.M., 230, 281
Henriques, G.R., 65 Innamorati, M., 256, 262
Henry, A., 92 Inoue, T., 266
Henry, C., 266 Insel, B.J., 159
Heppner, P., 69 Inskip, H.M., 233
Hershberger, S.L., 158 Ireland, M., 112
Hewitt, P.L., 13, 217 Isaac, M.K., 326
Hill, P.C., 124 Isometsa, E.T., 257–259, 263
Hillman, S.D., 232 Isometsä, E.T., 285, 288
Himmelfarb, N., 309
Hishinuma, E.S., 113 Jacobs, D., 194
Ho, L.W., 52 Jamieson, K.H., 331
Ho, T.P., 331 Jamieson, P.E., 331

360
Author Index

Jamison, K.R., 257, 263 Kennedy, R.D., 299


Jin, S., 113, 125 Kerkhof, A., 65, 75
Jobes, D.A., 16, 155, 162, 195, 198, Kerkhof, A.J.F.M., 71
207, 316 Kerkhof, J.F.M., 94
John, S., 108 Kessler, R.C., 231, 238, 240–241, 256,
Johnson, C.V., 121 264, 278
Johnson, J., 9 Keyes, C.L., 124
Johnson, J.G., 330–331 Kidwai, F., 173
Johnson, S.L., 256 Kim, C.D., 49
Johnson, W.B., 194 Kinderman, P., 4–5
Joiner, T., 6, 159, 238 King, C.A., 330
Joiner, T.E., 3, 29–30, 32, 34, 153, 155, King, E.A., 264
157–158, 196, 218 Kingsbury, S., 112
Joiner, T.E. Jr., 303, 317 Kingsbury, S.J., 162
Joseph, A., 42, 326 Kirmayer, L.J., 111
Joukamaa, M., 94 Kleespies, P., 154
Joyce, P.R., 234 Kleespies, P.M., 20
Judas, M., 262 Klein, P.S., 47
Judd, L.L., 256 Klempan, T., 43
Klimes-Dougan, B., 154
Kaizar, E.E., 160 Klingman, A., 334
Kalafat, J., 200, 333 Knox, K.L., 334
Kaltiala-Heino, R., 158 Kochman, F.J., 261
Kamal, Z., 121 Kockott and Feuerlein (1968), 234
Kamali, M., 46 Kocsis, J.H., 245
Kang, H., 309 Koivumaa-Honkanen, H., 156
Kaplan, A., 125 Koller, G., 240
Kaplan, H.I., 136, 139 Kopp, W. (coming as Knopp in text), 286
Kaplan, J.R., 48 Korn, M.L., 237
Kaplan, K.J., 32 Kotila, L., 154
Karasek, R., 301 Kovacs, M., 214
Karege, F., 47 Kposowa, A.J., 306
Katzman, D.K., 242 Kral, M.J., 108, 110, 125
Kavoussi, R.J., 48 Kramer, R.A., 334–335
Kaye, S., 235 Kranzler, H.R., 244
Kazarian, S.S., 108–109 Kreitman, N., 155
Kazdin, A., 217 Krieger, G., 44
Kazdin, A.E., 156 Krishnan, K.R., 269
Kehoe, N.C., 123 Krug, E.G., 91, 97
Kelemen, G., 144 Kubota, Y., 328
Kelleher, M.J., 327 Kumar, U., 91, 230
Keller, F., 285, 288 Kung, A.C., 280
Kelly, H.H., 101 Kung, H.C., 235

361
Suicidal Behaviour

Kuo, W.H., 329 Loewenthal, K.M., 121


Kupfer, D.J., 242 Lonnqvist, J., 122, 154, 234
Lonnqvist, J.K., 307
Lalovic, A., 48, 55 Lönnqvist, J.K., 238, 288
Lam, T.H., 331 Lopez, J.F., 44
Lau, S., 330 Lopez, P., 258
Lau, T.F., 330 Losada, M.F., 16
Lazerwitz, B., 115 Louma, J.B., 140
Leach, M.M., 113, 122, 125–128 Lozano, R., 91
Lee, C.S., 137, 327 Ludwig, J., 155
Lee, C.T.C., 174 Lund, E., 156
Lee, T.Y., 329–330 Luoma, J.B., 160–161
Leenaars, A.A., 6, 110, 114, 211 Luscomb, R.L., 238
Lehmann, L., 241 Lynskey, M.T., 235
Leibenluft, E., 257
Leiner, A.S., 157 Mabila, J.D., 235
Lesage, A.D., 281 MacLean, P.D., 7
Lester, D., 107–109, 114, 118, 121, 152, MacLeod, A.K., 12–13, 70
234, 260, 288, 304 Madge, N., 65
Lester (1987, cited in Zonda and Lester, Mahon, M.J., 298, 308–309
1990), 107 Maj, M., 258
Lesyna, K., 177 Mallya, G., 266
Le Tendre, G., 330, 334 Malone, K.M., 34, 236, 280
Lettieri, D., 231 Maltsberger, J.T., 26
Lettieri, D.J., 67 Mancinelli, I., 125
Leverich, G.S., 258 Mandal, M.K., 91
Levine, J., 242 Mann, J.J., 3, 31–32, 46, 56, 58, 154, 230,
Levine, R.E., 96 236, 245, 262, 280, 307, 316
Lewis, L., 328 Mansell, W., 13
Li, D., 51 Marecek, J., 118
Li, X., 47 Marion, M.S., 128
Liem, M., 95–96 Maris, R., 114
Lin, X., 280 Maris, R.W., 32, 49–50, 65, 115, 138,
Lind, L., 309 152, 196, 200, 245, 307–309
Linehan, M.M., 26, 29, 34, 67, 73, 155, Marker, H.R., 288
199, 205, 209–210, 316 Marneros, A., 269
Link, B.G., 94 Marsella, A.J., 109, 125
Linnoila, M., 46, 237 Marshall, H., 127
Linz, D., 183 Martin, B.C., 101
Lion, J.R., 94, 98, 102 Martin, G., 110–114, 120–122, 127, 157
Lipschitz, D.S., 154 Marttunen, M., 231, 308, 327, 334, 336
Liu, K., 327 Marttunen, M.I., 238
Loeber, R., 93 Marttunen, M.J., 281

362
Author Index

Marusic, A., 238 Miller, N.E., 100


Marzuk, P.M., 93–95, 235 Minnix, J.A., 34
Maser, J.D., 153, 261 Mishara, B.L., 108, 127–128, 143
Mason, B.J., 245 Misra, N., 230
Maxfield, M.G., 95 Mitra, S., 143
Mayer, P., 108, 141 Mittendorfer-Rutz, E., 50, 138–139
Mazza, J.J., 159 Mitterauer, B., 281
McAdams, C.R., 20 Miyamoto, R.H., 113
McAuliffe, C., 65, 75, 78–80 Modestin, J., 286
McAuliffe et al., 2005, 77 Molnar, B.E., 231
McBride-Chang, C., 329 Molock, S.D., 330
Mcbride-Chang, C.A., 324 Montgomery, A.M.J., 237
McCann, U.D., 237 Morgan, H.G., 279
McCormick, R.A., 241 Morrow, C.E., 20
Mc Cracken, L., 241 Mortensen, P.B., 137, 140
McGirr, A., 282, 284–285, 287 Moscicki, E., 43
McIntosh, J.L., 43, 110, 154 Moscicki, E.K., 136
McLeavey, B., 80 Moselhy, H.F., 238
McLeavey, B.C., 66–67, 72, 79, 81 Moskos, M., 158
McLeod, L.D., 206 Motto, J.A., 199
McManus, B.L., 163 Mowrer, O.H., 100
Mehlum, L., 297–298, 300–302, 304–306, Moyer, S., 335
308–309, 311–312, 314–315, 331 Mukherjee, S., 91
Meichenbaum, D., 194 Mukhopadhyay, P., 193
Meloy, J.R., 96 Mulder, R.T., 234
Melton, D.A., 47 Müller, M.J., 283
Meltzer, H.Y., 267 Muller-Oerlinghausen, B., 257
Meng, L., 115 Munchua, M.M., 208
Meninger, K., 230 Murphy, G., 234
Menninger, K., 92 Murphy, G.E., 139, 141, 243
Merchant, C.R., 330
Mercy, J.A., 91 Nace, E.P., 244
Metalsky, G.I., 218 Nahulu, L.B., 113
Metzger, R., 286, 289 Nasrallah, A., 284
Metzner, J.L., 154 National Crime Bureau, Ministry of Home
Michel, K., 141, 147–148 Affairs, Government of India, 42
Miklowitz, D.J., 256 National Crime Records Bureau, 325
Mikulincer, M., 304 National Institute of Aging, the, 124
Miles, C., 283 National Suicide Research Foundation,
Miller, A.L., 152, 158 65, 80
Miller, D.N., 161 Neeleman, J., 230
Miller, I.W., 202 Negron, R., 197, 207
Miller, J., 332 Nekanda-Trepka, C.J.S., 70

363
Suicidal Behaviour

Nelson, R., 298 Parker, G., 221


Nemeroff, C.B., 44 Parker, R.N., 92
Neuringer, C., 67, 79 Patsiokas, A.T., 67, 79, 238
Newman, C.F., 199 Paykel, E.S., 222
Nezu, A.M., 66–67, 81, 334 Payne, S., 138–143
Nga-man Leung, A., 324 Pearson, J.H., 280
Niehus, E., 288–290 Pearson, J.L., 140
Nishiguchi, N., 51–52 Peck, D.L., 99
Nock, M.K., 93–95, 144, 156, 218 Pelkonen, M., 327, 334, 336
Nordentoft, M., 65, 156 Pelles, D.J., 281, 285
Norstrom, P., 230 Penn, J.V., 154
Norstrom, T., 234 Penrod, S., 183
Norwegian Board of Health, 309 Perkins, D.F., 111, 113
Nrugham, L., 297, 329 Perry, J.L., 237
Persad, E., 108–109
O’Carroll, P.W., 24, 66, 194 Peterson, B.S., 262
O’Connor, D.B., 3, 7, 13 Peterson, L., 34
O’Connor, R., 12–13 Petrie, K., 78, 304
O’Connor, R.C., 3, 7, 9–10, 12–16, 211 Petronis, K., 235
O’Doherty, M., 236 Petronis, K.R., 281
O’Donovan, C., 266 Pettit, J.W., 34
Ohberg, A., 233–234 Pfeffer, C.R., 153, 155–156, 161–163, 206
Ohtani, M., 51 Pfennig, A., 44
Oldnall, A., 124 Philipp, M., 283
Olson, D.H., 220 Philips, M.R., 325, 328
Olson, Gorall and Tiesel, 2004, 220 Phillips, D.P., 177
Oquendo, M., 307 Phillips, M.R., 282
Oquendo, M.A., 258, 262 Pihl, R.O., 237
Orbach, I., 154, 222, 304, 334 Pihlgren, H., 285, 334
Ornstein, T.J., 238 Pillemer, K.A., 95
Osbourne, T.R., 119 Pirkis, J., 108
Osman, A., 156, 203 Pirkola, S.P., 233
Osvath, P., 136, 141, 144–145, 147–148 Platt, D.E., 280
Oyefeso, A., 236 Platt, J.J., 218
Platt, S., 155
Paight, D.J., 177 Plutchik, R., 34, 92, 210, 219, 237
Palermo, G.B., 96, 98–99, 102 Pohl, E., 331
Palmieri, G., 256 Pokorny, A.D., 284–285
Paloutzian and Ellison, 1982 cited in Pöldinger, W., 279
Ellison, 1983, 124 Pollock, L.R., 7–8, 65–67, 70
Pandey, G.N., 47, 55 Pompili, M., 256–257, 260–263
Paolucci, E.O., 95 Pope, K.S., 20
Pargament, K.I., 124 Potter, L.B., 34, 212

364
Author Index

Potter-Efron, R.T., 242 Rosenthal, P.A., 157, 159


Powell, K.E., 299 Rosenthal, R.N., 239, 244
Poznanski, E.O., 215 Rosenthal, S., 157, 159
Practice guideline for the assessment Rosovsky, H., 234
and treatment of patients with Rossello, J., 336
suicidal behaviors, 267 Rossow, I., 234, 302
Prescott, D., 95 Rotheram, M.J., 162
Preuss, U.W., 238 Rouillon, F., 257
Price, R.K., 306 Rounsaville, B.J., 240
Prigerson, H.G., 302 Roy, A., 29, 233, 236, 240, 279
Prudic, J., 267 Rubenowitz, E., 32, 158
Pryor, T., 242 Rubenstein, J.L., 158
Purselle, D., 236 Rucci, P., 267, 269
Rudd, M.D., 3, 20–21, 23, 28, 30, 32,
Qin, P., 137–138, 140, 156 34–37, 70–71, 74, 153–155, 205, 238,
245, 303, 312, 316, 333
Rajab, M.H., 70 Ruderman, A.J., 7
Rajkumar, S., 233, 282 Rule, B.G., 182
Raleigh, V.S., 111 Rumbold, B.D., 123
Range, L.M., 7, 108, 125, 128 Runeson, B., 154, 281
Rasmussen, S., 9 Rutz, W., 142, 285, 334
Ravandal, E., 240 Ryan, N.D., 257
Regier, D.A., 260 Ryff, C.D., 124
Renaud, J., 287
Resnick, M.D., 112 Saavedra, J.M., 55
Resnik, H.L.P., 231 Sabbath, J.C., 159
Rew, L., 112 Sabharwal, A., 230
Reynolds, W.M., 155, 161, 204, 214 Sackeim, H.A., 267
Rhimer, Z., 281, 285 Sadock, B.J., 136, 139
Ricaurte, G.A., 237 Saini, S., 109
Rice, J., 257 Sainsbury, P., 279
Rich, C.L., 231, 281 Sakinofsky, I., 65–66, 71–72, 76
Rihmer, A., 261 Sakinofsky, L., 136–138
Rihmer, Z., 137, 140–142, 148, 256–257, Salloum, I.M., 243
261, 264, 334 Samaraweera, S., 157
Rinck, M., 57 Sandin, B., 67, 69, 78
Ripley, H.S., 281 Sanislow, C.A., 154
Roberts, C.R., 110 Sansone, R.A., 157
Roberts, R.E., 110, 156 Santa Mina, E.E., 330
Roberts, R.S., 72 Sapienza, L., 256
Robins, E., 234, 281, 283 Sarason, I.G., 221
Romer, D., 175, 331 Sareen, J., 302
Romero, S., 263 Sargent, P.A., 47

365
Suicidal Behaviour

Sato, T., 259 Shneidman, E., 26


Schaller, E., 278 Shneidman, E.S., 6, 26, 304
Schamda, G., 102 Short, J., 92
Schlesinger, L.B., 92 Shroff, S., 143
Schmidtke, A., 65, 144 Siegel, M., 243
Schneck, C.D., 257 Sil, M., 195
Schneider, B., 278, 282–284, 288 Silveira, W.R., 112
Schotte, D., 3 Silverman, M.M., 24–25
Schotte, D.E., 67–68, 70–71 Simkin, S., 65
Schreurs, P.J.G., 77 Simon, R.I., 154–155
Schuckit, M.A., 233 Simon, T.R., 35
Schulsinger, F., 50 Sirota, P., 304
Schwebs, R., 300, 309, 311, 314–315 Sitharthan, T., 245
Scmidtke, A., 177 Skog, Ole-Jorgen, 234
Scott, T.A., 175–176 Slama, F., 263
Scoville, S.L., 298–299, 308, 316 Slee, N., 66, 73
Sears, R.R., 100 Sloan, K.L., 242
Segal, M., 48 Smith, E.G., 160
Séguin, M., 287 Smith, forthcoming, 156
Sentell, J.W., 299 Smith, M.D., 92
Serpi, T.L., 233 Snyder, C.R., 216
Serretti, A., 51 Soares, J.C., 262
Sexson, S.B., 128 Sokero, T.P., 285
Shaffer, D., 154, 158, 161, 163, 200, Somers-Flanagan, J., 36
281, 333 Somers-Flanagan, R., 36
Shaffi, M., 281 Sorenson, S.B., 111
Shafii, M., 231 Sorri, H., 122
Shafran, R., 13 Southwick, S.M., 302
Shafranske and Malony (1996), cited in Speilberger, C.D., 209
Johnson and Hayes, 2003, 119 Spirito, A., 109
Shahar, G., 307 Spivack, G., 218
Shaltiel, G., 52 Springer, L.S., 122
Shapiro, J.R., 159 Stack, S., 122, 156, 158, 173–175,
Sharma, R., 288 177–178, 233, 331
Sharma, V., 266 Stambolic, V., 47
Sharp, S.F., 110 Stanger, J.D., 331
Shea, S.C., 155, 162 Stanley, B., 30, 154
Sheehy, N.P., 13 Stanton, J., 124
Shek, D.T.L., 327 Stanton, R., 279
Shen, H., 111 Stark, E., 95
Sher, L., 245 Steadman, H.J., 94
Sherbourne, C.D., 211 Steen, M.D., 141
Shiang, J., 110 Steer, R.A., 202, 213

366
Author Index

Steiner, B., 284, 286, 289 Theorell, T., 301


Stepakoff, S., 157 Thibaut, J.W., 101
Stewart, A.L., 211 Thomas, N., 112
Stewart, S.M., 329 Thompson, S., 331
Stiles, T.C., 307, 315 Thoresen, S., 298–299, 301–303,
Stock, S.L., 242 305–306, 308
Stoppelbein, L., 128 Tierney, R.J., 334
Strakowski, S.M., 259 Tondo, L., 241, 257, 260, 263–264
Strauss, J., 329 Toros, F., 330
Stravynski, A., 34 Tortolero, S.R., 110
Strickland, P.L., 49 Townsend, E., 71, 73
Striegel-Moore, R.H., 242 Trivedi, J.K., 42
Strohbach, D., 57 Trovato, F., 110
Strosahl, K., 65, 156 Tsai, S.J., 51
Strube, M.J., 101 Tse, W.L., 330
Stuss, D.T., 42 Tseng, W.S., 108–109, 111
Substance Abuse and Mental Health Tsutsumi, A., 301
Services Administration, 241 Turecki, G., 43, 51, 287
Suicide Crisis Intervention Centre Turvey, C.L., 34, 156–157
(SCIC), 336 Tyrer, P., 73
Suicide Prevention Resource Center
(SPRC), 21 UC Atlas of Global Inequality, 42
Suokas, J., 234 Underwood, L.G., 124
Suominen, K., 264 Underwood, M.D., 45
Suris, A., 309 Unwin, C., 302
Sussex, J.N., 29 U.S. Food and Drug Administration
Swahn, M.H., 34 (FDA), 265
Swanson, J.W., 94 US Food and Drug Administration
Swinton, J., 124, 128 Public Health Advisory, 160
Szadoczky, E., 264
Vaglum, P., 240
Tabachnick, B.G., 20 Valtonen et al., 2005, 257
Takahashi, Y., 114 Valtonen et al. (2008), 259
Tanney, B.L., 43 van Heeringen, K., 4
Tarakeshwar, N., 124 Van Orden, K.A., 31, 152, 159, 312
Tarrier, N., 72 Van Praag, H., 237
Tassava, S.H., 7 Van Praag, H.M., 92
Tatarelli, R., 256 Varma, S., 42
Taylor, K., 72 Varnik, A., 138–139
Taylor, W.D., 262 Vega, W.A., 112–113
Teresi, J.A., 124 Velting, D.M., 162, 200
Thase, M.E., 243 Videtic, A., 51
Thatcher, W.G., 112 Vijayakumar, L., 108, 233, 282, 329

367
Suicidal Behaviour

Violato, C., 95 Wickramaratne, P.J., 257


Virkkunen, M., 46 Widom, C.S., 95
Vogel, R., 285, 289–290 Wiederman, M.W., 242
Volavka, J., 94 Wilcox, H.C., 233
Voros, V., 136 Williams, J.M., 71
Vörös, V., 144–145 Williams, J.M.G., 3, 6–8, 11–12, 65–67,
Vuorilehto, M., 286 70, 237
Willner, P., 11
Waern, M., 156 Wilson, A., 220
Wagner, B.M., 157, 317 Wilson, M., 97
Waldrop, A.E., 329 Wilson, Passik and Kuras, 1989, 220
Walker, R.L., 315 Wilson, S.T., 195
Walters, E.E., 231 Wingate, L.R., 30, 36, 154
Warheit, G., 113 Winkel, F.W., 330
Warheit, G.J., 112–113 Winokur, G., 284
Wasserman, D., 54, 234 Wintemute, G.J., 35
Wasserman, I.M., 177, 298 Wittchen, H.U., 278
Wasserman, L., 122 Witte, T.K., 154
Waternaux, C., 280 Wolfersdorf, Hole et al., 1990, 288–289
Watt, T.T., 110 Wolfersdorf, M., 278–281, 283, 285–286,
Webb, D., 119 288–290
Weiger, W.A., 98 Wolfgang, M., 98
Weimann, G., 178 Wong, A., 302, 308
Weiner, R.D., 268 Wong, B., 329
Weisaeth, L., 298, 302–304, 308 Worden, J.W., 211
Weissberg, M.P., 20 World Health Organization (WHO),
Weissman, A.D., 211 107, 183, 185–186, 230, 263
Weissman, M.M., 238 Woznica, J.G., 159
Wender, P.H., 50 Wright, J.C., 181
Wen-Hung, K., 156
Westermeyer, J., 96 Yaeger, D., 309
Westreich, L., 239, 244 Yahoo News, 332
Wetzler, S., 329 Yamada, A.M., 123
Whiteford, H., 108 Yang, B., 68, 71
Whitters, A.C., 233 Yazdani, A., 127
WHO, 91 Yerevanian, B.I., 265, 267
WHO Global Consultation on Violence Yip, P.S., 175, 327–328, 331, 335
and Health, 91 Young, E.A., 282–283, 285, 287
WHOQOL SPRB Group, 124 Ystgaard, M., 314, 330
WHO (World Health Organization) Yuen, N.Y., 113
Mortality Database, 324–325, 327 Yui-chi Fong, C., 324

368
Author Index

Zahl, D.L., 65 Zillmann, D., 101


Zeng, K., 330, 334 Zimmerman, R., 113
Zhang, J., 113, 125, 327, 329 Zimmerman, R.S., 112–113
Zhang, X., 139 Zonda, T., 107
Ziaian, T., 108 Zorko, M., 238
Zill, P., 52 Zwi, A.B., 91

369
Suicidal Behaviour

Subject Index

Acceptance-based behavioural therapy, Anger, 32


76 Antisocial personality disorder, 243
Acquired immune deficiency syndrome Anxiety, 32
(AIDS), 154 Anxiety disorders and substance abuse,
Active handling, 78 243
Active suicidal episode, 23 Apolipoprotein E4 (APOE4) gene, 55
Acute periods, of suicidal crises, 23 Asian adolescents, suicidal deaths
Acute risk, for suicide, 28–29, 31–34 patterns and trends, 112, 324–328
Adolescents and suicidal behaviour, preventive measures, 333–336
110–114. see also Asian adolescents, risk factors, 328–332
suicidal deaths warning signs for attempts, 332–333
predictors of death by suicide, Attitudes, towards suicide, 113
154–155 Attitude towards Youth Suicide scale
Adult Dispositional (Trait) Hope Scale, (AtYS), 121
216 Availability of means, 34–35
Adult Suicidal Ideation Questionnaire Avoidant personality disorder (PD), 243
(ASIQ), 197, 204
Affective disorder and substance abuse, Baseline risk, for suicide, 23, 28–31
240–242 Baumeister’s model of suicide, 5–6
African Americans, suicide rates among, Beck Depression Inventory (BDI), 198,
110 213
risk factors, 112 Beck Depression Inventory-II (BDI-II),
Age groups, of suicide victims, 42 198
Aggressiveness, as a predisposing factor, Beck Hopelessness Scale (BHS), 12,
100–101, 154 197–198, 215
and substance abuse, 237 Beck Scale for Suicide Ideation (BSSI),
A1438G variant, 52 202
Alcohol abuse and suicidal behaviour, 43, Behavioural contagion theory, 176
94, 113, 233–234, 240–242 Benzodiazepines, 147
Alcohol intoxication, 234 Bereavement, 154
Amok, 96–97 ‘Beyond the Front’ interactive video, 175

370
Subject Index

Binge eating disorder, 242 Child Suicide Potential Scale (CSPS), 206
Biopsychosocial model, of suicidal Cholecystokinin (CCK) gene, 55
behaviours, 4–6, 22–24 Chronic suicidality, 36
fluid vulnerability theory, 23 Chronic suicide, 231
importance to clinicians, 22–23 Chronological Assessment of Suicide
Bipolar affective disorder, 43 Events, 162
age of onset, 257 Clinical assessment, 194–195
antidepressant therapy, 266 Clinical responses, to risk levels, 37
antipsychotic therapy, 267 Clinical work, with suicidal patients,
depression, 257–260 26–27
electroconvulsive therapy (ECT), Clozapine, 267
267–268 Cocaine dependence and suicidal
lithium and mood stabilizer therapy, behaviour, 235–236
264–266 Cognitive-behavioural therapy (CBT)
prevalence, 256 interventions, 72–73
psychosocial interventions, 268–269 Cognitive behaviour therapy, 12
risks of premature mortality, 257 Cognitive flexibility impairment, 238
with substance abuse disorder, Cognitive impairment, due to substance
260–262 abuse, 238
suicide attempts, 262–264 Cohesion, impact on suicide prevention,
suicide risks, 257–259, 288 313–314
Collaborative Assessment and Man-
Borderline personality disorder (BPD),
agement of Suicidality programme
43, 73, 94, 243
(CAMS), 16, 162, 195, 316–317
Brady Act, 155
Collective identity, 298
Brain cancer, 154
Collective society, 298
Brain-derived neurotrophic factor
Columbia University Teen Screen
(BDNF), 53
Program, 161
Bulimia nervosa, 242
Co-morbidity, 43
Completed suicide, 231
Cannabis abuse/dependence and suicidal
Contents, in risk assessment
behaviour, 234–235 differential manifestations of risk,
Carbamazepine, 147 157–159
Case behaviour, of suicidal behaviour, factors unique to adolescents and
232 children, 160
Catechol-O-methyltransferase (COMT), risk factors, 153–155
53 warning signs, 155–157
Child and Adolescent Perfectionism Scale Coping behaviour, 69
(CAPS), 218 Copycat suicide, 174
Children’s Depression Inventory (CDI), Criminals and suicidal behaviour, 94–95
214 Cry of pain (CoP) model, of suicidal
Children’s Depression Rating Scale behaviour, 7–9
(CDRS-R), 215 communications from suicidal
Children’s Hope Scale, 216 patients, 9–11

371
Suicidal Behaviour

Cultivation theory, 178–180 stress and, 67–69


Cultural factors, of suicidal behaviour treatment outcomes and follow-up
Asian countries vs Western countries, measures, 81–82
108 Depression, role in suicidal behaviour,
Chinese women, 143 4–5
comparisons across Italian, Indian, CSF 5-HIAA level, 45
and Australian youths, 116–119 epidemiology, 283
discourses on meaning, 114–116 female depressive suicides, 141
ethnic differences, 110–114 Hawaiian students, 113–114
gender differences, 108. see also gender multiple suicide attempters, 30
differences, in suicidal behaviour during periods of acute
impacts, 107–110 symptomatology, 32
psychopathology, effects on, 109–110 psychopathology, 288–289
in risk assessment, 122–125 risk factors, 284–288
spirituality/religion, impact of, role of neurotransmitter serotonin
119–122 (5-HT), 45
Cyanoimipramine, 46 ST deficiency, 46
Cyclic AMP responses, 54–55 suicide prevention in, 289–290
Depressive-mood disorder, 43
Defeat Scale, 211 Deterrents, to suicidal behaviour, 33
Deliberate self-harm (DSH) Developmental issues, of suicide risk
association with optional thinking ability, assessment. see risk assessment, of
79–80 suicide
association with problem-solving Diagnostic and Statistical Manual-IV
process, 66–67, 69 (DSM-IV), 94
cognitive characteristics, 67 Dialectical Behaviour Therapy, 73
coping difficulties, 76 Diathesis, role in suicidal behaviour,
factors common to repeaters, 76 56–57
hopelessness and, 69–71 Diathesis-stress-hopelessness model, of
influence of early life experiences, suicidal behaviour, 68
68–69 Dopamine genes, synthesis of, 53
male repeaters, 76–77
methodological problems with iden- Eating disorder and substance abuse,
tification of risk factors, 74–75 242–243
motives for repeated, 75–76 Ego Function Assessment Scale-Modified
optional thinking ability and, 79–80 (EFA-M), 219
passive and avoidant problem Ego Function Test (EFA) test, 198
orientations, 77–79 Entrapment model, of suicidal behaviour,
problem-solving orientation among 7–9
repeaters, 77–81 Epigenetic Assessment Rating System
problem-solving style of repeaters (EARS), 220
of, 79 Escape-motivated suicides, 6–7
repeated, 65–66 Escape Potential Scale (EPS), 211

372
Subject Index

European Americans, suicidal behaviour, Generalised anxiety disorder (GAD), 243


113 Genetic factors, in suicidal behaviours, 29
European Parasuicide Study Interview apolipoprotein E4 (APOE4) and
Schedule, 75 cholecystokinin (CCK) genes, 55
Explicit intent, 33 cyclic AMP responses, 54–55
Extended Attributional Style Ques- family, twin, and adoption studies,
tionnaire (EASQ), 218 49–50
Extended suicide, 96 GABAergic and glutamatergic genes,
54
Familicide, 92 hypothalamo-pituitary-adrenal (HPA)
Family Adaptability and Cohesion axis gene dysfunction, 53, 55
Evaluation Scales (FACES IV), 220 markers, 50
Female self-poisoners, 72 mutations in nitric oxide synthase
Fenfluramine challenge test, 48–49 (NOS), 55
Filicide, 92 neurotrophic genes, 53–54
norepinephrine changes, 58
Firearms, suicide attempts using, 35
phoshpoinositide signalling systems,
Firestone Assessment of Self-Destructive
54–55
Thoughts (FAST), 212
serotonergic system, 50–52, 58
Fluid vulnerability theory (FVT), 23
shortcomings in studies, 55
Fratricide, 92
synthesis of noradrenaline and
Frustration–aggression hypothesis, 100
dopamine genes, 53
Future Thinking Task (FTT), 12 Glutaminergic systems, 54
GABAergic gene variations, 54 Glycogen synthase kinase-3-β, in suicide
Gender differences, in suicidal behaviour victims, 47
choice of methods, 139 Gotland project, 200
Group Interpersonal Problem-Solving
deaths using drugs, 147
Skills Training programme (PST), 80
family roles, protective effect of, 141
Guilt, 32
help-seeking behaviour, 141–142
incidence rate, 137 Hamilton Depression Rating Scale
marriage, protective effect of, 140–141 (HDRS), 213
masculinity vs femininity, 139–140 Hamilton Rating Scale for Depression
mental illnesses, 140 (HRSD), 198, 267
occupational factors, 138 Hamlet, 57
pregnancy, protective effect of, 140 Health services, for suicide prevention in
social support, 138 military personnel, 315
socio-cultural aspects, 142–143 Healthy worker effect, 298
socio-economic risk factors, 138–139 Heroin users and suicidal behaviour, 236
suicide attempts, 137, 144–147 Heterogeneity, 194
in therapy, 147–149 Homicide
‘typical’ female suicide attempter, classifications, 92–93
144–145 cultural differences, 97–101
urban living, 138–139 dynamic and interdisciplinary nature
of behaviour in, 92

373
Suicidal Behaviour

prediction based on developmental Language, related to suicidality. see


indicators, 93 standard terminology, related to
theories, 93 suicidality
vs suicide, 93–97 Leadership, impact on suicide prevention,
Homicide–suicide phenomenon, 95–97 309–312
Hong Kong, suicide rates, 175 Lethality, of a suicide attempts, 34–35
Hopelessness, 27–28 role of spirituality in preventing,
assesment of presence, severity and 121–122
duration, 33 Lethality Scales (LS), 211
as a predictor of suicidal behaviour, Levodopa, 53
156 Life events, influence on cognitive ability,
and problem-solving process, 69–71 68–69
SPP and, 14 Life Event Scales, 198
Hopelessness Scale For Children (HPLS), Life Experience Survey (LES), 221
217 Life Style Index (LSI), 219
Hostile aggression, 100 Limbic-hypothalamus-pituitary-adrenal
HPA axis dysfunction, 55 axis (LHPA Axis), relation with
5-HT 2 A receptors, role in suicidal suicidality, 44
behaviours, 47–48 Limbic-hypothalamus-pituitary-thyroid
Hungary, suicide rates, 145–147 axis (LHPT Axis), relation with
Hypothalamic-pituitary-adrenal (HPA) suicidality, 44
axis activity, in suicidal behaviour, 3 Lipid metabolisms, related to suicide, 48
gene, 53 Longitudinal studies, of suicidal
behaviour, 232
Imipramine, 46
Impact of Events Scale (IES), 221 Major depressive disorder, 43
Impulsive multiple attempters, 31 The Making of a Martyr, 180
Impulsive traits, as a predisposing factor, Meaning, of suicide, 114–116
28, 31, 154 Means–End Problem-Solving Procedure,
substance abuse, 237 218
Impulsivity Control Scale (ICS), 210 Measures, of suicide-related behaviours,
Imu, 96 196–199
Incarceration, 154 tools, 201–223
India, suicide rates, 42, 111 Media reporting, of suicides
Instrumental aggression, 100 arousal state, effects on, 183
Intent for suicide, 155 attentional and comprehension pro-
Intimate-partner violence, 95 cesses, effects on, 181
IS PATH WARM mnemonic, 155, 161 attitudes, effects on, 182–183
attributions and moral evaluations,
Jumping Frenchman, 96 effects on, 182
choice of methods, 177
Knowledge, of suicidal behaviours educating the public about risks for
biopsychosocial model of, 22–23 suicide, role in, 184–186
terminology and language, 23–25

374
Subject Index

emotional habituation, effects on, Models


183–184 Baumeister, 5–6
guideline on suicide reporting, biopsychosocial model, 22–23
185–186 clinical implications, 14–16
imitation effects from, 174–178 Kinderman, 4–5
media images, impact of, 180 neurobiological, 56–57
proactive role, 187 stress-diathesis, 58
social learning theory aspect of, 173 Williams and colleagues, 7–9
and suicide rate, 173–174 Modified Scale for Suicide Ideation
television viewing, effects of, 178–180 (MSSI), 197, 202
Meprobamate, 147 Monoamine oxidase A (MAOA), 52
Migration and suicide rates, 111 Motives for parasuicide questionnaire
Military personnel and suicidal behaviour (MPQ), 75
assessment and management, Multi Attitude Suicide Tendency Scale for
316–317 Adolescents, 222
contributing factors, 299–300 Multi-dimensional Hope Scale, 215
due to stringent requirements of Multi-dimensional Perfectionism Scale
military selection, 300 (MPS), 217
easy access to lethal means, 308–309 Multiple attempters, 23, 30
impact of combat trauma on women, Multiple sclerosis, 154
309
involuntary repatriation, 305–306 Negative life events and suicidal
behaviour, 4
lack of meaningfulness, 303–304
Neurobiological factors, for suicide
lack of social support, 304–305
fenfluramine, role of, 48–49
loss of individualism and conformity
5-HT2A receptors, 47–48
pressure, 301
lipid metabolisms, 48
and low levels of social integration,
models, 56–57
300–301
norepinephrine transporters, role
military lifestyle, 302
of, 49
peacekeepers from Norway, 298
transmitter non-specific neuro-
prevention strategies, 309–315
endocrine studies, 44
protective factors, 309–315
transmitter specific neuro-endocrine
psychopathology, 307–308 studies, 45–47
risk factors, 300–309 Neurotrophic genes, 53–54
traumatic events, exposure of, Nitric oxide synthase (NOS), mutations
302–303 in, 55
US Air Force, Army, Marine Corps Non-suicidal self-injurious behaviour
and Navy, 298 (NSSI), 153
Mindfulness-based cognitive therapy, Noradrenaline genes, synthesis of, 53
76 Norepinephrine, 49
Minnesota Multiphasic Personality Norwegian National Strategy for Suicide
Inventory (MMPI), 194 Prevention, 309

375
Suicidal Behaviour

Opiate users and suicidal behaviour, 236 family rigidity and adolescent abilities,
69
Parental Bonding Instrument (PBI), 221 hopelessness and, 69–71
Partoxetine, 46 influence of early life experiences,
Passive-avoidance, 77–78 68–69
Patient–clinician relationship, role in as a maintenance factor, 71–72
risk assessment. see therapeutic negative self-appraisal of, 70
relationship, role in risk assessment psychosocial stress and, 69
Paykel Suicide Items (PSI), 222 sequential model of social, 66–67
Perfectionism, 3 training for repeated DSH, 77–81
socially prescribed, 13–14 as a vulnerability factor for suicidal
Personality disorders behaviour, 67–69
and substance abuse, 240 Protective factors, 34
and violence, 94 Psychache Needs Questionnaire (PNQ),
Phoshpoinositide signalling systems, 198, 208
54–55 Psychiatric illness, and liability for
Physical illnesses, 154 suicide, 43, 279–283
PI3-K/Akt signalling pathway, in suicide Psychological autopsy studies, of suicidal
victims, 47 behaviour, 232
Platelet aggregation response, in suicidal Psychological factors, on suicidal
behaviours, 47–48 behaviour
Positive and Negative Suicide Ideation biopsychosocial model, 4–6
Inventory (PANSI), 203 clinical implications, 14–16
Positive future thinking, 12 cognitive ability, 3
Predictors, of suicidal behaviours, 33 cry of pain (CoP) or entrapment
hopelessness, 156 model of, 7–9
intent for suicide, 155 escape-motivated suicides, 6–7
severe anxiety/agitation, 156 personality traits, 3
sleep disturbances, 157 positive future thoughts, 12–13
social isolation, 156 socially prescribed perfectionism
Predispositions, to suicidality, 28–30 (SPP), 13–14
Preparatory behaviours, 33 spoken communications of suicidal
Prevention of Suicide in Primary- patients, 9–11
care Elderly: Collaborative Trial Psychopathology, effects of culture on,
(PROSPECT), 161 109–110
Problem orientation, defined, 66–67 Psychosis, 32
Problem-solving deficits, 154 Psychosocial stress and suicidal behaviour,
and substance abuse, 238 69, 238–239
Problem-solving process
definition of, 66 Reasons for Attempting Suicide Ques-
discussions, 83–84 tionnaire (RASQ), 206
efficacy in treating self-harm, 72–73 Reasons for Living Inventory (RFL),
197, 205

376
Subject Index

Religion, impact on suicide, 119–122 Scale for Suicide Ideation (SSI), 197, 201
Rescue Scale, 211 Scale for Suicide Ideation-Worst
Resilience, 12 (SSI-W), 197, 201
Reynolds Adolescent Depression Scale-2 Schizophrenia, 43, 94, 140
(RADS-2), 214 Screening for suicide, 161
Risk assessment, of suicide Selective serotonin reuptake inhibitors
acute, 31–34 (SSRIs), 160
APA guidelines, 21 Self-destructive pathways, 4–5
availability of means, 34–35 Self-esteem, 78
baseline, 29–31 Self-harm, 25
content. see contents, in risk Self-Inflicted Injury Severity Form
assessment (SIISF), 212
cultural factors, role of, 122–125 Self-inflicted unintentional death, 25
defining common goal of pain Self-injurious behaviours, 30
remediation, 26–27 Self Injury Implicit Association Test
empirically-supported areas for, 28 (SI-IAT), 218
identification of variables, 34 Self Monitoring Suicide Ideation Scale
process and context of risk, 160–163 (SMSI), 197, 203
risk categories, 35–36 Separation Anxiety Test (SAT), 209
risk levels and clinical responses, 37 Seppuku, 109
significance in clinical decision- Sequential Emotion and Event Form
making, 163 for Suicidal Adolescents (SEESA),
temporal factors influencing, 36 197, 207
therapeutic relationship, role of, Serotonergic system, 50–52
26–27 Serotonin (5-HT), role in suicidal
understanding suicide, 22–25 behaviour, 45–47
Risk categories, of suicidal patients, 35 Severe anxiety/agitation, 156
Risk factors, common across the lifespan, Sexual assault, 95
153–155 Shame, 32
antidepressant prescription as, 160 Siblicide, 92
childhood physical and sexual SLC6A4, 51
abuse, 157 Social isolation, 156
interpersonal conflict, 158 Socially prescribed perfectionism (SPP),
intimate partner violence, 157 13–14
lowered competence, 159 Social problem-solving capacity, in
mental disorders, 157–158 suicidal patients, 3
perceived expendability, 159 Social support, impact on suicide
personality traits, 158 prevention, 314–315
Risk-Rescue Rating Scale (RRR), 211 Sociopathic personality disorder, 43
Rorschach Inkblot test, 194 Spirituality, impact on suicide, 119–122
Rorschach Suicide Index Constellation Spiroperidol, 46
(RSIC), 220 Standard terminology, related to
suicidality, 23–25

377
Suicidal Behaviour

State Hope Scale, 216 Suicide plan, 25


State–Trait Anger Expression Inventory, Suicide prediction, 21
209 Suicide Probability Scale (SPS), 197,
Stress-diathesis model of suicide, 58 203
Stressor, role in suicidal behaviour, 56 Suicide Status Form (SSF), 207
Substance abuse, 43, 94, 113 Suicide threat, 25
aggressiveness, as a predisposing Suttee, 109
factor, 237
barriers to effective preventive man- Temporal factors, of suicide risk
agement, 247 assessment, 36
and cognitive impairment, 238 Thematic Apperception Test (TAT), 194
due to comorbidity of psychiatric Therapeutic relationship, role in risk
illnesses, 239–243 assessment, 26–27
impulsive traits, as a predisposing Therapy interfering behaviours, 26
factor, 237 Tryptophan hydroxylase (TPH), 52
preventive managament, 243–247 Tyrosine hydroxylase (TH), 53
and problem-solving impairment,
238 Undetermined suicide-related behaviour,
San Diego Suicide Study, 231 25
studies exploring the association US Army, suicide rate, 176
between suicide and, 232 US suicide rates, across the lifespan, 152
types of substances, 232–236 Utrecht Coping List (UCL), 77
Suicidal Behaviors Questionnaire Uxoricide, 92
(Revised) (SBQ-14), 210
Suicidal ideation, 33, 36, 57, 112, 195, 232 Violence
in childhood, consequences, 153 defined, 91
and religion, 120–121 in Diagnostic and Statistical Manual-
Suicidal Ideation Screening Questionnaire IV (DSM-IV), 94
(SIS-Q), 223 dynamic nature of, 95
Suicidality, 20 influence of early life, 95
Suicidal tendency, 279 neurochemical basis of, 98
Suicidal thoughts, assessment, 33–34 psychological perspective, 99–101
Suicide, definitions, 25, 91–92 sociological explanations, 98–99
Suicide attempts, 25, 112, 232
in childhood, consequences, 153 Warning signs, of suicidal crises, 155–157
Suicide Behaviours Questionnaire (SBQ), among Asian adolescents, 332–333
197, 209 WHO/EURO Multicentre Study on
Suicide contagion, 159, 174–178 Suicidal Behaviour, 75, 144
Suicide Ideation Scale (SIS), 205 objectives of, 77
Suicide intent, 196 Wihtico psychosis, 96
Suicide Intent Scale (SIS), 208 Willingness to disclose, on suicidal
behaviours, 162

378

You might also like