Hate Crime: Concepts, policy, future directions 9781843927792, 9781843927808, 9781315093109

Hate crime has become an increasingly familiar term in recent times as problems of bigotry and prejudice continue to pos

233 17 2MB

English Pages [285] Year 2010

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
Half Title
Title
Copyright
Dedication
Contents
List of tables, figures and case studies
Abbreviations
Acknowledgements
Notes on contributors
Future developments for hate crime thinking: who, what and why?
Part One: Developing Understandings of Hate Crime
1 The more things change ... post-9/11 trends in hate crime scholarship
2 The victimisation of goths and the boundaries of hate crime
3 Future challenges for hate crime policy: lessons from the past
4 Homophobic hate crime in Northern Ireland
5 Verbal and textual hostility in context
6 Hate crime offenders
Part Two: Developing Responses to Hate Crime
7 Law enforcement and hate crime: theoretical perspectives on the complexities of policing 'hatred'
8 From hate to 'Prevent': community safety and counter-terrorism
9 Hate crime victims and hate crime reporting: some impertinent questions
10 Racial aggravation or aggravating racism: overcoming the disjunction between legal and subjective realities
11 Healing harms and engendering tolerance: the promise of restorative justice for hate crime
Index
Recommend Papers

Hate Crime: Concepts, policy, future directions
 9781843927792, 9781843927808, 9781315093109

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Hate_Crime_PPCv6.qxd

15/4/10

19:29

Page 1

Concepts, policy, future directions Edited by

Neil Chakraborti

Hate crime has become an increasingly familiar term in recent times as problems of bigotry and prejudice continue to pose complex challenges for societies across the world. Although greater recognition is now afforded to hate crimes and their associated harms by academics, policy-makers and criminal justice agencies, the problem is still widespread and many key questions remain unanswered. Are we doing enough to protect vulnerable members of society? Are we doing enough to address the offending behaviour of hate crime perpetrators? Are there better ways of understanding and responding to hate crime?

It provides much-needed ways of taking the ‘hate debate’ forward as well as offering practical suggestions for developing both scholarship and policy in a more progressive manner. This book is written chiefly for students, academics and practitioners studying and working in the areas of ‘race’ and anti-racism; ethnicity; religious, gender and sexual identity; disability; equalities and human rights; victimology; offending behaviour; sociolegal studies; community safety and cohesion; social policy; crime prevention and reduction; policing; criminal justice.

Hate Crime Concepts, policy, future directions

Concepts, policy, future directions

This book brings together contributions from leading experts in the field to address these and other contested issues in this fascinating and often controversial subject area. Drawing upon innovative work being undertaken nationally and internationally, the book offers fresh ideas on hate crime scholarship and policy and in so doing enables readers to re-evaluate the concept of hate crime in the light of fresh research, theory and policy.

Hate Crime

Hate Crime

The editor

Nicole Asquith, Susan Bennett, Neil Chakraborti, Kris Christmann, Marian Duggan, David Gadd, Jon Garland, Nathan Hall, Carolyn Hoyle, Jack Levin, Hannah Mason-Bish, Jack McDevitt, Derek McGhee, Jim Nolan, Barbara Perry, Mark Walters, Kevin Wong.

Academic and Professional Publisher of the Year 2008 International Achievement of the Year 2009

www.willanpublishing.co.uk

Edited by

Contributors

Neil Chakraborti

Neil Chakraborti is Senior Lecturer in Criminology at the Department of Criminology, University of Leicester. He has researched and published widely in the fields of ‘race’, policing, victimisation and hate crime.

Edited by

Neil Chakraborti

Hate Crime

Hate Crime Concepts, policy, future directions

Edited by Neil Chakraborti

First published by Willan Publishing 2010 This edition published by Routledge 2015 2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN 711 Third Avenue, New York, NY 10017

Routledge is an imprint of the Taylor & Francis Group, an informa business © the editor and contributors 2010 All rights reserved; no part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the Publishers or a licence permitting copying in the UK issued by the Copyright Licensing Agency Ltd, Saffron House, 6–10 Kirby Street, London EC1N 8TS. First published 2010 ISBN 978-1-84392-779-2 paperback 978-1-84392-780-8 hardback British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library

Project managed by Deer Park Productions, Tavistock, Devon Typeset by GCS, Leighton Buzzard, Bedfordshire

In memory of Fiona Pilkington and Francecca Hardwick

Contents

List of tables, figures and case studies Abbreviations Acknowledgements Notes on contributors Future developments for hate crime thinking: who, what and why? Neil Chakraborti

ix xi xiii xv 1

Part One: Developing Understandings of Hate Crime 1

The more things change … post-9/11 trends in hate crime scholarship Barbara Perry

17

2

The victimisation of goths and the boundaries of hate crime Jon Garland

40

3

Future challenges for hate crime policy: lessons from the past Hannah Mason-Bish

58

4

Homophobic hate crime in Northern Ireland Marian Duggan

78

vii

Hate Crime

5

Verbal and textual hostility in context Nicole Asquith

6

Hate crime offenders Jack McDevitt, Jack Levin, Jim Nolan and Susan Bennett

99 124

Part Two: Developing Responses to Hate Crime 7

Law enforcement and hate crime: theoretical perspectives on the complexities of policing ‘hatred’ Nathan Hall

149

8

From hate to ‘Prevent’: community safety and counterterrorism Derek McGhee

169

9

Hate crime victims and hate crime reporting: some impertinent questions Kris Christmann and Kevin Wong

194

0 1

Racial aggravation or aggravating racism: overcoming the disjunction between legal and subjective realities David Gadd

209

11

Healing harms and engendering tolerance: the promise of restorative justice for hate crime Mark Walters and Carolyn Hoyle

228

Index

viii

249

List of tables, figures and case studies

Tables 4.1 4.2 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 6.1 6.2 6.3 9.1

Homophobically motivated incidents, crimes and   clearance rates from 2000/01 to 2008/09 All recorded crimes and clearance rates in Northern   Ireland from 2001/02 to 2008/09 Themes of verbal-textual hostility Examples of coding for themes of verbal-textual hostility Hate crime flag Hate crime location x hate crime flag Hate crime offence x hate crime flag Hate crime relationship x hate crime flag Reported and recorded verbal-textual hostility x   selected hate crime offences Themes of verbal-textual hostility x hate crime flag Themes of verbal-textual hostility x selected hate crime   offences Offender characteristics by type of victim targeted Characteristics of hate crimes by motivation of the   offender Culpability of individuals involved in hate crimes by   motivation of the offender British Crime Survey 2008/09: victims’ reasons for not   reporting crime to police

84 84 104 105 106 107 108 110 112 114 117 129 134 137 197

ix

Hate Crime

Figure 9.1

Schematic of hate crime reporting

202

Case studies 11.1 Homophobic abuse 11.2 Racial harassment (indirect mediation) 11.3 Racial harassment



236 238 242

Abbreviations

ACPO BME CDA CJINI CPS CRIS DCLG EHRC ESRC LASI LGBT MDA MLA MPS NIBRS NILT NPT OCSE ODIHR PPS PSNI PVEPF RJ RUC VOM

Association of Chief Police Officers Black and minority ethnic Critical Discourse Analysis Criminal Justice Inspectorate Northern Ireland Crown Prosecution Service Crime Report Information System Department for Communities and Local Government Equality and Human Rights Commission Economic and Social Research Council Lesbian Advocacy Services Initiative Lesbian, gay, bisexual or transgendered Mediated Discourse Analysis Member of the Legislative Assembly (Northern Ireland) Metropolitan Police Service (London) National Incident-Based Report System Northern Ireland Life and Times Neighbourhood Policing Team Organisation for Security and Co-operation in Europe Office for Democratic Institutions and Human Rights Public Prosecution Service Police Service of Northern Ireland Preventing Violent Extremism Pathfinder Fund Restorative justice Royal Ulster Constabulary Victim–offender mediation

xi

Acknowledgements

There are many people to thank for helping me to conceive of and complete this book: typically – and shamefully – I’ll forget to name them all so apologies in advance. First, I owe a considerable debt of gratitude to my colleagues at the Department of Criminology, University of Leicester for giving me the time and space needed to finish the book by covering my duties while I was on study leave. I’m especially grateful to Jon Garland who, despite cursing me repeatedly for disappearing on leave, has been a constant source of ideas, advice and pedantry, so thanks Jon. Thanks also to all of the contributors to this book: given the various demands made of our time, a contribution to an edited book can all too easily slip down the list of priorities so I’m deeply grateful to everyone for producing such excellent chapters. The impetus for this book came from the various exchanges that we and fellow hate crime scholars have had at various conferences and seminars over recent years and I look forward to continuing these exchanges in the years to come. Thanks too are due to Brian, Julia and the rest of the team at Willan Publishing for being such a pleasure to work with; to Tara for being such a pleasure to live with; and to my parents for their courage during difficult times.

xiii

Notes on contributors

Nicole Asquith is a senior lecturer in Applied Criminal Justice Studies at the University of Bradford, and associate research fellow at the Tasmania Institute of Law Enforcement Studies, University of Tasmania. Nicole has worked as a practitioner and researcher in the field of hate crimes since the early 1990s, and published on the role of hate speechtext in hate crimes in a variety of journals and book chapters, and Text and Context of Malediction (2008). Her current research is part of a larger project, which seeks to provide a comparative analysis of hate speech-text regulation in Commonwealth countries. Susan Bennett is a research associate at West Virginia University. Her current research focuses on inner city gangs, hate crimes, and community policing. She received graduate training at Northeastern University in Boston and University of California at Irvine. Neil Chakraborti is a senior lecturer in Criminology at the Department of Criminology, University of Leicester. He has researched and published widely in the fields of ‘race’, policing, victimisation and hate crime and is editor of Rural Racism (with Jon Garland, Willan Publishing, 2004) and author of Hate Crime: Impact, Causes and Responses (with Jon Garland, Sage, 2009). His current research projects include an analysis of police-recorded hate crime data; a study of faith communities’ attitudes towards the police; a systematic review of strategies employed in the policing of gang violence; and an evaluation of public authority responses to targeted harassment and violence. xv

Hate Crime

Kris Christmann is a research fellow at the Applied Criminology Centre, University of Huddersfield. He has been involved in researching racially and religiously motivated crime, especially offenders’ motivations in such offending, the nature and prevalence of racially motivated offending, and the dynamics of hate crime reporting. Other research areas have included examining offender programmes within the criminal justice system aimed at preventing hate crime or programmes looking to eradicate religious and political violence. Marian Duggan is a lecturer in Criminology at Sheffield Hallam University. Her research interests include examining homophobia and hate crimes towards members of lesbian, gay, bisexual and transgender (LGBT) communities in Northern Ireland. Her PhD, undertaken at Queen’s University Belfast, explored Northern Ireland’s homophobic culture and the impact of this on perceptions of lesbians and gay men as ‘acceptable’ victims of hate crime. David Gadd is a senior lecturer in Criminology at Keele University, Staffordshire, where he is currently the Deputy Director for Research in the Social Sciences. David’s research interests are in the fields of offender motivation, desistance, and narrative methodologies. Most of David’s work is concerned with the life histories of people who have committed violent crimes, including domestic abuse and racial harassment. David is well known for advocating for a psychoanalytically informed interpretive approach to criminology, and is the author of Psychosocial Criminology (with Tony Jefferson, Sage, 2007). David is currently editing the Sage Handbook of Criminological Research Methods with Steven Messner (SUNY) and Susanne Karstedt (Leeds University). In 2006 David completed an ESRC-funded project (with Bill Dixon and Tony Jefferson) about the perpetration of racist violence and harassment in North Staffordshire. A book about this research, entitled Losing the Race, has been published by Karnac. Jon Garland is a senior lecturer in Criminology at the Department of Criminology, University of Leicester. He has researched and published extensively in the fields of rural racism, identity, hate crime, policing and victimisation. His books include Racism and Anti-racism in Football (with Michael Rowe, Palgrave, 2001), Rural Racism (co-edited with Neil Chakraborti, Willan Publishing, 2004) and Hate Crime: Impact, Causes and Responses (with Neil Chakraborti, Sage, 2009). His work on hate crime is ongoing. xvi

Notes on contributors

Nathan Hall is a senior lecturer in Criminology and Policing at the Institute of Criminal Justice Studies at the University of Portsmouth. He has extensively researched hate crime, particularly in relation to criminal justice responses in England and Wales and in the United States. His first book, Hate Crime, was published by Willan in 2005, and he is joint editor of Policing and the Legacy of Lawrence, also published by Willan in 2009. Nathan has also recently completed a PhD thesis involving comparative research of the policing of hate crime in London and New York. In addition to working with a number of police services across the country, Nathan has acted in a consultative capacity to the Association of Chief Police Officers (ACPO), Her Majesty’s Inspectorate of Constabulary (HMIC), the Home Office, and is a member of the Race for Justice Independent Advisory Group. Carolyn Hoyle has been at the Oxford Centre for Criminology since 1991. Her earlier work focused on domestic violence and victims and her recent projects have been on restorative justice, including four years of research into restorative cautioning; a national evaluation of restorative justice schemes for juveniles; a comparative study of traditional and restorative measures for resolving complaints against the police by the public; and a comparative re-sanctioning study assessing the impact of restorative cautioning on offending. She has recently finished a book on restorative justice written with Chris Cunneen, under the new ‘Debating Law’ series, to be published by Hart Publishing in 2010. Other notable publications include; The Death Penalty (with Roger Hood, OUP, 2008); Negotiating Domestic Violence (1998); An Evaluation of the ‘One Stop Shop’ and Victim Statement Pilot Projects, (with Ed Cape, Rod Morgan and Andrew Sanders, 1998); Proceed with Caution (with Richard Young and Roderick Hill, 2002); and New Visions of Crime Victims (co-edited with Richard Young, 2002), as well as numerous articles and chapters on the restorative justice, the death penalty, domestic violence and victims. Jack Levin is the Irving and Betty Brudnick Professor of Sociology and Criminology at Northeastern University in Boston, where he codirects its Center on Violence and Conflict. He has authored or coauthored a number of books, including Hate Crimes, Why We Hate, The Functions of Prejudice, Hate Crimes Revisited, Serial Killers and Sadistic Murderers – Up Close and Personal, and The Violence of Hate. Levin has also published more than 150 articles in professional journals and major newspapers, such as The New York Times, Boston Globe, Dallas Morning News, Philadelphia Inquirer, Christian Science Monitor, Chicago xvii

Hate Crime

Tribune, Washington Post and USA Today. His most recent work focuses on two areas: animal abuse in relation to sadistic murder and the prosecution of hate offences. Hannah Mason-Bish is a lecturer in Criminology at Roehampton University. Her main research interests are in the development of hate crime policy, with a particular focus on age, gender and disability. She is currently working on a paper which examines the conceptualisation of disability hate crime and draws on her research. Hannah is a member of the Hate Crime Research Group, the Disability Hate Crime Network and the Roehampton University Gender and Crime Research Group. Jack McDevitt is the Associate Dean for Research and Graduate Studies at the College of Criminal Justice at Northeastern University, Boston, and directs the Institute on Race and Justice and the Center for Criminal Justice Policy Research. He is the co-author of three books, Hate Crimes: The Rising Tide of Bigotry and Bloodshed (with Jack Levin, 1993), Hate Crime Revisited: American War on Those Who Are Different (with Jack Levin, 2002) and Victimology (with Judy Sgarzi, 2003). He has also co-authored a number of reports on racial profiling including a monograph for the US Department of Justice and Statewide reports from Rhode Island and Massachusetts on the levels of disparity in traffic enforcement. Jack is Co-principal Investigator with Dr Amy Farrell of a national evaluation of the recent Police Integrity Initiative of the US Department of Justice’s Office of Community Orientated Policing Services (COPS), and has also testified as an expert witness before the Judiciary Committees of both US Senate and the US House of Representatives and as invited expert at the White House. Derek McGhee is Professor of Sociology at the University of Southampton. His primary research focus is on citizenship and ‘minority’ communities. He has published and conducted research on a range of inter-related topics from racism, homophobia and Islamophobia, and has examined policing practices, counter-terrorism legislation and immigration (and managed migration), integration and community cohesion policies. His most recent research focuses on Britain’s unfolding human rights culture. His books include Homosexuality, Law and Resistance (Routledge, 2001), Intolerant Britain? Hate, Citizenship and Difference (Open University Press, 2005), and

xviii

Notes on contributors

The End of Multiculturalism? Terrorism, Integration and Human Rights (Open University Press, 2008). His new book, Security, Citizenship and Human Rights – Shared Values in Uncertain Times will be published by Palgrave in 2010. Jim Nolan is an Associate Professor of Sociology and Criminology at West Virginia University. His research and teaching focuses on group relations, crime and social control. Dr Nolan was a police officer in Wilmington, Delaware for thirteen years and then worked for five years with the FBI. He earned his doctoral degree from Temple University in Philadelphia. Barbara Perry is Professor and Associate Dean of Criminology, Justice and Policy Studies at the University of Ontario Institute of Technology, and Visiting Professor at Nottingham Trent University in the UK. She has written extensively in the area of hate crime, including several books on the topic, including In the Name of Hate: Understanding Hate Crime and Hate and Bias Crime: A Reader. She has also published in the area of Native American victimisation and social control, including Silent Victims: Hate Crime Against Native Americans, a book based on interviews with Native Americans (University of Arizona Press, 2008). She has also written a related book on policing Native American communities: Policing Race and Place in Indian Country (Lexington Press, 2008). She is general editor of a five-volume set, Hate Crimes (Praeger), and editor of volume 3, Victims of Hate Crime, of that set. Dr Perry continues to work in the area of hate crime, and has begun to make contributions to the limited scholarship on hate crime in Canada. Mark Walters is a doctoral student at the Centre for Criminology, University of Oxford. Prior to this he completed an MSc in Criminology and Criminal Justice (Research Methods) at Oxford and an LLM specialising in criminal justice at the University of New South Wales, Sydney, Australia. Mark’s main research interests include hate crime and restorative justice. He has published various journal articles on the use of hate crime legislation and articles on community policing, victim impact statements and anti-discrimination law in Australian, international and English journals. Currently Mark is carrying out a year-long empirical study, as part of his DPhil, in London and Sussex which explores the use of restorative justice practices as a response to hate crime.

xix

Hate Crime

Kevin Wong is a consultant and researcher. He was previously an Assistant Director with NACRO, the principal crime reduction charity in England and Wales, where he worked with police services, local authorities and housing bodies, to develop strategies and interventions to address hate crime and provide support to victims. He has conducted research and written extensively about racially motivated crime and hate crime with a particular focus on policy development and implementation. Kevin is an associate lecturer at Manchester Metropolitan University and the University of Huddersfield.

xx

Future developments for hate crime thinking: who, what and why? Neil Chakraborti

Introduction As has been well documented in other recent texts, there is nothing especially new about the patterns of prejudice that give rise to what we now refer collectively as hate crime. Acts of bigotry directed towards marginalised and vulnerable communities are part of our historical fabric, rooted in the widespread and often culturally accepted demonisation of the Other, and we can all recall countless examples over time – be it high-profile cases of murderous hate, episodes of organised extremist violence or repeated acts of harassment, abuse and bullying – which have vividly illustrated the many harms of hate crime. In late 2009 those of us based in the UK were served a chilling reminder of these harms through the inquest verdict on the case of Fiona Pilkington. The case is particularly tragic – Fiona, a 38-year-old mother of an 18-year-old girl with learning difficulties, was driven to kill herself and her daughter in October 2007 by setting light to her car, with them both inside, near their home in Leicestershire following years of disablist abuse from local youths directed at her family – not least because it highlights in no uncertain terms the devastating cumulative impact of so-called ‘low-level’ incidents of othering, but also because the end result could so easily have been avoided. As the inquest verdict made clear, inaction from the police and local council was partially responsible for the victims’ plight, particularly in view of the fact that the police had been called out as many as 33 times in seven years without making a single arrest, while council workers 

Hate Crime

charged with tackling anti-social behaviour in the locality did not even record the case until almost seven years after it was initially reported to the police (Walker 2009a, 2009b). Of course, this case is by no means an isolated reminder of the persistence of bigotry and hate in this country. As I write this chapter, reports are filtering through of a vicious homophobic attack on James Parkes, an off-duty trainee police officer set upon by a group of up to 20 people in Merseyside; this attack follows soon after the homophobic killing of Ian Baynham, punched and kicked to death near Trafalgar Square on an evening out in central London by two teenage girls and a young man (BBC News 2009a, 2009b). Also of late, we have seen a series of anti-Islamic protests organised by the English Defence League and affiliated far right groups across various towns and cities designed to stoke up resentment against British Muslims and to damage community relations (Hamilton 2009), while in a similar fashion the British National Party (BNP) have launched fresh campaigns to cast Asian communities, and specifically Muslims, as the main perpetrators of hate crimes in the UK with white British people depicted as the victims of violent racism and institutional discrimination (BNP 2009). Equally alarming was the narrow avoidance of hate crimes of potentially devastating proportions that came with the arrest of Neil Lewington, an unemployed electrician from Reading who, in seeking to emulate the Oklahoma bomber Timothy McVeigh and the London nail bomber David Copeland, had been using his bedroom as a bomb-making factory to prepare a sustained hate campaign against Asian families (Press Association 2009).1 It should be stressed that these are merely selective snapshots of hate that have come to light at the time of writing; equally, one could turn to a whole host of graphic cases or disturbing developments here in the UK and further afield (see, inter alia, Perry 2009; ODIHR 2008; Gelber and Stone 2007) to illustrate the simmering tensions and manifestations of prejudice that underline the widespread presence of hate crime in contemporary society. However, while these problems show little sign of abating to any great extent, there is some degree of comfort to be sought in the improved response to hate crime from governments, practitioners and academics which, prima facie, suggests greater prioritisation and better understanding of the issues. Scholars in the UK, for instance, will be familiar with the measures taken post-Macpherson to combat hate crimes more effectively – measures that include the introduction of legislation governing hate acts and speech, the publication of national policy and guidance documents governing the policing and prosecution of hate crime, reviews of 

Future developments for hate crime thinking: who, what and why?

progress against the recommendations of the Stephen Lawrence Inquiry and a pronounced emphasis on community cohesion and interfaith dialogue (Home Office 2009) – and similar efforts have been made elsewhere across Europe to promote diversity and strengthen responses to violent and non-violent manifestations of hate (ODIHR 2008). The same can be said of the US where legislative protection for minority groups has been bolstered through the expansion of federal hate crime law to cover crimes committed on the grounds of gender, sexual orientation, gender identity and disability. Allied to these developments in the policy domain is the progress in what Iganski (2008: 5) refers to as the scholarly domain, where ‘there is an analytical coalition between scholars in once disparate fields of study concerned with oppression, discrimination and bigotry in various guises’. This surge of professional and scholarly interest in the subject is to be welcomed, with the concept of hate crime now firmly entrenched within the lexicon of academics, politicians and the public at large. Though the rationale for choosing the word ‘hate’ as a collective descriptor for the varied types of offence included within its framework may be contested, the fact remains that hate crimes – and the forms of victimisation associated with the label – are widely recognised as significant social problems and afforded greater priority upon research and policy agendas than ever before (Mason 2005; Chakraborti and Garland 2009).

Recurring questions Yet despite our collective acknowledgement of hate crime, it remains a contested area of study and policy. In part this is because of the ambiguity that surrounds its interpretation: although a number of hate crime scholars have sought to offer conceptual clarity and a coherent framework for criminal justice policy (see, inter alia, Perry 2001; Iganski 2002, 2008; Gerstenfeld 2004; Hall 2005a; Chakraborti and Garland 2009), there are still divisions over what the term really means and what its value is. However desirable the quest for conceptual consensus may be, our search for definitional clarity can sometimes overshadow attempts to develop more constructive lines of argument about how best to reduce levels of hate crime. Moreover, and as we know from the body of scholarship to have emerged over the past ten years or so, hate crime is a highly complex subject, and the deeper we delve to find solutions and answers the more likely we are to stumble across further problems and questions. Learning how 

Hate Crime

best to tackle these problems and questions is a difficult, ongoing task but one that should form a unifying theme within contemporary hate crime scholarship.

Who? The title of this chapter alludes to the myriad questions posed by hate crime that require careful consideration, and where appropriate rethinking, if we are to take the ‘hate debate’ forward constructively and decisively. This is by no means the first attempt to draw attention to unanswered questions or neglected areas of scholarship (see, most notably, Perry 2003); however, despite – or arguably because of – the growth of academic interest in hate crime, we find ourselves confronted with increasing numbers of avenues deserving of further enquiry. For instance, the ‘Who’ in the title itself raises a series of challenging questions. Who are victims of hate crime? Who should hate crime laws be designed to protect? Scholars in the UK might argue that the responses to such questions have already been addressed through the strategic guidance offered by the Association of Chief Police Officers (ACPO 2000, 2005, 20102) which earmarks hate crime as hate or prejudice motivated on particular grounds – race, faith, sexual orientation, transgender status and disability – and which governs the way in which hate crimes are conceived of and dealt with operationally. But equally, and as explored more fully elsewhere in this collection, there are some groups of Others who are targeted because of different identity characteristics from those specified above but whose victimisation is not recognised as amounting to a hate crime despite it bearing all the hallmarks of such an offence. Meanwhile, even within those recognised categories of hate crime victim, crass generalisations are often made to the effect of overlooking the dynamics, specificities and intersectionalities of victimisation within these broad-brush categories. Making generic assumptions about diverse communities at the expense of learning about the discrete experiences of those who are all too often subsumed through the labelling of such communities gives us insufficient information about who the victims of hate crime really are and the context behind their vulnerability (see also Perry 2003; Garland et al. 2006). Who commits hate crime? Again, this is a question that has been addressed in part by hate crime scholars who have sought to challenge the popular stereotype of organised hate groups or committed extremists being responsible for the majority of offences. But who 

Future developments for hate crime thinking: who, what and why?

then are these offenders? Are they, as Iganski (2008) suggests, ordinary people like ‘us’ – our friends, neighbours, colleagues – committing offences in the context of their ordinary, everyday lives? Are they ‘normal’ people acting on mainstream, ‘common-sense’ bigotries that encourage them to blame the Other for problems blighting their own lives? Or perhaps might far right organisations and hate groups be to blame after all, not necessarily directly as perpetrators but for the influence that their violent rhetoric can have on everyday individuals not committed to an extremist ideology? Similarly, when thinking about who hate crime perpetrators are, we need also to question their relationship to the victim. Are perpetrators strangers to victims whom they target on the basis of their membership of a particular group identity as opposed to any individual traits, or might they be more familiar to their victim either as an acquaintance, friend, family member, carer or partner? Perhaps the victim may know the perpetrator at some level and yet still feel emotionally distant from them, thereby allowing us to conceive of hate crimes as strangerdanger crimes even where they are committed by people familiar with their target? Again, and as emphasised by Mason (2005: 844), there are no ‘one size fits all’ solutions to these kinds of questions; rather, and as with our search for victims of hate, we need to resist from drawing neat, overly simplistic conclusions based around what we think we know and instead use these multiple realities to inform our understanding conceptually and empirically. As well as asking who the victims and perpetrators are, we must also ask who should take responsibility for shaping improved responses to hate crime. Is this primarily an issue for the police? Despite the significant strides taken to improve the policing of hate crime over the past ten years or so, problems remain, whether in the context of a legacy of mistrust among minority groups, problems of non-reporting or perceived and actual failures at an operational level. But what of other agencies? Modern-day policing requires support from a range of supplementary actors, be they community support officers, neighbourhood wardens and other types of municipal policing or forms of private security, and all share responsibility for responding to hate crime. This responsibility extends also to local authorities and other partner agencies as it does to different parts of the criminal justice system; indeed, while the bulk of scholarly attention in this context has focused on policing in its narrow sense, we have much to learn about the ways in which hate crimes are punished, how laws are enforced and how offending behaviour and



Hate Crime

motivations are addressed. Equally, the communities from which perpetrators are drawn share a collective responsibility for responding to hate crime, as indeed do states, not just for the effectiveness of their laws and other interventions to tackle hate crime, but for their role in legitimising forms of othering and creating an environment, wittingly or otherwise, in which hate crime can flourish. As is evident from contributions within this collection, these are all nascent areas of enquiry within contemporary hate crime scholarship.

What? In a similar fashion, the ‘What’ in the title is designed to flag up a series of questions pivotal to the development of debate and empirical study in this field. First of all, what should we class as hate crime? Or perhaps more pertinently, what shouldn’t we class as hate crime? As mentioned above, there is now a fairly substantial body of literature that grapples with this particular question and some degree of consensus over the types of offences included within the conceptual framework of hate crime. Nevertheless, this does not preclude recurring questions being asked of the seemingly odd state of affairs that sees much of what falls under the remit of hate crime as not being crime per se, nor necessarily being motivated by hatred. While there are compelling reasons for policy-makers to employ a broad interpretation of hate crime so as to give effect not just to violent acts that one might automatically associate with crimes of hate but to more subtle yet equally damaging and invasive expressions of prejudice, those charged with enforcing policy might be less familiar with the basis for this argument and might have a much narrower take on what is and what isn’t a hate crime. Tied in with this point comes another ‘what’: what issues do we know too little about? Doubtless there are many responses one could give to such a question, especially as it is nigh impossible to name a hate crime-related issue that we could conceivably know too much, or even enough about. Nevertheless, a review of existing literature would suggest that there are particular areas that demand further enquiry, including (to name but several) the emergence of xenoracism; violence against lesbian women; the victimisation of ‘invisible’ communities such as the Chinese or new migrant populations; minority-on-minority and minority-on-majority violence; the in terrorem effect of hate crime upon the victim’s wider community; the nature and impact of cyberhate; or the experiences of those who occupy multiple positions of culturally defined inferiority (see, for 

Future developments for hate crime thinking: who, what and why?

instance, Levin 2002; Perry 2003; Paterson et al. 2008; Adamson et al. 2009; Fekete 2009). In a similar vein it could be argued that we know far too little about the collective experiences of the homeless, the elderly, members of youth subcultures and other groups whose vulnerability extends beyond the boundaries of most hate crime policy and scholarly frameworks, nor have we paid anything like enough attention to the targeting of disabled and transgender people, despite these groups being recognised ‘beneficiaries’ of most official discourses on hate crime (Chakraborti and Garland 2009; Wachholz 2009). Indeed – and drawing for a moment on a recent personal experience – to hear myself and others having to justify during interviews with national and local media why the years of abuse directed towards Fiona Pilkington’s family warranted being treated as a hate crime brought home to me just how marginalised disablist victimisation has been from the ‘hate debate’ (and how rarely people equate supposedly ‘low-level’ bullying and harassment with hate crime, irrespective of the consequences that may ensue). Then there is the related question of what should be done to tackle hate crime more effectively. Again, there are all sorts of directions for us to choose from here. Among them we could call for more effective monitoring of the ways in which police officers operationalise strategic hate crime guidance in their response to hate incidents, or of the decision-making processes at the recording and prosecuting stages of the criminal justice response to hate crime. We could delve deeper into inter- and intra-agency working practices among statutory and voluntary organisations responsible for protecting vulnerable communities; we could investigate the deployment of third-party reporting systems, community engagement strategies or victim support mechanisms; or we could examine the scope for making better use of alternative modes of justice for dealing with hate crime perpetrators. Almost certainly readers will find many more avenues to pursue when thinking about what should be done to tackle hate crime more effectively, and the chapters that follow offer further insights in this regard.

Why? Finally, the ‘Why’ in the title refers essentially to one especially pertinent question: why do we need to think afresh about hate crime? Or to put it another way, given the increased prioritisation of hate crime, both nationally and internationally, and the associated series of academic publications, action plans, policy reviews and 

Hate Crime

guidance documents that have accompanied this prioritisation, why then do we need yet more debate about the subject? For me there are three main reasons for refining our ideas about and responses to hate crime. One relates to the inexorably high levels of hate crime that underline the ongoing marginalisation of vulnerable groups and the failings of existing policy and enforcement mechanisms. Despite the increased attention over the past ten years or so, the sheer number of hate crimes taking place within the UK and beyond should preclude us from feeling complacent or even satisfied with how far we have come. Indeed, while it is true to say that there is now much greater awareness than ever before about hate crime and its dynamics, the preceding discussion has hinted at just how much more there is to learn. Within the UK this awareness has in no small part been shaped through our response to tragic hate crimes – headline-hitting cases such as Rolan Adams, Rohit Duggal, Stephen Lawrence, Jody Dobrowski, Anthony Walker, Rikki Judkins, Brent Martin, Sophie Lancaster – and yet (and as the Pilkington case so clearly demonstrates), we should surely not by now have to wait for a tragedy to strike before the vulnerability of a particular group is recognised as relevant. Another reason for thinking again about hate crime relates to the continuing uncertainty surrounding what we mean by the term. As Phyllis Gerstenfeld suggested back in 2004, ‘hate crimes seem to be a topic of some interest to nearly everybody, and yet few people really know much about them’ (2004: xv), and arguably the same can still be said now despite the surge of interest in the subject in the years that have followed. Certainly, judging from what we hear and see at conferences, in journals and in media discussions it would seem that academics, practitioners and the general public are often talking in different tongues when it comes to our respective takes on what hate crime means. Some, particularly laypeople, will understandably adopt a more literal interpretation of the term in line with the more violent and extreme cases that receive public attention; scholars, meanwhile, tend to see hate crime as a social construct with no straightforward meaning and offer a set of defining characteristics that they regard as central to its commission;3 while practitioners are likely to follow a much less complex take, which requires few of the machinations evident within academic interpretations. Equally, there may be significant variations in the way in which hate crimes are conceived of in different countries. For instance, a glance at the literature from mainland Europe would suggest that the tradition of framing hate crime analyses in terms of organised far right (mainly 

Future developments for hate crime thinking: who, what and why?

racist) violence holds true today – which may partially explain why some European countries record relatively few hate crimes4 – while a similar argument could be extended to North American texts, many of which still, as Iganski (2008: 15) observes, ‘contain the seemingly obligatory chapter on far-right perpetrators of “hate crime” ’. The term hate crime has been widely adopted and used as something of a buzzword without there being consistency in its application, and this has implications for how we conceive of the offences grouped under its protective umbrella and the actors involved, be they victims, perpetrators or agencies of control. A third reason to extend our forays into hate crime relates to the uncertainty surrounding the effectiveness, or otherwise, of existing provisions to deal with hate crime and the enforcement of those provisions. There continue to be shortcomings evident throughout the criminal justice response to hate crime, whether we choose to speak in terms of policing, prosecutions or punishment (see, for instance, Hall 2005b; Chakraborti 2009), and solutions to these shortcomings must be sought if they are not to detract from other signs of progress within the context of criminal justice. Moreover, while most of us would in principle welcome the firmer legal footing for hate crime that has emerged in a number of countries through the introduction of various penalty enhancement and incitement laws, the legitimacy, scope and implementation of these laws have all come into question at one time or another, giving rise to suggestions of legal ambiguity, inconsistency in application and even tokenism in some quarters (Dixon and Gadd 2006; Gelber and Stone 2007; McGhee 2008; Phillips 2009). This is an important though highly contentious area of scholarship, and one that will no doubt assume a central role in future debates about hate crime.

About the book By now readers will, I hope, be convinced, were they not already, of the need to take our studies of hate crime further. A single volume of this nature cannot begin to do justice to the many and varied areas that demand enquiry, but the coalition of ideas put forward by its contributors can at least provide a starting point. Each of the authors in this edited collection is helping to shape hate crime scholarship and producing work that is highly relevant to national and international debates, and interestingly – when given an open brief to write about what they regarded as posing significant challenges to future policy 

Hate Crime

and practice – each chose to focus upon separate though inter-related themes from the others. Without question this is less an indication of sensible editorial planning and rather more reflective of the diverse and exciting work currently being undertaken within our field. These contributions have been grouped into two broad sections: those that seek to develop more nuanced understandings of hate crime, and those that seek to develop more nuanced responses to hate crime. Part One opens with the thoughts of Barbara Perry, who presents the findings from her extensive review of recent theoretical and empirical advances to highlight themes in need of further exploration. Her suggestions for future hate crime studies – born from her description of the field as ‘out of its infancy … [but] not yet matured into adulthood’ – raise important questions over the assumptions we often make about offenders, interventions and our classification of victims. This last point is a central theme within the next chapter, from Jon Garland, which calls for a closer look at the groups of victims protected within existing hate crime policy frameworks. By taking the victimisation of goths as its focal point, this chapter, and indeed the one that follows from Hannah Mason-Bish, underlines just how difficult and subjective a process it is to settle upon an acceptable threshold for deciding which forms of ‘othering’ to outlaw through hate crime legislation. In her review of the policy domain Mason-Bish warns against common-sense assumptions that can not only affect operational policing and prosecution but also can be used to automatically exclude particular groups of victim, and calls for us to give further thought to the message conveyed to those excluded from policy frameworks. Chapter 4 by Marian Duggan draws on related issues by outlining the marginalisation of homophobic hate crime in Northern Ireland and her analysis offers timely insights into problems that have barely featured on the academic radar. Equally, Nicole Asquith’s chapter examining the challenges posed by verbal-textual hostility takes traditional hate crime thinking in new directions by showing how the use of forensic linguistics can help us to evaluate the possible harms of hate speech. Developing a better understanding of what is said in hate crimes, and what effect this may have on the victim, has clear relevance to scholars and policy-makers, as does an understanding of what motivates people to commit hate crimes. This is the focus of the final chapter in Part One, by Jack McDevitt, Jack Levin, Jim Nolan and Susan Bennett. In this chapter the authors revisit their previous typology of hate crime offenders in the light of new data analysis to offer fresh ideas about perpetrators’ motivations, and in 10

Future developments for hate crime thinking: who, what and why?

so doing illustrate why this has important implications for criminal justice agencies. We then move into Part Two, which begins with Nathan Hall’s chapter on the complexities of policing ‘hatred’. It goes without saying that hate crime poses significant challenges for the police: not only are they responsible for recording offences and implementing related legislation, but also they will often be the primary point of contact for victims of hate crime, and through their behaviour, attitudes and policies can influence the way in which these kinds of issues are thought of by the state and general public. Hall examines the complexities at play in order to assess how best to judge police responses to hate crime and offers a series of important suggestions in this regard. Following on from this, Derek McGhee examines the complexities inherent to the policing of vulnerable, ‘at risk’ groups and pays particular attention to parallels and differences in the discourses of vulnerability used to shape the policing of LGBT5 and Muslim communities. We then move towards broader criminal justice issues in our search for improved responses to hate crime. Kris Christmann and Kevin Wong tackle the thorny issue of hate crime reporting and suggest that more complex narratives are required to recognise the reality of under-reporting and the unacknowledged deficiencies within current policy. The no less problematic issue of penalty enhancement legislation comes under scrutiny within the subsequent chapter by David Gadd whose analysis of racially aggravated offending calls into question the one-dimensional responses to offending behaviour so often put forward. Gadd suggests that hate crime penalties should be designed to help offenders gain an appreciation of why and how their behaviour was perceived as prejudiced and a similar message is conveyed in the final chapter, by Mark Walters and Carolyn Hoyle, which explores the potential of restorative justice in cases of hate crime. Walters and Hoyle draw upon theoretical and empirical research to highlight ways in which restorative processes can be utilised in the context of hate crime, and in so doing highlight the limitations of solely punitive responses in repairing the harms suffered by victims and in challenging the prejudices of offenders. Without any doubt, hate crime is an emotive and contentious subject area: any label that requires us to make qualitative distinctions between different forms of prejudice and vulnerability – and to find appropriate responses for victims and offenders – is likely to invite criticism and divide opinion. However, hate crime is also an extremely important subject area and one that requires complex solutions to the 11

Hate Crime

complex questions it poses. While this collection does not profess to offer all the solutions, what it does do is present a dynamic range of ideas from established and emerging hate crime scholars whose research is shaping conceptual and policy frameworks for the better. These are the kinds of ideas that can push forward the boundaries of the hate debate to inspire further exploration and intervention.

Notes 1 There was some degree of good fortune associated with the prevention of Neil Lewington’s attacks. Were it not for Lewington’s arrest for drunkenly abusing a female train conductor, he would quite conceivably have escaped the police radar; as it was, this relatively minor incident led to their discovery of components for incendiary devices on his person during a routine stop and search, which in turn triggered a more detailed investigation uncovering evidence of meticulously constructed explosives, planned attacks, white supremacist sympathies and videotapes of McVeigh’s and Copeland’s attacks (Press Association 2009). 2 Although unavailable at the time of writing, an updated guidance manual on hate crime is due to be published by ACPO in 2010. 3 This may include characteristics such as the group affiliation of the victim; the imbalance of power between perpetrator and victim; the relevance of context, structure and agency to the process of hate crime; or the notion of acts of hate being ‘message’ crimes designed to create fear within the victim’s broader community (Chakraborti and Garland 2009: 150). 4 For instance, according to the 2007 statistics on hate crimes taking place in the OSCE (Organisation for Security and Co-operation in Europe) region, countries as large as Italy and Poland recorded a total of only 148 and 125 hate crimes respectively for the entire year. Equally, criminal proceedings were initiated in as few as 170 hate crime cases in Russia, ten in Denmark and none at all in Iceland (ODIHR 2008). 5 LGBT is an abbreviation used to collectively represent lesbian, gay, bisexual or transgendered people.

References ACPO (2000) Guide to Identifying and Combating Hate Crime. London: ACPO. ACPO (2005) Hate Crime: Delivering a Quality Service – Good Practice and Tactical Guidance. London: Home Office Police Standards Unit. ACPO (2010) The ACPO Hate Crime Manual – 2010. London: Home Office Police Standards Unit. 12

Future developments for hate crime thinking: who, what and why?

Adamson, S., Cole, B. and Craig, G. (2009) Hidden from Public View: Racism against the UK Chinese Population. London: The Monitoring Group. BBC News (2009a) ‘Boys Bailed Over Anti-Gay Attack’, BBC News Online, 27 October. Available at http://news.bbc.co.uk/go/pr/fr/-/1/hi/england/ merseyside/8327214.stm [accessed 27 October 2009]. BBC News (2009b) ‘Man Dies after Homophobic Attack’, BBC News Online, 14 October. Available at http://news.bbc.co.uk/go/pr/fr/-/1/hi/england/ london/8306481.stm [accessed 14 October 2009]. BNP (2009) Racism Cuts Both Ways: The Scandal of Our Age. The British National Party’s Report on Hate Crimes Against White People. London: British National Party. Chakraborti, N. (2009) ‘A Glass Half Full? Assessing Progress in the Policing of Hate Crime’, Policing: A Journal of Policy and Practice, 3(2): 1–8. Chakraborti, N. and Garland, J. (2009) Hate Crime: Impact, Causes and Responses. London: Sage. Dixon, B. and Gadd, D. (2006) ‘Getting the Message? “New” Labour and the Criminalisation of “Hate” ’, Criminology and Criminal Justice, 6(3): 309–28. Fekete, L. (2009) A Suitable Enemy: Racism, Migration and Islamophobia. London: Pluto Press. Garland, J., Spalek, B. and Chakraborti, N. (2006) ‘Hearing Lost Voices: Issues in Researching “Hidden” Minority Ethnic Communities’, British Journal of Criminology, 46(3): 423–37. Gelber, K. and Stone, A. (2007) Hate Speech and Freedom of Speech in Australia. Sydney: Federation Press. Gerstenfeld, P. (2004) Hate Crimes: Causes, Controls and Controversies. London: Sage. Hall, N. (2005a) Hate Crime. Cullompton: Willan Publishing. Hall, N. (2005b) ‘Community Responses to Hate Crime’, in J. Winstone and F. Pakes (eds) Community Justice: Issues for Probation and Criminal Justice (pp. 198–217). Cullompton: Willan Publishing. Hamilton, F. (2009) ‘Street Fights as Police Try to Hold Back Muslims Defending Mosque’, The Times, 12 September, p. 18. Home Office (2009) Hate Crime: The Cross-Government Action Plan. London: Home Office. Iganski, P. (ed.) (2002) The Hate Debate: Should Hate be Punished as a Crime. London: Profile Books. Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Levin, B. (2002) ‘Cyberhate: A Legal and Historical Analysis of Extremists’ Use of Computer Networks in America’, American Behavioral Scientist, 45(6): 958–88. Mason, G. (2005) ‘Hate Crime and the Image of the Stranger’, British Journal of Criminology, 45(6): 837–59. McGhee, D. (2008) The End of Multiculturalism? Terrorism, Integration and Human Rights. Maidenhead: Open University Press. 13

Hate Crime

ODIHR (2008) Hate Crimes in the OSCE Region – Incidents and Responses. Annual Report for 2007. Warsaw: Office for Democratic Institutions and Human Rights. Paterson, S., Kielinger, V. and Fletcher, H. (2008) Women’s Experience of Homophobia and Transphobia: Survey Report. London: Metropolitan Police Service. Perry, B. (2001) In the Name of Hate: Understanding Hate Crimes. London: Routledge. Perry, B. (2003) ‘Where Do We Go From Here? Researching Hate Crime’, Internet Journal of Criminology. Available at www.internetjournalofcriminology. com/Where%20Do%20We%20Go%20From%20Here.%20Researching%20H ate%20Crime.pdf Perry, B. (2009) ‘Introduction: Filling in the Blanks’, in B. Perry (ed.) Hate Crimes, Vol. 3: Victims of Hate Crime. Westport, CT: Praeger (pp. xvii–xxi). Phillips, N. (2009) ‘The Prosecution of Hate Crimes: The Limitations of the Hate Crime Typology’, Journal of Interpersonal Violence, 24(5): 883–905. Press Association (2009) ‘Neo-Nazi Jailed for Planning Racist Bombing Campaign’, The Guardian, 8 September. Wachholz, S. (2009) ‘Pathways Through Hate: Exploring the Victimisation of the Homeless’, in B. Perry (ed.) Hate Crimes, Vol. 3: The Victims of Hate Crime. Westport, CT: Praeger (199–222). Walker, P. (2009a) ‘Bullying Family Still a Menace in Leicestershire, Inquest Told’, The Guardian, 23 September, p. 15. Walker, P. (2009b) ‘No Excuses for Abuse that Led to Double Death, Says Minister’, The Guardian, 29 September, p. 8.

14

Part One

Developing Understandings of Hate Crime

Chapter 1

The more things change … post-9/11 trends in hate crime scholarship Barbara Perry

In 2002 I presented an inaugural hate crime conference keynote address (Perry 2002), which was later published as ‘Where do we go from here? Future directions in hate crime scholarship’ (Perry 2003a). The paper was offered as a ‘synopsis of what we don’t know about hate crime’. I return to my musings from 2002 to see how far we have come since that time. I assess recent theoretical and empirical contributions that have emerged in the opening decade of the twenty-first century. The post-9/11 era has seen dramatic shifts in how we conceptualise and respond to hate crime, although many areas of enquiry remain underdeveloped. I tease out some of the scholarly advances, while also pointing out themes in need of further exploration. A great deal has changed since that paper was delivered, and many factors conducive to hate crime have emerged or been exacerbated in the intervening years. Terrorists struck again, most notably in London and Madrid. Far right political parties have gained substantial ground across Europe. The global economy has slumped into a devastating recession, leaving millions unemployed and bitter. The implications of these processes for hate crime are thus frequent themes in the emerging literature: the effects of the 9/11 attacks on anti-Muslim violence, the relationships between politics, culture and hate crime, the role of economic uncertainty on xenophobic violence – all are thus among the issues under the magnifying glass. In preparation for this chapter, I conducted an extensive online search of journals and texts addressing hate crime. In the end, I had before me over 120 pages that contained over 500 abstracts, which I 17

Hate Crime

then traced back to the source. Clearly, I didn’t read all 500 pieces in detail, but only closely enough to discern kernels at the core of each. I was able to identify four broad categories dominating the literature: making sense of hate crime; ‘categories’ of victimisation; hate groups; and responding to hate crime. In what follows, I try to offer a concise overview of the themes attendant with each of these categories. I conclude with a brief discussion of two new venues that have sought to pull together what might be called the field of hate studies. It should be noted at the outset that, in addition to the array of journal articles, a number of books related to the theme of hate crime have also been published in the intervening years. In particular, several anthologies that bring together diverse streams and perspectives have contributed to our knowledge of the issue at hand. Among these are Chakraborti and Garland (2009), Gerstenfeld (2004), Iganski (2002), Perry (2003b), Prum et al. (2007) and Winterdyk and Antonopoulos (2008). Each of these is well worth a peek. They contain analyses by leading scholars that range from the local to the global, the mundane to the extreme, the historical to the contemporary.

Making sense of hate crime When I was writing in 2002, I was most disturbed by the appalling lack of efforts to theorise hate crime in any sophisticated way. That, fortunately, has changed dramatically in the intervening years. There have been a number of very strong pieces that have advanced our conceptual understanding of the dynamics of hate crime. Many of these have drawn on psychological constructs, highlighting the individual motivations for engaging in such violence. Craig (2002) and Sullaway (2004) have both presented useful overviews of the ways in which psychology has contributed to our understanding. Hamer (2006), for example, draws on psychoanalytic notions of transference and projection to account for racialised violence specifically. The great value of this literature is that it typically lays out concrete clinical interventions intended to reduce hate crime. Social theory has also advanced the field of hate studies. Lyons (2007, 2008), for example, has published several tests of communitybased theoretical models, including social disorganisation and defended communities. Blazak (2001, 2004) has made very good use of strain theory and propaganda theories to ‘make sense’ of skinheads, especially. In 2005, Blee published an article entitled ‘“Racial violence in the United States’. This deceptively simple title disguises a very 18

The more things change … post-9/11 trends in hate crime scholarship

strong analysis that provides ‘a more rigorous exploration of the communicative, interpretive, and contextual nature of racial violence’. Increasingly, scholars are turning to cultural studies – especially identity theories – to account for hate crime. This has been especially the case for LGBT studies (Dunbar 2006), but also with respect to race and ethnic studies (Hoover and Johnson 2003/2004; Tomsen 2006). One approach that I find especially intriguing draws upon cultural geography. Flint (2004a) has made an explicit foray into the field with the publication of an innovative collection of essays specifically devoted to the geography of hatred and intolerance. Here contributors offer varied explorations of the ways in which organised and informal groups assert their territorial claims in efforts to purge their neighbourhoods, cities, regions or nation of the encroaching threats represented by people of colour and gay men and lesbians in particular. The authors share the recognition that ‘imperatives of the territorial defense of places and spaces result in the adoption of exclusionary visions and practices’ (Flint 2004b: 9). In the same year, Neil Chakraborti and Jon Garland (2004) published another related anthology, entitled Rural Racism. Here we see theoretical approaches to the ‘geography of hate’ as played out in rural contexts. The authors explore the dynamics of ‘insiders’ and ‘outsiders’, or those who are ‘out of place’ in the historically and prototypically white landscape. Careful attention is paid to the ways in which ‘community’ has come to be defined in rural settings, and how this subsequently excludes racial and ethnic minorities in cultural, social and political contexts. One of the most exciting hallmarks of recent hate crime scholarship has been the apparent maturation of empirical approaches within the field. There is an astounding diversity of methodologies currently being applied in efforts to more fully document and comprehend the problem. Sadly there are still what I would consider to be too many projects grounding their analyses in official data. I am often asked to review such papers, and generally reject them on the basis that they are ‘fatally flawed’ to the extent that they rely – uncritically – on police data and Uniform Crime Reports. Nonetheless, these are slowly being supplanted by more rigorous quantitative projects and more insightful qualitative projects. Psychologists have contributed to our knowledge in part through survey research on the immediate effects of hate crime on victims (Herek et al. 2002; Parks and Woodson 2002). Several experimental design studies have also emerged, most notably those exploring perceptions of hate crime. Some of these have revolved around how 19

Hate Crime

members of vulnerable communities interpret and react to hypothetical accounts of animus-motivated violence. The intent here is generally to understand the differential assessment of the seriousness of hate crime across groups. Not surprisingly, this literature tends to find that communities historically vulnerable to hate crime – Asians, Jews, sexual minorities and others – are more likely to classify relevant scenarios as hate crime, and to suggest harsher penalties (Boeckmann and Liew 2002; Lee et al. 2007). A related set of experimental studies has looked at how the features of a scenario shape perceptions. Such elements as the race of the perpetrator and victim or the severity of the act seem to shape whether an event is thought to constitute a hate crime, as well as sentencing responses (Cowan and Mettrick 2002; Saucier et al. 2008). An intriguing finding is emerging in some of this literature. For example, for her study of student perceptions of hate crime, Miller (2001) hypothesised that students of criminal justice would understand the concept better than other students and thus be more likely to identify incidents as hate crime. She was proven wrong, in that these students were less likely to define relevant scenarios as hate crime. Along similar lines, Olivero and Rodrigo (2001) studied law enforcement students and found that they were more homophobic than students from other majors, and thus less sympathetic to hate crime victims. Findings such as these raise serious questions about whether the relatively well-educated next generation of criminal justice practitioners will be any more effective in responding to hate crime than their peers. Indeed, without concentrated training in this area, they are likely to reproduce the culture of apathy and neglect. Alongside the large body of quantitative approaches are myriad qualitative explorations. Content analysis has been effectively applied to the study of hate groups, examining such elements as hate music (Grascia 2003), and especially internet websites (Stern 2001/2002; Gerstenfeld et al. 2003; Weatherby and Scoggins 2005/2006). The resultant scholarship has provided valuable insight into the cultivation of a sympathetic audience and enhanced membership. Surprisingly few scholars have used interview or focus group approaches in spite of the potential these have for providing rich and nuanced data. Among the hundreds of pieces I reviewed for this chapter, only a handful employed either or both of these techniques. I made extensive use of both in my interviews with Native Americans, interviewing nearly 300 participants and conducting focus groups in each community I visited (Perry 2008). Sandra Wachholz’s (2005) 20

The more things change … post-9/11 trends in hate crime scholarship

innovative study on violence against homeless people was grounded in 30 in-depth interviews, which revealed that hate crime in this context – like so many others – was intended to define ‘their place’. Generally, interview participants have been victims, or at least potential victims. Even fewer studies – quantitative or qualitative – focus on perpetrators. Brian Byers’ work on attacks against the Amish (Byers and Crider 2002) is an interesting exception, in that the analysis is based on the narratives of offenders. Also interesting are the few studies in which former members of white supremacist groups have shared their stories with researchers, shedding valuable light on recruitment, organisational structure, and exiting (Gadd 2006). Nonetheless, studies of this side of the victim–offender equation are rare, meaning that much of our conceptualisation of offender motives remains speculative. Another approach that lends deeper insight into the dynamics of hate crime is the case study, especially as they are used in the consideration of sensational incidents (Mason 2007). The technique is important both for humanising the victims and for putting the incidents into context. For the most part, these treatments of specific incidents of hate crime are not simply descriptive, but provide valuable analysis and insight into the broader social, cultural and political contexts in which they occur, such that they allow us to understand the enabling environment that allows hate crimes to persist. For example, a number of articles examining the Australian Cronulla Beach ‘race riots’ were published in 2006 and 2007. These bring to the fore the intersecting impacts of race (Hartley and Green 2006), gender (Ho 2007), moral panics (Lattas 2007), and the politics of multiculturalism (Halafoff 2006) on social and cultural relations in that country. More so than any other similar event, Cronulla laid bare the role of state rhetoric and practice in shaping a ‘culture of hate’ (Poynting 2006). In a similar vein, a handful of articles on the Matthew Shepard murder in Wyoming explore both the context and the effects of his killing. Two pieces are of particular interest. Karen Franklin (2003) makes the case that the murder may have played some role in the Lawrence v. Texas case in the US, which finally put paid to sodomy legislation. The decision, it is argued, was a public acknowledgement of the role that homophobic practices – including legislation – play in shaping anti-gay violence like that perpetrated against Shepard. Monique Noelle’s (2002) article examines a different class of impacts: that is, the psychological effects on the broader gay community. She observes a ‘vicarious traumatisation’ effect, whereby non-victims 21

Hate Crime

are in fact terrorised by awareness of anti-gay violence perpetrated against others. One final piece of the puzzle by which we ‘make sense’ of hate crime is the impact of animus-based violence. We have by now a sound inventory of the effects of hate crime on the immediate victim: for example, fear of additional victimisation (Parks and Woodson 2002); post-traumatic stress disorder (D’Augelli et al. 2006); and questioning self-identity (Dunbar 2006). Exciting work on the broader community effects of hate crime are just beginning to emerge. For example, Noelle (2002) applies assumptive world theory to the ‘ripple effects’ associated with hate crime. Using the Matthew Shepard murder as a focal point, Noelle situates a series of interviews with LGBT community members within this social psychological perspective. The interviews are mined for evidence that participants’ ‘fundamental assumptions’ about the way the world operates were challenged by their understanding of the Shepard murder. In short, she argues that in the face of vicarious traumatisation, other members of the targeted community also begin to question their long-standing perceptions of the world around them. Similarly, Nicole Asquith (2004) has explored the ways in which the everyday experience of hate speech and hate violence ‘police’ the behaviours of those deemed the Other. Both cultural and individual histories of bias-motivated violence keep members of these threatened communities in check, even to the extreme of attempting to hide their identities. This is a fruitful area of enquiry which can go a long way in highlighting the necessity of attending to hate crime not as an individual risk but as a public health issue.

Selective attention: naming the victims Perhaps one of the reasons that we have not attended more concretely to the broader impacts of hate crime is that we have not advanced very far in exploring the concrete experiences of discrete victim groups. Indeed, in the address that is the foil for this chapter, I made the claim that: To date, hate crime literature has tended to be very broad and non-specific in its focus. That is, little scholarship devotes attention to specific categories of victims. Extant literature has tended to discuss hate crime in generic terms, as if it was experienced in the same ways by women, by Jews, by gay men, 22

The more things change … post-9/11 trends in hate crime scholarship

by Latino/as, by lesbians. Even racial violence is collapsed into one broad category, as if all racial and ethnic groups experienced it in the same way. Consequently, we do not have a very clear picture of the specific dynamics and consequences that may be associated with victimization on the basis of different identity positions. (Perry 2003a) Plus ça change, plus c’est la même chose. Indeed, I would not dramatically alter my position on this today. With a couple of notable exceptions, we have still failed to engage in either theoretical or empirical work that does justice to the specificity of the experiences of diverse communities. There is still the stubbornly annoying tendency in too much of the hate crime literature to elide the experiences of racialised minorities rather than taking seriously the specificity of victimisation and offending within and across groups. Thus, scholars write about ‘racially motivated crimes’ as if the concrete target is immaterial. Perhaps even more troubling is the tendency to conflate ‘race’ with blackness and/or whiteness. Lyons’ (2008) otherwise intriguing test of multiple theoretical approaches is one such example. Here, he examines only the community dynamics of anti-black and anti-white hate crime. Keep in mind that the study was conducted in Chicago, arguably a very diverse city. Similarly, D’Alessio et al. (2002) claim to explore the relationship between racial threat and interracial crimes, but they too confine their analysis to the traditional black/white binary. Of course, important exceptions exist. Jannson’s (2006) Home Office report explicitly separates out the experiences and perceptions of blacks, whites and Asians, and in fact makes even finer distinctions within these groups. So too with Parks and Woodson’s (2002) analysis of ‘ethnic minority male survivors’ of racially motivated hate crime, which addresses hate crime victimisation across five different groups, including biracial/multiracial youth. Moreover, there are also some very good treatments of specific racialised communities such as Native Americans (more below). Ironically, while black people are the most frequent victims of hate crime, especially in the US, the UK and Canada, there is very little scholarship devoted to this community as a distinct community aside from their treatment as the ‘normative’ victim group. The same might be said for anti-semitic violence. Again, it is relatively common – perhaps increasingly so – yet surprisingly under-explored. The exception is the ongoing work by Paul Iganski and his colleagues in the UK (Iganski 2007). North American scholars are virtually silent on this. 23

Hate Crime

Other ‘invisible’ target communities are Latinos and Asians. I am also unaware of any literature that takes on the issue of antiwhite violence, with a couple of exceptions that critique hate crime legislation as something likely to be taken advantage of as much by whites as non-whites (Chorba 2001; Adams and Toth 2006). I have exchanged a few emails with a colleague in the UK on this point, and neither of us has been able to identify relevant studies. Here is an area ripe for exploitation. Granted, it is also an area fraught with political minefields! Recent work by Chris Cunneen (2001) in Australia, and Barbara Perry (2008), and Andrea Smith (2005) in the US have made important inroads in establishing a field of scholarship on hate crime victimisation in Aboriginal communities. Importantly, these authors highlight the links between processes of colonisation and racist violence, whereby the explicitly genocidal practices of early settlers as well as the more ‘subtle’ forms of ethnocide in the twentieth century leave a legacy of violence that resonates to this day. Indeed, it is impossible to fully comprehend the current plight of Aboriginal communities without paying attention to the colonial history of the relationship between European ‘settlers’ and native peoples around the world. One of those things that has not changed in the past decade or so has been the relatively heavy concentration on hate crime within LGBT communities (Janoff 2005). So extensive are the empirical and theoretical analyses of anti-LGBT violence that one hardly knows where to begin an attempt to summarise the literature. One means to do so is to recognise that there are three key ‘stages’ addressed: motivation, incidents and impacts. A sophisticated body of scholarship theorising violence against LGBT communities has been established. Typically, these focus on identity and/or gender role theories as a means of understanding the impetus behind such violence (Alden and Parker 2005; Dunbar 2006; Tomsen 2006). A second body of literature continues to build our empirical understanding of the concrete characteristics of anti-gay violence – incidence, prevalence, vulnerability, location, for example (Herek et al. 2002; Willis 2004). A third set attends to the impacts of anti-LGBT violence, with an emphasis on individual psychological and behavioural factors (Herek et al. 2002; Otis 2007), although there have been some recent efforts to conceptualise the community impacts as well (Noelle 2002). Aside from these areas, an additional notable trend is the relatively novel interest in violence against transgender/transsexual people as distinct

24

The more things change … post-9/11 trends in hate crime scholarship

from that of lesbians and gay men, to the extent that trans-identities add an additional layer of vulnerability. This will be an especially exciting field to watch as it matures. There is, of course, one very notable new development in the victim literature. In the aftermath of 9/11, it is perhaps not surprising that anti-Muslim violence has become an important focus for scholars (Kaplan 2006; Sheridan 2006; Byers and Jones 2007). Immediately following these attacks, backlash violence against those perceived to be Muslim escalated dramatically, resulting in assaults, arsons, even racially motivated murders, across the world and especially in nations aligned with the United States. Even in Canada, one of the western nations that did not support the 2003 US invasion of Iraq, anti-Muslim violence rose, as anti-Muslim practices were instituted and intensified by the state (Poynting and Perry 2007). Dozens of related articles and a handful of scholarly books have appeared in the past few years. A final new area of enquiry is that which explores violence against people with disabilities (McMahon et al. 2004). Ironically, while this group is typically among the largest ‘visible minority group’, their victimisation experiences have not been the subject of extensive scholarship (Zaversek 2003). The seminal piece in this area is probably the 2005 article by Grattet and Jenness in the Journal of Criminal Law and Criminology. Here they situate the utility of including disability as a protected category within the broader debates around the ‘dilemma of difference’. Having established that bias-motivated violence against people with disabilities ‘fits’ the hate crime model, they nonetheless go on to question whether, in fact, this is desirable. On the one hand, recognising people with disabilities in the legislation is segregating in that it ‘entails affording “special” treatment to those with disabilities’ (Grattet and Jenness 2005: 695). On the other hand, it is also integrative and inclusive in that it treats all vulnerable communities ‘the same’. Their argument is that the latter is problematic because it is quite likely that violence against people with disabilities is not the same as violence against Latinos, or against Jews. Each, in fact, has its own dynamics. This reaffirms my own positive evaluation of the growth of scholarship devoted to discrete communities – contra popular and political opinion, it is not the case that a crime is a crime is a crime. The concrete motives and impacts are discrete and should be understood in their specificity. Thus, I call for a renewed emphasis on research focusing on the distinct or even comparative experiences of narrowly defined communities.

25

Hate Crime

Organised hate Organised hate groups continue to draw the attention of hate crime scholars. Sadly, there is little that is new or innovative in much of this work. There is sustained interest in skinheads (Watts 2001; Hicks 2004) and in white supremacist symbolism (Etter 2001; Bradley 2007). However, there is some reason for optimism in two novel areas of enquiry that have recently come to the fore, drawing attention to important and intriguing dimensions of the hate movement that had previously gone un/underexplored. The first is captured in Abby Ferber’s (2004) Home Grown Hate: Gender and Organized Racism. Here the gendered dynamics of the movement are front and centre. Some authors consider the involvement of women in the contemporary hate movement (Blazak 2004; Blee 2004), while others focus on the gendered and patriarchal assumptions that underlie their rhetoric and practice (Dobratz and Shanks-Meile 2004; Perry 2004). This is an important advancement that forces us to acknowledge the intersections of race, gender and sexuality within the white supremacist imagination. The second fruitful area of enquiry has been what is now referred to as cyberhate. Increasingly, scholars are examining the ways in which the internet allows the hate movement to retrench and reinvent itself as a viable collective. The many electronic means available to the movement – blogs, newsgroups, ’zines – allow an ease of communication and dissemination of their views never before possible. Internet communication facilitates the creation of the collective identity that is so important to movement cohesiveness (Bocji and McFarlane 2003). Clearly, this has strengthened the domestic presence of these groups in countries like the United States, Canada, Germany and Sweden. In addition, due to the nature and the configuration of the internet, it allows the hate movement to extend its collective identity internationally, thereby facilitating and giving fuel to a potential ‘global racist subculture’.

Responding to hate crime If there is one topic that has consistently engendered wide-ranging dialogue in the hate crime literature, it is the question of ‘legislating hate.’ Indeed, even the US National Institute of Justice acknowledged this in 2007, with the publication of a report entitled Hate Crime in America: The Debate Continues (Shively and Mulford 2007). Likewise, Paul Iganski’s (2002) collection The Hate Debate: Should Hate Be 26

The more things change … post-9/11 trends in hate crime scholarship

Punished as a Crime indicates that the controversy is not confined to American navel-gazing, but rages on both sides of the Atlantic. While hate crime provisions have been adopted in the US, Canada, the UK and some EU countries, the presumed legitimacy of such controls remains a hotly contested notion that generally boils down to the perennial weighing of the values of free speech and equality. The debate has clearly been most prolonged and turbulent in the US, due in large part to the sanctity of libertarian values embedded in the First Amendment. There, the chronology of policy and politics around hate speech, and later, hate crime, began nearly a century ago, in the 1920s (Levin 2002) when the American Civil Liberties Union was first established. Since that time, policy-makers and justices have had to grapple with the uneasy balance between the equality provisions of the Fourteenth Amendment, and free speech protections of the First Amendment. The latter has typically prevailed. To support the privileging of expression over equality, proponents generally argue that the ‘offensiveness’ of hate speech is not sufficient reason to restrict its use as it does not typically rise to the level of threat. Proponents of regulation of hate speech counter that there is harm inherent in offensive speech acts, to the extent that they silence those to whom such speech is directed. Among the most active scholars engaged in this controversy are S. B. Gellman and Frederick Lawrence. In 2004, they collaborated on an article – ‘Agreeing to Agree: A Proponent and Opponent of Hate Crime Laws Reach for a Common Ground’ – which aired both sides of the argument. In the end, the two do ‘agree to agree’, coming to a consensus on a ‘third way’ approach that might satisfy both camps. The debate appears to be much more muted in Europe. This is not to say that statutory provisions have been adopted without challenge, or that they have been uniformly adopted. Indeed, the revision of the UK’s Crime and Disorder Act whereby racial violence is legally recognised has not been without its detractors (Burney 2003; Tausz and Ormerod 2007). Moreover, not all western nations have followed suit in confronting hate crime via legislative routes. Bleich’s (2007) analysis makes this abundantly clear, as does a recent edited volume entitled Racist Victimisation: International Reflections and Perspectives (Winterdyk and Antonopoulos 2008). For example, Bleich observes the distinctions between the UK, Germany and France, which have invoked, respectively, statutory, community activist and educational interventions. Obviously, statutory provisions require the subsequent mobilisation of police and prosecutorial resources to fully activate their potential. 27

Hate Crime

Arguably, it is law enforcement practices that most directly affect the recording of hate crime. In the past six years or so this has become a very popular area of enquiry. Much scholarship has, in fact, focused precisely on the ways in which police negotiate the ambiguities of the law (Stanko 2001; Nolan et al. 2004; Grattet and Jenness 2005). Specifically, research has focused on the institutional factors that shape the enactment of hate crime legislation on the street. These are characterised by an array of internal structural factors such as the presence of formal protocol, training initiatives, support of leadership, and police resources (Balboni and McDevitt 2001; Nolan and Akiyama 2002). Parallel work has focused on environmental factors that also shape enforcement: state policies, political party in power, public opinion, and perceptions of violent crime rates, for example (HaiderMarkel 2001; Soule and Earl 2001; Jenness and Grattet 2005). Sadly, very few scholars have examined the role of police subcultural biases in the enforcement of hate crime provisions, thereby neglecting the impact of ‘societal power dynamics and social inequalities’ on police decision-making (Franklin 2002). In contrast to the growing literature on policing, there have been relatively few attempts to evaluate subsequent prosecution. Some have argued that there is a very practical reason for this: few cases make it that far. Given that police – as gatekeepers – are uneven at best in their willingness or ability to identify hate crime as such, it is thus not surprising that hate crime prosecutions are rare (Franklin 2002; Burney 2003; Hate Crimes Community Working Group 2006). The impact of police actions and decisions comes to the fore in this literature. However, equally important are the facts that prosecutors face a difficult tension in balancing ‘hatred’ against free speech protections, and that they find it exceedingly difficult to ‘prove’ motive (McPhail and Jenness 2005/2006). One particular field of hate studies has initiated a great deal of discussion on legislative controls: the regulation and control of cyberhate. In many cases, existing legislation has been invoked, as in the case of defamation laws, incitement to hatred policies, and human rights legislation. Bailey (2006) urges the innovative application of intellectual property law, libel law, and even the filing of union grievances in work places (such as libraries) that are exposed to online hate. More broadly, however, scholars and activists point to the necessity of international initiatives and the global harmonisation of laws in order to effectively address hate crime. As mentioned earlier, cyberhate knows no boundaries: its perpetrators are anonymous and fluid; its messages globally available. Any single nation is powerless 28

The more things change … post-9/11 trends in hate crime scholarship

to battle the movement on its own. Only by sharing resources and regulatory strategies and personnel can they make inroads to curb cyberhate. Moreover, the law is not the only – or perhaps even the most effective – weapon available to counter cyberhate. Bailey (2006) suggests four key mechanisms that can supplement legal intervention: filtering, monitoring organisations (such as the Simon Wiesenthal Center, the Southern Poverty Law Center), hate speech hotlines, and ISP self-regulation. There are lessons to be learned from the nature of the cyberhate scholarship, to the extent that strong emphasis is placed on strategies that lie outside the realm of statutory regulation. This is an area that is only just emerging in the broader literature on hate crime. Sadly, relatively little attention has been paid to either developing or evaluating social policy initiatives. The primary cause for optimism has been the work on school-based anti-hate programming. Moreover, there is a tendency in this literature to link, empirically and theoretically, hate crime, harassment and bullying. In fact, Englander (2007) convincingly draws three parallels between hate crime and bullying: focus on ‘difference’ as a motivator; lack of respect for and justification of violence against others; and concentration of such behaviours among youth. The natural conclusion, then, is that both sets of behaviours may be amenable to similar interventions that focus on fostering positive interpretations of difference. This is not an atypical response to school-based hate crime. Indeed, Englander’s analysis is shared by the likes of Wessler and DeAndrade (2006), and Cobia and Carney (2002), who identify verbal bullying and harassment (real-time and online) as precursors to hate crime. They too argue for initiatives that directly address bias and its attendant manifestations through school-wide programmes promoting respect, responsibility and tolerance. Interestingly, there are those who advocate a similar approach at the community level. The focus here is on community organising as a means of resisting the encroachment of racist groups and individuals. Wassmuth and Bryant (2001/2002) and Rabrenovic (2007; see also Levin and Rabrenovic 2004) are illustrative. Both profile specific community reactions against emerging patterns of racism and hate violence, emphasising the need for enhancing public awareness of both the violence and the need to respond to it. Their work highlights the role that the media can and should be encouraged to play in representing difference in a positive light. Beyond these areas of enquiry, there is very little recent scholarship on non-traditional interventions. Mark Umbreit and his colleagues 29

Hate Crime

have made some forays into the use of restorative justice models in the context of hate crime (Umbreit et al. 2003; Coates et al. 2006). And there have been a handful of articles addressing victim services (Dunbar 2001; Craig-Henderson and Sloan 2003; Jalota 2004). In contrast, there is virtually no literature that explores direct nonpunitive interventions with hate crime offenders. Thus, if there is any single area in need of further development it is most certainly the wide open field of anti-hate initiatives that lie outside the criminal justice system. We would be well advised to engage in significantly more programme development and assessment.

Conclusion: promoting hate studies My review of the post-9/11 scholarship on hate crime leads me to conclude that, while out of its infancy, the field has not yet matured into adulthood. There is still considerable ground to cover. I don’t see this necessarily as cause for despair so much as it is inspiring for those of us working in the field. It means that there are niches that we can carve out, contributions that are still to be made. There are, I would argue, several areas in which there is great scope for development, beginning with consideration of the broader impacts of hate crime. It is ironic that one of the key assumptions underlying the discrete consideration of hate crime – that it affects more than just the immediate victim – has not been the focus of concentrated attention. There is a great deal to be uncovered about how this affects not just victims’ communities, but the common weal generally. There is also much more to learn about the discrete experiences of different victim groups. This not only opens doors for work within distinct communities, but also comparative work across communities. What are the differences between groups? What are the similarities? How do victims make sense of their victimisation? How does it alter their sense of self, or their sense of belonging within the group and within their national context? On the other side of the equation is work on offenders. We assume a great deal about motivations, or about the need for empowerment, but we don’t really have a great deal of empirical evidence to support those contentions. As difficult as it can be to sample victim populations, it is perhaps even more difficult to solicit participants who self-identify as hate crime offenders. Sampling incarcerated offenders is difficult on at least two counts: lack of relevant convictions, and lack of access to institutions. We must get creative in identifying meaningful samples of 30

The more things change … post-9/11 trends in hate crime scholarship

offenders, especially those who are not affiliated with organised hate groups. Last on my wish-list is more scholarship and activity around nonpunitive interventions. Purely punitive responses have the potential to be counter-productive. Franklin (2002: 166) observes that: It remains an open question as to whether penalty enhancements will lead to increased tolerance of minorities among the general public. Some critics argue that the laws, although well intentioned, may actually increase the social divisions they are designed to ameliorate … They cite the popular belief that hate crime laws are an example of certain groups receiving special rights not accorded to other citizens. Most offenders are youth, and especially young men who are responding to what they see as a threat – to their community, to their neighbourhood, or to their self-esteem. Often, these threats are more imagined than real. It may be more effective, then, to challenge those myths, and to thus ‘humanise’ the victims and their communities. Incidents of hate crime can be taken as a starting point for education and healing rather than simply punishment. Consequently, communitybased responses represent valuable alternatives. We have not come very far at all in creating such initiatives, let alone evaluating them. Incentive for developing nascent or even missing areas of enquiry may come from the welcome establishment of the Journal of Hate Studies, which was initiated in 2001. The mission of the journal is to ‘promote the sharing of interdisciplinary ideas and research relating to the study of hate, where hate comes from, and how to combat it. It contains the works of authors in the fields of psychology, religious studies, information science and technology, human rights activism, and law.’ It has already proven to be a fruitful venue for diverse publications on hate crime. Moreover, it has begun a dialogue on the movement towards an identifiable ‘field’ of hate studies through publication of articles on precisely this in three of the last four volumes (Stern 2003/2004; Blitzer 2005/2006; Mohr 2007/2008). In the first of this series of papers, Stern (2003/2004) calls for the disciplines to begin to talk to one another to enable a richer, more sophisticated understanding of a very complex phenomenon. As Stern (2003/2004: 32) insists, the development of an interdisciplinary field of hate studies

31

Hate Crime

… might create a common vocabulary among the various academic disciplines, thus encouraging an integrated system of knowledge and research. It might help us gain a more complete understanding of the various components of hatred, and provide testable theories to guide the actions of individuals, groups, and institutions – including governments. It might spur woefully needed research on education, and not only help debunk myths on which we may well be wasting great resources and energy, but also help define the approaches that work and refine, improve, and institutionalize them. At the risk of shameless self-promotion, I would like to mention a recent project of my own that I hope will in some way also contribute to the development of such a field. I was approached by Praeger to develop a five-volume set on hate crime. Together with an outstanding team of volume editors (Brian Levin, Paul Iganski, Randy Blazak and Frederick Lawrence), I was able to pull together an extensive set of readings that highlights many of the issues discussed here. The simply titled Hate Crimes offers interested readers a comprehensive collection of original articles surveying this phenomenon we have come to know as hate crime. The contributors represent a variety of disciplines, including law, sociology, criminology, psychology and even public health. Moreover, since it is also a global phenomenon, we invited not just American scholars but international contributors as well. This comparative/cross-cultural approach adds an important element to the set. It reminds readers that hate crime is a universal problem, and that approaches taken elsewhere might be of use to North Americans. As I was developing this project, and thinking about possible contributors, I was already beginning to identify some of the themes that have emerged in this chapter. It is hoped that the articles included in the set have brought together in one place the most current thinking on both well-established and newly emerging issues. It is also hoped that it furthers the dialogue begun within the pages of the Journal of Hate Studies that will ultimately culminate in a discrete field capable of responding to unanswered questions.

References Adams, M. and Toth, C. (2006) ‘Unanticipated Consequences of Hate Crime Legislation’, Judicature, 90(3): 129–34. 32

The more things change … post-9/11 trends in hate crime scholarship

Alden, H. and Parker, K. (2005) ‘Gender Role Ideology, Homophobia and Hate Crime: Linking Attitudes to Macro-Level Anti-Gay and Lesbian Hate Crimes’, Deviant Behavior, 26(4): 321–43. Asquith, N. (2004) ‘In Terrorem: ‘With their Tanks and their Bombs, and their Bombs and their Guns, in Your Head’, Journal of Sociology, 40(4): 400–16. Bailey, J. (2006) ‘Strategic Alliances: The Inter-Related Roles of Citizens, Industry and Government in Combating Internet Hate’, Canadian Issues, Spring: 56–9. Balboni, J. and McDevitt, J. (2001) ‘Hate Crime Reporting: Understanding Police Officer Perceptions, Departmental Protocol, and the Role of the Victim: Is There Such a Thing as a “Love” Crime?’, Justice Research and Policy, 3(1): 1–27. Blazak, R. (2001) ‘White Boys to Terrorist Men’, American Behavioral Scientist, 44(6): 982–1000. Blazak, R. (2004) ‘ “Getting It”: The Role of Women in Male Desistance from Hate Groups’, in A. Ferber (ed.) Home-Grown Racism. New York: Routledge (pp. 161–80). Blee, K. (2004) ‘Women and Organized Racism’, in A. Ferber (ed.) HomeGrown Racism. New York: Routledge (pp. 49–74). Blee, K. (2005) ‘Racial Violence in the United States’, Ethnic and Racial Studies, 28(4): 599–619. Blee, K. (2007) ‘The Microdynamics of Hate Violence: Interpretive Analysis and Implications for Responses’, American Behavioral Scientist, 51(2): 258–70. Bleich, E. (2007) ‘Hate Crime Policy in Western Europe: Responding to Racist Violence in Britain, Germany, and France’, American Behavioral Scientist, 51(2): 149–65. Blitzer, J. (2005/2006) ‘Toward an Interdisciplinary Field of Hate Studies: Developing a Framework’, Journal of Hate Studies, 4: 139–49. Bocji, P. and McFarlane, L. (2003) ‘Cyberstalking: The Technology of Hate’, Police Journal, 76(3): 204–21. Boeckmann, R. and Liew, J. (2002) ‘Hate Speech: Asian American Students’ Justice Judgment and Psychological Responses’, Journal of Social Issues, 58(2): 363–81. Bradley, M. (2007) ‘Symbolizing Hate: The Extent and Influence of Organized Hate Group Indicators’, Journal of Crime and Justice, 30(1): 1–15. Burney, E. (2003) ‘Using the Law on Racially Aggravated Offences’, Criminal Law Review, January: 28–36. Byers, B. and Crider, B. (2002) ‘Hate Crimes against the Amish: A Qualitative Analysis of Bias Motivation using Routine Activities Theory’, Deviant Behavior, 23(2): 115–48. Byers, B. and Jones, J. (2007) ‘The Impact of the Terrorist Attacks of 9/11 on Anti-Islamic Hate Crime’, Journal of Ethnicity in Criminal Justice, 5(1): 43–68. Chakraborti, N. and Garland, J. (2004) Rural Racism. Portland, OR: Willan Publishing. 33

Hate Crime

Chakraborti, N. and Garland, J. (2009) Hate Crime: Impact, Causes and Responses. London: Sage. Chorba, C. (2001) ‘The Danger of Federalizing Hate Crimes: Congressional Misconceptions and the Unintended Consequences of the Hate Crimes Prevention Act’, Virginia Law Review, 87(2): 319–79. Coates, R. B., Umbreit, M. S. and Vos, B. (2006) ‘Responding to Hate Crime through Restorative Justice Dialogue’, Contemporary Justice Review: Issues in Criminal, Social and Restorative Justice, 9(1): 7–21. Cobia, D. and Carney, J. (2002) ‘Creating a Culture of Tolerance in Schools: Everyday Actions to Prevent Hate-Motivated Violent Incidents’, Journal of School Violence, 1(2): 87–104. Cowan, G. and Mettrick, J. (2002) ‘The Effects of Target Variables and Setting on Perceptions of Hate Speech’, Journal of Applied Social Psychology, 32: 247–63. Craig, K. (2002) ‘Examining Hate-Motivated Aggression: A Review of the Social-Psychological Literature on Hate Crimes as a Distinct Form of Aggression’, Aggression and Violent Behavior, 7(1): 85–101. Craig-Henderson, K. and Sloan, L. R. (2003) ‘After the Hate: Helping Psychologists Help Victims of Racist Hate Crime’, Clinical Psychology: Science and Practice, 10(4): 481–90. Cunneen, C. (2001) Conflict, Politics and Crime: Aboriginal Communities and the Police. Crow’s Nest, NSW: Allen and Unwin. D’Alessio, S., Stolzenberg, L. and Eitle, D. (2002) ‘The Effect of Racial Threat on Interracial and Intraracial Crimes’, Social Science Research, 31(3): 392–408. D’Augelli, A., Grossman, A. and Starks, M. (2006) ‘Childhood Gender Atypicality, Victimization, and PTSD among Lesbian, Gay, and Bisexual Youth’, Journal of Interpersonal Violence, 21(11): 1462–82. Dobratz, B. and Shanks-Meile, S. (2004) ‘The White Supremacist Movement: Worldviews on Gender, Feminism, Nature and Change’, in A. Ferber (ed.) Home-Grown Racism. New York: Routledge (pp. 113–42. Dunbar, E. (2001) ‘Counseling Practices to Ameliorate the Effects of Discrimination and Hate Events: Toward a Systematic Approach to Assessment and Intervention’, Counseling Psychologist, 29(2): 279–307. Dunbar, E. (2006) ‘Race, Gender, and Sexual Orientation in Hate Crime Victimization: Identity Politics or Identity Risk?’, Violence and Victims, 21(3): 323–37. Englander, E. (2007) ‘Is Bullying a Junior Hate Crime? Implications for Interventions’, American Behavioral Scientist, 51(2): 205–12. Etter, G. (2001) ‘Totemism and symbolism in the White Supremacist Movement: Images of an Urban Tribal Warrior Culture’, Journal of Gang Research, 8(2): 49–75. Ferber, A. (2004) Home-Grown Hate: Gender and Organized Racism. New York: Routledge. Flint, C. (ed.) (2004a) Spaces of Hate: Geographies of Discrimination and Intolerance in the USA. New York: Routledge. 34

The more things change … post-9/11 trends in hate crime scholarship

Flint, C. (2004b) ‘Introduction: Spaces of Hate: Geographies of Discrimination and Intolerance in the USA’, in C. Flint (ed.) Spaces of Hate: Geographies of Discrimination and Intolerance in the USA. New York: Routledge (pp. 1–19). Franklin, K. (2002) ‘Good Intentions: The Enforcement of Hate Crime PenaltyEnhancement Statutes’, American Behavioral Scientist, 46(1): 154–72. Franklin, K. (2003) ‘Homophobia and the “Matthew Shepard Effect” in Lawrence v. Texas’, New York Law School Law Review, 48(4): 657–95. Gadd, D. (2006) ‘The Role of Recognition in the Desistence Process: A Case Analysis of a Former Far-Right Activist’, Theoretical Criminology, 10(2): 179–202. Gellman, S. B. and Lawrence, F. (2004) ‘Agreeing to Agree: A Proponent and Opponent of Hate Crime Laws Reach for Common Ground’, Harvard Journal On Legislation, 41(2): 421–48. Gerstenfeld, P. (2004) Hate Crimes: Causes, Controls and Controversies. Thousand Oaks, CA: Sage. Gerstenfeld, P., Grant, D. and Chiang, C. (2003) ‘Hate Online: A Content Analysis of Extremist Internet Sites’, Analyses of Social Issues and Public Policy, 3(1): 29–44. Grascia, A. (2003) ‘White Supremacy Music: What Does It Mean to our Youth?’, Journal of Gang Research, 10(2): 25–31. Grattet, R. and Jenness, V. (2005) ‘Examining the Boundaries of Hate Crime Law: Disabilities and the “Dilemma of Difference” ’, Journal of Criminal Law and Criminology, 91(3): 653–98. Haider-Markel, D. (2001) ‘Implementing Controversial Policy: Results From a National Survey of Law Enforcement Department Activity on Hate Crime’, Justice Research and Policy, 3(1): 29–61. Halafoff, A. (2006) ‘UnAustralian Values’, Proceedings of UNAUSTRALIA Conference, University of Canberra. Hamer, F. M. (2006) ‘Racism as a Transference State: Episodes of Racial Hostility in the Psychoanalytic Context’, Psychoanalytic Quarterly, 75(1): 197–214. Hartley, J. and Green, J. (2006) ‘The Public Sphere on the Beach’, European Journal of Cultural Studies, 9(3): 341–62. Hate Crimes Community Working Group (2006) Addressing Hate Crime in Ontario. Toronto: Attorney General and Minister of Community Safety and Correctional Services. Herek, G., Cogan, J. and Gillis R. (2002) ‘Victim Experiences in Hate Crimes based on Sexual Orientation’, Journal of Social Issues, 58(2): 319–39. Hicks, W. (2004) ‘Skinheads: A Three Nation Comparison’, Journal of Gang Research, 11(2): 51–73. Ho, C. (2007) ‘Muslim Women’s New Defenders: Women’s Rights, Nationalism and Islamophobia in Contemporary Australia’, Women’s Studies International Forum, 30(4): 290–8. 35

Hate Crime

Hoover, K. and Johnson, V. (2003/2004) ‘Identity-Driven Violence: Reclaiming Civil Society’, Journal of Hate Studies, 3(1): 83–94. Iganski, P. (2002) The Hate Debate: Should Hate be Punished as a Crime. London: Profile Books. Iganski, P. (2007) ‘Too Few Jews to Count? Police Monitoring of Hate Crime against Jews in the United Kingdom’, American Behavioral Scientist, 51(2): 232–45. Jalota, S. (2004) ‘Supporting Victims of Rural Racism: Learning Lessons from a Dedicated Racial Harassment Project’, in N. Chakraborti and J. Garland (eds) Rural Racism. Portland, OR: Willan Publishing (pp. 143–60. Jannson, K. (2006) Black and Minority Ethnic Groups’ Experiences and Perceptions of Crime, Racially Motivated Crime and the Police: Findings from the 2004/05 British Crime Survey. London: Home Office Research Development and Statistics Directorate. Janoff, D. (2005) Pink Blood: Homophobic Violence in Canada. Toronto: University of Toronto Press. Jenness, V. and Grattet, R. (2005) ‘The Law-In-Between: The Effects of Organizational Perviousness on the Policing of Hate Crime’, Social Problems, 52(3): 337–59. Kaplan, J. (2006) ‘Islamophobia in America? September 11 and Islamophobic Hate Crime’ Terrorism and Political Violence, 18(1): 1–33. Lattas, J. (2007) ‘Cruising: “Moral Panic” and the Cronulla Riot’, Australian Journal of Anthropology, 18(3): 320–35. Lee, Y. T., Vue, S, Seklecki, R. and Ma, Y. (2007) ‘How Did Asian Americans Respond to Negative Stereotypes and Hate Crime’, American Behavioral Scientist, 51(2): 271–93. Levin, B. (2002) ‘From Slavery to Hate Crime Laws: The Emergence of Race and Status-Based Protection in American Criminal Law’, Journal of Social Issues, 58(2): 227–45. Levin, J. and Rabrenovic, G. (2004) Why We Hate. Amherst, NY: Prometheus. Lyons, C. (2007) ‘Community (Dis)organization and Racially Motivated Crimes’, American Journal of Sociology, 113(3): 815–63. Lyons, C. (2008) ‘Defending Turf: Racial Demographics and Hate Crime against Blacks and Whites’, Social Forces, 87(1): 357–85. Mason, G. (2007) ‘Hate Crime as a Moral Category: Lessons from the Snowtown Case’, Australian and New Zealand Journal of Criminology, 40(3): 249–62. McMahon, B., West, S., Lewis, A., Armstrong, A. and Conway, J. (2004) ‘Hate Crimes and Disability in America’, Rehabilitation Counseling Bulletin, 47(2): 66–75. McPhail, B. and Jenness, V. (2005/2006) ‘To Charge or Not to Charge? That Is the Question: The Pursuit of Strategic Advantage in Prosecutorial Decision-Making Surrounding Hate Crime’, Journal of Hate Studies, 4(1): 89–119. 36

The more things change … post-9/11 trends in hate crime scholarship

Miller, A. (2001) ‘Student Perceptions of Hate Crimes’, American Journal of Criminal Justice, 25(2): 293–305. Mohr, J. (2007/2008) ‘Hate Studies through a Constructivist and Critical Pedagogical Approach’, Journal of Hate Studies, 6(1): 65–80. Noelle, M. (2002) ‘The Ripple Effect of the Matthew Shepard Murder’, American Behavioral Scientist, 46(1): 27–50. Nolan, J. and Akiyama, Y. (2002) ‘Assessing the Climate for Hate-Crime Reporting in Law Enforcement Organizations: A Force-Field Analysis’, Justice Professional, 15(2): 87–103. Nolan, J., McDevitt, J. and Cronin, S. (2004) ‘Learning to See Hate Crime: A Framework for Understanding and Clarifying Ambiguities in Bias Crime Classification’, Criminal Justice Studies, 17(1): 91–105. Olivero, J. M. and Rodrigo M. (2001) ‘Homophobia and University Law Enforcement Students’, Journal of Criminal Justice Education, 12(2): 271–81. Otis, M. (2007) ‘Perceptions of Victimization, Risk and Fear of Crime among Lesbians and Gay Men’, Journal of Interpersonal Violence, 22(2): 198–217. Parks, C. and Woodson, K. (2002) ‘Anxiety Symptoms Among Sexually Abused Ethnic Minority Male Survivors of Racially Motivated Hate Crimes: An Exploratory Study’, Family Violence and Sexual Assault Bulletin, 18(2): 13–19. Perry, B. (2002), ‘Where Do We Go From Here? Future Directions in Hate Crime Scholarship’, Keynote Address, First Annual Hate Crime Conference, Nottingham University, February. Perry, B. (2003a) ‘Where Do We Go From Here? Future Directions in Hate Crime Scholarship’, Internet Journal of Criminology. Available at www. internetjournalofcriminology.com/Where%20Do%20We%20Go%20From%2 0Here.%20Researching%20Hate%20Crime.pdf Perry, B. (2003b) Hate and Bias Crime: A Reader. New York: Routledge. Perry, B. (2004) ‘ “White Genocide”: White Supremacists and the Politics of Reproduction’, in A. Ferber (ed.) Home-Grown Racism. New York: Routledge (pp. 75–96). Perry, B. (2008) Silent Victims: Hate Crime against Native Americans. Tucson AZ: University of Arizona Press. Perry, B. (2009) (ed.) Hate Crimes, five vols, Westport, CT: Praeger. Poynting, S. (2006) ‘What Caused the Cronulla Riot?’, Race and Class, 48(1): 85–92. Poynting, S. and Perry, B. (2007) ‘Climates of Hate: Media and State Inspired Victimisation of Muslims in Canada and Australia since 9/11’, Current Issues in Criminal Justice, 19(2). Prum, M., Deschamps, B. and Barbier, M. C. (2007) Racial, Ethnic and Homophobic Violence: Killing in the Name of Otherness. New York: RoutledgeCavendish. Rabrenovic, G. (2007) ‘When Hate Comes to Town: Community Response to Violence against Immigrants’, American Behavioral Scientist, 51(2): 349–60. 37

Hate Crime

Saucier, D., Hockett, J. and Wallenberg, A. (2008) ‘The Impact of Racial Slurs and Racism on the Perceptions and Punishment of Violent Crime’, Journal of Interpersonal Violence, 23(5): 685–702. Sheridan, L. (2006) ‘Islamophobia Pre- and Post-September 11th, 2001’, Journal of Interpersonal Violence, 21(3): 317–36. Shively, M. and Mulford, C. (2007) ‘Hate Crime in America: The Debate Continues’, National Institute of Justice Journal, 257: 8–13. Smith, A. (2005) Conquest: Sexual Violence and American Indian Genocide. Boston, MA: South Side Press. Soule, S. A. and Earl, J. (2001) ‘The Enactment of State-Level Hate Crime Law in the United States: Intrastate and Interstate Factors’, Sociological Perspectives, 44(3): 281–305. Stanko, E. (2001) ‘Reconceptualizing the Policing of Hatred: Confessions and Worrying Dilemmas of a Consultant’, Law and Critique, 12(3): 309–29. Stern, K. (2001/2002) ‘Hate and the Internet’, Journal of Hate Studies, 1(1): 57–108. Stern, K. (2003/2004) ‘The Need for an Interdisciplinary Field of Hate Studies’, Journal of Hate Studies, 3(1): 7–36. Sullaway, M. (2004) ‘Psychological Perspectives on Hate Crime’, Psychology, Public Policy, and Law, 10(3): 250–92. Tausz, D. and Ormerod, D. (2007) ‘Whether Using Words “Bloody Foreigners” and “Go Back to Your Own Country” has Effect of Turning Offence contrary to s. 4A of the Public Order Act 1986 into Racially Aggravated Offence Contrary to s. 31(1)(a) of the Crime and Disorder Act 1998’, Criminal Law Review, July: 579–81. Tomsen, S. (2006) ‘Homophobic Violence, Cultural Essentialism and Shifting Sexual Identities’, Social and Legal Studies, 15(3): 389–407. Umbreit, M. S., Lewis, T. and Burns, H. (2003) ‘A Community Response to a 9/11 Hate Crime: Restorative Justice through Dialogue’, Contemporary Justice Review, 6(4): 383–91. Wachholz, S. (2005) ‘Hate Crimes against the Homeless: Warning-out New England Style’, Journal of Sociology and Social Welfare, 32(4): 141–63. Wassmuth, B. and Bryant, M. J. (2001/2002) ‘Not in Our World’, Journal of Hate Studies, 1(1): 109–32. Watts, M. W. (2001) ‘Aggressive Youth Cultures and Hate Crime – Skinheads and Xenophobic Youth in Germany’, American Behavioral Scientist, 45(4): 600–15. Weatherby, G. and Scoggins, B. (2005/2006) ‘A Content Analysis of Persuasion Techniques used on White Supremacist Websites’, Journal of Hate Studies, 4(1): 9–31. Wessler, S. and DeAndrade, L. (2006) ‘Slurs, Stereotypes, and Student Interventions: Examining the Dynamics, Impact, and Prevention of Harassment in Middle and High School’, Journal of Social Issues, 62(3): 511–32. Willis, D. (2004) ‘Hate Crimes against Gay Males: An Overview’, Issues in Mental Health Nursing, 25(2): 115–32. 38

The more things change … post-9/11 trends in hate crime scholarship

Winterdyk, J. and Antonopoulos, G. (eds) (2008) Racist Victimisation: International Reflections and Perspectives. Aldershot: Ashgate. Zaversek, D. (2003) ‘Invisible Violence: Normativity and Normalisation of Violence against People with Motor, Sensorial and Intellectual Disabilities’, Revija Za Kriminalistiko in Kriminologijo, 54(1): 2–13.

39

Chapter 2

The victimisation of goths and the boundaries of hate crime Jon Garland

Introduction A number of disturbing murders of people from minority backgrounds have hit the headlines in the United Kingdom in the last five years or so, including that of black teenager Anthony Walker in 2005 in Liverpool, openly gay Jody Dobrowski in London in the same year, and the fatal assault upon a man with mental health issues, Brent Martin, in Sunderland in 2007. The brutal nature of these incidents shocked other members of the victims’ minority communities and the police that investigated them. The three killings were also generally viewed as being clear-cut examples of different forms of hate crime, whether racist, homophobic or disablist in nature (Chakraborti and Garland 2009). Another similar case was that of the horrific murder of Sophie Lancaster in Lancashire in the late summer of 2007. Lancaster, a member of the goth subculture, was attacked and killed by a group of youths for no other reason than her hair, make-up and clothes made her stand out from the norm. She was ‘different’, and her startling appearance precipitated her victimisation. Her goth boyfriend, Robert Maltby, was badly beaten in the same incident. As Mark Hodkinson posits, ‘Sophie Lancaster did not die because of her race, religion or sexuality. She died because she was a goth’ (Hodkinson 2008: 29).1 What makes this case especially perplexing for hate crime scholars is that at the trial of Lancaster’s attackers the judge, Anthony Russell QC, stated that not only were Lancaster and Maltby targeted solely 40

The victimisation of goths and the boundaries of hate crime

‘because their appearance was different’, but he also made a point of labelling the murder a ‘hate crime’ (BBC News 2008a). He was therefore suggesting that for someone to be considered a victim of a hate crime they need not necessarily be a member of one of the ‘established’ and generally recognised hate crime victim groups, such as the minority ethnic, gay or disabled communities, that have a history of marginalisation and discrimination (Perry 2001). Instead, it was enough for the judge that Lancaster and Maltby had been deliberately singled out because of their actual or perceived difference, even if this difference was due to their membership of a subcultural grouping rather than them being from a minority ethnic or sexual background, for example. This chapter explores this viewpoint in further detail by assessing whether the harassment and victimisation of goths or those who engage in ‘alternative’ subcultures2 can actually be bracketed with other crimes, such as racist, religious or homophobic assaults and abuse, which scholars and criminal justice agencies routinely categorise as hate crimes. It begins with necessarily brief discussions of the hate crime framework that will be used in the rest of the chapter and of the characteristics of the goth subculture itself, including notions of collective goth identity. It then moves on to consider the Lancaster case at greater length and places it within the broader context of the targeted victimisation of goths. Following this, the chapter draws parallels between this type of victimisation and other forms of hate crime, including the othering of ‘despised’ societal outgroups and the ‘fear of difference’ that may explain some of this harassment. The chapter concludes that it may be time to develop a definition of hate crime that is not predicated on the victim being a member of a historically marginalised and disadvantaged group but instead prioritises the reasons behind their victimisation.

The hate crime framework The use of the term ‘hate crime’ to describe the prejudicial victimisation of minority communities has a more established history in the United States than it has in the United Kingdom. It is generally accepted that its roots in the US are traceable to the civil rights campaigns of various minority groups in the 1960s and 1970s when it was used as a way of drawing out the commonalities of those struggles shared between different oppressed groups (Grattet and Jenness 2003). However, in the UK it has gained common usage among academics 41

Hate Crime

and practitioners only in the last decade or so, largely as a result of the bombing campaign of neo-Nazi David Copeland in 1999 and also the publication of the Macpherson Report in the same year, which brought issues of diversity and discrimination to the forefront of criminal justice policy (Rowe 2004). Despite the rise in currency of the hate crime concept there is little agreement on what actually constitutes a ‘hate crime’. For authors like Gerstenfeld the primary characteristic of such a crime is not that it is motivated by hate per se but that the victim is targeted because of their group affiliation (Gerstenfeld 2004). Hate crimes are thus ‘stranger danger’ occurrences in which the perpetrator does not know the victim as an individual at all; they are picked out solely on the basis of the victim’s actual or perceived membership of a social grouping that the perpetrator despises. They therefore damage the self-worth, confidence and feelings of security of the victim more than ‘ordinary’ crimes because victims are targeted due to their intrinsic identity: something that is central to their sense of being and that they cannot alter (Iganski 2008). Hate crimes have another important facet: they are enacted to subordinate not just the victim but their wider grouping; they are thus ‘message crimes’ designed to intimidate and frighten whole communities that are different, in some way, from the norm (Perry 2001). They are perpetrated to maintain society’s hierarchical power relations, with the more powerful victimising the less in order to maintain their privileged position. Such harassment reflects broader social attitudes and values that reproduce and maintain this inequality, as Perry (2009: 71) asserts: … hate crime provides a context in which the perpetrator can reassert his or her hegemonic identity and, at the same time, punish the victim for the individual or collective performance of his or her identity. In other words, hate-motivated violence is used to sustain the privilege of the dominant group and to police the boundaries of the group by reminding the Other of his or her place. Perpetrators thus recreate their own masculinity, or whiteness, for example, while punishing the victims for their deviant identity performance. Perry offers a broader and more structural notion of hate crime that suggests that offenders perpetrate such crimes to reinforce their own dominant social position while reinforcing the subordinate position of ‘othered’ groups. Her ideas of the punishment of identity performance 42

The victimisation of goths and the boundaries of hate crime

and the ‘policing’ of boundaries are particularly instructive, and thus her notion of hate crime will be adopted and then debated within the context of the victimisation of goths. It is to a discussion of this subculture that this chapter now turns.

Goth subculture As mentioned above, Sophie Lancaster was victimised because of her striking ‘alternative’ and gothic appearance. Goth itself emerged as a distinct ‘spectacular’ youth culture in the late 1970s and early 1980s, developing a rather doom-laden style that ‘revolved around a general emphasis on artefacts, appearances and music deemed suitably dark’ (Hodkinson 2002: 41). The nascent goth scene adopted bands such as Bauhaus, Siouxsie and the Banshees and, later, Sisters of Mercy, Fields of the Nephilim, The Mission and The Cure, who attracted a growing and mixed-gender audience noted for their ‘deathly pallor, backcombed or ratted black hair, ruffled Regency shirts, stovepipe hats, leather garments, spiked dog collars, the ensemble accessorized with religious, magical or macabre jewellery’ (Reynolds 2005: 423). Paul Hodkinson (2002) emphasises goth’s most important stylistic features: the ‘sombre and the macabre’ (denoted by black, sometimes vampiric, clothing, stark make-up and sinister music), its ‘femininity and ambiguity’ (typified by the androgynous appearance of many males and the ‘hyperfeminine’ look of females3) and its ability to adapt other youth cultures, such as punk, metal, industrial, new romantic and indie, to evolve several different stylistic, but still clearly gothic, strands of its own. The goth subculture enjoyed a peak of popularity in the mid to late 1980s but has since maintained a significant, if more underground, presence within broader popular culture (Brill 2007). Its followers, typically white middle class but split evenly between male and female (Hodkinson 2002: 70), have created and sustained a scene that is distinctive, tight-knit and proud of its ‘difference’, one that balances expressions of individuality within a collective sense of shared cultural norms, in a similar fashion to other distinctive subcultures such as punk (Worley 2010). It revolves around specialist clothes shops, clubs, pubs and websites, and has as a central attraction the biannual gathering of goths at the ‘Whitby Goth Weekender’, held in the coastal Yorkshire town famously associated with the fictional arrival of the vampire Dracula in Bram Stoker’s novel of the same name (Hodkinson 2002). 43

Hate Crime

Interestingly, and in contradiction to some of the ‘postmodern’ takes on contemporary subcultures which suggest that levels of commitment to them are often low (see Muggleton 2000), over the last two decades goth has become, for many of its devotees, less of a youth subculture and more of a ‘way of life’ that permeates almost every aspect of their daily activities and that stays with them for decades (Telford 2008). This chimes with Paul Hodkinson’s (2004) assertion that goths have a strong and well-developed sense of group subcultural identity that transcends local and national boundaries. While this identity may not involve shared political goals or beliefs, for example, it nevertheless possesses, as an important aspect of its collective values, a tolerance of those who express their individuality through performing unconventional gender roles or those who belong to minority sexualities (Brill 2008). For many goths, this tolerance of difference, coupled with their non-aggressive outlook, demarcates them from those in the mainstream, the so-called ‘trendies’ or ‘townies’ who are regarded as dull, intolerant and violent. Indeed, Hodkinson notes that this demarcation can be marked by the drawing by goths of distinct boundaries separating them from these ‘trendies’ and indeed all outsiders, who are often viewed with ‘intense suspicion’ (Hodkinson 2005: 135). Contemporary goth activity therefore operates mostly autonomously from mainstream and other stylistic subcultures, reinforcing the significant sense of collective affiliation between goths and their distance from ‘conventional others’. Thus, for Hodkinson (2002) goth is characterised by the combination of the externally visible difference between those within the scene and those without, the shared sense of identity between goths and the high levels of commitment to that subculture demonstrated by its followers. He sees goth subculture as being something that can both reflect and dominate participants’ identities and sense of self, as one of his male goth interviewees claimed: ‘It is the most important thing in my life, there’s no doubt about it, it is the most important thing in my life – I couldn’t fathom existing without it at all’ (2002: 73). The above factors, if taken together, would appear to indicate that goth is more than merely an ephemeral youth fad that young people might adopt for a while before moving on, and instead is something more long term and meaningful than that. Goths see themselves as part of a unified and long-standing culture that stands apart from the norm. It is this sense of distinctiveness, strong group identification and history that arguably allows those that belong to this subculture to consider themselves as part of a ‘community’ that, like disabled or 44

The victimisation of goths and the boundaries of hate crime

transgender communities, could be classified as a hate crime victim group. Similarly, it is goths’ sense of separation, and their distinctive and challenging appearance, that makes them an easily identifiable ‘outgroup’ that is prone to victimisation. Interestingly, though, Sophie Lancaster and Robert Maltby, while almost always wearing black and sporting clothes and make-up clearly identifiable as ‘goth’, did not like to think of themselves as belonging to a rigid goth subculture (Hodkinson 2008). However, their startling appearance caused them to be victimised for their ‘difference’ on more than one occasion before the violent assault upon them in August 2007. This chapter will now discuss the killing of Sophie in more detail before outlining the broader victimisation of goths.

The murder of Sophie Lancaster In summer 2007 Robert Maltby (aged 21) and Sophie Lancaster (20), who had been together as a couple for over two years, were living in King Street in the town of Bacup, Lancashire (Court of Appeal 2008). On the night of 10 August, after spending the evening at a friend’s house in another part of the town, they decided to walk home, stopping off at a 24-hour petrol station on the way. It was there that they fell into good-natured conversation with a group of local youths, with whom they strolled to the skate-ramp area of nearby Stubbylee Park. Suddenly, and without warning, one of the youths hit Maltby, precipitating a full-scale, ferocious assault upon him by five of the group. The Court of Appeal (2008: 3) described what followed: [Maltby] was brought to the ground by punches and kicks. When on the ground he was kicked viciously to the head and body, and at least one of his assailants stamped on his head … With remarkable courage Miss Lancaster rushed to give whatever assistance to him she could, and as he lay prone, she cradled her boyfriend’s head in her lap, calling for help and shouting at the appellants to leave him alone. [Ryan] Herbert and [Brendan] Harris turned their attention to her and she, too, was subjected to a sustained and vicious attack which involved kicking and stamping until, she too in her turn, was beaten unconscious. A 14-year-old girl described the attackers as ‘just booting them [Maltby and Lancaster]. They were making loads of noises, screaming noises. They were booting them in the head, swinging the leg back 45

Hate Crime

properly booting them, like jumping on them’ (Wainwright 2008: 15). The horrific injuries that the victims sustained were vividly illustrated during a 999 call that a female witness made, which was played to the court at the subsequent trial of the assailants. Crying hysterically throughout, the witness began by pleading for an ambulance: ‘We need an ambulance at Bacup Park, this mosher’s just been banged because he’s a mosher4 … It’s a mosher just been banged for no reason. His girlfriend is on the floor as well. They’re still breathing but they are full of blood. Please just send an ambulance quick. She’s choking on her blood, please will you help us quick.’ After being instructed to try to stop the bleeding, the panicking girl stated: ‘No, it’s all over their hands, all coming out of their eyes, all out their nose and everything. Please just help us quick, please, please’ (BBC News 2008a). After the assault the killers boasted to their friends that they had ‘done summat good’, saying: ‘There’s two moshers nearly dead up Bacup park – you wanna see them – they’re a right mess’ (Wainwright 2008: 15). When paramedics arrived they found the victims lying side by side, covered in blood and unconscious. They had been beaten so severely that paramedics could not gauge what sex they were (Wainwright 2008). Both were rushed to hospital, where Maltby was found to have 22 sites of injury; Lancaster had 17. However, while Maltby began gradually to recover, and was released from hospital on 24 August, Lancaster never regained consciousness. She died in her mother’s arms on the same day Maltby left hospital (Court of Appeal 2008: 4–5). At the subsequent trial Michael Shorrock QC, for the prosecution, said that Maltby and Lancaster had been ‘singled out not for anything they had said or done but because they dressed differently’ (Jenkins 2008: 25). Intriguingly, the judge, Anthony Russell QC, felt that the attack was actually a hate crime, telling the youths: ‘This was a hate crime against these completely harmless people targeted because their appearance was different to yours’ (BBC News 2008b). The two ringleaders during the assault, Brendan Harris and Ryan Herbert, were found guilty of murder and given prison sentences of 18 and 16 years respectively. The other gang members, brothers Joseph and Danny Hulme and Daniel Mallett, were sentenced to between four and five years each for grievous bodily harm with intent upon Maltby. However, in the months after the attack Robert Maltby struggled to come to terms with events, becoming a virtual recluse and undergoing treatment for the ‘serious psychiatric disorder’ caused by the attack (Court of Appeal 2008: 5). 46

The victimisation of goths and the boundaries of hate crime

In a small town like Bacup, economically deprived, geographically remote and almost mono-ethnic, a sense of isolation and a purveying social conservatism means that those whose appearance is radically different stand out starkly, whether they be goths, those from other alternative subcultures, or the small number of families from visible minority ethnic backgrounds. Journalist Mark Hodkinson’s investigation into the nature of the town discovered that minority ethnic families ‘get firebombed out of their houses and given a whack with a baseball bat to make sure they get the message’ to leave (Hodkinson 2008: 31), and suggested that this fear of difference and resentment of ‘outsiders’ was demonstrated through aggressive behaviour towards goths. This hostility towards the ‘other’ characterises many such isolated and monocultural communities and can manifest itself in acts of abuse and violence against minority groups (Garland and Chakraborti 2006). The victimisation of Maltby and Lancaster was part of a wider localised pattern of harassment. On the surface, therefore, the attack upon Maltby and Lancaster appears to share some of the characteristics of those crimes commonly viewed as hate crimes. For instance, the victims were members of a societal ‘outgroup’ that is repeatedly targeted for abuse: indeed, Maltby and Lancaster had been victimised on previous occasions because of their appearance (Purdy 2008), just as the harassment of victims of hate crimes is, in many cases, part of a longer process of victimisation (Bowling 1999). They did not personally know their attackers (meaning that the assault conformed to the notion of hate crimes as ‘stranger danger’ crimes) and the ferocious nature of the assault had many similarities with violent homophobic hate crimes (Chakraborti and Garland 2009). Also, like other forms of hate crime, the surviving victim was deeply affected psychologically and the incident also impacted upon the social confidence of the victims’ wider community (in this case, goths (Purdy 2008)). This chapter will now go on to explore evidence that suggests that the horrific beating of Maltby and Lancaster was part of a distinctive nationwide pattern of victimisation of goths.

The victimisation of goths Following Sophie Lancaster’s death, fellow goths and ‘alternatives’ began a campaign to get attacks against them recognised as hate crimes and included in wider hate crime legislation (Gough 2009). Part of this campaign involved the instigation of an online petition, 47

Hate Crime

which attracted over 7,100 signatures (Prime Minister’s Office, 2008a). The petitioners claimed that: Most gothic/subcultural people regularly complain of being harassed, abused, spat on or attacked for the way they look, having done or said nothing to otherwise provoke this. People who prefer ‘darker’ fashions; who sport facial piercings, alternative makeup or dyed hair regularly experience being singled out on these grounds, by the less tolerant. (Prime Minister’s Office 2008a) The sizeable number of signatures to the petition provides some evidence of significant concern among goths and ‘alternatives’ that they are being repeatedly targeted because of their appearance and group membership. The advent of the ‘Alternatives Have Rights Too’ website5 (that offers a forum for victims of similar crimes to share their experiences) also indicates that this concern may be widespread (Richardson 2008). The wording of the petition, and in particular its emphasis upon the victimisation of those ‘who prefer “darker” fashions’, is indicative of the worry that it is indeed goths and ‘alternatives’, rather than those who are members of other subcultures that do not starkly and visibly stand out from the norm (such as casuals or mods), that are most at risk of attack. The fact that goths are widely perceived as ‘passive’ may also increase their chance of assault as they may be viewed as a ‘soft target’ who will not retaliate, as a goth spokesperson suggests: Goth teenagers can tend to be bookish, quiet and thoughtful. As a result they can sometimes become distanced from the majority and become targeted. In my experience of over 20 years, as a general rule, you have to search hard to find a goth who is violent in nature and even harder still to find one who has not been targeted in some way. These packs so often see a slightly built guy on his own, wearing make-up and to them, he is fair game, an easy target to bully. (Purdy 2008) However, evidence of the precise nature and frequency of the victimisation of goths is hard to come by. A spokesperson for the Stamp Out Prejudice, Hatred and Intolerance Everywhere campaign (SOPHIE), instigated after Lancaster’s death with the purpose of promoting tolerance of difference and the changing of hate crime law, estimated that 70 per cent of people who had contacted the campaign 48

The victimisation of goths and the boundaries of hate crime

had been a victim of some form of attack due to their membership of an alternative subculture (Hodkinson 2008: 32).6 There is also plenty of anecdotal evidence to suggest that the problem is commonplace. During his extensive research into the nature of goth subculture, many of Paul Hodkinson’s interviewees reported suffering ‘prejudice and occasional violence’ from members of the public they termed ‘trendies’, due to their ‘unconventional appearance’ (2002: 74). This assertion is substantiated by Purdy (2008), who cites claims from goths that they are abused on a daily basis, especially by people acting in gangs. As one Sheffield-based goth stated: We live in a world that doesn’t accept this kind of difference and young people are continuously mocked, attacked and sometimes even killed for the way they look … There are hundreds of stories about the abuse suffered by our city’s ‘different’ young people. We need to stop the prejudice. (The Star 2008)

Comparisons with other forms of hate crime The frequency of verbal abuse, and the fact that perpetrators act in gangs, has some similarities with other forms of hate crime, such as racist or homophobic, in which victims also report frequent verbal victimisation, often from youths acting in groups (Chakraborti and Garland 2009). Physical assaults upon goths are much rarer but still very impactful. Such incidents are typified by the experience of a goth friend of Lancaster’s who recounted being threatened with a meat cleaver by a stranger while on a night out, and that his goth friends had also been spat at and punched (Hodkinson 2008: 32). Almost a year after Lancaster’s murder another extreme assault occurred in a public park, where a 26-year-old goth, Paul Gibbs, was the victim of a ‘brutal and sustained gang attack’ during which he lost an ear and was placed in a coma (Bullough 2009). Two other goths were hospitalised as a result of the same Leeds-based incident. It was reported that before the violence one of the assailants told his friends, ‘I’m a chav and I’m going to get some moshers’, and other witnesses heard the three perpetrators screaming ‘dirty moshers and goths’ at their intended victims (Gardner 2009). Interestingly, as Brill (2008) notes, some of the harassment of goths can be homophobic in nature (even though its recipients are predominantly heterosexual), drawing parallels between this form of victimisation and that suffered by gay communities who are routinely 49

Hate Crime

considered to be a hate crime victim group. This abuse is especially directed at some male goths, whose androgynous and effeminate appearance can provoke ‘frequent’ homophobic abuse or assault from members of the public, with some of those within the subculture implying that this places male goths at a higher risk of prejudiced victimisation than females, who with their typically ‘hyperfeminine’ gothic appearance instead tend to be recipients of sexual harassment. As Brill (2007: 121) argues: For a straight man, the adoption of effeminacy with its consequent risk of homophobic abuse can be seen as a deliberate – and hence courageous – step towards the marginalized position women and gay men occupy in our society. For women, by contrast, at least the milder forms of sexual harassment are a fairly normal experience; they simply come with the territory of the inferior social position women have been assigned in our culture. There is also the suggestion that, like other forms of hate crime such as disablist (see, for example, Quarmby 2008), violent assaults upon goths and major ‘signal’ events, such as the attack upon Maltby and Lancaster that gained national publicity, can be especially impactful on those affiliated to the subculture, creating the impression among sections of the ‘alternative’ community that such attacks are frequent and may even be increasing (Denham 2009). Thus the fear of becoming a victim of abuse or violence may influence feelings of safety and security in that community, causing many goths and ‘alternatives’ to choose to avoid certain areas of town and city centres, as well as venues where they feel they might be vulnerable to harassment (Hodkinson 2002). It is interesting, therefore, that two of the most violent incidents discussed in this chapter occurred in public parks in the home towns of the victims, where presumably they felt reasonably safe and yet were brutally victimised due to their ‘difference’. Central to many discussions of hate crimes more broadly is the prejudice that perpetrators have against identifiable outgroups, such as Muslims, disabled people or homosexuals, whom they deem to be different from the acceptable norm. The attitude of the attackers in the Lancaster case – whose behaviour was described as ‘very aggressive, intolerant and callous and violent’ by the original trial judge (Court of Appeal 2008: 6) – was in many ways mirrored by that of the Leeds assailants, whose language during the incident betrayed their own bigotry and small-mindedness. Such attacks on members of youth subcultures are not new – Savage (1991), for example, refers to the 50

The victimisation of goths and the boundaries of hate crime

frequency with which punks were assaulted by ‘ordinary’ members of the public in the mid-1970s – but the anecdotal evidence discussed here may suggest that assaults against goths may be more frequent than those against members of other, less conspicuous or more accepted subcultures.

Conclusion This chapter has discussed the victimisation of goths and those from ‘alternative’ subcultures with a view to assessing whether these attacks should be understood as hate crimes. It examined the characteristics of the subculture, including the striking, sometimes shocking appearance of many goths, the androgynous dress of some males and the eclectic but tightly knit nature of the current goth scene. It was suggested that a significant proportion of participants are very dedicated to their lifestyle, seeing it as a central part of their daily lives and core identities. It was also noted that there is a strong sense of kinship among goths, who readily identify with each other while being, in a lot of cases, wary of those outside of their subculture. For many, there exists a distinct boundary that separates them from the ‘mainstream’. Attacks upon goths, perpetrated due to prejudice against their strikingly different appearance, were then reviewed, followed by a discussion of the assault upon Robert Maltby and Sophie Lancaster that resulted in the hospitalisation of the former and the death of the latter, which concluded that its apparent motivation was hatred, on behalf of the attackers, of their gothic and alternative image. There was also evidence that this assault was part of a broader pattern of the harassment and abuse of goths in that town of Bacup and that Maltby and Lancaster had previously been victimised there. The broader victimisation of goths was then examined; this appeared to indicate that they are subject to frequent verbal abuse and harassment, including on rare occasions physical assault. Some of these attacks, such as the one that severed the ear of a Leeds goth, have been ferocious in nature, bringing to mind incidents like the homophobic assault upon Jody Dobrowski in 2005 that was investigated and then prosecuted as a hate crime. It was also noted that the frequency and pattern of this abuse – in that the most common form of harassment is verbal and that single incidents tend to be part of a broader and more long-term pattern of victimisation – may also have some similarities to hate crimes more generally. 51

Hate Crime

Intriguingly, it was also suggested that some of the victimisation of male goths is homophobic in nature as they are targeted because of their gender-ambiguous appearance (Brill 2008). Moreover, according to Perry (2001), groups that become victims of hate crimes are those that are somehow different from the accepted norm, which is something that could be applied to goths, whose dress codes and lifestyles clearly mark them out as ‘different’. She also argues that each of the hate crime ‘outgroups’ has particular traits attached to them by sections of society that are constructed in negative, oppositional terms (Perry 2001). Again, it could be argued that this applies to goths, who in the eyes of some are weird and ‘deviant’, and whose presence can be unsettling and challenging especially for those in monocultural, socially conservative and often isolated communities. As Saucier et al. (2006) argue, hate crimes can also have severe consequences for the victim and their wider minority community, generating feelings of anxiety, fear and depression, and the chapter has suggested that there is some evidence of this among goth communities following the murder of Sophie Lancaster. However, despite this, it is extremely difficult to know which groups should be included under the hate crime ‘umbrella’ and which should not. For Perry (2001: 10), hate crimes are ‘directed towards already stigmatized and marginalized groups’, tying in with the idea discussed earlier that the concept of hate crime has its roots in the civil rights struggles in the US of forty or so years ago, as it was seen as a way of underlying the commonalities (such as deprivation, discrimination and victimisation) shared in the lived experiences of the campaigning groups. If this idea is accepted, then it could be argued that the inclusion of subcultural groups such as goths among those that have been historically oppressed (such as minority ethnic or disabled populations) may, from the viewpoint of the latter, trivialise or belittle their own history of marginalisation and their own experiences as victims of hate. One reason it may do this is because a subcultural style is adopted by someone, rather than them being born that way. Certainly this was the view of the UK government expressed in its response to the online petition (discussed above) that called for attacks on goths to be included in hate crime legislation: [UK] legislation focuses on hate crimes on the basis of race, faith, sexual orientation, disability and gender identity – and it is these categories which are currently monitored. We do not plan to extend this to include hatred against people on the 52

The victimisation of goths and the boundaries of hate crime

basis of their appearance or sub-cultural interests. These are not intrinsic characteristics of a person and could be potentially be very wide ranging, including for example allegiance to football teams – which makes this a very difficult category to legislate for. (Prime Minister’s Office 2008b) While it is accepted here that those in alternative or goth subcultures are not part of a community that has a significant and longstanding history characterised by struggles for equality (like those from minority ethnic backgrounds, for example), it is nevertheless suggested that for many goths their subculture is an intrinsic and vital part of their lives. Being a goth governs not just their leisure time but also their everyday activities, and is something that far from being an ephemeral fashion statement becomes a fixture in their lives for decades (Hodkinson 2005). It involves high levels of long-term commitment that develops a sense of solidarity and community with other goths. Attacks against goths, and especially the highprofile murder of Sophie Lancaster, therefore impact on their wider community: they really do ‘hurt more’, just as hate crimes such as the murders of Stephen Lawrence or Brent Martin also do. This issue highlights just how problematic and contentious the process of framing the boundaries of hate crime can be. However, as mentioned earlier, Perry (2009: 72) asserts that ‘hatemotivated violence is used to sustain the privilege of the dominant group and to police the boundaries of the group by reminding the Other of his or her place’. It is less clear how the victimisation of goths fits into this structural idea of hate crime that sees powerful social groups perpetrating hate acts against marginalised and despised ‘outgroups’ in order to keep them in their ‘place’. Instead, it appears that it is the fear of difference, and of the despised ‘other’, that drives much of the anti-goth harassment; as in the attacks in Leeds and Lancashire discussed above, it may also be the case that much of this abuse and violence is actually perpetrated by those from socially and economically deprived backgrounds against those (goths) who, as Paul Hodkinson (2002) suggested, are predominantly middle class. As the judge in the original trial of the Bacup attackers stated, Maltby and Lancaster were targeted because of their difference, which he regarded as a ‘serious aggravating feature’ which should ‘be equated with other hate crimes such as those where people of different races, religions, or sexual orientation are attacked because they are different’ (Court of Appeal 2008: 5). Indeed, there is a 53

Hate Crime

suggestion that sentencing guidelines to judges in cases like these may be altered, to encourage them in future to impose longer sentences upon those found guilty of similar acts (Watkinson 2009). It could therefore be argued that Perry’s structural view of hate crimes masks their most fundamental aspect, which is that the victim is targeted solely because of their group affiliation (Gerstenfeld 2004). In other words, it is the fact that victims are targeted because of who they are that is the most important facet of such incidents, as it is their membership of a despised social outgroup that sparks abuse and assaults against them. Such harassment is therefore less to do with keeping the victim in their subordinate place within the social structure as it is motivated by a more base and unthinking instinct: the fear or hatred of difference. It could also be argued, though, that the most important aspect of victimisation is the pain, suffering and fear that it causes to the victim and their wider community: something very evident in the aftermath of the Lancaster case and seemingly an ever-present facet of life as a goth or ‘alternative’. To be targeted due to one’s difference, in whatever form this may take, and the fear this may cause, could well be the most important facet of these discussions. If this is the case, then some academic definitions of hate crime may need to be rethought.

Notes 1 To clarify, there are two separate Hodkinsons cited in this chapter – Mark Hodkinson (a journalist, responsible for the 2008 article cited) and Paul Hodkinson (an academic, author of all of the others). 2 In this case, ‘alternative’ subcultures are considered to be those involving individuals with ‘extreme’ appearances, including spiked or vividly dyed hair, visible tattoos and multiple body-piercings, and clothes, such as those worn by goths, which are considered to be outside of regular high street fashions. 3 Goth subcultural ideas of male and female appearance are characterised by this somewhat contradictory attitude towards male gender performance (where a rebellious androgynous look is encouraged) and that of females (where ‘traditional criteria of feminine attractiveness’ are preferred (Brill 2007: 118)). 4 ‘Mosher’ is another term for a goth or a rock fan of ‘alternative’ appearance. 5 ‘Alternatives Have Rights Too’ can be found at www. alternativeshaverightstoo.co.uk/. 54

The victimisation of goths and the boundaries of hate crime

6 The campaign has now become a charity, the Sophie Lancaster Foundation – see www.sophielancasterfoundation.com/.

References BBC News (2008a) ‘Attack Jury Played Desperate Call’, BBC News Online. Available at http://news.bbc.co.uk/go/pr/fr/-/1/hi/england/ lancashire/7292869.stm [accessed 20 August 2009]. BBC News (2008b) ‘Goths Plea Following Park Attack’, BBC News Online. Available at http://news.bbc.co.uk/go/pr/fr/-/1/hi/england/ west_yorkshire/7592714.stm [accessed 25 August 2009]. Bowling, B. (1999) Violent Racism: Victimisation, Policing and Social Context. Oxford: Oxford University Press. Brill, D. (2007) ‘Gender, Status and Subcultural Capital in the Goth Scene’, in P. Hodkinson and W. Deicke (eds) Youth Cultures: Scenes, Subcultures and Tribes. London: Routledge (pp. 111–26). Brill, D. (2008) Goth Culture: Gender, Sexuality and Style. Oxford: Berg. Bullough, C. (2009) ‘Gang Jailed for Barbaric Attack on Rothwell Goth; Goth Had His Ear Sliced Off’, Wakefield Express, 5 June. Chakraborti, N. and Garland, J. (2009) Hate Crime: Impact, Causes and Responses. London: Sage. Court of Appeal (2008) Judgment: Before The Lord Chief Justice Mr Justice Owen and Mr Justice Christopher Clarke Between: R v Herbert (1), Harris (2), Joseph Hulme (3), Danny Hulme (4), and Daniel Mallett (5). London: Royal Courts of Justice, EWCA Crim 2501. Denham, C. (2009) ‘Don’t Abuse Us Because of the Way We Look’, Kentish Express, 7 April. Gardner, T. (2009) ‘Rothwell Park Attack Goth Victim Interview’, Rothwell Today, 9 June. Available at www.rothwelltoday.co.uk/news/Rothwellpark-attack-Goth-victim.5348497.jp [accessed 24 August 2009]. Garland, J. and Chakraborti, N. (2004) ‘England’s Green and Pleasant Land? Examining Racist Prejudice in a Rural Context’, Patterns of Prejudice, 38(4): 383–98. Garland, J. and Chakraborti, N. (2006) ‘Recognising and Responding to Victims of Rural Racism’, International Review of Victimology, 13(1): 49–69. Gerstenfeld, P. B. (2004) Hate Crimes: Causes, Controls and Controversies. London: Sage. Gough, C. (2009) ‘Sophie Mum’s Hate Crime Bid’, Whitby Gazette, 9 May. Grattet, R. and Jenness, V. (2003) ‘The Birth and Maturation of Hate Crime Policy in the United States’, in B. Perry (ed.) Hate and Bias Crime: A Reader. London: Routledge (pp. 389–408). Hodkinson, M. (2008) ‘On 11 August 2007 a Young Goth Died at the Hands of a Brutal Teenage Gang: One Year On, Thousands of Her Supporters the 55

Hate Crime

World Over Have United in the Name of Tolerance – and the Girl Who Dared to be Different’, Observer Magazine, 3 August: 28–34. Hodkinson, P. (2002) Goth: Identity, Style and Subculture. Oxford: Berg. Hodkinson, P. (2004) ‘The Goth Scene and (Sub)Cultural Substance’, in A. Bennett and K. Harris (eds) After Subculture: Critical Studies in Contemporary Youth Culture. London: Palgrave (pp. 135–47). Hodkinson, P. (2005) ‘Insider Research in the Study of Youth Cultures’, Journal of Youth Studies, 8(2): 131–49. Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Jenkins, R. (2008) ‘Boy of 15 Who Attacked Woman For Being a Goth Is Convicted of Murder’, The Times, 28 March, p. 25. Muggleton, D. (2000) Inside Subculture: The Postmodern Meaning of Style. Oxford: Berg. Perry, B. (2001) In the Name of Hate: Understanding Hate Crimes. London: Routledge. Perry, B. (2009) ‘The Sociology of Hate: Theoretical Approaches’, in B. Levin (ed.) Hate Crimes, Vol. 1: Understanding and Defining Hate Crime. Westport, CT: Praeger (pp. 55–76). Prime Minister’s Office (2008a) E-Petitions. Available at http://petitions. number10.gov.uk/goth-hatecrimes/ [accessed 16 June 2009]. Prime Minister’s Office (2008b) Goth Hate Crimes – E-Petition Response. Available at www.number10.gov.uk/Page15282 [accessed 16 June 2009]. Purdy, C. (2008) ‘The Darker Side of Life as a Goth’, BBC News Online. Available at http://news.bbc.co.uk/go/pr/fr/-/1/hi/england/lancashire/7314306. stm [accessed 27 April]. Quarmby, K. (2008) Getting Away With Murder: Disabled People’s Experiences of Hate Crime in the UK. London: Scope. Reynolds, S. (2005) Rip It Up and Start Again: Post-punk 1978–84. London: Faber. Richardson, L. (2008) ‘Youth Forum Aims to End Hate Crimes’, The Northern Echo, 29 September, p. 31. Rowe, M. (2004) Policing, Race and Racism. Cullompton: Willan Publishing. Saucier, D. A., Brown, T. L., Mitchell, R. C. and Cawman, A. J. (2006) ‘Effects of Victims’ Characteristics on Attitudes Toward Hate Crimes’, Journal of Interpersonal Violence, 21(7): 890–909. Savage, J. (1991) England’s Dreaming: Sex Pistols and Punk Rock. London: Faber. Stamp Out Hatred and Intolerance Everywhere (SOPHIE). Available at www. myspace.com/inmemoryofsophie. The Star (2008) ‘Goths to March Through Sheffield in Bid to Stop Abuse’, The Star, 18 July. Telford, W. (2008) ‘We Wear Black – and We’re Proud’, Evening Herald, 4 September, p. 6. Wainwright, M. (2008) ‘Woman Died After Drunken Gang Attacked Couple Dressed as Goths’, The Guardian, 13 March, p. 15. 56

The victimisation of goths and the boundaries of hate crime

Watkinson, D. (2009) ‘Sophie Lancaster’s Mum Set to Win Tougher Sentences for Hate Crimes’, Lancashire Telegraph, 7 May. Worley, M. (2010) ‘One Nation Under the Bomb: The Cold War and British Punk to 1984’, Journal for the Study of Radicalism, forthcoming.

57

Chapter 3

Future challenges for hate crime policy: lessons from the past Hannah Mason-Bish

In September 2009 a cross-government Action Plan to combat hate crime was launched. The long-awaited report was published after lengthy consultation with a variety of stakeholders and aimed to identify key areas of concern in tackling hate crime. Fundamentally the plan is about victims – identifying them, helping them and improving their confidence in the criminal justice system (Home Office 2009: 15). In the foreword of the report, Home Secretary Alan Johnson noted that it came ten years after the Stephen Lawrence Inquiry, which ‘serves as a reminder to us of where we have come from, what we have learned, what we have achieved and how much further we have to go’ (2009: 3). This chapter seeks to serve as a similar reminder, by examining the development of hate crime policy so far and identifying future challenges. In Great Britain we have seen the emergence of legislation that identifies different types of hate crime, based on the victim’s race, religion, sexual orientation or disability, and stipulates that the perpetrator should be punished more harshly. The criminal justice system has also issued an array of policy related to hate crime, designed to define hate crime and explain why offences motivated by hostility should be treated differently from other crimes. These developments have been set against a backdrop of campaign group activism, and the victim categories included within the hate crime policy domain represent those who have successfully utilised the concept of hate crime, been involved with advisory groups and monitored implementation of criminal justice provisions. This chapter examines the future challenges facing hate crime policy by reviewing 58

Future challenges for hate crime policy: lessons from the past

the efforts of campaigners and criminal justice practitioners who have been involved with shaping the policy domain so far. It look at the usefulness of the concept of hate crime and its future potential. Importantly, as hate crime legislation has been defined by what victim groups are included within it, future policy complications in this area are examined with a particular focus on the categories of age and gender. An examination of the future challenges facing hate crime policy in Britain is a timely pursuit. In 1998 the Crime and Disorder Act created a list of racially aggravated offences, which would require an enhanced punishment if the perpetrator demonstrated hostility based upon the victim’s actual or presumed membership of a racial group. In 2001, this was extended to encompass religiously aggravated crimes under the terms of the Anti-Terrorism, Crime and Security Act. In 2003 the Criminal Justice Act included a clause that provided a statutory aggravation for crimes where hostility towards sexual orientation or disability was a factor. This did not create specific offences on a par with those for racially and religiously aggravated crimes but rather allowed the judiciary to have discretion in adding an enhanced sentence. At the time of writing, Scottish hate crime laws only include crimes motivated by racist, religious or sectarian prejudice. The Scottish Executive is currently reviewing proposed legislation that would extend hate crime legislation in Scotland to cover crimes motivated by ‘malice or ill-will based on a victim’s actual or presumed sexual orientation, transgender identity or disability’ (Scottish Executive 2004).1 Related policy initiatives are also subject to ongoing changes. The Crown Prosecution Service (CPS) in England and Wales produced guidance on prosecuting disability hate crime in February 2007 (CPS 2007a) and the Association of Chief Police Officers (ACPO) have again been working on an improved hate crime manual – their third in eight years, which provides a new definition of hate crime that all police services have to use (Home Office 2009: 52). Such changes illustrate the way that hate crime policy is in an ongoing state of flux, with victim groups and definitions being debated. As this chapter demonstrates, these developments are likely to continue to face challenges in the future.

Hate crime policy development: research findings This chapter is based upon the findings of the author’s own research which will be used to demonstrate how past developments in hate 59

Hate Crime

crime policy might shape its future. The research aimed to explore why provisions for racially aggravated offences were established in 1998 and to explain their subsequent expansion to include religion, sexual orientation and disability. As such, it was specifically focused upon the development of legislation that punished existing crimes more severely where it was seen to be motivated by hostility. It also examined the categories that have been explored as potential hate crime victim areas – age and gender. Fifty semi-structured interviews were carried out with a diverse array of respondents, ranging from social movement activists to criminal justice practitioners. The research revealed some interesting findings that are fundamental to understanding the future difficulties that hate crime policy will face. The role of campaign groups has been central to the way that policy has developed. The victim groups included in hate crime legislation represent those who have activists and campaigners lobbying on their behalf, gathering data and engaging with criminal justice monitoring and advisory groups. These campaigns have been initiated because those working with victims of hate-based violence have felt that the police and criminal justice system were not taking victims seriously. It was essentially a quest for recognition and about getting practical change. Hate crime was a useful banner under which to frame these claims and to highlight the similarities between different forms of victimisation. As many academics have noted, being included in hate crime legislation was designed to send a positive message to specific victim groups (Iganski 1999; Hall 2005). For the police it was a useful way to engage communities and to win their ‘trust and confidence’ (ACPO 2000: 18). For the Crown Prosecution Service it allowed them to consult with community representatives and to reshape their approach. However, this research revealed two themes that represent some fundamental challenges relating to the role of campaign groups in the shaping of hate crime policy. The first is that by its very nature, legislation that is accompanied by a list of victims is both inclusive and exclusive. Many of the groups included felt that there was a victim hierarchy whereby more ‘established’ hate crime victims received preferential treatment within the criminal justice system. These questions will be addressed in some depth throughout this chapter. A second key theme throughout the research concerned the use of the concept of hate crime itself. Much of the scholarly attention given to this subject has centred upon the usefulness of ‘hate crime’ as an operational term (Iganski 1999, 2008; Perry 2004). This research found that conceptual difficulties over its use have been fraught with contradictions that are yet to be resolved. Some campaigners have 60

Future challenges for hate crime policy: lessons from the past

become reconciled to using the term, feeling that they must comply with existing bureaucratic and administrative language. Others were keen to utilise the power of the concept. For example, activists campaigning against homophobic and disablist violence deliberately shifted towards using the term hate crime because it told the victim that what was happening to them was unacceptable. It also allowed them to draw comparisons between the experiences of different victim groups by being a useful banner which connected them as victims of prejudice. However, although campaigners might exercise a degree of control over whether or not to utilise the concept in their lobbying efforts, these decisions were accompanied by a large amount of conceptual baggage. In their study of racist hate crime, Ray and Smith were critical of the term, suggesting that it might convey an image of a stranger, attacking at random in an impersonal way (Ray and Smith 2001). Other researchers have concurred, with Kielinger and Stanko (2002) criticising the suggestion that hate crimes were always motivated by clear-cut hatred. Importantly, these misconceptions have had a direct impact on policy-making decisions. The categories of age and gender have been routinely debated as possible hate crime victim groups in parliamentary discussions and consultations. Yet they are excluded because they are thought to be crimes of a different order. Policy-makers and criminal justice practitioners explained during interviews that gender violence was personal – hate crime was perpetrated by strangers. Age-related violence, such as elder abuse, took place in private, whereas hate crime was a public event. Even for those victim groups included in hate crime policy, such as disability, doubt was cast over their legitimacy because of preconceptions about hate crime. In a recent piece of research, the Equality and Human Rights Commission investigated what it called ‘disability targeted violence’, arguing that hate crime was a confusing and misleading term that had had its day (EHRC 2009). This chapter therefore also focuses upon the future challenge of hate crime as a type of policy and questions whether it has any use in future initiatives.

Defining victims Victim hierarchies One of the key features of hate crime policy to date has been the need to define a list of victims. It is this requirement that has meant that some victim groups will inevitably be excluded from policy 61

Hate Crime

and legislative protection. What was apparent from the research was that hate crime policy creates a perceived victim hierarchy whereby some victims are deemed worthy of inclusion and others are left out. However, the story is even more complex than that because hate crime policy has been formed through the work of lobbying and advisory groups who have had quite narrow remits, often focusing exclusively on one area of victimisation. This has contributed to a hierarchy within hate crime policy itself, whereby some identity groups seem to receive preferential treatment in criminal justice responses to hate crime. This is seen most acutely in race and disability hate crime. One of the findings of the research was that disability campaigners felt that disability was at the bottom of the list of hate crime victim identities. Criminal justice professionals were also acutely aware of this fact. One senior police officer explained that ‘in the hierarchy it is clearly race, homophobia, faith and disability. We are conscious that race is nearly “done” because people are so aware of race that we tend to concentrate on the others coming through.’ Another police officer described the policing response to disability hate crime as ‘absolutely woeful’. The perception of a victim hierarchy also explains why age and gender have not been included in hate crime policy. When discussing the concept of hate crime with a policy officer from an elder abuse organisation, one interviewee explained: If calling it hate crime meant that we as a society viewed it with the same level of seriousness as we do other areas of hate crime – like racially motivated hate crime – then we would tactically support that … It’s all violence and I think the silos and hierarchies don’t actually help. He went on to explain that if age was included in hate crime policy, it would be at the bottom of the pile of victim categories. So the perception of differential treatment between victim groups has had a direct impact on campaigning choices of key charities and organisations. Many respondents worked directly with victims of hate crime and were relaying their fears that a perceived hierarchy also had a negative effect on victim reporting. Therefore, a future challenge for policy-makers and practitioners will be to ensure that if policy expands to include other hate crime victims, they are not seen to be at the bottom of a long list. While perceived discrepancies in the treatment of hate crime victims might impact on reporting of crime there is a further complication. It 62

Future challenges for hate crime policy: lessons from the past

could be argued that an ongoing problem of hate crime policy is that it creates a ‘competition for suffering’ between victim groups. Such concerns should not be overstated and the author would not go as far as Jacobs and Potter who have argued that hate crime legislation ‘balkanises’ groups by fragmenting laws into simplistic victim and offender groups (Jacobs and Potter 1998: 130). They suggest that the exclusion of gender, for example, might cause a rift between different sections of society. What this research revealed was not animosity or resentment, but concern about the way that the criminal justice system was dealing with hate crime in its policy initiatives. Race campaigners did not dispute that the disabled were attacked and they did not question the harms that they suffered. What they were worried about was that policy would move on and draw a line under race. This represented a real concern for some race activists who felt that they had campaigned for a long time to have a usable and effective law. October 2007 saw the formation of the new Equality and Human Rights Commission, established to ‘eliminate discrimination, reduce inequality, protect human rights and to build good relations, ensuring that everyone has a fair chance to participate in society’. It covers six areas of diversity: race, sexual orientation, religion, gender, disability and age (EHRC 2007). Previously separate organisations – the Disability Rights Commission, Commission for Racial Equality and Equal Opportunities Commission – had had responsibility for issues of disability, race and gender. Some of the race campaigners interviewed thought that the new development could lead to the marginalisation of race and racist incidents. Concerns were raised, even among those who felt that it was a good idea in principle, about the potential for important dynamics to be overlooked, as illustrated below in the comments from one interviewee: It is very important that you have one commission and bring in disability and human rights. It’s a very good move. The problem that I do have is that race is different from the others. All of the six groups have some defined consumer groups that can badger and argue and make the case and campaign … When it comes to racism and when it’s lumped in with the other five kinds of discrimination then in a way it becomes diluted because there is something special about racism that needs it to be very high on the agenda. This much is true. There are a large number of disability and gay rights groups who will remain in existence and will continue to lobby 63

Hate Crime

on behalf of their victims. The majority of disability campaigners interviewed for my research thought that they would have a lot to gain working closely with other groups. One senior police officer also said that there were tensions and that identity groups ‘often have an identity that is very strong and of which they are very proud, or wedded to, and that cross-understanding can be quite difficult’. He drew particular attention to a meeting between the Muslim community and the gay advisory group whereby the differences were more apparent than the similarities. One of the objectives of hate crime policy has been to facilitate greater community cohesion and trust. The perceived disparity between the treatments of victim groups in the criminal justice system might affect this as different sections of the community feel that their form of hate crime is not taken seriously. As hate crime policy continues to develop and new groups of victims might be added, these concerns of preferential treatment will grow ever more acute.

Victim silos and diverse identities A further concern for the development of hate crime policy is the suggestion that to date it has been accompanied by rather simplistic notions of victimhood. Separating victims via crude titles – race, religion or disability, for example – does not always help us to understand the unique harm that a victim might experience. Such a concern has been noted by an increasing number of academics in relation to research on diverse communities. To quote directly from a senior CPS official, ‘there are black women, gay people who have disabilities and life doesn’t come in nicely packaged departments’. Garland et al. (2006) have noted that researchers investigating hidden communities must be aware of the diversity of group experiences. Hate crime policy has tended to divide victims into clear groups which do not take account of this diversity. For example, one independent activist said: ‘I am disabled, gay and a woman. If I am targeted am I supposed to say which aspect was the most hurtful and damaging?’ A clear example of this simplification of victimhood and the need to label it can be seen in the treatment of transphobic hate crime. The presence of transphobic hate crime in policy and legislation is patchy. The CPS expanded their policy on homophobic hate crime to become homophobic and transphobic hate crime (CPS 2007b) and the police have also included it in their hate crime policy. However, to date there is no legislation in England and Wales that provides for sentence enhancement. During my research there were 64

Future challenges for hate crime policy: lessons from the past

lengthy discussions with practitioners and campaigners alike who were concerned about the need to accurately ‘place’ transphobic hate crime within policy. They were unsure about whether it should fit within issues of sexual orientation or gender or whether it should be a category on its own. A senior police officer interviewed for this research observed that when organising hate crime advisory groups, some members of the transgender community were offended at being ‘lumped in’ with the gay community, but felt that it was more about gender. Others felt that they should be linked to gay-related issues and included with homophobic hate crime policy. Issues of group classification are important because hate crime policy defines a list of victims who will then be provided with support, and local help groups will receive funding. Sometimes their perpetrator will be punished more harshly, depending on what type of hate crime it is. However, it is also about the symbolic recognition of hate crime victims and their group identity. After the creation of the provisions for racially aggravated offences under the Crime and Disorder Act 1998, there was an increasing campaign in Britain for Islamophobia to be recognised as a form of hate crime. The government initially rejected the calls to expand race hate crime provisions to encompass religion, arguing that most Islamophobic attacks were at least partly motivated by racism (Straw 1998). The government felt that existing race provisions could be used to punish religious hate crime offenders but they failed to acknowledge the symbolic importance of recognising victim identities. The response to the campaign against Islamophobic hate was eventually to simply add religion to legislation in 2001, rather than review the intricate nature of the victims’ experiences. Respondents in the research who worked with victims or had experienced hate crime directly wanted policy that would reflect their diverse experiences. They felt that once they had been attached to a label of ‘homophobic hate crime’, for example, they would then be directed towards an established set of measures and resources that was not flexible enough to meet their unique requirements. One example given was that of a series of homophobic attacks against black men in East London. The victim often lived a heterosexual life but was involved in secretive gay sexual behaviour. Victims were dealing with complex issues of sexuality, race and gender, and having to choose which hate crime label to apply did not assist their recovery. In the future this silo approach will need to be addressed, as an expanding and separate list of victims could become unworkable. As such it could be argued

65

Hate Crime

that hate crime policy has been unable to cope with victimisation which is complex and might cross different diversity strands.

Campaign groups and influence One key research question was to ask whether or not there was a pattern in how campaigners tried to get a victim group included in hate crime policy and legislation. It became clear that the development of hate crime policy in Britain has required a struggle and the mobilisation of different activists. Therefore, the four key areas of hate crime victimisation recognised in policy and legislation represent those who have organised campaign groups and whose victimisation has been proven via the use of empirical evidence and case studies. However, there are clearly other groups who are victimised but who are not included in the policy domain. For example, other groups who might warrant inclusion could be the homeless and sex workers: both groups are targeted because of their identity and experience prejudice and hatred. During an interview, a senior police officer who was working on the revised ACPO hate crime manual explained how they decided which victim groups would be included and excluded from new policy proposals. He listed four factors, namely: 1 Whether such cases occurred in meaningful numbers. 2 If the offences were likely to be motivated by hatred rather than vulnerability. 3 If the offence has an impact on cohesion and safety within the community. 4 If the offence has an increased impact on the victim. Arguably the homeless and sex workers would still seem to fit within this list, so questions could be asked as to why they – and other groups – are excluded. Academics such as Jo Morgan have asked why some of the most oppressed groups in society are not afforded the status of hate crime victims. Her study examining victimisation of sex workers and paedophiles showed that inclusion in hate crime policy is not just about proven victimisation and campaign group activism. She observed that ‘competition to be counted and the political clout required to be counted has not only frozen out disorganised groups and individuals that experience hate crime, it has also led to infighting between social movements’ (Morgan 2002: 32). Society – or government – has to care about and recognise the legitimacy of that group (Morgan 2002). In other words, hate crime policy relies upon a 66

Future challenges for hate crime policy: lessons from the past

group of victims being seen as sympathetic and having an organised campaign base to fight their corner. This is an ongoing problem of hate crime policy. It defines which groups are worthy of inclusion based upon the successful lobbying efforts of a campaign. A future challenge will be for us to think about the extent to which groups who do not have a powerful lobby are deemed unworthy of inclusion in hate crime policy. This is a troubling message. While issues of campaign group influence are a concern for groups currently excluded from policy, they are of continuing relevance to those already represented in hate crime policy. The research revealed the ongoing tensions between social movement activists and largescale campaign bodies that Jo Morgan notes in her article. One lobbyist, working for a grassroots community group on homophobic abuse, said about their relationship with Stonewall: ‘There isn’t one [a relationship], they don’t really listen to us.’ As the largest national body with significant parliamentary lobbying influence, Stonewall has much more funding and resources than many of the smaller, local groups. This was also the case with disability campaigners. During interviews with five people who could best be described as independent lobbyists on disability rights, all expressed slight annoyance that disability hate crime was only ‘on the agenda’ when the big, well-funded charities ‘noticed it’ in 2007. This was further complicated by the fact that many of the large campaign charities are not run by disabled people, yet ‘purport to speak on behalf of disabled people’. As one independent activist said, ‘They get the funding, they do a bit of lobbying and they get the credit.’ Sometimes there are contradictions between campaign groups and charities. For example, in terms of disability hate crime two independent activists interviewed were keen that a definition of disability hate crime should include crimes in the home and ‘private’ abuse. However, this does not necessarily fit with the definition being put forward by the large charities such as Mind and Mencap. This inconsistency necessarily affects the implementation of hate crime policy and will continue to do so. A further issue regarding the role of activists in the construction of hate crime policy is the ‘volume’ of the voices and weight that they carry with policy-makers. A senior policy officer for the Crown Prosecution Service said: One of the interesting things from our perspective is that different communities have different volumes at which they

67

Hate Crime

speak. For a multitude of reasons, those arguing from the groups and communities concerned with race issues are very vocal and articulate and have been able to contest and argue their cause very effectively. Part of this relates to time and issue creation. Monitoring organisations and advisory groups with a focus on racist violence have historical links with criminal justice agencies and as such have ‘honed’ their ability to make the most of opportunities to shape policy. The Metropolitan Police Service Disability Advisory Group was established in 2004 and as yet might not be expected to have ‘perfected’ the art of arguing their case effectively. This also applies to policy guidance. The CPS guidance for homophobic hate crime is now in its second form, having been first devised in 2002. One might expect that it is more developed and has been reviewed more thoroughly than the 2007 first policy on disability hate crime. Within each identity group there might also be tensions. Two policy officers from a leading learning disability charity were concerned that people with mental health issues might be left out of discussions on disability issues. Those working with learning disability groups were similarly keen that they were not overlooked in favour of campaigners representing people with physical disabilities. While this might simply mean that activists become more ‘pushy’ in their attempts to influence hate crime policy construction, it can also translate directly into inclusion in hate crime policy. In Scotland, the Executive’s Violence Against Women Unit was not supportive of the inclusion of gender in hate crime policy (Scottish Executive 2004: 32). Despite the efforts of other, smaller lobbying groups to challenge this position, theirs carried more weight and contributed significantly to the exclusion of gender hate crime from the recommendations of the Working Group on Hate Crime. This also happened with age; ‘big’ organisations like Age Concern and Help the Aged did not support the inclusion of age and theirs was the opinion that mattered. The involvement of campaigners is crucial to the ongoing construction of hate crime policy to challenge and attempt to expand definitions. Much of this now takes place within the criminal justice system, where practitioners consult and involve campaigners in the policy-making process. This will continue to present challenges into the future.

68

Future challenges for hate crime policy: lessons from the past

Understanding hate crime Challenging assumptions The ongoing role of campaigners in shaping hate crime policy and the problems of defining lists of victims has been assessed in this chapter so far. It is now prudent to move on to examine the idea that common-sense perceptions of hate crime are used by policymakers and campaigners alike to define the boundaries of hate crime and what groups should be included in policy in the first place. In the future these will continue to be present in policy debates and undoubtedly new victim groups will be identified and potentially rejected from the policy list. Age and gender have been the two most consistently debated categories which have yet to find a home in the policy domain. In England and Wales, each time the ACPO hate crime manual has been revised age and gender are again discussed as potential categories and eventually omitted. One of the fundamental reasons for this has been because crimes against women and older people are seen as somehow inherently different from hate crime and are excluded from policy on this basis. Many campaigners and criminal justice practitioners interviewed for the research felt that age and gender-related crimes were an abuse of trust rather than hatred. One policy officer working for an age charity explained why they do not consider elder abuse to be a hate crime: ‘We have an uneasy relationship with the notion of hate crime because most of our work is guided by the relationship between the perpetrator and the victim and there has to be a clear expectation of trust.’ In cases of domestic violence or elder abuse, the female victim has a personal relationship with her attacker, which means that the victims are not interchangeable. However, Weisburd and Levin (1994: 36) have disputed the importance of interchangeability as criterion for excluding gender in hate crime policy: The lack of victim interchangeability does not prevent the characterisation of a crime as a bias crime with respect to the other categories … To rigidly adhere to victim interchangeability as the dispositive criterion only in gender cases and not in other traditional bias crime cases imposes an inequitable gender biased double standard. Even if it were the case that in many instances of gender-based violence there was a personal relationship between the victim and perpetrator, 69

Hate Crime

Weisburd and Levin question whether this should be the basis for excluding gender altogether. Kristin Taylor feels that in many hate crimes the perpetrator may be a neighbour or co-worker (Taylor 1996: 598). Taylor looks at domestic violence and while acknowledging the personal relationship, suggests that if a man has a history of beating his partners, then surely he is a repeat hate crime offender as he always attacks women (Taylor 1996: 599). Even so, a growing body of scholarly work has examined the relationship between victims and perpetrators and disputed the idea of hate crime as ‘stranger danger’ anyway (Iganski 2008; Mason 2005a, 2005b, 2005c). No hate crime policy or legislation makes any reference to the random nature of it or the need for the perpetrator to be a stranger. So using this as a way of deciding which groups to include is hard to justify and will continue to be so as age and gender will be debated in the future. A second crucial perception of hate crime is that it is motivated by clear-cut hatred. Specifically, a distinction is made between crimes motivated by hatred and those motivated by vulnerability and this is directly influencing policy implementation. One senior police officer said during an interview for my research: Are they being attacked because they are easy targets or because someone has a hatred of old people? With hate crime we went along the lines of the perpetrator. Why are they targeting that victim? Is it because they are easy pickings or because they don’t like that person because of who they are? The decision was made that we wouldn’t have age or gender as targeted victims of hate crime. Crucially, these views are also expressed by campaigners who excluded themselves from policy discussions. In their consultation, the Working Group on Hate Crime in Scotland invited women’s organisations to participate. The Scottish Executive’s Violence Against Women Unit, one of the most influential in this area, submitted their response, which said that ‘domestic violence is abuse of power within a relationship, whereby the man seeks to exert his power over his female partner but does not generally abuse other women’ (Scottish Executive 2004: 27). A representative of the working group explained that the concept of gender hate crime had caused a lot of debate because it was difficult to think of examples where ‘gender might be considered’ and that ‘domestic abuse might not be perpetrated because of the hatred of somebody’s gender’. In terms of age-related hate crime, this too caused problems in terms of establishing the 70

Future challenges for hate crime policy: lessons from the past

motivation of the offender. The working group felt that there was a difference between an opportunistic crime against a vulnerable person and a hate crime. A senior police officer from ACPO supported this view and stated that ‘if someone is old and frail and gets attacked by someone in the street, they were probably motivated by the opportunism of taking someone’s bag’. In terms of motivations of crime there were two issues: whether there were crimes where the perpetrator ‘hated’ women or older people, and the difficulty in proving that hatred in a court setting. Aside from the difficulties of proving hatred as a motivation for age-related and gendered hate crimes, the research revealed an ‘all or nothing’ approach that was applied by campaigners and those working in criminal justice. Put simply, age and gender were seen as difficult areas of victimhood and their complex nature was used as justification for their exclusion from the hate crime policy domain. A senior police officer, who has been involved with updating hate crime guidance for the criminal justice system, explained the lengthy discussions that were had in their advisory groups about the exclusion of age: We also talked about what crimes we are talking about and consistently people perceived two types of crime. One was abuse in care relationships, either by a family member or by a professional and the second was about bogus distraction burglaries. Both need addressing and I couldn’t begin to think which would be motivated by hatred … So my view was and our recommendation was that we shouldn’t include it for those reasons. The same officer also mentioned this idea in relation to gender: The problem is that not all rape is gender hatred. Some of it is misreading the signs, drunken behaviour. How do you separate the ones that are and aren’t? How do you know unless you have convicted someone? Then there were others like domestic violence. Some included forced marriage honour crime, but it’s a struggle. Age and gender are excluded from the hate crime policy domain on the grounds that it would be hard to establish which incidents were motivated by hatred and which by other factors. It is rarely heard that the difficulty in establishing motive in relation to racist and 71

Hate Crime

homophobic hate crime is justification for their omission from policy. This suggests an oversimplified understanding about the motivations of offenders in other forms of hate crime.

Explaining exclusion It is easy to critique the conceptual arguments that justify the exclusion of gender and age from hate crime legislation. However, a crucial aspect to consider in the future is why the campaign groups representing older people and women do not seem to want to be included in hate crime policy. It is most interesting to note that when discussing the exclusion of age and gender from hate crime policy most of the reasons centre around two issues. The first is the way that practical change could not result from their inclusion. Discussions with respondents from charities and the criminal justice system mostly highlighted either the fear of being at the bottom of a list of victims or the need to implement existing policy before any new ideas were devised. Looking slightly deeper, it is possible to understand why there was no perceived practical benefit from the addition of gender and age to hate crime policy. Elder abuse groups thought that hate crime was a simplistic term that would miss the variety of factors in crimes against older people. One campaigner from an elder abuse charity said: You end up with a subjective opinion on the motive of perpetrators. Is it a hate crime if someone thinks that they can attack a victim with Alzheimer’s because they can get away with it? The worry for us with the way that the CPS is going is that all crimes against the elderly will be seen as motivated by hate. There was a fear that something would be lost if crimes against the elderly were linked to hate crime. This was a particular worry for respondents who felt that there was already a large amount of policy and guidance in place that merely needed implementing. Creating something new that would lead to additional measures might overcomplicate procedures. Aside from placing a further layer of policy demands on practitioners, it was felt that the victims would not gain anything from being labelled as hate crime victims. This was partly because campaigners were concerned that elder abuse was an easier term for victims – and the criminal justice system – to understand. 72

Future challenges for hate crime policy: lessons from the past

As stated throughout this chapter, the presence and efforts of campaign groups has been of central importance to the development of hate crime policy. The absence and resistance of women’s groups in the hate crime debate is most notable. An independent lobbyist was interviewed for this research who had been involved with attempting to include disability in hate crime policy in 2003. Originally he had wanted to include gender as well, but could not get any support from women’s organisations. He said: I also talked to the leading gender organisations and just couldn’t get the support and they just didn’t want to take it up and I had to take the rational decision that if I couldn’t get the support of women’s movement then I wasn’t going to get anywhere with it. I would be stupid to fight for gender on my own. You can only take the women’s movement to the water, you can’t make them drink. Part of the explanation for this could be the perceptions of hate crime as a concept. Horvath and Kelly (2007: 4) have critiqued it as a term, arguing that it focuses attention on the individual pathology of the offender rather than looking at wider structural processes that maintain the position of women as ‘normal’ targets of abuse. However, there are wider historical and practical factors of relevance here. Some respondents pointed towards the huge numbers of cases of violence against women, which would swamp statistics and provisions for other victims of hate crime. Furthermore, domestic violence units pre-date hate crime provisions in the criminal justice system. For women’s groups there are violence against women units and resources devoted to dealing with domestic violence and rape. One police officer spoke of how much ‘further ahead’ the police were in understanding and tackling violence against women. Campaigners do not want to pursue that policy because victims would not get a better criminal justice ‘service’. This goes to the heart of the future of hate crime policy. It might be seen as a useless term, with policy that cannot cope with the complexity of the victim experience. Instead of focusing on definitions and classifications of hate crime, debates about the merits of hate crime policy per se might be the main challenge for the future.

73

Hate Crime

Conclusion This chapter has provided a wide-ranging analysis of the problems facing hate crime policy. To examine the first key theme – the issue of hate crime victimisation and policy – it has been suggested that initiatives in this field have created the appearance of a hierarchy of victims and that notions of victimhood appear to be simplistic. Furthermore, concerns have been expressed over policy that requires a group to have campaign support and public sympathy. However, many of these criticisms of hate crime policy have been more about how that policy is implemented than about its existence. There are often issues of funding and how local advisory groups and campaign bodies represent victims rather than a critique of what hate crime policy is attempting to achieve. There are also grounds for optimism where this is concerned. The criminal justice system appears to be moving towards a more nuanced understanding of victimhood. In July 2008, the CPS issued guidance on ‘Prosecuting Crimes Against Older People’. It was a comprehensive policy and includes hate crime, abuse in a residential setting, distraction burglary and fraud, neglect and harassment (CPS 2008a, 2008b). In November 2007 Baroness Scotland, the Attorney-General for England and Wales and Northern Ireland, addressed the Tackling Hate Crime in Europe conference and used it as an opportunity to highlight the work of the Race for Justice project which was seeking to define and monitor hate crime. As part of that speech she said: I want to look for ways in which we can work with those groups who condemn all hate crimes regardless of their victims’ origins. This is why I am delighted to launch the Race for Justice declaration. It is an opportunity for all bodies, whether governmental or community led to make clear that they will not condone hate crimes against others. Within the conference literature was a declaration containing a statement that each organisation should sign as a commitment against all forms of hate crime. The title Race for Justice was defended by a senior police officer as meaning that the ‘race was on’ for justice, rather than it prioritising race as a form of hate crime. Even so, some respondents involved with the project were resentful about the title, showing that the difficult relationships between victim groups

74

Future challenges for hate crime policy: lessons from the past

will be hard to overcome. The criminal justice system seems to be moving towards policy that identifies the many different ways that a particular section of society might be victimised. The new crossgovernment Action Plan on Hate Crime specifically noted that the ‘complex interactions and combinations of the recognised hate crime strands and aspects of people’s identities can affect the nature, prevalence and reactions to … hate crime’ (Home Office 2009: 2). This does not mean that the concept of hate crime will become obsolete, but instead that hate crime policy might increasingly become aware of the intersections between different victim experiences and that none should be prioritised. As such, a victim is recognised as an individual victim with specific needs. The issue of the use of the concept of hate crime itself remains rather complex. The research found that many campaign groups from different diversity strands used it to attempt to highlight the damage caused by bigoted violence. It was also a tool to connect different forms of hate and a bureaucratic term that policy-makers understood. Campaigners did not use the term lightly, but in many cases – for disability and homophobic hate crime, for example – specifically began using it to ‘take their campaign to the next level’. However, all were aware that it was accompanied by what could be labelled ‘conceptual baggage’. That baggage consists of common-sense assumptions about hate crime that might not only affect operational policing and prosecution, but have also been used as justification for the exclusion of certain victim groups. Age and gender-related offences have not been included in hate crime policy because they are seen as crimes of a different order. Perpetrated mostly in private settings, by known offenders and without an interchangeable victim, they do not fit with common-sense assumptions of hate crime. These ideas influence policy-making decisions about what groups to include and also when campaigners decide about the merits of being included in hate crime policy. The author is not suggesting that age and gender – or any other specific groups – need to be embraced within the hate crime policy domain. Instead, there must be a discussion and debate about what policy is trying to achieve and what message is sent to excluded groups. Hate crime policy has always been about sending a symbolic message to victims – but that message goes both ways. You are either worthy or not worthy of inclusion. This dilemma will continue to be a challenge for hate crime policy and to the campaigners and practitioners who help to create it.

75

Hate Crime

Note 1 In 2003 the Scottish Executive arranged for a consultation on hate crime policy and legislation. The Working Group on Hate Crime met eight times with a variety of interest groups and practitioners. It eventually recommended extending legislative provisions to include disability, sexual orientation and gender identity. In 2005 the Scottish Executive rejected its recommendations.

References ACPO (2000) Guide to Identifying and Combating Hate Crime. London: Association of Chief Police Officers. ACPO (2005) Hate Crime: Delivering a Quality Service. London: Association of Chief Police Officers. CPS (2007a) Guidance on Prosecuting Disability Hate Crime. London: Crown Prosecution Service. CPS (2007b) Guidance on Prosecuting Homophobic and Transphobic Crime. London: Crown Prosecution Service. CPS (2008a) CPS Publishes Policy on Prosecuting Crimes Against Older People, press release, 15 July. London: Crown Prosecution Service. CPS (2008b) Guidance on Prosecuting Crimes Against Older People. London: Crown Prosecution Service. EHRC (2007) Equality and Human Rights Online. Available at www. equalityhumanrights.com [accessed 16 September 2008]. EHRC (2009) Disabled People’s Experiences of Targeted Violence and Hostility. London: Office for Public Management. Garland, J., Spalek, B. and Chakraborti, N. (2006) ‘Hearing Lost Voices: Issues in Researching Hidden Minority Ethnic Communities’, British Journal of Criminology, 46(3): 423–37. Hall, N. (2005) Hate Crime. Cullompton: Willan Publishing. Home Office (2009) Hate Crime: The Cross-Government Action Plan. London: Home Office. Horvath, M. and Kelly, L. (2007) From the Outset: Why Violence Should be a Priority for the Commission for Equality and Human Rights. London: End Violence Against Women. Iganski, P. (1999) ‘Why Make “Hate” a Crime?’ Critical Social Policy, 19(3): 386–95. Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Jacobs, J. and Potter, K. (1998) Hate Crimes: Criminal Law and Identity Politics. New York: Aldine De Gruyter. Kielinger, V. and Stanko, B. (2002) ‘What Can We Learn from People’s Use of the Police?’ Criminal Justice Matters, 48: 4–5.

76

Future challenges for hate crime policy: lessons from the past

Mason, G. (2005a) ‘Can You Know a Stranger? Racist and Homophobic Harassment in the United Kingdom’, Current Issues in Criminal Justice, 17(1): 185–201. Mason, G. (2005b) ‘Hate Crime and the Image of the Stranger’, British Journal of Criminology, 45(6): 837–59. Mason, G. (2005c) ‘Being Hated: Stranger or Familiar?’, Social and Legal Studies: An International Journal, 14(4): 585–605. Morgan, J. (2002) ‘US Hate Crime Legislation: A Legal Model to Avoid in Australia’, Journal of Sociology, 38(1): 25–48. Perry, B. (2004) ‘The Semantics of Hate’, Journal of Hate Studies, 4: 121–37. Ray, L. and Smith, D. (2001) ‘Racist Offenders and the Politics of Hate Crime’, Law and Critique, 12(3): 203–21. Scottish Executive (2004) Working Group on Hate Crime Report. Edinburgh: Scottish Executive. Straw, J. (1998) House of Commons Debates, 23 June 1998, col. 894. Taylor, K. (1996) ‘Treating Male Violence against Women as a Bias Crime’, Boston University Law Review, 76: 575. Weisburd, S. and Levin, B. (1994) ‘On the Basis of Sex’, Stanford Law and Policy Review, 2: 21–47.

77

Chapter 4

Homophobic hate crime in Northern Ireland Marian Duggan

Introduction In 2008, Democratic Unionist Party (DUP) MP Iris Robinson, who at the time was both Chair of the Stormont Health Committee and the wife of the First Minister Peter Robinson, stated live on a national breakfast radio interview that homosexuality was an ‘abomination’ which ‘nauseated’ her, suggesting that homosexuals could be ‘cured’ with psychiatric treatment, while promoting the services of a ‘very nice’ psychiatrist who helped to ‘reorientate’ homosexuals back to heterosexuality (Young 2008). She defended her statements claiming that she believed it was the ‘duty of Government to uphold God’s law’ (Belfast Telegraph 2008a). Despite a significant public outcry and a police investigation into the comments, no action was taken against Mrs Robinson either by her party or the police. However, the incident proved significant in showing the cultural differences that surround attitudes towards homosexuality in Northern Ireland and the impact these can have on addressing and combating homophobia. Homophobic hate crime towards members of lesbian, gay, bisexual and transgender (LGBT) communities in England is an issue that has come to the fore of UK public policy following the high-profile murder investigations of victims David Morley in 2004 and Jody Dobrowski in 2005. Less attention, however, has been paid to the violence incurred by members of LGBT communities in Northern Ireland. Subsequently, theories of hate crime in Northern Ireland remain significantly absent when compared to those advanced in Great Britain or other comparable western societies (see Comstock 78

Homophobic hate crime in Northern Ireland

1991; Herek and Berrill 1992; Mason and Palmer 1996; Mason and Tomsen 1997). While some existing theories may be attributable to similar incidents in Northern Ireland, it may also be the case that there are culturally relevant factors that need to be taken into consideration as well. Although Mrs Robinson’s comments highlighted the problem of homophobia in Northern Ireland, little other literature or research exists that explores this cultural prejudice. Northern Irish LGBT persons have been as marginalised by academic researchers as they have by some members of their own society. The Police Service of Northern Ireland (PSNI) officially began recording homophobically motivated incidents in 2000 and homophobically motivated crimes in 2004. Around the same time, research into LGBT people’s experiences of homophobia began to emerge (Quiery 2002; Jarman and Tennant 2003; McNamee 2006). Official statistics, supplemented by these studies, indicate that homophobic hate crime is a form of prejudice experienced by many LGBT people but reported by few. This chapter aims to highlight some of the culturally relevant issues surrounding homophobic hate crime in Northern Ireland. Taking into consideration Northern Ireland’s recent sociopolitical history when analysing this form of prejudice, the theories offered may account for both the high likelihood of experiencing homophobia and the low level of reporting of this particular form of prejudice. The examination of homophobic hate crime in this chapter begins with an overview of the ongoing transition towards a ‘new’ Northern Ireland. Part of this process involves ensuring the equality and protection of members of minority identity groups through policies and legislation. An important part of this legislation has been recognising the existence of ‘hate crime’. The analysis of homophobic hate crime in Northern Ireland juxtaposes official statistics (incidents and crimes recorded by the PSNI as well as clearance and prosecution rates) with unofficial research (studies conducted with LGBT people) to illustrate the problems with this information. Namely, while police figures suggest that the likelihood of discrimination is low, studies with LGBT respondents reveal that fears and experiences of this form of prejudice remain high. Some reasons for this discrepancy are explored with particular reference to PSNI procedures in response to homophobic hate crime. The chapter then moves on to take a broader perspective of prejudice against LGBT people in Northern Ireland. Traditional conceptualisations of homophobia cite it as being predicated upon a perceived threat to the family and society (Weinberg 1972). It is 79

Hate Crime

then suggested that these feelings are enhanced in Northern Ireland, where prejudicial attitudes towards members of LGBT communities have been widely documented (NILT 1998, 2004, 2005; LASI 2006; Borooah and Mangan 2007; Equality Commission Northern Ireland 2009). There follows an examination of how homophobic attitudes both inform and exemplify the morally conservative nature of Northern Irish society. This in turn illustrates the influence of religious doctrine on political discourses as a particularly pertinent cultural factor informing these negative attitudes. Finally, the chapter concludes by returning to assess the sociopolitical positioning of lesbians and gay men in the ‘new’ Northern Ireland.

The ‘new’ Northern Ireland Northern Ireland is a society undergoing a period of economic, social and political transition after three decades of an often violent ethnopolitical conflict, known as the Troubles (McKittrick and McVea 2001). Societies in Northern Ireland were divided along sectarian lines, with many communities incurring regular episodes of violence. Fighting between the segregated Catholic and Protestant populations and the British security forces resulted in 40,000 casualties and 3,500 deaths by the time a peace agreement was reached the mid-1990s (Fay et al. 1999; McKittrick et al. 1999; Amstutz 2004). Given Northern Ireland’s population of just 1.7 million, the figures indicate the severity of the conflict and the importance of the peace process currently under way. The Belfast (Good Friday) Agreement 1998 signalled an end to the worst years of the Troubles and the start of the ongoing peace process. Some of the changes that have characterised this new postconflict era include several significant amendments to policing and legislation. Developments in policing objectives from regulating to including the community are part of the post-Patten era of institutional change in the police service. The Patten Report, published following the signing of the 1998 Agreement, recommended a number of changes, most designed at harmonising members of the community traditionally divided across sectarian lines. Negative connotations associated with the Royal Ulster Constabulary (RUC) centred on the inherent British ethos and suggestions of brutality towards members of Catholic communities. To address this, the RUC was renamed the PSNI and began actively recruiting members of Catholic communities, as well as women, ethnic minorities and members of the LGBT community. 80

Homophobic hate crime in Northern Ireland

The enactment of various laws and policies enshrining equality and promoting good relations, trust and mutual respect has set the scene for Northern Ireland currently being at its most culturally diverse, benefiting largely from migration as a result of a growing European Union. These values were enshrined in the first of the Programme for Government documents, which offered guidelines to oversee the implementation of such policies that would embed these values in the ‘new’ Northern Ireland. The enactment of laws concerning the commission of crimes motivated by prejudice – or ‘hate crimes’ – were included in this attempt to ensure greater and more visible equality. The key organisations tasked with implementing equality and combating hostility included the Office of the First Minister and Deputy First Minister (OFMDFM), based at the Stormont Assembly, and to a lesser degree the Northern Ireland Office (NIO) based at Westminster (CJINI 2007). Stormont is a devolved administration. Politicians, known as Members of the Legislative Assembly (MLAs), do not currently have power over criminal justice policies. Therefore amendments to enact hate crime legislation have been implemented by Westminster and not directly as a result of requests made by Northern Irish politicians or the Northern Irish electorate. While this has proved advantageous to members of LGBT communities in Northern Ireland, the implementation of these policies has failed to address the underlying prejudices to homosexuality evident in Northern Ireland’s history.

Homosexuality and the law in Northern Ireland As with the rest of the UK, homosexuality was at one time criminalised in Northern Ireland. However, the recommendations made in the Wolfenden Report, which led to the decriminalisation of homosexuality in England and Wales in 1967, were not extended to Northern Ireland. Fierce moral opposition to these proposals were headed by the Reverend Ian Paisley, who mounted a temporarily successful ‘Save Ulster from Sodomy’ campaign in the early 1970s. As a result, homosexuality remained illegal and the RUC continued to bring prosecutions for gross indecency against gay men under the Criminal Law (Amendment) Act 1885. It was this continued persecution that led to Belfast citizen Jeffrey Dudgeon and a cohort of other gay men to take their case for discrimination under article 8 of the European Convention of Human Rights (which covers the right to private life) to the Strasbourg Court in the late 1970s. After 81

Hate Crime

a lengthy process, the Court found the UK to be in violation of this right, leading to the repeal of homosexual criminalisation in Northern Ireland in 1982.1 Few other developments occurred until the swathe of equality legislation that was implemented as part of section 75 of the 1998 Agreement. This section specifically seeks to promote ‘good practice and due regard’ for minority groups in Northern Ireland. What this means is that whereas in Great Britain public authorities are advised, but not obliged, to engage in minority consultation processes, in Northern Ireland such consultations are mandatory. The requirement is important in that LGBT communities are now officially recognised as stakeholders in social and political initiatives (Feenan et al. 2001; Loudes 2003). As a result, LGBT rights are now protected by laws such as the Employment Equality (Sexual Orientation) Regulations (Northern Ireland) 2003, the Civil Partnership Act 2004, and the Equality Act (Sexual Orientation) Regulations (Northern Ireland) 2006. While these laws are beneficial to members of LGBT communities in Northern Ireland, a top-down approach to addressing minority victimisation may overlook underlying social factors contributing to this form of prejudice which may require more than mere legal reform.

Developing hate crime legislation For a society affected by decades of hostility largely predicated on the basis of identity, the notion of legislating for ‘hate crime’ in Northern Ireland may, at first, seem anomalous. However, the enactment of laws indicated an important factor that had long been overlooked in Northern Ireland as a result of the ongoing conflict. Prejudice-based crimes against people on the basis of their race, religion and sexuality do occur in Northern Ireland. However, very little information is known about these incidents as for decades they were rendered as less important than the more pressing political and social issues incurred through sectarian divisions. As with the rest of the UK, the development of hate crime legislation can be traced to the Crime and Disorder Act 1998, which created racially aggravated offences in England and Wales. This was largely due to the 1993 murder of teenager Stephen Lawrence and subsequent criticisms surrounding the Metropolitan Police Service’s handling of the investigation. Sexual orientation as a basis for prejudice and inclusion in hate crime legislation was initially recognised following the 1998 murder of a young gay man, Matthew Shepard, in Wyoming, USA. 82

Homophobic hate crime in Northern Ireland

Homophobically motivated crime in Northern Ireland is covered by the Criminal Justice (No. 2) (Northern Ireland) Act 2004. Like similar legislation in England and Wales (outlined in section 146 of the Criminal Justice Act 2003), this aggravating factor, if proved, must be taken into account by the court during sentencing. However, unlike cases aggravated by racial or religious hostility, perpetrators of crimes motivated by sexual orientation hostility are not charged with a specific sexual orientation offence. Instead, perpetrators of such hate crimes are charged with an existing offence, such as assault, and the sexual orientation motivation is taken into account during sentencing. The additional tariff imposed is discretionary and based on the individual decision of the judge overseeing the sentencing. Hate crime legislation also covers incitement to hatred, or ‘hate speech’. Whereas lengthy debates took place in England and Wales as to whether or not to extend laws to cover incitement to hatred on the basis of sexual orientation, in Northern Ireland this issue was resolved some years earlier.2 The Criminal Justice (Northern Ireland) Order 2004 amended the Public Order (Northern Ireland) Order 1987 to include sexual orientation to the categories covered by incitement to hatred. This Order specifies that hatred can be incited ‘against a group of persons’ or that a person can be charged with ‘arousing fear’ or ‘fear of persons’. This Order supersedes the equivalent Act in Great Britain on two grounds. First, the additional inclusion of the ‘arousing fear’ element theoretically affords greater protection to members of LGBT communities in Northern Ireland. Second, unlike comparable legislation in England and Wales, there is no requirement to prove ‘intent’ for a successful prosecution. Instead, the Order is worded to say: ‘having regard to all the circumstances hatred is likely to be stirred up or fear is likely to be aroused thereby’. However, as demonstrated by the inaction following Iris Robinson’s comments, there is a great deal of confusion over when, how and against whom this ‘incitement to hatred’ law can, or will, be used in Northern Ireland. On the other hand, the use of the homophobic hate crime legislation against offenders has proved to be somewhat less complicated for the PSNI to implement.

Recording homophobic hatred The PSNI officially began recording homophobic incidents in 2000. For the first four years, on average 50 homophobic incidents were reported annually. In 2005 this number rose significantly to 196. From 2005 to 2008, on average 183 incidents were reported to the 83

Hate Crime

PSNI annually. In 2004, the PSNI began recording homophobically motivated crimes as well as clearance rates for these crimes. A breakdown of these figures is shown in Table 4.1. To give an indication of where homophobically motivated crime sits in relation to all forms of recorded crime in Northern Ireland, Table 4.2 presents statistics for all crimes and clearance rates from the period 2001/02: Although homophobic hate crimes make up a small proportion of the total crime recorded in Northern Ireland, this ought not to mean that they are taken any less seriously. Iganski (2008) has shown that hate crimes harm not only the individual but also their immediate and wider communities more than other non-prejudiced crime. A further breakdown of these statistics shows that on average, half of the homophobically motivated crimes recorded by the PSNI involved a physical assault to the person, often serious enough to incur some degree of wounding. In comparison to the information on crimes against other minority communities in Northern Ireland, homophobically motivated crimes are more likely to involve violence (PSNI 2008). Recognising this trend, official reports speculate that clearance rates for homophobically motivated hate crimes are higher than for other forms of hate crimes due to this larger percentage of violent assaults where a perpetrator may be more easily identified (CJINI 2007). While these figures are low in comparison to Table 4.1  Homophobically motivated incidents, crimes and clearance rates from 2000/01 to 2008/09 Type

00/01 01/02 02/03 03/04 04/05 05/06 06/07 07/08 08/09

Incident 57 40 Crime N/A N/A Clearance   rate (%) N/A N/A

35 71 N/A N/A

196 151

220 148

155 117

160 114

179 134

N/A N/A

22.5

32.4

23.1

15.83

21.6

Table 4.2  All recorded crimes and clearance rates in Northern Ireland from 2001/02 to 2008/09 Type

01/02

02/03

03/04

04/05

05/06

06/07

07/08

08/09

Crime 139,786 142,496 127,953 118,124 123,194 121,144 108,468 110,094 Clearance   rate (%) 20.0 23.0 27.4 28.2 30.6 23.6 20.5 23.0 84

Homophobic hate crime in Northern Ireland

general crime figures, research undertaken with members of LGBT communities in Northern Ireland has indicated that they illustrate a wider culture of under-reporting. Several pieces of research have sought to show the ‘view from below’ in terms of experiences of homophobic victimisation, owing to the recognition of low levels of reporting such incidents. In the first of these, conducted in 1996 by Stonewall, a UK LGBT pressure group, 39 per cent of respondents reported experiencing homophobic violence, 36 per cent experienced homophobic harassment and 67 per cent had been verbally abused (Mason and Palmer 1996). In 2003 the Institute for Conflict Research, an investigative body established in light of the Agreement, conducted the first major study into homophobic violence in Northern Ireland (Jarman and Tennant 2003). The research surveyed 186 people from LGBT communities in Northern Ireland, revealing that homophobia was a serious problem for the respondents. A significant proportion, 82 per cent, had experienced harassment while 55 per cent had been subjected to physical violence. Overall, the study showed that the percentage of respondents who had experienced harassment and violence in Northern Ireland was higher than comparable surveys in Great Britain and Ireland. Not only was the proportion of experiences higher, but the nature of the incident was seen to incur a greater deal of violence to the victim. One respondent in the study observed that ‘there [is] now a greater use of violence and a greater propensity to use violence in [homophobic] attacks’ (Jarman and Tennant 2003: 65). Only 42 per cent of those who experienced homophobic harassment had reported the incident to the PSNI. Many chose not to report their experience due to the belief that the police would not or could not help in any way, that the incident was too trivial, that the police were homophobic themselves or because people were reluctant to reveal their sexual orientation. While this study remains the largest study of homophobia in Northern Ireland, other more recent pieces of research have highlighted particular areas in which Northern Irish lesbians and gay men have been exposed to varying levels of victimisation. The Lesbian Advocacy Services Initiative (LASI), a Belfast-based organisation focused on lesbian and bisexual women and their families, regularly undertake research with women in order to highlight the specific forms of discrimination incurred as a result of their gender. Lesbian women are notoriously difficult to access, particularly women in Northern Ireland, as they may fear coming or 85

Hate Crime

being out if they have children, are in jobs where homosexuality has traditionally been suspect (such as youth work, nursing or teaching), or have prejudiced friends/family members/work colleagues (Quiery 2002, 2007). Despite increases in visibility and representation in the community, the fact that these fears still remain contravenes assumptions that life is now ‘easier’ for lesbian women to be open about their sexuality to others around them. Almost half of the 160 women involved in Quiery’s study (2002) had experienced varying levels of discrimination at work. Virtually all the women questioned recounted experiences of ‘lower level’ harassment and abuse – such as verbal victimisation – with some indicating that this had become a regular occurrence. However, more worrying was the fact that one in five women had experienced at least one violent assault that they perceived as homophobic. Moreover, not one of these women had reported their experiences to the police. Instead, the women internalised these feelings and adapted their identities to the point where they felt they had a greater level of freedom (or invisibility). This position has been outlined by Berrill (1992) who noted that lower reporting rates among lesbian women may be a result of their reduced visibility or the complexity in distinguishing homophobic violence, which is specifically antilesbian, from misogynistic violence, which is generally anti-woman.

Accounting for official low figures Efforts to encourage greater reporting have been commended in reports on hate crime in Northern Ireland but the significant level of under-reporting was noted in a review of the criminal justice system (CJINI 2007). The PSNI follows the Association of Chief Police Officers (ACPO) guidelines on the recording of incidents and crimes that may have been motivated by hostility towards the victim’s identity or perceived community affiliation.4 However, the 2007 report by the Criminal Justice Inspectorate Northern Ireland suggested that three years after the implementation of hate crime legislation some officers still required training to understand the subjective perception element of hate crimes, which need not be substantiated by evidential requirements at the time of reporting (CJINI 2007). While hate crime definitions were seen to be generally understood by PSNI officers, further investigations into their application, coupled with the issue of how the incident progressed and the retention of the aggravating element, highlighted some inherent problems. Subjective perceptions by the police, together with the need to meet clearance rate targets, 86

Homophobic hate crime in Northern Ireland

meant some discrepancies in how incidents were recorded on the police computer systems, which in turn affected the information being passed on for prosecution (CJINI 2007). Complaints by members of the Public Prosecution Service (PPS) about incomplete or contradictory information being provided by the PSNI led them to query some police officers’ understandings of what constitutes a hate crime for the purposes of recording incidents. Prosecutions for homophobic hate crime are notoriously low in Northern Ireland. As of 2007, there were just 66 prosecutions for summary offences and 27 prosecutions for indictable offences where hostility on the basis of sexual orientation was an aggravating factor. While ‘aggravation’ may be a difficult element to prove, it has been noted that there is a problem in tracking the course of a hate crime through the criminal justice system in order to assess exactly where the attrition is occurring (CJINI 2007). Furthermore, the high level of rejections by the PPS of offences recorded as being motivated by hostility to the victim’s sexual orientation was of particular note to the Criminal Justice Inspectorate, as the figures stood out as far higher than for any other form of prejudice in comparison. The low level of prosecutions, coupled with the high levels of the hate crime element disappearing before conviction, suggests that perceptions of the existence or severity of homophobic hostility in a given crime may be too subjective. Although the ACPO guidelines were intended to be utilised in a manner that facilitated the incorporation of prejudice if it was felt to have shaped a particular incident or crime, the vagueness of these instructions may be having an adverse effect in a society where homophobic hostility is more visible and less challenged, as in Northern Ireland.

Understanding homophobia in Northern Ireland In Northern Ireland there may be specific cultural factors informing and sustaining homophobic ideologies that are less evident in comparable societies. Discourses of homosexuality have historically fluctuated between ‘nature/nurture’ debates, most of which have attempted to account for this form of sexual ‘deviancy’ rather than the hostility incurred (Plummer 1975; Weeks 1977). At different points in UK history, these theories have informed criminal and medical treatment programmes designed to ‘cure’, contain or control homosexuality (Kinsman 1996; Moran 1996). While such negative responses dominated ideologies on same-sex desire for centuries, 87

Hate Crime

it was not until the mid-twentieth century and a shift from seeing homosexuality as a biological issue to social construction paved the way for alternative perspectives on the victimisation of homosexuals. In other words, rather than situate the focus on the homosexual individual as ‘deviant’ or ‘wrong’, analyses sought to examine how this individual and their identity was responded to by society and why this might incur such negativity and fear. An important attribute to this shift was the coining of the term ‘homophobia’ by Dr George Weinberg in 1972. His intention was to depict the socially informed fear or hatred that some people held towards homosexuality and/or homosexuals: I coined the word homophobia to mean it was a phobia about homosexuals … It was a fear of homosexuals which seemed to be associated with a fear of contagion, a fear of reducing the things one fought for – home and family. It was a religious fear and it had led to a great brutality as fear always does. (Weinberg, cited in Herek 2004: 7) Weinberg’s key concern was that notions of ‘homophobia’ should focus explicitly on redefining the problem as a heterosexual intolerance of homosexuality. In other words, the existence of homosexuality was not in itself the problem; it was the responses it elicited and the normalisation of these that required attention. Weinberg therefore queried the basis for this hostility, in particular the socially constructed nature of homosexuality in a given society that gave rise to such fear or hatred. Weinberg’s groundbreaking perspective has since been adopted in many western studies into homophobia and homophobic violence which seek to situate this form of prejudice within the individual’s specific social environment. For example, Herek (1991) attributed social psychology to incidences of homophobia, invoking the role of cultural stereotypes and negative ideologies in violence towards lesbians and gay men. Narrowing this down, Comstock (1991) highlighted the importance of deconstructing hegemonic masculinity among young groups of males in his research on violence towards lesbians and gay men in America. In Australia, Mason and Tomsen (1997) analysed the impact of HIV/AIDS on the fears attributed to gay men and the impact of these on elevated levels of homophobic violence. However, while these studies sought to situate the prejudice within the individual perpetrators, other analyses of homophobia have sought to show that isolated incidents may be endemic of 88

Homophobic hate crime in Northern Ireland

unaddressed wider social issues, locating minority prejudice within the specific sociopolitical environment (Perry 2001; Herek 2004; Hall 2005; Iganski 2008). Perry (2003) has suggested that the violent victimisation of minority groups may be an indication of particular perpetrators ‘doing difference’: that is, acting upon the prejudices held by many rather than the few. Perry suggests that even extreme violence is, in some cases, a manifestation of prejudices held by communities, which in turn demonstrates the social structures of labour, power, sexuality and culture. This may account for why some victims are seen to incur more sympathy and justice than others. Perry (2003) has also argued that exclusions and cultural ideologies expose particular members of minority groups to hostility, simultaneously constructing them as vulnerable yet legitimate targets. She suggests examining the connections between cultural oppression and personal experiences of violence through taking a broader look at how minority groups have been constructed in any one particular society.

Morally conservative prejudices Opposition to homosexuality in Northern Ireland has been documented in several studies that have aimed to show a snapshot of opinions over a cross-section of the population since equality legislation was implemented. For example, in 2006 LASI questioned over a thousand Northern Irish people on their general attitudes to lesbians and gay men. Of these respondents, 88 per cent were supportive of the principle that lesbians and gay men should not be discriminated against (LASI 2006); 75 per cent claimed that they were either ‘quite accepting’ or ‘very accepting’ of lesbian and gay people in society. Two-thirds of respondents thought that in Northern Ireland sexual minorities were generally ‘not very’ or ‘not at all’ accepted. While these reports were informative at a general level, the reasons behind low rates of acceptance were not probed. A year later, in 2007, a study into western bigotry cited Northern Ireland (tied with Greece) as the most homophobic country out of the 23 countries and 32,000 people surveyed (Borooah and Mangan 2007). In one of the questions, the researchers asked respondents to pick from various different groups of people those they would least like to have as their next-door neighbour. Over a third of the Northern Irish respondents chose gay people, from a list that also included minority religious and ethnic groups. A survey by the Equality Commission in 2009 demonstrated that 35 per cent would mind having a close relative 89

Hate Crime

in a relationship with a lesbian, gay or bisexual person, and 23 per cent would mind a gay, lesbian or bisexual person living next door (Equality Commission Northern Ireland 2009). Other studies have given more in-depth analyses of where prejudice lies in society by providing the demographics of these types of opinions. The Northern Ireland Life and Times (NILT) survey is one such resource. This was established in 1998 to provide an insight into people’s lives, attitudes and perspectives on a range of social and political issues. The survey has occasionally included questions relating to attitudes towards lesbians and gay men, in particular whether homosexual sex is seen as ‘wrong’. In 1998, 72 per cent of those asked agreed it was always wrong no matter what the circumstances (NILT 1998). Similarly high proportions of Catholics (67 per cent) as Protestants (78 per cent) believed that homosexual sex was wrong. By 2004, when the same question was asked again, the proportion of people who believed that homosexual sex was ‘wrong’ had dropped overall to 61 per cent with a more significant reduction in Catholic opposition (51 per cent) than for Protestants (73 per cent) (NILT 2004). Following the 2004 Civil Partnerships Act, in 2005 respondents were asked whether lesbians and gay men should have the right to marry; 35 per cent of the total sample said ‘yes’, women approved more often than men, and Catholics approved more often than Protestants (NILT 2005). The importance of including the differences in religious demo­ graphics is specifically attributable to Northern Ireland’s culturally relative history. In the new post-conflict era, measures are taken to account for views from both sides of the traditional divide. However, Catholicism and Protestantism, though both Christian denominations, differ significantly in their understanding and acceptance of homosexual identities. Catholic teaching understands homosexuality through a natural law perspective and papal direction, specifically condemning ‘homogenital expression’ (Nugent and Gramick 1989). The fundamental tenets of Catholicism centre upon a forgiving and repentant framework; as a result, some lesbians and gay men in Northern Ireland have cited it as being the easier religion in which to be a sexual minority (Jarman and Tennant 2003; Kitchin and Lysaght 2004; McNamee 2006). However, this forgiveness can be seen as a failure to accept the person’s sexuality, instead hoping to guide them towards a heterosexual orientation. Protestant perspectives on the whole rely more heavily on biblical testimony and classify homosexuality as

90

Homophobic hate crime in Northern Ireland

part of the whole being (Kitchin and Lysaght 2004).5 Within the many facets and denominations of Protestantism, the individual is perceived as a sinner from birth; abortion and homosexuality are set apart as particularly abhorrent sins. Particular individuals are designated as the moral judges over other, less moral individuals, with the central concept of salvation through repentance, and rejection, of sin. Rarely are discourses heard where the homosexual in his or her entirety is welcomed, unquestioningly, into the mainstream parish. This discrepancy is widely understood as inherent in Northern Ireland culture, as was outlined by one respondent in Kitchin and Lysaght’s (2004: 89) study: I know from the [lesbian and gay] Helpline where people do have problems, where people raise religion and have difficulty reconciling their sexuality, their religion tends more often to be Protestant, Presbyterian, Free Presbyterian … rather than Catholic … You don’t have this sort of fire and brimstone kind of speeches in the Catholic Church or anything like that. In other words, while Catholicism generally welcomes the person so long as they abstain from engaging in ‘homosexual acts’, Protestantism can really only accept the person once they have renounced both the homosexual identity and the desire to engage in such acts. The guilt imposed about sexuality can be overcome in Catholicism with abstinence, whereas in Protestantism an outright rejection is necessary. This is a significant issue for lesbians and gay men in Northern Ireland given the permeation of Christianity in society and the possibility of their own ascription to a faith that they may wish to remain part of. Whether or not religion is a reliable indicator of social values in contemporary Northern Irish society, these arguments are used politically to advance policy issues. For example, opposing legislation to give women in Northern Ireland the right to an abortion is one area where traditional religious and political discourses claim to be speaking for the electorate. The 1967 Abortion Act, which decriminalised abortion in England and Wales, has not been extended to Northern Ireland despite several attempts (Fegan and Rebouche 2003).6 This level of moral conservatism has been strongly visible in political discourses about homosexuality in recent years, with the most vocal opposition to homosexuality in Northern Ireland societies being from those who ascribe to a Protestant denomination.

91

Hate Crime

Politically informed condemnations The biblical literalism of the Christian Religious Right has been illustrated in political condemnation of homosexuality in Northern Ireland. The degree to which political and legal policy is heavily influenced by religious doctrine has become more evident with the onset of (and opposition in part to) equality legislation, particularly in relation to sexual orientation equality. While politics aims to become more mainstream, active resistance to modernisation is still advocated in religious circles. The interpretation of biblical discourses by politicians renders their conservatism more fundamental in nature than in other areas of the UK. Furthermore, the general culture in which this conservatism is displayed often allows them to feel comfortable in publicly denouncing lesbians and gay men without fear of repercussions. Over 16,000 people signed an online Downing Street petition urging that Mrs Robinson be reprimanded for her statements. The petition cited the amendment provided in the Public Order (Northern Ireland) Order 1987, which covers incitement to hatred or stirring up fear on the grounds of sexual orientation. However, Prime Minister Gordon Brown dismissed suggestions that she had breached ministerial protocol, stating that MPs were ‘accountable to their electorate for their own comments’ (BBC News 2008). The Assembly Ombudsman later agreed that her comments had not breached the Ministerial Code of Conduct, stating that ‘while Members must be entitled to express their opinions, as public representatives they have a particular responsibility for the manner in which they express such beliefs’ (Gordon 2009). In response to complaints, the PSNI claimed that ‘there were no reasonable grounds to believe an offence had been committed’ (Gordon 2009). This was despite later revelations that Mrs Robinson had stated that ‘there can be no viler act, apart from homosexuality and sodomy, than sexually abusing innocent children’ (Belfast Telegraph 2008b) and had compared homosexuals to murderers (Henry 2008). The controversy surrounding Mrs Robinson’s defamatory comments illustrated an opportune moment to rectify the damage done by previously negative ideologies. However, the failure to reprimand her spoke volumes about her political party’s continuing attitude to sexual minorities and cast doubt on their ability to effectively implement equality measures designed to protect members of LGBT communities. Mrs Robinson’s comments were not the first to call into question MLAs’ accountability to their LGBT electorate. In 2007, 92

Homophobic hate crime in Northern Ireland

former DUP Junior Minister, Ian Paisley Jr, who was then Minister for Equality, was quoted in Hot Press magazine as saying: I am, unsurprisingly, a straight person. I am pretty repulsed by gay and lesbianism. I think it is wrong. I think that those people harm themselves and – without caring about it – harm society. That doesn’t mean to say that I hate them. I mean, I hate what they do. (O’Toole 2007) Had either Mr Paisley or Mrs Robinson been prosecuted, then the symbolic nature of this may have proved to be an important factor in the battle against hate crime in Northern Ireland. This was outlined in the Criminal Justice Inspectorate’s report, which specifically stated that ‘prosecutions need to attract widespread media attention to act as a deterrent to others that hate crime will be robustly tackled by the criminal justice system’ (CJINI 2007: 46). Therefore, the failure to take any action following Mrs Robinson’s comments could be seen as another missed opportunity to show that this type of prejudice will not be tolerated in society. Alternatively, the lack of action may be construed by some as meaning the exact opposite, namely that certain members of society continue to constitute ‘acceptable victims’ of such prejudice (Jarman and Tennant 2003).

Conclusion Discussions around hate crime in Northern Ireland have increased in prominence as a result of numerous racist attacks on members of the Roma community (McAleese 2009a) and the sectarian murder of Kevin McDaid (McAleese 2009b). Political representatives are usually quick to condemn such violence and the intentions of the attackers in an eager bid to show that Northern Ireland is indeed moving on from the problems of its past. However, a deafening silence surrounds similar attacks on members of LGBT communities. The message appears to be that homophobic prejudices in Northern Ireland are not regarded with the same seriousness as racial, religious or sectarian forms of discrimination. Unopposed statements that continue to portray lesbians and gay men as engaging in deviant lifestyle choices negates their social positioning as legitimate citizens or potential victims of violent prejudice. This in turn has implications for interaction with members of the criminal justice system, be they making, passing or enforcing 93

Hate Crime

laws designed to protect the persecuted. Paying lip-service to equality directives may tick the necessary political boxes but will do little to change the quality of people’s lives in the long term. Herein lies a vicious circle: for lesbians and gay men to be recognised as equal citizens in Northern Ireland and dispel popular myths, such as those displayed by the above MPs, they will have to become more visible in everyday life. However, in doing so, they may be exposing themselves to increased risk and vulnerability, knowing that their experiences of violence may then be less prioritised by those charged with protecting them. The mixed messages currently being sent out – that lesbians and gay men are targets of hate yet are complicit in their own victimisation – mean that the focus continues to detract from dispelling the negative imagery that impedes LGBT integration, instead fuelling debates around whether or not homosexuality is ‘wrong’ and against ‘God’s will’. While reports on hate crime in Northern Ireland have highlighted the need for a closer, more reciprocal, relationship between the PPS and the PSNI when dealing with incidences of hate crime, there is still much to be done outside of the criminal justice system (CJINI 2007). Efforts to progress Northern Irish society away from the problems of its past may be too eager to overlook the potential remnants of this past on current prejudices. Indeed, the impact of three decades of violent conflict have been highlighted as culturally informing contemporary prejudices towards those deemed ‘other’ in society (Steenkamp 2005; Knox 2002; CJINI 2007; Duggan 2008). While economic viability is promoted as beneficial to all, this eagerness to promote a ‘new’ Northern Ireland may be overlooking the costs to particular individuals in society. As part of learning the lessons of the past, members of the Westminster government, the Northern Ireland Assembly and the PSNI have recognised that only a joint denouncement of issues such as sectarianism and racism will allow Northern Ireland to be seen as an inclusive and progressive society. However, while homophobic victim-blaming ideologies prevail, prejudiced members of society in Northern Ireland may continue to act – with impunity – upon the divisive views and negative imagery in a similar manner to that characterised during the worst years of the Troubles.

Notes 1 Dudgeon v. U.K. 45 Eur. Ct. H.R. (ser. A) at 14 (1981). 94

Homophobic hate crime in Northern Ireland

2 Section 74 and Schedule 16 of the Criminal Justice and Immigration Act 2008 amended the Public Order Act 1986 to include the offence of stirring up hatred on the grounds of sexual orientation. This covers conduct – namely words or behaviour – or material that is threatening in nature and is intended to stir up hatred against a group of people who are defined by their sexual orientation. 3 The Home Office significantly restricted the clearance types available to the police from 1 April 2007, which meant that virtually all clearances resulting in ‘no further police action’ (i.e. non sanction clearances) could no longer be claimed as a valid clearance. This change affected the way statistics were collated, appearing to lead to a lower clearance rate. 4 ACPO defines hate incidents in the UK as: ‘Any incident, which may or may not constitute a criminal offence, which is perceived by the victim or any other person, as being motivated by prejudice or hate’; and hate crimes as: ‘Any incident, which constitutes a criminal offence, perceived by the victim or any other person, as being motivated by prejudice or hate’ (ACPO 2005). 5 An exception to the Protestant militancy against homosexuality is the Church of Ireland which is more liberal and accepting of lesbians and gay men and seeks to promote this perspective within Christianity. The Church of Ireland works closely with the group Changing Attitude Ireland, who describe themselves on their website as ‘A network of people, gay and straight, lay and ordained, working for the full affirmation of lesbian and gay Christians within the Churched in Ireland’ (www. changingattitudeireland.org/index.html?_ret_=return). 6 Similarly, much opposition surrounded the lowering of the age of consent from 17 years of age to 16 for both homosexual and heterosexual sexual intercourse in Northern Ireland. This brought the law in line with the rest of the UK, through the Sexual Offences (Northern Ireland) Order 2007. Also, applications for divorce in Northern Ireland may only be made after two years of marriage, as opposed to one year in England and Wales and no time limit at all in Scotland.

References ACPO (2005) Hate Crime: Delivering A Quality Service. London: Home Office Police Standards Unit. Amstutz, M. (2004) The Healing of Nations: The Promise and Limitations of Political Forgiveness. Lanham, MD: Rowman and Littlefield. BBC News (2008) ‘No Brown Rebuke over Gay Comments’, BBC News Online, 14 August. Available at news.bbc.co.uk/1/hi/northern_ireland/7560747. stm [accessed 10 June 2009]. Belfast Telegraph (2008a) ‘Iris Robinson: It’s Government’s Duty to Uphold God’s Law’, 18 July. Available at: www.belfasttelegraph.co.uk/news/ 95

Hate Crime

local-national/iris-robinson-itrsquos-governmentrsquos-duty-to-upholdgodrsquos-law-13912046.html. Belfast Telegraph (2008b) ‘Iris: Gays More Vile than Child Abusers’, 21 July. Available at www.belfasttelegraph.co.uk/news/local-national/iris-gaysmore-vile-than-child-abusers-13913517.html. Berrill, K. (1992) ‘Anti-Gay Violence and Victimisation in the United States: An Overview’, in G. Herek and K. Berrill (eds) Hate Crimes: Confronting Violence Against Lesbians and Gay Men. London: Sage. Borooah, V. and Mangan, J. (2007) ‘Love Thy Neighbour: How Much Bigotry is There in Western Countries?’, Kyklos, 60(3): 295–317. Comstock, G. (1991) Violence Against Lesbians and Gay Men. Manhattan, NY: Columbia University Press. CJINI (2007) Hate Crime in Northern Ireland: A Thematic Inspection of the Management of Hate Crime by the Criminal Justice System in Northern Ireland. Criminal Justice Inspectorate Northern Ireland. Available at www.cjini. org/CJNI/files/02/0272e50a-2218-482a-87e0-66a243a27900.pdf. Duggan, M. (2008) Theorising homophobic violence in Northern Ireland, Papers from the British Criminology Conference, Vol. 8, pp. 33–49. Available at www.britsoccrim.org/volume8/3Duggan08.pdf. Equality Commission Northern Ireland (2009) Equality Awareness Survey 2008. Available at www.equalityni.org/archive/pdf/ECSurvey2008.pdf. Fay, M., Morrisey, M. and Smyth, M. (1999) Northern Ireland’s Troubles: The Human Costs. London: Pluto Press. Feenan, D., Fitzpatrick, B., Mazwell, P. and O’Hare, U. (2001) Enhancing the Rights of Lesbian and Gay People in Northern Ireland. Belfast: Human Rights Commission. Fegan, E. and Rebouche, R. (2003) ‘Northern Ireland’s Abortion Law: The Morality of Silence and the Censure of Agency’, Feminist Legal Studies, 11(3): 221–54. Gordon, D. (2009) ‘MLA Discipline Code Didn’t Cover Complaints’, Belfast Telegraph, 7 July. Available at www.belfasttelegraph.co.uk/news/politics/ mla-discipline-code-didnrsquot-cover-complaints-14387944.html [accessed 30 June 2009]. Hall, N. (2005) Hate Crime. Cullompton: Willan Publishing. Henry, L. (2008) ‘Robinson: No Regrets over Gay Comments’, Belfast Telegraph, 10 June. Available at www.belfasttelegraph.co.uk/news/local-national/ robinson-no-regrets-over-gay-comments-13507741.html [accessed 30 June 2009]. Herek, G. (1991) ‘Stigma, Prejudice, and Violence against Lesbians and Gay Men’, in J. Gonsiorek and J. Weinrich (eds) Homosexuality: Research Implications for Public Policy. California: Sage. Herek, G. (2004) ‘Beyond “Homophobia”: Thinking about Sexual Stigma and Prejudice in the Twenty-First Century’, Sexuality Research and Social Policy, 1(2): 6–24. Herek, G. and Berrill, K. (eds) (1992) Hate Crimes: Confronting Violence Against Lesbians and Gay Men. California: Sage. 96

Homophobic hate crime in Northern Ireland

Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Jarman, N. and Tennant, A. (2003) An Acceptable Prejudice? Homophobic Violence and Harassment in Northern Ireland. Belfast: Institute for Conflict Research. Kinsman, G. (1996) The Regulation of Desire: Homo and Hetero Sexualities. London: Black Rose Books. Kitchin, R. and Lysaght, K. (2004) ‘Sexual Citizenship in Belfast, Northern Ireland’, Gender, Place and Culture, 11(1): 83–10. Knox, C. (2002) ‘ “See No Evil, Hear No Evil”: Insidious Paramilitary Violence in Northern Ireland’, British Journal of Criminology, 42(1):164–85. LASI (2006) LASI Ipsos Mori Survey. Ballymena: Lesbian Advocacy Services Initiative. Available at www.lasionline.org/research.htm. Loudes, C. (2003) Learning to Grow Up: The Multiple Identities of Young Lesbians, Gay Men and Bisexual People in Northern Ireland. Belfast: Northern Ireland Human Rights Commission. Mason, A. and Palmer, A. (1996) Queer Bashing: A National Survey of Hate Crimes Against Lesbians and Gay Men. London: Stonewall. Mason, G. and Tomsen, S. (eds) (1997) Homophobic Violence. Sydney: Hawkins Press. McAleese, D. (2009a) ‘Families Flee after Attacks by Racist Mob’, Belfast Telegraph, 15 June. Available at www.belfasttelegraph.co.uk/news/localnational/families-flee-after-attacks-by-racist-mob-14339286.html [accessed 30 June 2009]. McAleese, D. (2009b) ‘Casualty of Sectarian Attack Remains on Life Support’, Belfast Telegraph, 29 May. Available at www.belfasttelegraph.co.uk/ news/local-national/casualty-of-sectarian-attack-remains-on-life-support14319288.html [accessed 30 June 2009]. McKittrick, D. and McVea, D. (2001) Making Sense of the Troubles. Belfast: Blackstaff Press. McKittrick, J., Kelters, S., Feeney, B. and Thornton, C. (1999) Lost Lives: The Stories of the Men, Women and Children Who Died Through the Northern Ireland Troubles. Edinburgh: Mainstream Publishing. McNamee, H. (2006) Out On Your Own: An Examination of the Mental Health of Young Same-Sex Attracted Men. Belfast: Rainbow Project. Moran, L. (1996) The Homosexual(ity) of Law. London: Routledge. NILT (1998) Northern Ireland Life and Times Survey: Religious Observance. Available at www.ark.ac.uk/nilt/1998/Religious_Observance/SEXHOMO. html. NILT (2004) Northern Ireland Life and Times Survey: Religious Observance. Available at www.ark.ac.uk/nilt/2004/Religious_Observance/SEXHOMO. html. NILT (2005) Northern Ireland Life and Times Survey: Gender and Family Roles. Available at www.ark.ac.uk/nilt/2005/Gender_and_Family_Roles/ MARVIE24.html Nugent, R. and Gramick, J. (1989) ‘Homosexuality: Protestant, Catholic, and Jewish Issues: A Fishbone Tale’, Journal of Homosexuality, 18(3): 7–46. 97

Hate Crime

O’Toole, J. (2007) ‘The Junior Minister has his Say about Gays’, Hot Press, 20 June. Available at www.hotpress.com/archive/2931251.html [accessed 30 June 2008]. Perry, B. (2001) In the Name of Hate: Understanding Hate Crimes. London: Routledge. Perry, B. (2003) ‘Accounting for Hate Crime: Doing Difference’, in B. Perry (ed.) Hate and Bias Crime: A Reader. London: Routledge. Plummer, K. (1975) Sexual Stigma: An Interactionist Account. London: Routledge. PSNI (2008) Statistical Report No. 3: Hate Incidents and Crime. Belfast: PSNI. Available at www.psni.police.uk/3._hate_incidents_and_crimes-4.pdf. Quiery, M. (2002) A Mighty Silence: A Report on the Needs of Lesbians and Bisexual Women in Northern Ireland. Ballymena: LASI. Quiery, M. (2007) Invisible Women: A Review of the Impact of Discrimination and Social Exclusion on Lesbian and Bisexual Women’s Health in Northern Ireland. Ballymena: LASI. Steenkamp, C. K. (2005) ‘The Legacy of War: Conceptualizing a Culture of Violence to Explain Violence after Peace Accords’, The Round Table, 94(2): 253–68. Weeks, J. (1977) Coming Out: Homosexual Politics in Britain from the Nineteenth Century to the Present. London: Quartet Books. Weinberg, G. (1972) Society and the Healthy Homosexual. New York: St Martin’s Press. Young, D. (2008) ‘Gay lifestyle is “abomination” not a mental disorder: Iris’, Belfast Telegraph, 1 July. Available at www.belfasttelegraph.co.uk/ news/local-national/gay-lifestyle-is-lsquoabominationrsquo-not-a-mentaldisorder-iris-13507744.html [accessed 30 July 2008].

98

Chapter 5

Verbal and textual hostility in context Nicole Asquith1

Introduction Many democratic nations have responded to hate crime with a variety of policy and practice initiatives. Yet few have wrestled with, or solved the challenges raised by verbal-textual hostility, whether as stand-alone acts of violence, or where these acts are part of more conventional forms of hate crime such as assault, vandalism or harassment. In this chapter, approaches employed in forensic linguistics – such as Critical Discourse Analysis (CDA) – will be used to assess the illocutionary force and perlocutionary effects of hate speech-text recorded by the Metropolitan Police Service in their hate crime files. Over the last ten years, the field of forensic linguistics has developed beyond the ad hoc use of linguists in court proceedings – often to analyse the author of verbal or written evidence – into a sophisticated field of study. Contemporary forensic linguistics aims to provide a coherent theoretical and methodological framework with which to understand the force and effects of speech in crime and criminal proceedings (Coulthard and Johnson 2007). While many hate crime scholars acknowledge that verbal-textual hostility plays a significant role in victims’ subjective perceptions of hatred and police officers’ assessment of a hate crime, to date, the role of hate speech-text in hate crime has been largely uninterrogated by criminologists or linguists. This chapter aims to assess and evaluate the forensic possibilities contained in a closer reading of the words used in hate crimes. Through a Critical Discourse Analysis of incident characteristics and officers’ narratives of incidents, this chapter maps 99

Hate Crime

out how key hate speech-text indicators may assist to better evaluate the force and effects of hate crimes, and as a consequence, assist in the development of risk-assessment instruments for use by front-line officers in the evaluation of hate crimes and the harms caused by verbal-textual hostility. At the core of this research are the 99,727 case files collated by London Metropolitan Police Service (MPS) from January 2003 to December 2007. Conditions imposed by the Data Protection Act 1998 meant that a limited number of de-identified variables were made available for detailed analysis; the most important of these is the abridged narrative of the incident provided by reporting officers. In approximately 16 per cent of the MPS case files, the reporting officer recorded the verbatim verbal-textual exchange between victims, perpetrators, witnesses and/or officers. This selective sample offers important insights into hate crime reporting mechanisms, front-line officers’ construction of hate crime – and the importance they place on forensic evidence such as the words used in hate crimes – and the patterns of verbal-textual hostility used in hate crimes. It is expected that this type of contextual analysis will lead to the development of more sophisticated risk assessment tools and more targeted serviceenhancements for victims of hate crimes.

The approach At the centre of debates about the verbal-textual hostility used in hate crime is the habituated defence of free speech. From the perspective of liberal political theory – as epitomised in Mill’s On Liberty (1972 [1859]) – with free expression abridged, democracy is worthless, and incapable of generating the human progress envisioned from this type of political system. According to traditional Millian theory, free expression of the individual is central to the performance of democracy, the core values of democracy inherently assume that free expression is indivisible, and free debate informs an education in tolerance that is necessary for representative democracy in a plural society (Ten 1980: 124). According to this approach, verbal-textual hostility is ‘mere words’ that add to the ‘marketplace of ideas’ in the democratic system, which cannot – should not – be unequally regulated (no matter the unequal consequences of that speech-text). It also serves as a onesided education in tolerance whereby the marginalised must tolerate the unequal practices of the dominant, in order for the dominant to be aware of the failings of the democratic process. 100

Verbal and textual hostility in context

In the 1980s and 1990s, the fundamentally particularistic nature of free speech provisions began to be questioned by legal, political and linguistic philosophers, especially those working on the issues of pornography and ‘hate speech’. In an attempt to short-circuit the limitations imposed by the First Amendment of the US Constitution upon the regulation of ‘mere words’, MacKinnon (1993) and critical race theorists (such as Matsuda et al. 1993) have reconstructed the speech acts of pornography and ‘hate speech’ as conduct. Central to the conversion from speech to conduct was the redeployment of Austin’s analysis of perlocutionary and illocutionary performative speech acts. In How to Do Things with Words, originally published in 1955, he argues: We first distinguished a group of things we do in saying something, which together we summed up by saying we perform a locutionary act, which is roughly equivalent to uttering a certain sentence with a certain sense and reference … Second, we said that we also perform illocutionary acts such as informing, ordering, warning, undertaking … i.e., utterances which have a certain (conventional) force. Thirdly, we may also perform perlocutionary acts: what we bring about or achieve by saying something, such as convincing, persuading, deterring, and even, say, surprising or misleading. (Austin 1980: 109)2 From this classification of speech and the ways in which speech can be action, theorists such as Bourdieu and Thompson (1991), Butler (1997) and Langton (1993), have developed systems for assessing the force and effects of subordinating and silencing speech-text. However, particularly for Bourdieu and Langton, it is Austin’s deeper analysis of the forms of illocutionary speech that provides a basis for their claims about the authority to speak, and the power of authorised speech. Austin (1980: 163) details five classes of illocutionary utterances: verdictives (exercise of judgement), and exercitives (exercise of power) being the two most significant classes for the study of verbal-textual hostility. These have been reconstituted by Langton (1993: 305) as authoritative illocutions, as both require speech actors to be authorised to speak, or to speak with authority. While verdictives aim to deliver an official or unofficial finding (for example, ‘you’re a filthy Paki’), exercitives are statements about how the future should look: an advocacy, judgement or threat of things to come (‘You’re dead’). Understanding the harm generated out of efficacious 101

Hate Crime

illocutionary speech acts allows for a deeper analysis of how words wound.3 Equally, understanding the consequential – or perlocutionary – effects enables us to understand the process of incitement and the power of infecting others’ minds, and perhaps their actions. Therefore, Austin’s (1980) speech act theory, and its reworking by Bourdieu and Thompson (1991), Butler (1997) and Langton (1993), offers a theoretical approach that privileges neither action over speech nor force over effects. It also offers a framework that acknowledges that authorisation to act with hatred is both institutionally bound and socially contingent. It is difficult to understand the role of verbal-textual hostility in hate crime without getting access to what is said during the incident. Unfortunately, without an audio recording, effective forensic linguistics relies on the memory of victims or witnesses, and the verbatim recording of this evidence by police officers. At both points, the quality of the data can deteriorate: in the first instance by the trauma of violence (which can lead to memory loss and distortion) and in the latter by the legislation, policies and practices of policing hate crime (which may prioritise physical evidence over linguistic evidence, perhaps because of the first issue of data distortion). Despite these problems in working with forensic linguistics, police case records are unique documents that offer an insight into the role of verbal-textual hostility in hate crime. Just as important, critical analyses of hate crime records – as instruments of measurement and as snapshots of policing practice – can lead to the development of critical policy and practice options, which may contribute to more effective policing and prosecution of hate crimes. In order to ‘get at’ the meaning of verbal-textual hostility in hate violence, this research uses CDA to triangulate between hate violence context characteristics and hate speech-text content characteristics, and the legislative framework created to manage both (such as the Public Order Act 1986, Crime and Disorder Act 1998, Anti-Terrorism, Crime and Security Act 2001 and Criminal Justice Act 2003). CDA has developed out of the growing influence of social sciences on linguistics, especially since the emergence of sociolinguistics as an orthodox approach to language and discourse analysis. However CDA, and its methodological companion Mediated Discourse Analysis (MDA), differ significantly from sociolinguistics in that they do not begin from the point of what is said. In particular, unlike sociolinguistics, practitioners using CDA and MDA do not reduce social encounters to the words used, nor privilege language use as an ideal representation of social action. Language, as such, is not 102

Verbal and textual hostility in context

foregrounded; rather, language is constructed as one of many contexts to social action (Wodak 2001). Van Dijk (1987: 384) provides a six-step process which enables an interrogation of the contextual factors that, if not determinative, at least predispose particular intersubjective relationships between the dominated and dominant. He argues that we need to consider the following 1 2 3 4

What do people actually say? How do people talk about others? What are the communicative sources of verbal-textual hostility? What and how does such talk express or signal underlying structures and strategies of prejudice in social cognition? 5 What are the real or possible effects of prejudiced talk? 6 What are the social contexts of such talk? Researchers working with this methodological approach triangulate data across historical, cultural and political discourses in order to understand the conventions of specific intersubjective encounters. The primary hate crime discourses triangulated in this research are the subjective experiences of hate crimes (from police case files, and including victims, witnesses and responding officers), sociocultural analysis of verbal-textual hostility (such as the historicity and conventions of language use), and a mapping of institutional responses to hate violence (including legislation, policies and practices). Using CDA means that the sociohistorical roots of the words used – and how they are used – in verbal-textual hostility can be highlighted along with the institutional factors that determine the use of abuse against particular marginalised groups.

The source Despite legislative, policy and practice limitations on access to case files, in September 2009 the MPS granted limited access to 99,727 hate crime case records dating from January 2003 to December 2007. This chapter provides a detailed investigation of the 20,979 Metropolitan Police Service case files (16,623 incident files) from 2003. The five fields of deidentified data provided were: • Hate Crime Flag • Location 103

Hate Crime

• Offence • Relationship between Perpetrator and Informant • Abridged Narrative of Incident (< 250 characters) While not being a complete record, the last of these fields offers up a rich variety of additional information about the incident.4 Most importantly for this research, in 16 per cent of incidents, the abridged narrative includes a record of the verbal-textual exchanges between the perpetrator(s) and victim(s)/witness(es). This subset of the data is a manageable, largely representative, selective sample of the larger database. Prior to undertaking a CDA of the MPS data, the 2,617 incident files that included a verbatim recording of the verbal-textual hostility were coded across nine themes (see Table 5.1). Eight of the nine themes were identified in earlier research into verbal-textual hostility in Australia (Asquith 2004, 2008). As this earlier research investigated only anti-semitic and homophobic violence, an additional category was inductively developed during this coding process to capture what appears to be a practice dominant to racism – that of expatriation. To illustrate how the data were prepared for analysis, examples of this coding process are provided in Table 5.2. Table 5.1  Themes of verbal-textual hostility. Theme

Number of Frequency Frequency speech acts (Cases) (Responses)

Interpellation 2,049 78.3% 39.3% Pathologisation 124 4.7% 2.4% Demonisation 345 13.2% 6.6% Sexualisation 370 14.1% 7.1% Criminalisation 17 0.6% 0.3% Expatriation 408 15.6% 7.8% Terrorisation 565 21.6% 10.8% Profanity 1,196 45.7% 22.9% Other 145 5.5% 2.8%

104

Explanatory Notes Naming the other; calling the other into being Dirt and disease Devils, demons and mongrels (turning people into animals) Sexual organs, sexual acts Liars, cheats and criminals Exile from space, neighbourhood, nation Threats of violence and death Cursing and swearing Silly, stupid, ugly, etc.

Verbal and textual hostility in context

‘Here comes the nigger’ P ‘White trash’ P P ‘Fucking faggot, piece of shit’ P P P ‘Chinese bitch, I hope you   get SARS’ P P P P P ‘Brown rats, go back to your   own country’ P P P ‘Get out, Gypsy whore’ P P P ‘Fucking Paki cunt’ P P P ‘Fuck off, you lying Jewish   bastard’ P P P P ‘Fuck off, you black nigger,   you’re just a slave’ P P ‘I’m going to have you,   you fucking Paki’ P P P ‘Fucking lesbian, fucking   dyke’ P P ‘Die cunt’ P P P

Other

Profanity

Terrorisation

Expatriation

Criminalisation

Sexualisation

Demonisation

Pathologisation

Interpellation

Table 5.2  Examples of coding for themes of verbal-textual hostility.

P

The context Before discussing the linguistic properties of the verbal-textual hostility recorded in hate crimes reported to the MPS, it is important to understand the context of this violence. Hate crime regulation in the UK was initially developed in a response to a series of controversial hate crimes, particularly the murder of Stephen Lawrence in 1993. The Macpherson Inquiry into Stephen Lawrence’s death facilitated the introduction of hate crime offences (Hall et al. 2009), which created a ‘policy career’ (Jenness and Grattet 2001) of hate crime regulation. This career informs what is popularly understood about hate crime, and in turn informs the ways in which hate crime is policed through the lens of racist hate violence.5 The ‘template’ of racist hate violence is clearly illustrated in the frequency of each of the forms of hate violence captured in the 2003 MPS data (see Table 5.3). If taken at face value, the heightened frequency of racist hate violence gives the 105

Hate Crime

Table 5.3  Hate crime flag. Flag frequencies Number Racist Anti-semitic Homophobic Faith Total

14,721 299 1,398 280 16,698

Per cent (Cases)

Per cent (Responses)

90.5% 1.8% 8.6% 1.7% 102.7%

88.2% 1.8% 8.4% 1.7% 100.0%

impression that this form of hate crime is more prevalent than that of anti-semitic, faith-based or homophobic violence. However, it is important to remain mindful of the mediating factors in reporting crimes to the police. In particular, the Lawrence Inquiry clearly demonstrated to black and minority ethnic (BME) communities that the government was concerned about facilitating the reporting of racist hate violence. In creating a dual system of justice – with racist (and later, faithbased) hate crimes named under the Crime and Disorder Act 1998, and all other forms of hate violence covered only by sentencing legislation (such as Criminal Justice Act 2003) – the government and police may have contributed to a perception that other victims are not deserving of the heightened attention that comes with specific hate crime provisions. Further, some victims of hate crimes have – and continue to have – an acrimonious or estranged relationship with the police. This is especially the case for gay men, who have had their sexuality criminalised in the past, and who continue to face discrimination within the criminal justice system (GALOP 2008). It is also the case for those UK residents who have been identified as Muslim, and, as a consequence, become subject to a range of disciplinary practices relating to counter-terrorism measures in the UK (such as those introduced through the Anti-Terrorism, Crime and Security Act 2001, Prevention of Terrorism Act 2005 and Terrorism Act 2006). The criminalisation of identity in both of these cases could lead to a decreased willingness to report hate crimes to the police. It is therefore important not to assume that BME communities in the UK experience heightened levels of hate violence. Rather, it may demonstrate that these communities have developed a conditional trust in the police and their capacities to remedy the injustices raised by racist hate violence.6

106

Verbal and textual hostility in context

Hate crime location Public spaces (such as streets, parks and footpaths) are the primary sites of reported hate crimes, with 40–45 per cent of all incidents occurring within this environment. Despite this shared pattern, the template of racist violence informs the contours of the consolidated results. There is only a slight variation between the total database and the experiences reported by BME communities in relation to racist violence, with only a slight increase in the likelihood of racist violence occurring in a commercial or business environment, such as corner and takeaway shops (see Table 5.4). There is a much higher likelihood for faith-based and antisemitic hate violence to occur in and around public buildings. In the majority of these cases, the site is a religious institution such as a mosque, temple or religious community organisation. It is, however, surprising – given the tenacity of anti-semitic financial conspiracies – to find that there is a decreased likelihood of anti-semitic hate violence occurring within the business or commercial environment. The business environment constitutes 8.8 per cent of the reported Table 5.4  Hate crime location7 x hate crime flag.8 Location x flag Commercial/   business Public building Public space Residential Transport Other location Total

Racist Anti-semitic Homophobic Faith Total n = 14313 n = 284 n = 1337 n = 263 database n = 15784 2,433 17.0% 761 5.3% 6,417 44.8% 4,905 34.3% 428 3.0% 447 3.1% 14,313 100.0%

25 8.8% 36 12.7% 133 46.8% 95 33.5% 9 3.2% 5 1.8% 284 100.0%

159 11.9% 40 3.0% 619 46.3% 545 40.8% 24 1.8% 51 3.8% 1,337 100.0%

34 12.9% 53 20.2% 105 39.9% 80 30.4% 4 1.5% 10 3.8% 263 100.0%

2,601 16.5% 826 5.2% 7,082 44.9% 5,502 34.9% 455 2.9% 498 3.2% 15,784 100.0%

Note: p = .000

107

Hate Crime

anti-semitic violence, which is significantly lower than the overall rate of 16.5 per cent. Clearly, while some of the strongest antisemitic propaganda relates to the financial and business world, the perpetrators of this violence may not. Table 5.4 also highlights the differences faced by gay men and lesbians. There is an increased likelihood for homophobic violence to occur in and around the home of the victim. While this increased likelihood is marginal, it is nonetheless significant when considered in light of the social prohibition of public displays of same-sex relationships. With only the private space of the home left for gay men and lesbians to demonstrate their sexuality freely, an increased prevalence of the home as a site of violence creates the impression that there is no place safe from homophobic hate violence.

Hate crime offence The differences between forms of hate violence are also highlighted when the correlation turns from the hate crime flag to the offence initially recorded by MPS officer. While 65 per cent of all incidents involve violence against people (including for example, harassment, common assault, ABH/GBH), anti-semitic and faith-based hate Table 5.5  Hate crime offence9 x hate crime flag. Offence x flag crosstab Violence against   people Violence against   objects Verbal & textual   hostility Theft/burglary Other notifiable   offence Other offence Total Note: p = .000 108

Racist Anti-semitic Homophobic Faith Total n = 14,277 n = 291 n = 1349 n = 273 database n = 15,762 9,427 66.0% 2,137 15.0% 2,170 15.2% 400 2.8% 155 1.1% 217 1.5% 14,277 100.0%

151 51.9% 62 21.3% 67 23.0% 9 3.1% 5 1.7% 1 0.3% 291 100.0%

808 59.9% 158 11.7% 290 21.5% 69 5.1% 26 1.9% 27 2.0% 1,349 100.0%

139 50.9% 70 25.6% 46 16.8% 7 2.6% 6 2.2% 6 2.2% 273 100.0%

10,278 65.2% 2,337 14.8% 2,501 15.9% 475 3.0% 182 1.2% 247 1.6% 15,762 100.0%

Verbal and textual hostility in context

violence is significantly less likely to include these offences, and correspondingly, more likely to include violence against objects (such as dwellings, cars and other buildings; see Table 5.5). This matches the increased likelihood of these forms of hate violence occurring in and around public buildings; that is, attacks against religious institutions and organisations. Further, both anti-semitic and homophobic violence were marginally more likely to include only verbal-textual hostility (23 per cent and 21.5 per cent respectively, compared to an overall 15.9 per cent). Finally, a practice that appears to be largely related to homophobic violence is the use of theft/burglary/robbery particularly against gay men. The targeting for this type of violence is perhaps a result of the feminisation of gay men, a misperception that they are easy marks,10 and a perceived unwillingness on the part of victims to report to police because of homophobia or a desire to remain closeted.

Hate crime relationship(s) The final context characteristic to inform the role of verbal-textual hostility in hate violence is the relationship between perpetrators and victims. Early definitions of hate crime (such as Mason 1993; Cunneen et al. 1997) constructed this type of violence as a matter of ‘stranger danger’. However, recent research on police case files has clearly demonstrated that there is a division between the ‘strangerdanger’ model of motivated hate crime and the more domesticated nature of aggravated hate crime (see, for example, Mason 2005; Moran 2007; Iganski 2008). There is also ambiguity around whether ‘known’ perpetrators can be considered as such when they are only ‘known’ in passing (in the street, shared hallways or shops: Mason 2005; Iganski 2008). In spite of these caveats, in the MPS data, someone known to the victim perpetrated 35.1 per cent of all recorded incidents of hate violence. There are distinct differences in this experience; not only in terms of the proportion of cases with known perpetrators, but also the type of ‘known’ perpetrator. Homophobic violence is marginally more likely to be perpetrated by a known person (38.8 per cent), and anti-semitic violence is less likely to include a known perpetrator (25.1 per cent). In the case of the latter form of hate violence, this is perhaps a result of a majority of these incidents involving anonymous criminal damage. In the case of the former, however, this is perhaps the result of the higher likelihood of the violence being perpetrated by a family member. As with victims of some forms of interfaith hatred, gay men and lesbians 109

Hate Crime

at times face extreme violence at the hands of their family members, including parents and siblings (see Table 5.6). Gay men and lesbians are twice as likely to experience violence at the hands of family members, and victims of faith-based violence are 2.5 times more likely to experience violence at the hands of family. However, neighbours were less likely to be perpetrators in faithbased violence when compared with other forms of hate crimes (34.5 per cent and 56.9 per cent respectively). Perhaps this is due to the heightened frequency of this form of hate crime occurring in and around public buildings rather than residential spaces. This contextual analysis of hate crimes reported to the MPS offers an insight into the unique patterns of racist, anti-semitic, homophobic and faith-based violence. However, more importantly, it demonstrates the ground shared between those who experience hate violence. There are shared experiences despite the great differences in identity and identity-formation, and the legislative and social boundaries of both. Reported hate crime – no matter who is the subject of this violence – is predominantly perpetrated against the person, in and around their homes or local public spaces, and of those perpetrators known to the informant/victim, the violence is most likely to be carried Table 5.6  Hate crime relationship11 x hate crime flag. Relation x flag crosstab Acquaintance/   friend Business/work   colleague Family member Neighbour/ landlord/REA Known other Other relationship Total Note: p = .000 110

Racist Anti-semitic Homophobic Faith Total n = 5,128 n = 75 n = 543 n = 87 database n = 5,703 672 13.1% 376 7.3% 269 5.2% 2,958 57.7% 882 17.2% 59 1.2% 5,128 100.0%

12 16.0% 9 12.0% 5 6.7% 35 46.7% 14 18.7% 3 4.0% 75 100.0%

81 14.9% 35 6.4% 72 13.3% 287 52.9% 75 13.8% 6 1.1% 543 100.0%

16 18.4% 8 9.2% 14 16.1% 30 34.5% 18 20.7% 1 1.1% 87 100.0%

761 13.3% 420 7.4% 349 6.1% 3,243 56.9% 965 16.9% 67 1.2% 5,703 100.0%

Verbal and textual hostility in context

out by neighbours or property owners. If there is a habitus – or convention – of hate, surely this offers governments and victimised groups shared visions of everyday life without hate.

The text While there is a shared contextual pattern in hate crimes, the force and effects of this violence varies considerably. In this section, the variation in force and effects is highlighted through a Critical Discourse Analysis of the words used in hate crime to give life to the (sub)conscious intentions of perpetrators. Addressors may not clearly understand the locutionary force of their ‘words that wound’ (Matsuda 1993). However, the force of these speech acts passes from generation to generation, and with an accretion of meaning through social reproduction – including the reproduction of the knowledge of being hated – each generation learns what these words do and how they are best employed. In this way, while young people may not understand what it means to be gay, they are in no way mistaken about the force and effect this label has on the sense of self – and gender norms – of the addressee. Therefore, in addition to understanding the contextual factors of hate violence, it is vitally important to understand the words used by perpetrators before, during and after incidents to enhance the force of their violence.

Reporting and recording verbal-textual hostility Reporting officers indicated in the abridged narrative of MPS files that approximately 70 per cent of incidents included some form of verbal or textual hostility, and 16 per cent included a verbatim recording of the speech-text used in these incidents. The four forms of hate crime analysed in this chapter, however, do not equally share this prevalence. Importantly, incidents of homophobic and faith-based hate violence were less likely to include reported verbal or textual hostility (64 per cent and 57 per cent, respectively). These forms of hate violence were also less likely to include a verbatim recording of the verbal-textual hostility used before or during incidents (13.5 per cent and 11.4 per cent respectively). These results may represent a differential experience of verbal-textual hostility. That is, perpetrators may be less likely to use verbal or textual hostility in these forms of hate crimes. However, as this was not the case for faith-based hate crimes (see Table 5.5), this difference could be attributed to reporting 111

Hate Crime

officers not recording the details of hate speech-text as often in this form of hate violence. Differential patterns of reported and recorded verbal-textual hostility also emerge when the analysis turns to selected ‘signal’ crimes. Incidents that included physical violence against people (such as ABH, GBH and Common Assault) or violence against objects (such as Criminal Damage) were significantly less likely to contain reported and recorded verbal-textual hostility (see Table 5.7). This could be due to perpetrators of these types of incidents preferring not to use verbal or textual hostility. In the case of the latter, damage to buildings and objects rarely require verbal or textual hostility – the exception being graffiti. In the case of physical violence against the person, there is a clear drop in the prevalence of reported and recorded verbal-textual hostility between the lower offence of common assault (60 per cent and 15 per cent, reported and recorded, respectively) and the higher offences of aggravated and grievous bodily harm (ABH/GBH) (43 per cent and 9 per cent). As the physical violence increases, the need for verbal-textual hostility appears to decrease. Table 5.7  Reported and recorded verbal-textual hostility x selected hate crime offences. Verbal- Other Harass- Criminal Common ABH/ Total textual accepted ment damage assault GBH database hostility n = 2,290 n = 5,953 n = 2,298 n = 2,690 n = 1,302 n = 16,263 Reported Recorded

87.3% 19.3%

91.4% 19.7%

36.2% 9.1%

60.0% 14.6%

43.4% 9.4%

69.5% 16.1%

This, however, raises concerns about how victims/informants ‘know’ that this physical violence and criminal damage is hate motivated or aggravated. In cases without verbal-textual hostility, an ongoing pattern of violence between the perpetrator and victim, which included reported/recorded verbal-textual hostility at some time, would be necessary to understand these incidents as hate crimes. However, as Iganski (2008: 12) details, a majority of respondents to the British Crime Survey stated that they perceived crime as having a hate component purely on the basis of the victim’s race/ country of origin. The differential practice of reported and recorded verbal-textual hostility could also reflect reporting officers’ belief that proving the substantive offence of physical violence is more important than providing forensic evidence of the hate motivation or 112

Verbal and textual hostility in context

aggravation. This is highlighted in incidents that included offences such as criminal harassment and ‘other accepted crime’,12 which were significantly more likely to include reported and recorded verbaltextual hostility (91 per cent and 20 per cent, 87 per cent and 19 per cent, respectively). In these types of offences, the reporting officer is more likely to record the speech-text used in these incidents, as verbal-textual hostility constitutes the substantive crime. The number of recorded speech acts in each incident also illustrates differential practices in the use of verbal-textual hostility. Homophobic violence is significantly more likely than all other forms of hate violence to include a single speech act (58 per cent and 39 per cent, respectively). This single speech act consists predominantly of naming the addressee (see Table 5.8 below), and perhaps illustrates the force and effect of calling someone gay or lesbian. A single speech act is also more likely to occur at the hands of family members (64 per cent), in and around the family home (45 per cent). At the other end of the scale, incidents involving four or more speech acts were more likely to be perpetrated against BME communities – with twelve cases of racist violence including five speech acts in a single incident. These incidents with four or more speech acts were predominantly perpetrated by neighbours or property owners, and were more likely to be textual rather than verbal. The number of speech acts in each incident also decreases with an increase in the level of physical violence, with the majority of incidents involving four or more speech acts being assigned to ‘speech crimes’ such as ‘other accepted crime’ and criminal harassment (6 per cent and 8 per cent respectively, compared with 4 per cent for common assault). To get a better understanding of the force and effects of hate speechtext, it is also vital to understand how these characteristics correlate with the actual words used in hate violence. In this sense, while violence perpetrated by neighbours or family members, in and around the family home, using extended speech acts may lead to heightened harm, equally, a well-timed, well-placed single speech act can create crippling consequences. For example, in a case file managed by the Lesbian and Gay Anti-Violence Project (Sydney, Australia), occupants of a passing car called the victim ‘poofta’. In itself, this speech act meant little to the victim (who identified as bisexual); however, other people standing next to him took this as a statement of truth and pushed him under a passing bus.13 This example illustrates not only the inefficacy of some illocutionary speech acts on the addressee (the verdictive illocution of interpellation), but correspondingly, the consequential force and effects of the same single speech act on 113

Hate Crime

other members of the hate speech-text audience (the perlocutionary incitement generated out of the inefficacious illocution). It is therefore important to be cognisant of how speech acts bring about harm in the saying/writing, and as a consequence of the saying/writing.

Victim/informant identity and verbal-textual hostility As was shown in Table 5.1, central to the verbal-textual hostility recorded in the MPS incident files is the practice of naming the addressee – with approximately 78 per cent of incidents including the use of interpellation, and 21 per cent of incidents consisting of interpellation only. Further, as with Stokoe and Edwards’ (2007) research, apart from incidents involving interpellation only, the most common form of verbal-textual hostility experienced by all victims is profane naming. That is, the use of interpellation, commonly intensified by simple profanity (such as the use of the intensive, ‘fucking’). Profane naming constitutes 16 per cent of those incidents with recorded verbal-textual hostility. However, the prevalence of naming the addressee – or profane naming – varies considerably in terms of the form of hate violence. Homophobic violence was more likely to include only interpellation, and racist and homophobic violence employs interpellation more often in combination with other speech acts (see Table 5.8). As mentioned above, in the case of homophobic violence, the use of only interpellation may indicate that naming someone as gay or lesbian may be sufficient to give life to the addressor’s hatred or Table 5.8  Themes of verbal-textual hostility x hate crime flag. Profane naming Interpellation only Interpellation Profane naming

Racist n = 2,350

Anti-semitic n = 30

Homophobic n = 167

Faith n=8

19.8% 79.2% 16.0%

16.7% 56.7% 13.3%

38.9% 74.9% 12.0%

12.5% 37.5% 0.0%

19.5% 22.0% 4.9% 43.9%

28.6% 0.0% 14.3% 71.4%

Other themes of verbal-textual hostility14 Demonisation Sexualisation Expatriation Terrorisation Note: p = .000 114

21.3% 22.6% 26.2% 32.6%

4.8% 14.3% 9.5% 71.4%

Verbal and textual hostility in context

hostility. However, in the case of racist violence this may be due to addressors identifying the most obvious physical feature of addressees that mark them as different. This act contrasts with anti-semitic and faith-based violence, where difference may not be so obvious or naming may not be sufficiently hostile to instantiate the addressor’s hatred or hostility. While interpellation or profane naming can create a hostile environment, the dominance of this single form of verbaltextual hostility masks more substantive speech acts. Ironically, it is in the absence of the conventional form of hate speech-text (that of, profane naming) that we get a better understanding of the force and effects of verbal-textual hostility. As such, in secondary data analysis, the variables of interpellation and profanity were removed from the calculations. Beyond the conventional form of profane naming, the verbaltextual hostility used by perpetrators of hate crimes reported to the MPS draws upon four main themes: demonisation, sexualisation, expatriation and terrorisation.15 Each of these additional speech acts does more than name individuals and groups. These substantive speech acts reduce individuals to their most private corporeal experiences, incite fear of ‘evil incarnate’, and seek the exile, extermination or elimination of the ‘other’. As such, addressing marginalised individuals using these speech acts may create stronger illocutionary force and perlocutionary effects. However, the efficacy of these speech acts pivots on the addressor’s clear identification that specific individuals and groups are the object of the hostility. In this way, through convention, the names attributed to addressees become reductive versions of the stronger themes of hate speechtext, while simultaneously, the stronger themes are without context unless the addressee is named, or their identity can be read from their appearance, attire or cultural practices. There are clear differences in the prevalence of each of the additional themes. Most stark is the heightened use of the theme of terrorisation in anti-semitic and faith-based hate crimes. In the UK, the terrorisation of BME, and gay and lesbian communities was the impetus for increased regulation of hate crime.16 Yet, Jews and Muslims, in particular, are more likely to experience threats of violence and death in hate crimes. In the case of anti-semitic violence, this could be due to the efficacy of ‘calling down’ the Shoah as a promotion of things to come (such as ‘Hitler had the right idea with the final solution’, or the use of the swastika in the majority of criminal damage cases of anti-semitic violence). However, it is important to note that while terrorisation is more likely to occur in faith-based 115

Hate Crime

hate crimes, this speech act was the most prevalent practice – apart from profane naming – in all forms of hate crimes. The harm caused by terrorisation is intensified by the higher likelihood of this speech act being employed by family members, in and around the family home (69 per cent and 43 per cent respectively, compared to the 34 per cent overall frequency). While terrorisation appears to be a dominant practice in faith-based violence, racist violence is more likely to include calls for the exile of the addressee; that is, exile from the space shared by victim and perpetrator, whether that is the neighbourhood or the nation. Seeking the exile of difference from our shared space may appear to be less harmful than the extermination sought in terrorisation, yet it is only marginally so. Oliver Cromwell Cox (1970, cited in Fraser 1995: 77), in comparing anti-semitism with white supremacy, suggested that when the other is an abomination then the aims of the addressor are expulsion, forced conversion or extermination. That is, a continuum exists between expatriation and terrorisation.17 This continuum leads from an unheeded direction to ‘fuck off’, to a warning of violence unheeded, to physical violence.

Hate crime offences and verbal-textual hostility There are clear differences in the experience of hate speech-text when considered in light of the identity of victims/informants. In order to secure a better understanding of the harms generated out of these speech acts, it is also important to consider how these themes are employed in ‘signal’ hate crimes such as ‘other accepted crime’, harassment, criminal damage, common assault and ABH/GBH (see Table 5.9). Contrary to expectations, victims were more likely to be subject to profane naming in violent crimes, such as common assault and ABH/GBH (including 100 per cent of the GBH incidents). Further, there was significantly lower likelihood of name-calling in cases of criminal damage. However, cases of interpellation only were more likely to occur in ABH/GBH and criminal damage. In the former this may be due to the body – particularly the fists – doing the talking, and in the latter the tendency for criminal damage to relate to graffiti on private residences, which only names the subjectivity of the residents (such as ‘Jew’ spray-painted on the front door of a family home). With name-calling removed from the data set, clear patterns and markers emerge around particular offences. At the violent end of hate crimes (ABH/GBH), there is a spike in the practice of sexualising 116

Verbal and textual hostility in context

Table 5.9  Themes of verbal-textual hostility x selected hate crime offences. Profane naming Interpellation only Interpellation Profane naming

Other accepted n = 441 20.0% 74.1% 10.9%

Harass- Criminal Common ABH/ ment damage assault GBH n = 1,172 n = 209 n = 392 n = 122 21.0% 84.4% 16.7%

27.3% 64.1% 11.0%

21.2% 88.5% 25.8%

32.8% 92.6% 18.0%

13.2% 16.3% 20.9% 62.8%

26.4% 26.9% 21.6% 13.5%

18.3% 41.7% 18.3% 16.7%

Other themes of verbal-textual hostility18 Demonisation Sexualisation Expatriation Terrorisation

24.6% 16.4% 32.8% 28.5%

21.5% 26.2% 28.4% 27.9%

Note: p = .001

addressees. This is in large part the result of perpetrators using the word ‘cunt’ more often in these higher offences than in any other signal crime. Further, there is a significant drop in the practice of terrorisation in these higher offences, as there is no need for putting into words that which the perpetrator achieves through physical means. At the nexus of violent and speech crimes lie property crimes such as criminal damage. In criminal damage, the primary speech act is that of terrorisation. Directing their violence against objects and structures (rather than human bodies) may lead perpetrators to risk a mediated, perhaps even an opportunistic, threat. At the other end of the hate crime spectrum is ‘other accepted crime’. In these offences (which contain nothing more than verbaltextual hostility), the primary practice is expatriation. This heightened prevalence is perhaps the outcome of the extended exchanges necessary to constitute hatred without physical violence or criminal damage. A disturbing anomaly found in incidents of ‘other accepted crime’ was the presence of terrorisation. Terrorisation – or its incitement – is the only form of hate speech-text that constitutes a criminal (rather than civil) offence, in and of itself, whether committed as a hate crime or not. Approximately 180 incidents (1.2 per cent of total database) were assigned as ‘other notifiable offence’ – the MPS category for threats of violence. In a closer analysis of the lower offence of ‘other accepted crime’, an additional 88 incidents were found to include threats of violence. Either these incidents have been incorrectly assigned the 117

Hate Crime

lower offence, or the perceived threat or harm from these speech acts have been underestimated by the reporting officer, informant or victim. The largest problem with assigning the higher offence stems from the ‘reading’ of events by both the victim/informant and the reporting officer. If neither believes that the threat is capable of being given life, then it is less likely to be assigned the higher offence (rather than ‘other accepted crime’ or harassment). The ambiguity between ‘other accepted crime’ and ‘other notifiable offence’ highlights the important role that key tools of forensic linguistics could play in assisting front-line officers in judging the harm, force and effects of some speech acts.

Conclusion Contrary to the platitudes of our parents, names can and do hurt, and under the right circumstances can result in increased harm and distress. However, most people – no matter their personal, social or religious characteristics – face this situation at some time in their lives, especially at the hands of the schoolyard or workplace bully. What brings us the greatest offence or causes the greatest harm and distress will have as much to do with who we believe ourselves to be as it does with how others perceive us. In this sense, naming one’s subjectivity can be a positive act of nomination and an act of subordination, depending on who is naming, and who has the power to name. Subordinating the other through naming creates a hierarchy of subject positions. Equally, for those assigned a subordinate position, the performative act of naming is essential to participation in the polity, and consequentially, the opening up of opportunities for moving out of a subordinate position. Therefore, the force and effects of naming is always ‘out of our control’ and up for negotiation (Butler 1997). However, in cases of ongoing incidents of verbal harassment, the harm and possible effects increase, and must be managed much more closely by policing services. In single acts of verbal or textual hostility, however, it may be dangerous to censor or chill profane naming. It is therefore important that resources be created to ensure that these speech acts are met with a proportionate response from criminal justice agencies, including the police. Understanding what is said in hate crimes assists us in understanding when hate is a crime, and, necessarily, the responsibility of policing services. However, we need to be wary that in regulating harmful verbal-textual hostility we do not assist free speech absolutists who 118

Verbal and textual hostility in context

seek out unnecessarily chilled speech to use in rallying against any regulation, including incitement to hatred and terrorisation. A more critical use of forensic linguistics, as illustrated in this analysis, can assist us to develop sophisticated instruments with which to judge the possible harm, force and effects of hate speech-text. This is especially the case when assessing whether pure ‘speech’ crimes constitute the lower offence of ‘other accepted crime’ or the higher offence of ‘other notifiable offence’. Further, hate speech-text practices in cases of ABH/ GBH, and to a lesser extent common assault – such as the decreased number of speech acts and the ‘softer’ themes of hate speech-text employed by perpetrators – underscore the way in which verbaltextual hostility acts as a surrogate for physical violence in lower hate crime offences. The role of surrogate is also illustrated by the dominance of terrorisation in lower offences such as harassment and criminal damage, and the overwhelming dominance of this speech act in racism, anti-semitism, homophobia and faith-based violence. There are clear differences in the text and context of hate crimes when considered through the lens of victim-status. This differential practice is especially highlighted in the increased likelihood for homophobic hate crimes to be more violent (with a higher prevalence of ABH/GBH, murder and rape), and the higher likelihood of the use of terrorisation in faith-based violence. However, more binds victims of hate crime than separates them. The conventional form of hate crime in the UK in the pre-7/7 era consists of violence against the person, in public spaces in and around the family home, with neighbours using verbal-textual hostility, harassment and common assault to instantiate their hatred and intolerance. Their primary speech acts consist of profane naming and terrorisation. This conventional form of verbal-textual hostility in hate crimes translates into the Austinian illocutionary speech act of in terrorem (Iganski 2002: 28–9) – or more precisely, ‘I terrorise thee’. These are not mere words, and they are not easily shaken off. Beyond the discriminatory action represented in these illocutions, these speech acts make addressees more vulnerable to a series of perlocutionary psychosocial and biomedical responses that harm the individual and community long after the incident of hate crime. The capacity to respond effectively to these harms is not equally shared among victims; yet, this is not only due to how people identify themselves or are identified by perpetrators and policing services. It is also about the individual, social and institutional resources available to seek and find justice. Surely, shared experiences provide enough common ground for shared responses and resources, and a unitary system for regulating hate crime that does not privilege 119

Hate Crime

one victim of hate over another solely because of identity. Evaluating who is harmed in hate crimes must be a matter of text and context. Employing forensic linguistics to this task ensures that the matter of power – and the symbolic violence adhered to language use in hate crime – is not forgotten.

Notes 1 The data analysis provided in this chapter would not be possible without the support and approval of the Metropolitan Police Service. In particular, I would like to thank Professor Betsy Stanko for approving this research, Vicki Kielinger and Susan Patterson (Citizen Focus Directorate) for their support and advice, and, most importantly, the staff at the Violent Crime Directorate (especially, Detective Sergeant David MacNaghten) for their time and insights into hate crime policing practices. I would also like to thank Dr Neil Chakraborti, Christopher Fox and Rebecca Moran for their comments on early drafts of this chapter. 2 Austin’s illocutionary speech acts are epitomised by the first-person singular present indicative verb statements such as ‘I promise’, ‘I thee wed’, ‘I condemn thee’ or ‘I name this ship’. 3 See Asquith (2009) for an extended discussion of speech act theory and the harms generated out of illocutionary and perlocutionary speech acts. 4 The complete narrative of the incident would have been a more ideal source for a critical analysis of verbal-textual hostility; however, this field was unavailable due to the presence of private and personal information of perpetrators, victims, witnesses and informants. 5 It is important to note that this situation is not unique to the UK; the ‘policy career’ of hate crime legislation and policing practices is very similar across western democracies, with racist hate crime gaining the most attention of policy-makers. 6 This trust is conditional in large part because of the continued differential policing practices in relation to stop and search powers and the collection of DNA, and the failure of policing services to meet their Lawrence Inquiry commitments on the recruitment and retention of BME officers. In all three practices, policing services in the UK continue to provide an unequal service; under-representation on the latter, and over-representation on the former (EHRC 2009). 7 The six location variables in this table are user-defined recodes, which reduced 131 separate MPS locations into workable categories that highlight the role of hate speech-text. 8 Table totals in this, and Tables 5.5 and 5.6 relate to incidents, while row and column totals to locations, offences and relationships respectively. As a single incident can occur in multiple locations, with multiple offences and multiple relationships between suspect and perpetrator, 120

Verbal and textual hostility in context

9 10 11 12 13 14

15 16

17 18

row and column totals exceed the total incidents recorded in this table. In this respect, total incidents are provided as a comparative rather than summative figure. The six offence variables in this table are user-defined recodes, which reduced 29 separate MPS offences into workable categories that highlight the role of hate speech-text. Similar constructions of failed-masculinity have been mooted in relation to recent violence against Indian students in Melbourne, Australia. The six relationship variables in this table are user-defined recodes, which reduced 57 separate MPS relationships into workable categories that highlight the role of hate speech-text. This is the MPS CRIS code for verbal-textual hostility. Incidents assigned to this offence do not include any other criminal offence. This was a case managed by the author in her role as client advocate at the Lesbian and Gay Anti-Violence Project in 1994/95, and is part of the data set analysed in Asquith (2008). Calculations in the bottom half of this table are based on forms of verbaltextual hostility without the categories of interpellation and profanity. Further, only four themes were found to be significant in reported hate crime. Chi square tests found pathologisation and criminalisation to have low significance in relation to the Hate Crime Flag. See Asquith (2004; 2007; 2008) for an extended discussion of these themes. For example, the 25 BME men – in additional to Stephen Lawrence – who were murdered in racially motivated hate crimes between 1991 and 1999, and the nail bomb attacks of 1999 and the murders of Anthony Walker and Jody Dobrowski in 2005. Ironically, given these results, Cromwell Cox argued that racism did not seek the expulsion, conversion or extermination; rather, the goal of white supremacy was exploitation. Calculations in the bottom half of this table are based on forms of verbal-textual hostility without the categories of interpellation and profanity. Further, only four themes were found to be significant in reported hate crime. Chi square tests found pathologisation and criminalisation to have low significance in relation to selected Hate Crime Offences.

References Asquith, N. L. (2004) ‘In Terrorem: “With their Tanks and their Bombs, and their Bombs and their Guns, in your Head”’, Australian Journal of Sociology, 40(4): 400–16. Asquith, N. L. (2007) ‘Speech Act Theory, Maledictive Force and the Adjudication of Vilification in Australia’ in J. Ensor, I. Polak and P. Van 121

Hate Crime

Der Merwe (eds) New Talents 21C: Other Contact Zones. Perth: Network Books (pp. 179–88). Asquith, N. L. (2008) Text and Context of Malediction. Saarbrücken: VDM. Asquith, N. L. (2009) ‘The Harms of Verbal and Textual Hatred’, in P. Iganski (ed.) Hate Crime, Vol. 2: The Consequences of Hate Crime. Westport, CT: Praeger (pp. 161–74). Austin, J. L. (1980) How to Do Things with Words (2nd edn). London: Oxford University Press. Bourdieu, P. and Thompson, J. B. (1991) Language and Symbolic Power. Cambridge: Polity Press. Butler, J. (1997) Excitable Speech: A Politics of the Performative. New York: Routledge. Coulthard, M. and Johnson, A. (2007) An Introduction to Forensic Linguistics: Language in Evidence. London and New York: Routledge. Cunneen, C., Fraser, D. and Tomsen, S. (1997) ‘Introduction: Defining the Issues’, in C. Cunneen, D. Fraser and S. Tomsen (eds) Faces of Hate: Hate Crime in Australia. Sydney: Hawkins Press (pp. 1–14). EHRC (2009) Police and Racism: What has been Achieved 10 Years after the Stephen Lawrence Inquiry Report? London: Equality and Human Rights Commission. Fraser, N. (1995) ‘From Redistribution to Recognition? Dilemmas of Justice in a “Post-Socialist” Age’, New Left Review, 312: 68–93. GALOP (2008) Filling in the Blanks: LGBT Hate Crime in London. London: GALOP. Gelber, K. (2002) Speaking Back: The Free Speech versus Hate Speech Debate. Amsterdam and Philadelphia: John Benjamins Publishing Company. Hall, N., Grieve, J. and Savage, S. P. (2009) ‘Introduction: The Legacies of Lawrence’, in N. Hall, J. Grieve and S. P. Savage (eds) Policing and the Legacy of Lawrence. Cullompton: Willan Publishing (pp. 1–24). Heyman, S. J. (ed.) (1996) Hate Speech and the Constitution. New York and London: Garland Publishing. Iganski, P. (2002) ‘How Hate Hurts’, in G. Csepeli and A. Orkeny (eds) Gyulolet es Politika (Hate and Politics). Budapest: Friedrich Ebert Stiftung – Minoritas Alapitvany Kisebbsegkutato, Intezet (pp. 25–35). Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Iganski, P. (2009) ‘How Hate Crimes Hurt More: Evidence from the British Crime Survey’, in P. Iganski (ed.) Hate Crime, Vol. 2: The Consequences of Hate Crime. Westport, CT: Praeger (pp. 1–14). Jenness, V. and Grattet, R. (2001) Making Hate a Crime: From Social Movement to Law Enforcement. New York: Russell Sage Foundation. Langton, R. (1993) ‘Speech Acts and Unspeakable Acts’, Philosophy and Public Affairs, 22(4): 293–330. MacKinnon, C. A. (1993) Only Words. Massachusetts: Harvard University Press. Mason, G. (1993) ‘Violence against Lesbians and Gay Men’, Violence Today, 2: 1–10. 122

Verbal and textual hostility in context

Mason, G. (2005) ‘Hate Crime and the Image of the Stranger’, British Journal of Criminology, 45(6): 837–59. Matsuda, M. (1993) ‘Public Response to Racist Speech: Considering the Victim’s Story’, in M. Matsuda, C. R. Lawrence, R. Delgado and K. Williams Crenshaw (eds) Words that Wound: Critical Race Theory, Assaultive Speech and the First Amendment. Boulder, CO: Westview Press (pp. 17–52). Matsuda, M., Lawrence, C. R., Delgado, R. and Williams Crenshaw, K. (eds) (1993) Words that Wound: Critical Race Theory, Assaultive Speech and the first Amendment. Boulder, CO: Westview Press. Mill, J. S. (1972 [1859]) Utilitarianism, On Liberty and Considerations on Representative Government. London: Dent. Moran, L. J. (2007) ‘“Invisible Minorities”: Challenging Community and Neighbourhood Models of Policing’, Criminology and Criminal Justice, 7(4): 417–41. Stokoe, E. and Edwards, D. (2007) ‘“Black This, Black That”: Racial Insults and Reported Speech in Neighbour Complaints and Police Interrogations’, Discourse and Society, 18(3): 337–72. Ten, C. L. (1980) Mill on Liberty. New York: Oxford University Press. Van Dijk, T. A. (1987) Communicating Racism: Ethnic Prejudice in Thought and Talk. London: Sage. Wodak, R. (2001) ‘What CDA is About – A Summary of its History, Important Concepts and its Development’, in R. Wodak and M. Meyer (eds) Methods of Critical Discourse Analysis. Londons: Sage (pp. 1–13).

123

Chapter 6

Hate crime offenders Jack McDevitt, Jack Levin, Jim Nolan and Susan Bennett

Introduction Since the 1980s researchers have sought to understand the phenomenon of hate or bias crimes. These are described as crimes ‘motivated entirely or in part by a person’s difference’ (Levin and McDevitt 1993). Research to date has produced useful information regarding the characteristics of and psychological impacts on victims targeted by hate; at both an individual and group level (Perry 2003; Dunbar and Creveocoer 2005; Levin 2007). Additionally, much research has focused on the existence of hate crime as a legal concept. Some have argued that hate crimes are so damaging as to require a specific statutory response (Levin and McDevitt 1993; Lawrence 1999), while others have argued that hate crimes are a ‘social construction’ that require no additional legal protection (Jacobs and Potter 1998). While this debate continues, an area where research has lagged involves understanding hate crime offenders, particularly what motivates these offenders to engage in acts of hate-related violence. Research on offenders has been limited by the paucity of information on those individuals who commit hate crimes. In the United States, the research into offenders has focused largely on offender characteristics gleaned from official records, most commonly the National Hate Crime Statistics published annually by the Federal Bureau of Investigation (FBI). These data as well as data from individual police agencies, including New York, Boston and Los Angeles, have been used to examine characteristics of offenders who have been identified by the police. These data are limited in two 124

Hate crime offenders

major ways: first, they only include information on the 11 variables required by the FBI as part of their Hate Crime Reporting System, and second, they contain little information about the motivation of offenders. As noted by Shively and Mulford in their review of the literature, ‘A large body of research exists on prejudice and bias but it does not address why prejudice prompts people to commit a hate crime’ (Shively and Mulford 2007: 4). In addition to these large data sets, a number of additional studies have focused on a smaller number of offender records and interviews (Franklin 2000; Dunbar and Creveocoer 2005). These studies have contributed considerably to our understanding of the motivation of hate crime offenders but are limited in their small sample sizes, thereby reducing our ability to determine how representative these cases are of the general population of hate crime offenders. It is important to continue to learn more about hate crime offenders in order to organise that knowledge to help victims’ service providers and law enforcement who seek to assist victims and identify and arrest offenders. This chapter has two main goals: to add to our knowledge about offenders through the use of the FBI’s National Incident-Based Report System (NIBRS), a relatively underutilised data source, and to reconsider what is known about offender motivation in light of new data and events.

Using NIBRS to understand hate crime offenders The new reporting system, NIBRS, was added by the FBI to their ongoing Uniform Crime Reporting programme in the 1980s. This data system allows participating police agencies to submit additional information for each crime reported. The system collects details for each criminal incident, an improvement over the existing system which simply counts the number of incidents. Regarding hate crimes, NIBRS added the question ‘whether this crime was motivated by bias and if so what kind?’ in 1997, which allows all crimes to be evaluated on whether hate was a motivation (Strom 2001). The new system has increased the amount of information collected on each case from 11 data elements in the original system to 57 in the NIBRS system, producing far more information about each bias incident. A particular shortcoming of NIBRS is that it is both costly and timeconsuming to implement, and therefore a number of law enforcement agencies have yet to adopt the system. To date, only a small portion 125

Hate Crime

of law enforcement agencies participate in NIBRS, and a number of the largest US cities are excluded. While this is a limitation in the utility of the data, the system still provides the ability to obtain additional information from thousands of law enforcement agencies which can significantly enhance our knowledge about hate crimes and hate crime offenders. The first review of hate crime data from NIBRS was published in 2001 by the Bureau of Justice Statistics (Strom 2001). This report reviewed hate crimes reported via the NIBRS system and identified 2,976 incidents reported between 1997 and 1999. From this data, we find support for information established in nearly all prior literature, that the vast majority of hate crimes with known offenders are committed by white males (62 per cent), followed by black males (20 per cent), followed by white females (11 per cent). Not surprisingly, in most incidents where females were identified as the offender/s females also tended to be the victims. When we look at each individual group targeted in hate incidents – for example, blacks, Latinos, or Jews – more than 80 per cent of the offenders were white. Some recent work has raised questions about the prior finding that the majority of hate crime offenders are strangers to their victims. Research conducted on hate crimes reported in London raised the question that while the victim may categorise the offender as a stranger while reporting the incident to the police, the offender may not be completely unknown to the victim (Mason 2005). The literature has previously found that most hate crime offenders are strangers to their victims (Levin and McDevitt 1993). The NIBRS data similarly shows that many attacks are committed by strangers. However, of the violent incidents reported in NIBRS, 38 per cent were committed by acquaintances and 26 per cent were committed by strangers. According to NIBRS, in only 7 per cent of incidents is the offender found to be a friend or relative of the victim. This strengthens the assertion that hate crime offenders maintain tenuous relationships with their victims. In many cases the victim and offender live in the same neighbourhood, go to the same school, or work in the same organisation. The NIBRS data seem to support this interpretation that hate crime offenders may know their victim, for example, as the black tenant that just moved into the neighbourhood or the Latino employee who was just hired. Cases may be rare but certainly still occur where hate crime offenders have no idea about the identity of their victim. The NIBRS data seem to offer additional support for this contention: nearly two-thirds of the offenders in hate crimes who were arrested 126

Hate crime offenders

lived in the town where the crime occurred. This counters the notion in some circles that groups of hate crime offenders ride from town to town looking for unsuspecting victims. It appears that most hate crime offenders stay in their local area and look to victimise individuals with whom they have some familiarity. From a public policy point of view it is useful for police who are investigating hate crimes to know that the offender is likely to be from the same town where the crime occurred. Hate crime offenders are most likely to target a single victim. NIBRS indicates that offenders in violent hate crimes targeted a single victim in 83 per cent of cases. This speaks to the perceived vulnerability of the victim by the offender. Offenders, either alone or in groups, seem to look for victims they perceive as unlikely to fight back and one way they do this is to look for victims who are alone. This focus on such victims may be part of the reason why arrests are so difficult in hate crimes, since no one other than the victim is available to help identify the offenders. About three-quarters of the hate crimes reported in NIBRS involved a single offender. While the overall hate incident may include multiple participants, including encouraging friends or bystanders who do not participate in the violence but do participate in the incident, it appears that many acts involve a sole offender attacking a single victim. As shown below, although many hate crimes involve a single offender, hate-motivated assaults are more likely to involve multiple offenders than non-hate-motivated assaults. Although we know that in many of these crimes the offender received encouragement from his friends, the exact role of that encouragement in facilitating the attack is unaddressed in most of the research literature. Clearly, much more needs to be understood about the dynamics of hate-motivated assaults. Another area where NIBRS data support prior research is that offenders in hate crimes seldom use weapons. In NIBRS only 18 per cent of violent hate crime offenders used weapons and only 4 per cent used a firearm. The lack of weapon use may reflect the fact that many hate crimes arise spontaneously given a particular situation. Offenders may encounter a potential victim while they are out, and the situation develops rather than groups planning their attack far in advance. As we have written previously (Levin and McDevitt 2002), if it is true that most hate incidents are rather spontaneous, it may be more difficult for police to intervene prior to the attack. However, since the offenders seem to have less of a long-term commitment to hate-motivated violence, hate incidents should be a little easier to deter from a public policy perspective. 127

Hate Crime

For the present analysis, we used more recent NIBRS data to compare assaults that were hate-motivated to those assaults where hate was not reported as a motive. An interesting pattern emerges when hate-motivated assaults are compared with all assaults reported in NIBRS. First, hate-motivated assaults are significantly more likely to involve juveniles than other assaults reported to NIBRS. This is very useful information since there has been some debate in the literature about how often juveniles are involved in hate crimes. This analysis finds that 25 per cent of hate-motivated assaults were committed by juveniles while only 16 per cent of all other assaults were committed by juveniles. It is also interesting to note that hate-motivated assaults by juveniles are most likely to be committed after school, between 2 and 4 p.m., while hate-motivated assaults by adults are most often committed between midnight and 2:00am. This seems to indicate that while, overall, most hate-motivated assaults are committed by adults, juveniles are more likely to be involved in hate-motivated assaults than other assaults. Of the assaults reported to NIBRS, hate-motivated assaults were more likely to result in physical injury. Approximately 9 per cent of hate-motivated assaults resulted in injury, while only 5 per cent of non-hate-motivated assaults resulted in injury. This is interesting in that so few of the assaults reported to NIBRS resulted in injury overall, while hate-motivated assaults were almost twice as likely to produce injury. Similarly, hate-motivated assaults were more likely to be categorised by police as aggravated assaults, with 34 per cent of the hate-motivated assaults being categorised as aggravated while only 21 per cent of the non-hate-motivated assaults were categorised as aggravated. Another question in the literature involves the number of offenders typically involved in hate offences; researchers have tried to determine whether hate crimes are more likely to involve multiple offenders. As indicated above, most hate-motivated incidents were reported to involve only one offender; however, compared to non-hate-motivated assaults, hate-motivated assaults were almost three times more likely to involve multiple offenders (15 per cent as against 6 per cent). It is interesting to note how few assaults of either kind involved multiple offenders in the NIBRS data.

Characteristics by target of attack Table 6.1 illustrates the offender characteristics by type of victim targeted. While no clear pattern develops from this analysis, a few 128

Hate crime offenders

Table 6.1  Offender characteristics by type of victim targeted Race Religion Ethnicity Sexual orientation Juvenile offender Juvenile victim Serious injury Aggravated (not simple   assault) Offender arrested Offender used knife or   blunt instrument Victim was stranger Victim was acquaintance   or otherwise known Multiple offenders on   one victim Multiple offenders involved Multiple juvenile offenders Juvenile or young adult   (up to 25 years)

24% 31% 9%

32% 39% 7%

35% 38% 6%

22% 20% 10%

35% 38%

35% 49%

34% 36%

30% 31%

14% 26%

6% 16%

13% 24%

11% 16%

20%

39%

36%

39%

17% 27% 11%

10% 26% 16%

15% 32% 17%

11% 19% 8%

56%

63%

63%

62%

interesting differences can be seen. Juvenile offenders are more likely to be involved with hate crimes motivated by the victim’s religion or ethnicity than their race or sexual orientation. Interestingly, we see little difference in the level of injury across various categories of victims. This may be due to the fact that we are looking at all assaults as reported in NIBRS. Prior research has indicated that anti-semitic hate crimes are more likely to be property crimes, while hate crimes based upon a person’s sexual orientation are more likely to be assaults (FBI 2004). Given this prior research, it interesting to note that hate-motivated assaults targeting victims due to their sexual orientation are no more likely to involve injury or to be categorised as aggravated assaults. It may be that hate crime offenders are more likely to assault certain groups of victims but the level of injury resulting from the assault is not very different. Finally we can see that across all victim groups, most offenders in hate crimes are juveniles or young adults. Nearly two-thirds of all hate-motivated assaults, regardless of the type of victim targeted, are committed by youths under 25 years of age.

129

Hate Crime

Criminal justice intervention Similar to many other crimes, hate crime offenders are rarely arrested. According to the 1997–1999 NIBRS data, only about 20 per cent of the hate crime incidents resulted in arrest (Strom 2001). While the fact that eight out of ten hate crime incidents do not result in an arrest may be distressing enough, the pattern gets worse when the dynamics of the arrest are taken into account. Almost 40 per cent of the arrests occur at the scene of the incident; a person calls the police, they arrive, and the offender is still in the area, often still involved in the attack. When we remove these arrests from the data, we realise that if an arrest is not made at the scene, the likelihood that a hate crime offender will be arrested drops to only 17 per cent. While there are many reasons for this low arrest rate, including the fact that many offenders are strangers or acquaintances, this does indicate that, at present, the risk of criminal sanction for hate crime offenders is very low. While the NIBRS data indicates that hate-motivated assaults were more likely to involve physical injury, multiple offenders, and to be committed by juveniles, it is interesting to note that when comparing hate-motivated assaults to assaults where hate was not a motivation, hate-motivated assaults were less likely to result in arrest. Thirtyseven per cent of the hate-motivated assaults resulted in arrest while 47 per cent of the non-hate-motivated assaults resulted in arrest. This may be due to the larger proportion of hate-motivated assaults that involve strangers.

Summary of NIBRS offender analysis The analysis of NIBRS data supports much of the previous literature about hate crime offenders. These offenders are most likely to be white males, who may be acquaintances or neighbours of the victim but are not friends with the victim. Hate crime offenders are most likely to attack a victim who is alone and seldom use weapons. Additionally, when we look at all assaults reported in NIBRS and compare hate-motivated assaults to non-hate-motivated assaults, we find that hate-motivated assaults are more likely to involve juveniles and are twice as likely to result in injury to the victim. Finally, hate crimes in general are unlikely to result in any arrest of the offenders, and hate-motivated assaults are less likely to result in arrest than non-hate-motivated assaults.

130

Hate crime offenders

Motivation of hate crime offenders While the kind of descriptive information discussed above is helpful to police, prosecutors, victims’ service providers and public policymakers, in 1993 and again in 2002 a typology of hate crime offenders was proposed to help these groups better understand and react to hate crimes that occur in their jurisdiction (Levin and McDevitt 1993; McDevitt et al. 2002). Recently, the typology has received some comment and critique from scholars, replicating the original research with alternative data sets. This portion of the chapter reviews this typology, addresses some of the major critiques to date, and offers some reflection on how the original typology continues to evolve. In 1993 Levin and McDevitt put forward a typology of hate crime offenders in Hate Crimes: The Rising Tide of Bigotry and Bloodshed. This typology, based on analysis of data from the Community Disorders Unit of the Boston Police Department and from interviews with police officers who have investigated these crimes, as well as a limited number of interviews with hate crime victims and offenders, led us to develop and put forth a three-category offender typology: offenders who act for the thrill of it, those who perceive themselves as defending their turf, and those offenders who are on a mission to ‘rid the world of evil’. In 2002, McDevitt, Levin and Bennett tested this typology with additional data from the Boston Police Department and found that the typology could be expanded by adding an additional category, of retaliatory violence, thus making the typology a fourcategory typology. In discussions both during this period and subsequently, law enforcement officials expressed a lack of understanding regarding how to differentiate hate-motivated crimes from other crimes (McDevitt et al. 2004). Officers often stated their difficulty in identifying the motivation in potential bias offences and that both police officers and prosecutors had difficulty convincing juries and judges that these crimes were different from non-hate-motivated crimes. It was specifically to address this ambiguity in identifying and prosecuting hate-motivated crimes that we hoped the typology might offer some assistance. We had hoped that the typology might group together some similar types of hate-motivated incidents, reflect general patterns that we saw from a review of the descriptive data on incidents to date, and more importantly reflect the reality of police officers, prosecutors and others who deal with offenders on a regular basis. This was the primary goal of the original and revised typologies. 131

Hate Crime

The hate crime offender typology has received widespread acceptance since it was originally proposed in 1993. The typology has been embraced by the FBI and has been included in the hate crime training curriculum that is offered as part of the FBI training in Quantico, VA. Additionally, the typology was a key component of the national hate crime training curriculum developed for the United States Department of Justice – Responding to Hate Crime (McLaughlin et al. 2000). Additionally the typology has been included in the hate crime training provided by the American Prosecutors Research Institute, A Local Prosecutor’s Guide for Responding to Hate Crimes (2000). The widespread acceptance of the typology has also been noted in the academic literature (Franklin 2000; and Fisher and Salfati 2009; Phillips 2009)

Hate crime offender typology As noted above, the original offender typology included three categories. We added a fourth category to reflect a more refined version of the typology that we felt more accurately reflected the motivation we were seeing in the Boston data and the experiences of police and prosecutorial officials. As discussed previously, hate crime offences differ significantly in their defining characteristics from other crimes not motivated by hatred. For example, hate-motivated assaults are more likely to result in physical injury when compared to other assaults. Furthermore, hate crimes are generally perpetrated by strangers on lone victims, in acts that can often appear to be random, senseless, or irrational. Victims are selected based on their group affiliation, not personal attributes. Finally, hate-motivated assaults are perpetrated by multiple offenders more often than in the case of non-hate assaults (Levin and McDevitt 2002). Hate crime perpetrators may be somewhat distinct in comparison to other criminals. For example, in a study of undergraduate per­ ceptions of hate crime victims and perpetrators, participants viewed perpetrators of hate-motivated crime as being more culpable than perpetrators of non-hate crime (Rayburn et al. 2003). Further, in a survey of law enforcement, the majority of hate crime investigators indicated that they viewed hate-motivated incidents as more serious than similar crimes not motivated by hatred (McDevitt et al. 2000). Typologies have been established to characterise the motivation of hate crime perpetrators. Building on an earlier typology 132

Hate crime offenders

conceptualisation (Levin and McDevitt 1993), McDevitt et al. (2002) proposed a method of characterizing four unique hate crime perpetrator motivations: thrill, defensive, retaliatory, and mission. In a review of 169 hate crime cases investigated by the Boston Police Department, thrill was found to be the most frequent motivation, distinguishing well over half of all hate incidents. Thrill crimes are characterised by a desire for excitement and may be typified by an immature desire for power. Thrill offences often involve groups of teenage or young adult offenders, with offences often occurring on the victim’s ‘turf’. These crimes are committed for ‘excitement’ or ‘thrills’ by youth who are bored and looking for something to do. In comparison to other perpetrators, there is often less of a commitment to hatred in ‘thrill’ offenders (McDevitt et al. 2002). In many of these cases, young men looking for excitement or thrills decide to attack someone who they perceive as different. Based on messages they have received from our culture, these young criminals do not think anyone will care if they attack a member of one of these target groups. In many cases, these young people incorrectly believe others in the community will support their actions, even if they would not act out themselves. Defensive hate crimes represent the next most common type, according to the 2002 analysis. These crimes are committed when perpetrators target victims under the perception that the perpetrator is protecting valuable resources or defending threats to his or her neighbourhood, workplace or school. As with thrill offences, defensive crimes are often perpetrated by teenagers or young adults, but in contrast, most defensive hate crimes occur in the offender’s neighbourhood, not the victim’s. It is the offender’s ‘turf’ that is being protected. A common example of defensive hate crimes involves harassment suffered by a black family who moves into an all-white neighbourhood (McDevitt et al. 2002). Many hate crime victims tell the police that their attackers told them that they did not belong in ‘their’ neighbourhood, in the course of the attack. The third most common hate crime motivation is that of retaliation. retaliatory offences occur in reaction to a perceived hate crime. Here, it is not important whether in fact an assault occurred, only that the offender believes it took place. Retaliatory offenders are likely to act individually, often seeking out a victim to target in the victim’s own territory. These crimes differ from thrill or defensive hate crimes because the offenders see themselves as reacting to a specific incident involving the victim or a member of the victim’s group. These retaliatory crimes can be followed by other retaliatory crimes by the 133

Hate Crime

victim’s group, a cycle of attack and counter-attack that can involve many innocent victims. In certain cities this cycle of retaliatory hatemotivated violence has resulted in large increases in the number of hate crimes over a relatively short period of time. Finally, the least common, but potentially most critical motivation for hate crime offenders is that of mission. Mission offenders perceive themselves to be crusaders whose lives are completely committed to hatred and bigotry. Mission offenders may operate in groups (in affiliation with an organised hate group) or alone (such as in the example of Timothy McVeigh1) (McDevitt et al. 2002). Table 6.2 illustrates some of the characteristics of these crimes by differing offender motivation. Finally, as indicated in Table 6.2, it may be the case that offenders involved in hate crimes differ in their potential to be deterred from future hate-motivated violence. It has been suggested that ‘thrill’ offenders may have the lowest commitment to bias of all hatemotivated offenders. In this case, if offenders believe that the police and other public officials were likely to react negatively to their acts of violence, it may be that many of these offenders could be deterred. Table 6.2  Characteristics of hate crimes by motivation of the offender. Attack Thrill characteristics

Defensive

Retaliatory

Mission

Number of Group offenders

Group/single offender

Single offender

Single offender

Age of offender(s)

Teens–young Teens–young adults adults

Teens-young Young adults– adults adults

Location Victim’s turf Offender’s turf Victim’s turf Victim’s or offender’s turf Weapon

Hands, feet, rocks

Hands, feet, rocks

Hands, feet, Bats, guns rocks

Victim– Acquaintance Previous acts Often no offender of intimidation history history

None

Commitment Little to bias Deterrence

134

Likely

Moderate

Moderate

Full

Unlikely

Unlikely

Most unlikely

Hate crime offenders

On the other hand, when dealing with ‘mission’ offenders, who have developed a world-view based upon bias and hatred, it may be much more difficult to deter these individuals with the threat of criminal justice sanctions.

Levels of culpability among hate crime offenders Overall, typologies categorise perpetrator motivation and can assist law enforcement and other agencies to better detect hate-motivated crime when it occurs. These typologies also provide guidance for more empirically based research addressing the etiology of hate crimes and intricacies that may exist among diverse perpetrators. Ultimately, a better understanding of motivation for hate crimes will lead to stronger policy and prevention strategies. As a supplement to the original offender typology in 2002 we also suggested that offenders may play various roles in those hate crimes that involve multiple offenders. We suggested that offenders may be seen as leaders, fellow travellers, unwilling participants and heroes. The leader in a hate crime group is the individual who is most likely to initiate the idea of the crime, often suggesting that the group ‘go out looking for some action’. We had previously suggested that the leader would be the person who initiated the violence when the group encounters a potential victim. As a result of the information on the number of attackers described above, the leader in an incident may be the only person involved in an attack, or the person primarily involved in the attack. In this way, while others may play additional roles in a hate crime incident, from the victim’s perspective there may have been only one offender. We suggested that a second role for offenders in a hate crime was that of a fellow traveller. This role, originally suggested by Watts (1998), would include those offenders who may have never thought or had the nerve to engage in violence if they were alone but would participate in an attack if it was initiated by the leader. These individuals, who are primarily young males, may be deterred from involvement if they view the stakes to be high, such as a high likelihood of investigation and arrest by the local police. The third group that we identified was the unwilling participant. These individuals may be involved with a group of young men who talk about or initiate an act of violence. They may disagree with the sentiments being espoused by the leader or other members, but due to their fear of being criticised by the group they keep silent. They may hold back and never participate in any violence, but 135

Hate Crime

the fact that they do not express their reservations can be taken as their acquiescence by others in the group. Sometimes young women involved in group attacks, such as the girlfriends of others involved, may take this role in an incident because of not wanting to anger or disappoint their boyfriends. The final role for those involved in hate crimes was that of the hero. This individual actively attempts to stop his or her friends from committing a violent act. These individuals can act on their beliefs in a variety of ways ranging from alerting authorities to the potential for a hate crime to directly confronting the leader or others in the group about the folly in their thinking and plans. We used the term hero to illustrate the considerable personal risk these individuals incur as they actively attempt to stop particular hate crimes before they occur. In Table 6.3 we attempt to apply these culpability roles to the four categories of motivation identified above. We see the culpability roles as being very similar across the categories of our typology, but we note that the role of hero may change across the categories of the crime. In thrill hate crimes the hero may try to dissuade his or her friends from engaging in actions that would end up in violence. In defensive hate crimes the hero may also alert the victim since he or she may know the identity of the potential victim. In retaliatory hate crimes the hero may try to dissuade fellow offenders from taking on acts of retaliation and may also alert officials, for example school officials or police, to the potential for future violence. In mission hate crimes the hero may have little likelihood of reversing the plans of other members of the group. They may, at considerable personal risk, report the existence of the group to the police and work with the police to develop information that might be used to prosecute the members of the group.

Expansion of the offender culpability roles Based on the analysis presented above, particularly the information detailing the victim’s perspective, there is often only one offender identified involved in the hate crime. We are suggesting expanding our culpability categories. While many victims may only identify a single assailant, much previous research indicates that hate crimes are not often committed by an individual acting entirely on his own. As suggested above, the individual acting as the leader in an incident may be the one committing the final assault, while he is 136

Does not actively participate in crime but does not attempt to stop crime or warn victim

Actively or hesitantly participates in crime

Does not actively participate in crime but does not attempt to stop crime or help victim

Fellow traveller

Unwilling participant

Colour

Key = Highly amenable to change = Moderately amenable to change = Change highly unlikely

Hero Attempts to stop crime Warns victim

Actively participates in planning or crime

Encourages others

Defensive offenders

Leader

Level of Thrill Culpability

Attempts to stop crime and warns victim

Does not actively participate in crime but does not attempt to stop crime or warn victim

Actively or hesitantly participates in crime

Retaliatory

Table 6.3  Culpability of individuals involved in hate crimes by motivation of the offender.

Reports group behaviour to police

Does not actively participate in crime but cannot withdraw from the group

Actively participates in hate group and crime

Mission

Hate crime offenders

137

Hate Crime

frequently encouraged, empowered, or aided by others. The role of these others has been identified by Jack Levin as ‘sympathizers and spectators’, those who stand by and are silent or at worst provide verbal encouragement to the eventual offender. We originally proposed the culpability roles to describe the potential roles of each individual involved in multiple offender hate crimes. We are now expanding these roles to include single offender incidents. In these incidents we believe that the culpability role continues to apply but in different fashions. The leader is still the individual who commits the eventual crime, but others involved assume slightly different roles. The fellow travellers in single offender crimes are those who encourage and help plan the incident. In these cases, the fellow travellers may be the students in the high school or college cafeteria who encourage the leader ‘to go ahead and teach one of them a lesson’ or ‘show them whose country this is’. The unwilling participant is the youth who disagrees with the racial or ethnic slurs being used but fails to speak up, again for fear of losing friends or having their friends think less of them. Here again, we have heroes: those who stand up to the leader and the fellow travellers, telling them that what they are saying or planning is wrong, and who alert authorities when they will not be dissuaded. We believe these expanded roles can help school officials and other adults dealing with young men to see how they may intervene in situations that may develop into hate crime over time. By cultivating heroes in a variety of groups of young men and women, these officials may be able to prevent, through their intervention, hate crimes that would otherwise have been committed.

Critiques of the typology In recent years a few academic scholars have tested the original typology using different data sets (Fisher and Salfati 2009; Phillips 2009). While some of the original categories were replicated with these new data sources, these authors have not found the same patterns of offenders as those indicated in the original analysis of Boston Police Department data. We would like to address a couple of the points made by these authors before we suggest an evolution of the original typology. These authors have suggested that the categories of the typology are not mutually exclusive and we agree (a point we made in our 2002 article). It is the case that it is difficult to find a single motivation 138

Hate crime offenders

when offenders perpetrate hate-motivated crimes. It may be the case, for example, that a group of potential offenders are out for excitement one night and looking to attack someone they perceive as different. While driving around they encounter the son of a black family who recently moved into the neighbourhood, and attack him saying that he was not wanted in ‘their’ neighbourhood. This crime could be categorised as either a defensive hate crime (protecting their turf) or a thrill hate crime (out for excitement). In our original analysis we would have included this incident as a thrill hate crime since initially that was the group’s original motivation, but a reasonable argument can be made that it was a defensive crime. We suggest that this situation of multiple motivations may be rather common in hate crimes. We offer two reactions to this level of ambiguity, first involving data limitations and second the reality of the complexity of human motivations. First, much of the existing research utilises data from official police sources to try to infer motivation. This data is notoriously inadequate in this regard. It has been suggested that crime scene variables might be able to identify motivation. We think that if researchers are utilising police files for data, they must use information from the investigation not simply from the crime scene to have any chance of identifying offender motivation. Often at the crime scene, the victims are unsure about the motivation, and it is only through a competent investigation that the real motivation for the crime can be discerned. In their official records police generally attempt to document two ‘facts’: one, that the person they arrested was in fact the person who perpetrated the crime; and two, that in hate-motivated crimes that bias played some role in the offender’s motivation. As seen in the research by McDevitt et al. (2000), police most often utilise racial, ethnic or anti-semitic slurs as their indication of bias motivation. Most police reports do not provide sufficient detailed information that would allow researchers to make the kind of subtle distinctions necessary to properly categorise the example provided above. This research team was very fortunate in the late 1980s and early 1990s to be working with a highly unusual set of police data. At the time, the Boston Police Department’s Community Disorders Unit was considered a national model for hate crime investigative units. The unit was led by a true pioneer of the hate crime investigations, Sgt Detective Billy Johnston. In a period when many still questioned the existence of hate crimes, Sgt Detective Johnston was known for having some of the most well-documented case files in the entire 139

Hate Crime

Boston Police Department. Sgt Detective Johnson required that detailed case notes be provided any time a detective questioned anyone in connection with a hate crime case. So in each case there could be as many as a dozen or more memos describing each of the interviews conducted on the case. All these files were made available to the research team. In addition, Sgt Detective Johnston allowed the members of the research team access to the detective involved with each case. In this way the research staff could further enhance the information in the case files with the recollections of the investigators involved. This set of data was highly unique and, we expect, almost unheard of in many of today’s record management systems and detective case management systems. A second and more important reason why categories of the typology might not be mutually exclusive is that the offenders themselves may be functioning from multiple motivations. As indicated in the example provided above, offenders themselves may change their motivation during the course of a single incident – from looking for excitement to defending their turf, for example. Offenders, particularly those in groups, may have multiple motivations and may reflect different commitment to bias across the group. We never expected that the categories of the typology would offer mutually exclusive categories of offender motivation, since after trying to code these cases we realised how often offenders may be operating from multiple motivations. Our hope was that we could utilise the information available to construct a primary motivation for the criminal event. Even though our typology may not be able to cleanly differentiate one hate crime offender from another, knowing the primary motivation for the crime has been and, we believe, will continue to be useful to those public officials charged with responding to hate crimes. Knowing the primary offender motivation remains helpful to investigators attempting to narrow or focus their investigation. If the victim is new to the neighbourhood and there are indicators that the crime is defensive in nature – for example, harassing phone calls prior to the incident – then the police would be advised to focus their investigation on residents of the immediate neighbourhood. If, on the other hand, the crime appears to be motivated by the youth looking for thrills or excitement, they may reach out to the school resource officer in the local high school. Since it has been noted previously that the offenders in thrill crimes often brag about their crime to other students (Levin and McDevitt 2002), police could look for heroes who want to help the victim and prevent further harassment as a source of information. We do not suggest that police 140

Hate crime offenders

focus their investigations only in these locations, but that this seems like a reasonable approach for their initial investigative focus. Based on our typology, the police would not necessarily be able to rule out a particular offender characteristic (such as the location of the crime scene or their age), but they could prioritise a number of such characteristics based on their likelihood of occurrence. Similarly, prosecutors who are trying to convict offenders who do not look like the stereotype of a racist skinhead, for example, can use the categories from the typology to help paint a picture to a jury of the kind of offender that commits this kind of crime. Judges and juries do not see many hate crimes in court and are influenced by the most sensational hate crimes played out in the media. By describing the motivation and culpability of the offenders involved, the prosecutor may be able to paint a more realistic picture of the crime for those judicial decision-makers.

Evolution of the original typology Upon reflection on the original typology, we have some thoughts about how it may be evolving. All suggestions of ways to characterise human behaviour are limited in various ways and need to continually evolve and adapt to new data and new insights. The hate crime offender typology is no exception. We offer here a few suggestions for how the hate crime offender typology may be evolving to meet and reflect new data and new cultural situations.

Categories of offenders Based on developing research, by both the authors of this chapter and others, we do not see the need to add or remove any of the four existing categories of the typology. We still believe that thrill, defensive, retaliatory and mission seem to capture the motivations of the majority of hate crime offenders. One important caveat identified earlier in this chapter is that we do not think that these categories of offenders are mutually exclusive. For the reasons outlined above we believe that offenders may act out of multiple motivations, but we suggest that the typology we put forward is a way of identifying and categorising the primary motivation. The typology remains focused on helping law enforcement, and service providers conceptualise the motivations and levels of culpability of hate crime offenders. 141

Hate Crime

Cases in each category may have shifted. Since the original typology was developed, based upon data in the early 1990s when the crime rate was so much higher, we believe that the four motivations are similar; however, it is an open research question as to whether the proportion of incidents that are in each category remain the same. For example, based on crime data from the 1990s, we found that the majority of incidents were thrill crimes, but since this was a time when there were more juveniles committing crimes, we may not observe the same patterns today. In the aftermath of the attacks of September 11, 2001 it may be that more defensive or retaliatory hate crimes are being committed and fewer thrill hate crimes. Also, as recently noted by the Southern Poverty Law Center, since 9/11, with high unemployment rates, and the election of an African-American President, there may have been an increase in mission hate crimes committed by members of organised hate groups. It may also be the case that future research will be able to identify pathways where hate crime offenders move from property-based hate crimes to violent hate crimes. For example, it is important to know why one group of thrill offenders spray-paints racial or anti-semitic slurs on a building and another group of thrill offenders physically beat their victim. Would these differing levels of violence be due to different characteristics of the incident (numbers of offenders, for example) or different experiences of the offenders (such as level of bias)? Police could utilise the typology to predict future acts of violence. The categories of the typology should help law enforcement pre­ pare for and anticipate hate crimes that may result from events or demographic shifts at the neighbourhood level. Some of the best police agencies already do this. Police agencies may anticipate hate-motivated violence when there are demographic shifts within neighbourhoods. Also, in the aftermath of a hate crime police should anticipate retaliatory violence by youth from the victimised group. Additionally, local police should monitor any local organised hate group for actions that they might take to incite violence from other similarly inclined individuals. Additionally, when hate crime offenders are arrested prosecutors can utilise the categories of the offender typology as well as the culpability roles to help judges and juries understand the complex motivation in hate crimes. The application of the offender motivation categories should be expanded to include roles that others play in the preparation and execution of a hate-motivated attack. As we consider the roles that

142

Hate crime offenders

others play in encouraging or discouraging hate-motivated crimes, we think that broadening the offender role categories may be helpful to law enforcement and others involved in deterring and reacting to hate crimes. We originally conceptualised the role of leader, fellow traveller, unwilling participant, and hero as they might be applied to participants in a particular crime. We now think it might be helpful to expand this conceptualisation to include those individuals who encourage the violence before the attack occurs, those who give support to the offenders after the attack and those who stand up to the offenders before or after the crime. The first group has been labelled as ‘sympathisers’ and ‘spectators’ in previous work (Levin 2007). In this way school officials, for example, may find a more direct link between some of the rude and biased behaviour of youth in their school and subsequent violence committed by students or their friends.

Conclusion Less is known about the characteristics and motivations of hate crime offenders than the characteristics of the events and the impact on victims. This chapter has reviewed an under-utilised data source (the National Incident-Based Reporting System) to offer some additional insight into the characteristics of hate crime offenders. Additionally, the authors have reconsidered the original hate crime typology put forward in 1993 and 2002 and have suggested ways that the original hate crime offender typology may be evolving to fit changing historical and cultural events. Identifying and reacting to hatemotivated violence remains the most crucial act in combating further acts of hate committed in a community. This chapter has sought to add to our understanding of those who perpetrate hate crimes in an attempt to improve our efforts to detect and deter these acts of violence and bigotry.

Notes 1 Timothy McVeigh was convicted of bombing the Murrah Federal Building in Oklahoma City in 1995 killing 168 people. His anti-government attack is considered the deadliest act of domestic terrorism in US history.

143

Hate Crime

References APRI (2000) A Local Prosecutor’s Guide for Responding to Hate Crimes. Alexandria, VA: American Prosecutors Research Institute. Dunbar, E. and Creveocoer, D. (2005) ‘Assessment of Hate Crime Offenders: The Role of Bias Intent in Examining Violence Risk’, Journal of Forensic Psychology Practice, 5(1): 3–29. FBI (2004) Uniform Crime Reports: Hate Crime Statistics 2004. Washington, DC: Federal Bureau of Investigation. Fisher, C. and Salfati, G. (2009) ‘Behavior of Motivation: Typologies of Hate Motivated Offenders’, in B. Perry and R. Blazak (2009) Hate Crimes, Vol. 4: Hate Crime Offenders. Westport, CT: Praeger. Franklin, K. (2000) ‘Antigay Behaviors among Young Adults: Prevalence, Patterns, and Motivators in a Non-Criminal Population’, Journal of Interpersonal Violence, 15(4): 339–262. Jacobs, J. B. and Potter, K. A. (1998) Hate Crime: Criminal Law and Identity Politics. New York: Oxford University Press. Lawrence, F. (1999) Punishing Hate: Bias Crimes Under American Law. Cambridge, MA: Harvard University Press. Levin, J. (2007) The Violence of Hate. Boston, MA: Allyn and Bacon. Levin, J. and McDevitt, J. (1993) Hate Crimes: The Rising Tide of Bigotry and Bloodshed. New York: Plenum. Levin, J. and McDevitt, J. (2002) Hate Crimes: America’s War on Those Who are Different. New York: Plenum. Mason, G. (2005) ‘Hate Crime and the Image of the Stranger’, British Journal of Criminology, 45(6): 837–59. McDevitt, J., Levin, J. and Bennett, S. (2002) ‘Hate Crime Offenders: An Expanded Typology’, Journal of Social Issues, 58(2): 303–17. McDevitt, J., Balboni, J., Bennett, S., Weiss, J., Orchowsky, S. and Walbolt, L. (2000) Improving the Quality and Accuracy of Bias Crime Statistics Nationally. Final Report to Bureau of Justice Statistics. Washington, DC: US Department of Justice. McDevitt, J., Cronin, S., Balboni, J., Farrell, A., Nolan, J. and Weiss, J. (2004) Bridging the Information Disconnect in National Bias Crime Reporting. Final Report to Bureau of Justice Statistics. Washington, DC: US Department of Justice. McLaughlin, K. A., Malloy, S. M., Brilliant, K. J. and Lang, C. (2000) Responding to Hate Crime: A Multidisciplinary Curriculum for Law Enforcement and Victim Assistance Professionals. Newton, MA: National Center for Hate Crime Prevention, Education Development Center, Inc. Perry, B. (2003) ‘Defenders of the Faith: Hate Groups and Ideologies of Power’ in B. Perry (ed.) Hate and Bias Crime: A Reader. London: Routledge (pp. 301–18). Phillips, N. (2009) ‘The Prosecution of Hate Crimes: The Limitations of the Hate Crime Typology’, Journal of Interpersonal Violence, 24: 883–905. 144

Hate crime offenders

Rayburn, N., Earleywine, M. and Davison, G. (2003) ‘Base Rates of Hate Crime Victimization Among College Students’, Journal of Interpersonal Violence, 18(10): 1209–21. Shively, M. and Mulford, C. (2007) ‘Hate Crime in America: The Debate Continues’, NIJ Journal, 257. Strom, K. (2001) Hate Crimes Reported in NIBRS, 1997–99. Washington, DC: Special Report Bureau of Justice Statistics. Watts, M. (ed.) (2008) Cross-Cultural Perspectives on Youth and Violence. New York: JAI Press.

145

Part Two

Developing Responses to Hate Crime

Chapter 7

Law enforcement and hate crime: theoretical perspectives on the complexities of policing ‘hatred’ Nathan Hall

Introduction Over the past thirty or so years the enactment of hate crime legislation in the United States has represented one of the most lively and significant trends in criminal law-making at both the federal and state level (Levin 2002; Streissguth 2003; Hall 2005). England and Wales have followed suit with various legislative provisions against crimes motivated or aggravated by racial or religious prejudice, sexual orientation and disablism, and the incitement of racial and religious hatred. More recently, other countries have implemented, or are in the process of implementing, various forms of hate crime legislation. The development of legal responses to the hate and discrimination ‘problem’ is evident from the records of the Office for Democratic Institutions and Human Rights (ODIHR), a part of the Organization for Security and Co-operation in Europe (OCSE), who collate data on the existence and development of anti-discrimination and hate crime laws in member states (see www.legislationline.org for the latest legislative developments). At the time of writing, ODIHR’s records show the existence or development (albeit in different ways and to greatly varying extents) of ‘hate crime’ laws in 55 different countries. Jacobs and Potter (1998) suggest that hate crime laws generally fall into one or more of four categories. First, there are those that enhance sentences for hate-motivated offences; second, those that redefine existing criminal behaviours as a ‘new’ crime or as an aggravated form of an existing crime; third, those that relate specifically to civil rights issues; and finally, those that concern themselves solely 149

Hate Crime

with matters of reporting and data collection. One or more of these categories are generally reflected in legislative provisions adopted by countries enacting ‘hate crime’ laws. The rise of hate crime as a contemporary social issue has been driven by a number of different factors. In many countries, particularly western democracies, a combination of the extent and nature of hate crimes, their seemingly increasing upward trend, coupled with increased public tolerance to issues of diversity and sensitivity to prejudice, and the influence of identity politics (the political processes by which certain groups might come to be recognised as disadvantaged), has forced ‘hate’ crimes onto the statute books. Although their creation has been, and remains, a contentious issue, in essence the rationale behind hate crime laws are fairly commonsensical. Proponents argue that specific legislation represents an official recognition of an apparent emerging and increasing threat to society, and signifies the importance attached by government (at least in theory) to combating this threat. Where provisions exist, the potential for the increased punishment of the offender signifies an appreciation of the apparent disproportionate harm that hate crime can have on the victim and wider communities, and provides a deterrent to potential offenders by clearly stating that hate crimes will not be tolerated and that firm action will be taken against those that perpetrate them. Hate crime laws also promote social cohesion by officially declaring that the victimisation of ‘different’ groups is not acceptable, and the very existence of law arguably provides a symbolic message designed to force social change in relation to prejudiced attitudes and behaviours. Given the expansion of legislative provisions in many countries designed to combat the apparent growing problem of hate crime, this chapter considers theoretical issues relating to the law and law enforcement in relation to hate crimes. By drawing predominantly on the UK and the US experience, it is argued that because hate crimes are complex events, they pose significant challenges to lawmakers, the police and prosecutors. The chapter suggests that there are many issues both internal and external to those individuals and agencies responsible for law enforcement, and over which they have varying degrees of control, that inevitably impact upon their ability to enforce the law and to respond effectively and to provide a service appropriate to the needs of victims and wider communities.

150

Law enforcement and hate crime

The limits of effective legal action Before examining the enforcement of law in this area, it is necessary to consider, albeit briefly, some of the sociological issues concerning the purpose of law and its relationship to social change. Grossman and Grossman (1971) suggest that social change can be viewed in three stages. First, social change may be incremental, in which patterns of individual behaviour may be altered. Second, it may be comprehensive, in which group norms or patterns of relations between groups may be altered. Third, social change may be revolutionary, in which society’s values may be altered. As such, much social research on law has largely concentrated on two related issues: the ability of law to influence social change and, conversely, the ability of social change to influence law. Within this literature there has been a particular emphasis on the former, with the latter often viewed (erroneously) as ‘too obvious to require discussion’ (Cotterrell 1992: 49). The powerful effect of social change on shaping hate crime legislation is evident from the influence of identity politics (see Jacobs and Potter 1998) Opinion within the literature concerning the ability of law to influence social change is, however, divided. Some argue that law serves important educative and symbolic functions (Lester and Bindman 1972) that promote social integration and unity (Arnold 1935) and can influence people’s beliefs as well as their behaviours (Berger 1952). Conversely, others suggest that the ability of the law in this regard is grossly exaggerated (Erlich 1936; Kahn-Freund 1969) and that the direct use of law in promoting social change is problematic. Others have argued that law may indirectly influence social change by shaping social institutions (including the institutional framework and policy priorities and the legal duties of such institutions) that have a direct influence on the rate and character of social change (Dror 1959). In sum, Cotterrell (1992: 61) suggests that it is best to view legal strategies ‘as part of a long-term process of negotiation of attitudes and perceptions of interests in which political and legal action constitute only one element in a complex network of influences on social change’. It is also important for the context of this chapter to consider some issues relating to the wider limitations of legal effectiveness. Cotterrell (1992) suggests that modern studies of the limits of effective legal action can be traced to the work of Pound (1917) who laid down a number of principles that, despite being the best part of a century old, remain pertinent to a consideration of hate crime legislation and

151

Hate Crime

its effective enforcement. First, Pound (1917) suggested that law can only deal with the outside and not the inside of people. In other words, for a host of reasons including problems relating to proof, Pound argues that law cannot attempt to control attitudes or beliefs, only observable behaviour. As will become evident in this chapter, this point has served as the crux of the debate concerning the legitimacy and efficacy of hate crime laws. A second relevant point relates to the fact that law necessarily requires an external agency (in this case the police) to put its machinery in motion that itself (in most cases) requires invocation from the public. In other words, in this context, hate crime laws do not enforce themselves. Rather they are enforced by the police who in turn frequently require the input of the public (often as victims or witnesses) to start and pursue the process. This in turn raises some important prerequisites for the effectiveness of law, most notably that it must be, and must be seen to be, in the interests of those upon whom the law depends for either its invocation or its enforcement to set the process in motion (Cotterrell 1992). For hate crime laws to be effective, victims must have enough confidence in the law (and its enforcement agencies) to achieve the ends for which it is intended to make it worthwhile reporting crimes committed against them. In addition, law enforcement agencies, and the individuals working within, must have both the ability and the desire to enforce the law. Another important point raised by Pound (1917) is the notion that while there are interests and demands that it might be desirable for the law to recognise, the reality of such demands means that by their very nature they cannot be safeguarded by law. In this regard Pound refers specifically to issues of clarity concerning legal precepts and the limitations of law which arise from the difficulty of ascertaining the facts on which law is to operate. In other words, to be effectively enforced by state officials in state agencies, law requires a high degree of clarity, particularly concerning issues of proof (Cotterrell 1992). As will become clear in due course, clarity of proof in hate crime cases where evidence is frequently problematic is another issue of concern. Finally, Pound (1917) also notes that in some areas of social life the sanctions of state law appear useless and can disrupt rather than repair social relations. More recent writers in this field have drawn on this point by suggesting that the law may simply serve to bureaucratise social and moral relations and in so doing create uncertainty, hostility and distrust (Cotterrell 1992). Such issues are particularly acute if the purpose of law is held to be the creation of social change, but 152

Law enforcement and hate crime

arguably less so if the opposite is true. Either way, each of Pound’s points hold particular relevance for effectively legislating against hate crimes. Furthermore, as Pound (1917) suggested, law does not enforce itself and therefore the points he makes extend to law enforcement agencies. Such concern necessarily brings us to the role of the police upon which, as the gatekeepers of the criminal justice system, so much subsequently depends.

The problem of motive The vast majority of hate-motivated incidents are breaches of existing criminal law. In this sense, hate crimes are in essence like any other crime. However, to accept such a view is to be ignorant of the nature of hate crime (see Perry 2001; Hall 2005), and of the potentially devastating impact these crimes can have on the victim and the wider community. As Holdaway (1996: 45) states, ‘the question of how far a [hate] motive changes the nature of an assault or act of criminal damage brings to the fore wider issues about the social context within which people from minority … groups live and the appropriate response of legislators and policy.’ As noted in the introduction to this chapter, the response of lawmakers in many countries, but most notably in England and Wales and the US, to growing concerns about hate crime has been to introduce new hate-based offences to the statute books. Consequently, the expression of hate motivation by an offender when committing existing offences attracts stiffer sentences from the courts, and in doing so places hate crimes in a unique position in terms of the law. Prior to such legislation the determination of criminal liability for all offences had essentially concentrated upon an act and upon whether the defendant intentionally, recklessly or knowingly committed it. At this point it is important to be aware of the difference between motive and intent. While the motive gives rise to the intent, it was the latter and not the former that made an act criminal. Thus, suspects were only convicted for intentionally committing an act and not for having a motive. In the past, therefore, the law has punished acts and not motivations, and as such the issue of motive has been entirely distinct from intent or purpose. However, the inclusion of specific hate offences in criminal law changes this situation. Not only are new crimes created, but the most significant aspect of these offences is the motivation behind their commission. Notwithstanding issues concerning the parameters 153

Hate Crime

of legislation determined by case law (see Streissguth 2003; Hall 2005), the criminalisation of hate motivation arguably represents a shift in the focus of the criminal law from deed to thought. As a result, the courts are now required to establish the offender’s motives and therefore ‘what is really being punished is a criminal’s thoughts, however objectionable they may be. The actions – incitement, vandalism, assault, murder – are already against the law’ (Haberman 1999: 2). This significant change represents, at least in principle, the importance attached by policy-makers to tackling hate crime. This is important because, as Holdaway suggests, ‘both victims and offenders learn from the messages contained in the attack and from the nature and quality of legal and policy responses by institutions’ (1996: 47). The impact that this has on policing is crucial because in order for legal and policy responses to achieve their aims, the police are required to provide evidence of both motive and intent. Without proof of motive, a hate crime is simply an ordinary crime. This necessarily raises serious questions concerning the ability of the police to accurately identify and subsequently provide evidence of the motivation behind an offence – a task that they are not required to do for any other crime (Bell 2002). Although some cases are clear-cut, the successful prosecution of most hate crimes is difficult because of the need to prove beyond reasonable doubt that the offender was motivated by prejudice, and also because of the difficulty in proving that prejudice was the sole (or at least a substantial) causal factor in the commission of the offence. These two issues, Jacobs and Potter (1998) suggest, often make hate crime cases more difficult to prosecute than crimes committed without the hate element. In England and Wales the Crown Prosecution Service (CPS 2003) also recognise the inherent difficulty in proving bias motivation (see also Burney and Rose 2002). For the police, the problem of accurately identifying motive is problematic. Existing literature on perpetrators demonstrates that a whole host of motivations might be present in the commission of a hate offence. The literature further suggests that no two hate crimes are ever the same in terms of the motivation behind them (Craig 2002). As Green et al. (2001: 27) suggest, ‘It might take the better part of a lifetime to read the prodigious research literature on prejudice … yet scarcely any of this research examines directly and systematically the question of why prejudice erupts into violence.’ This raises an interesting question: if decades of research into prejudice and hate has failed to conclusively identify the motivations 154

Law enforcement and hate crime

behind these crimes, can we reasonably expect a police officer to do so to the extent that motivation is proved beyond reasonable doubt? Assisting officers in making these decisions is therefore crucial. However, neither the law in the US nor in England and Wales provides any guidance concerning what hate motivation might look like in practice; officers simply have to work it out for themselves, albeit often with assistance from prosecutorial bodies.

The police as decision-makers As the ‘gatekeepers’ of the criminal justice system, the decisions made by police officers, and in particular those of the lower ranks, are crucial in determining what, and how much of what, ultimately comes to the attention of the rest of the justice system. Given the amount of discretion inevitably afforded to officers of lower ranks, their decisions concerning both whether and how to enforce the law in individual cases (decisions that for the most part remain unchecked by others in the justice system) become crucial (Lipsky 1980; Bell 2002). While the decision-making of officers is relevant to all types of incident in terms of determining how much further official attention that incident receives, it is acutely important in hate crime cases. This is because, as noted above, officers not only have to make decisions concerning the crime, but decisions also have to be made concerning the motive behind the crime. It is the decisions made concerning the latter that ultimately determine whether a crime becomes a hate crime. In short, a crime is not officially a hate crime until the police decide that it is. This is certainly the case in the US, and also remains the case in England and Wales, despite definitional changes that allow anyone to apply the hate label for reporting and recording purposes. Ultimately at some point a police officer has to make decisions concerning motivation (is this a crime or a hate crime?) and then decide whether to charge a suspect with an ordinary offence or a hate-motivated or aggravated offence. In so doing it is the police who officially apply the hate label, although in most cases in England and Wales this now occurs further down the policing process than was the case prior to the post-Lawrence definitional changes, and with guidance from the CPS. The nature and workings of police discretion and police decision-making are therefore crucial to the success of legal and policy responses to hate crime. In relation to hate crimes, Cronin et al. (2007) suggest that the decision-making of patrol officers, whose decisions concerning 155

Hate Crime

potential motivations for an offence to a large extent determine whether a hate crime is ever officially recognised as such, is affected by issues of ambiguity (where multiple motivations might be evident), uncertainty (where only limited information about an incident might be available), and infrequency (where hate crimes are so infrequent that officers may never gain experience in responding to them). Any of these issues, they argue, can affect the accuracy of hate crime classifications made by the police, as of course can the attitudes, beliefs and practices of individual investigating officers (Franklin 2002). In terms of shaping police decision-making, Grimshaw and Jefferson (1987) have highlighted the existence of a ‘hierarchy of police relevance’ that officers use, often subconsciously, which determines their response to any given incident. At the top of the hierarchy are what Bowling (1999) terms ‘good crimes’. These are clear criminal offences with innocent, reliable and credible victims, perpetrators who are ‘real’ criminals, and ones that offer a clear opportunity of detection and arrest and a good result in terms of securing a conviction. Further down the hierarchy are ‘rubbish’ crimes, in which the ‘quality’ of the victim and perpetrator may be poor, where there is a much reduced chance of detection and arrest, or where there is an increased likelihood that the victim will withdraw the allegation at a later date. At the bottom end of the hierarchy are what Bowling terms ‘disputes’ or ‘disturbances’, which are frequently perceived to be legally ambiguous and of limited police relevance. In his interviews with police officers Bowling (1999) found that with the exception of certain crimes (such as robbery, assault and theft) where the legal relevance to the police is clear, racist offences tended to be viewed as being at the lower end of the hierarchy of police relevance. Bowling found that the perpetrators of racist hate crimes were largely perceived by the police as being ‘yobs’ or ‘hooligans’, and that racist incidents were regarded as acts of ‘yobishness’. In this sense both the offence and the offender often fall into the category of ‘rubbish’, and receive less attention from the police than ‘real’ crimes. Similarly, Bowling’s research highlighted the extent to which, from a police perspective, racist incidents routinely fall short of the criteria required to classify them as ‘good crimes’ and are regarded as ‘disputes’ or ‘disturbances’. In other words, in all but the most serious crimes, racist incidents tend to appear at the lower end of the hierarchy of police relevance, and as such receive less police attention than crimes where the legal duty to respond is unambiguous. Crucially, Bowling’s research 156

Law enforcement and hate crime

illustrated that this placing of racist incidents on the hierarchy, most notably by rank-and-file officers, remained consistent despite changes in force policy and force prioritisation of racist offences. In assuming that the majority of hate offences are simply minor incidents, the disproportionate impact that these can have on victims is often not appreciated and is overlooked by the police. The result of this clash of perceptions is that victims may feel that their victimisation is not taken seriously by the police. The extent of police discretion and the rudimentary decisionmaking process, predominantly by rank-and-file officers, of course has important implications for the number of hate crimes that are ultimately recorded and investigated. The hierarchy of relevance indicates a historical propensity for racist incidents not to be recorded, or treated appropriately, except where the offences are particularly serious. In turn this has helped to create a lack of trust in the police among minority groups that has subsequently led to a disinclination to report offences committed against them (Macpherson 1999). This, of course, is crucial because if hate crimes are not reported to the police, then for the most part they cannot be responded to by them. Other influences on police decision-making in this area relate to the nature of both individuals and organisations. There is a plethora of historical evidence to suggest that the police in both England and Wales and the US have, and indeed still do, exercise their discretion disproportionately and unfavourably with regard to minority groups. Traditional explanations for this discrimination include the individual (Scarman 1981) and institutional (Macpherson 1999) prejudice of police officers and police organisations, and the occupational culture of the police, research into which has consistently highlighted issues of police prejudice particularly along race, gender and sexuality lines (Skolnick 1966; Reiner 1992; Chan 1997; Waddington 1999). The importance of these biases is noted by Lipsky (1980: 85) who suggests that: Routines and simplifications are subject to workers’ occupational and personal biases, including the prejudices that blatantly and subtly permeate the society. These biases expressed in streetlevel work may be expected to be manifested in proportion to the freedom workers have in defining their work life and the slack in effective controls to suppress those biases. Since streetlevel bureaucrats have wide discretion about clients, are usually free from direct observation by supervisors or the general public, 157

Hate Crime

and are not much affected by client preferences, their routines and simplifications deserve considerable scrutiny. With regard to the police, almost without exception, police services in England and Wales and the US have been, and remain, white, maledominated organisations with attitudes that largely reflect those held by mainstream society; after all, historically police officers have been, and remain, predominantly drawn from the majority population. Given the historical and contemporary societal prejudice held by the mainstream towards minority groups (Hall 2009) it should therefore not be too surprising that minority groups’ experience of and trust in policing is qualitatively different to that experienced by the majority group, and this is reflected in the views historically articulated by minority groups (Bradley 1998; Macpherson 1999). Furthermore, there is evidence from both England and Wales and the US that police officers are often reluctant to enforce hate crime legislation (Bell 2002; John 2003; Hall 2005, 2009; AVP 2007). While this might be a product of individual, institutional or cultural discrimination, there are a number of other important issues that are specific to hate crime. First, it is important to remember that the enforcement of hate crime laws effectively reverses long-documented stereotypical police perceptions about minority groups, and in particular minority ethnic groups. Research has highlighted the historically grounded stereotype held by British police officers that black males are disproportionately involved in crime as perpetrators (Gordon 1983), and American studies of racial profiling and excessive use of police force suggest similarly (Bell 2002; Johnson 2003). Indeed, the origins of some American police departments have been traced to a goal of controlling the ‘problematic’ black population both during slavery and after emancipation (Bolton and Feagin 2004). In theory at least, however, the concept of hate crime necessarily reverses this stereotype in the sense that minorities are more likely to be victims than perpetrators, a situation that does not sit comfortably with the findings from studies of police occupational culture. Given the extent of discretion afforded to police officers, this has potentially significant implications for both the amount and quality of service afforded to hate crime victims by the police. Second, police officers may be reluctant to make decisions about hate motivation because, in effect, they may feel that they simply don’t have to, or that their job would be easier if the crime didn’t have the hate label applied to it. Hate crimes are already criminal offences, and without the hate element are, of course, categories of 158

Law enforcement and hate crime

offences that officers are more used to dealing with. The decision not to label a crime as a hate crime may therefore enable officers to bypass many of the aforementioned problems associated with identifying motivation, while still being able to get a ‘result’ for the underlying offence. Furthermore, officers may decide against applying the hate label simply because they may fail to understand or recognise the qualitative differences between a crime and a hate crime in terms if the potential impact on victims (Hall 2009). A third issue that may affect police decision-making relates to the politicisation of hate crime. In the current climate in England and Wales and in the US in particular, hate crimes are potentially subject to disproportionate and/or excessive public scrutiny. For example, a hate-motivated assault is likely to generate greater public and/ or media interest than a comparable regular assault. Even where incidents of hate crime do not attract attention in this way, given the impact that they can have on the wider community, the police may still find themselves under added pressure from that community, or from advocacy groups, to label a crime as hate motivated and to act accordingly. Conversely, there may also be conflicting pressures not to label too many incidents as motivated by hate for fear of portraying an image of intolerance and disharmony in their jurisdiction. As Bell (2002: 4) points out with regard to the policing of hate crime in the US: The ‘politics of hate’ influences nearly every stage of the process, including what gets reported, the help the [police] receive to investigate incidents, whether witnesses come forward, whether the community and City Hall pressure the [police] to investigate or not to investigate, and finally, the action the district attorney will take on a case. If an incident makes it through the investigation process without being dropped or classified as unfounded, it must then make it over a different set of hurdles when prosecutors and the courts come into play.

The police as policy-makers The ability of police officers to make key decisions concerning individual cases means that not only are they decision-makers but they are also, in effect, policy-makers. Lipsky (1980) states that the policy-making roles of ‘street-level bureaucrats’ such as the police are built upon two inter-related facets of their positions, namely relatively 159

Hate Crime

high degrees of discretion coupled with relative autonomy from organisational authority. He argues that ‘although they are normally regarded as low-level employees, the actions of most public service workers actually constitute the services “delivered” by government. Moreover, when taken together the individual decisions of these workers become, or add up to, agency policy’ (1980: 3). He further suggests that ‘unlike lower-level workers in most organisations, street-level bureaucrats have considerable discretion in determining the nature, amount, and quality of benefits and sanctions provided by their agencies’ (1980: 13). Of course, the issues described above are not the only factors that influence police decision-making. As Bell (2002) suggests, any police investigation is a complicated process in which officers use the law, official policies and procedures, prior training, routines, and practices to shape their decisions and to establish ‘fact’ in a given case. Lipsky (1980) acknowledges the importance of policies, rules and regulations in restraining, at least to some degree, the discretion of street-level bureaucrats, but points to what he sees as the impossibility of severely reducing or removing discretion from this type of work. This is because, he argues, the work of street-level bureaucrats involves complex tasks for which the elaboration of rules, guidelines or instructions cannot circumscribe the alternatives. The situations in which street-level bureaucrats work are too complicated to be reduced to programmatic formats, and these situations often require responses to the human dimensions of situations that are responsive to the unique circumstances of individual clients. The need to be compassionate and flexible on the one hand, yet at the same time to be impartial and rigidly apply the rules on the other, represents for Lipsky a dialectic of public service reform. As a consequence of the issues discussed here, it is widely accepted that there will be slippage between orders and the carrying out of orders, as a result of either poor communication, inadequate resourcing, inadequate sanctions for non-compliance, or workers’ disagreement with organisational goals and objectives and their consequent withholding of co-operation in the exercise of their duties (Lipsky 1980). The issue of slippage between policy and practice is of particular interest. The problem is, from the issues discussed here, that changes to prescriptive policy are not a guarantee that success will be achieved ‘in the real world’. Indeed there is substantial historical evidence that suggests that the transformation of police policy into effective practice is a complex and vulnerable process, particularly in the 160

Law enforcement and hate crime

field of police–community relations (Grimshaw and Jefferson 1987). Bowling (1999) suggests that during the 1980s and 1990s in Britain very little changed in policing practice despite numerous policy changes aimed at improving the police response to racist incidents. Part of the reason for this can be found in the problems that have been outlined already in this chapter. There is a long-held belief among minority groups, based on experience, that the police are either unable or unwilling to effectively investigate crimes against them, and that their cases are not taken seriously (Macpherson 1999; Victim Support 2006). This situation is compounded by research evidence that historically points to widespread and discriminatory ‘over-policing’ (Hunte 1966; Landau 1981; Willis 1983). The extent to which such ineffective policing is the result of individual or institutional prejudice and discrimination, inadequate resources, occupational culture, poor policing at lower levels or bad management at upper levels, or a product of the nature of hate crime itself is difficult to determine and has been the subject of much debate (Human Rights Watch 1997; Reiner 1997). A further factor that influences both decisions and the exercise of discretion is the issue of resourcing. As Lipsky (1980) suggests, there are several ways in which street-level bureaucracies such as the police characteristically provide fewer resources than necessary for workers to do their jobs adequately, the two most important being first, the ratio of workers to clients, and second, time to deal adequately with each client in what are characteristically large caseloads. Part of the problem of caseload size is a product of increased demand for services over time. In policing terms, hate crime has emerged as a relatively recent social problem, effectively meaning that there is political and public pressure for the police to respond that previously would not have existed. The most obvious example relates to the huge increase in recorded racist hate crimes (in effect, increased demand for police services) in England and Wales as a result of the post-Lawrence definitional change (Hall 2005). In response to this, following the Lawrence Inquiry, police forces and in particular London’s Metropolitan Police Service increased resourcing in order to meet Macpherson’s recommendations and to help challenge the problems the police faced at that time. At face value it would seem logical to expect that increases in funding and resources in this area should result in improvements in police performance. It is, however, a feature of public services that this is not likely to be the case, certainly in the long term. As Lipsky (1980: 33) explains: 161

Hate Crime

A distinct characteristic of the work setting of street-level bureaucrats is that the demand for services tends to increase to meet the supply. If additional services are made available, demand will increase to consume them. If more resources are made available, pressures for additional services utilising those resources will be forthcoming. In other words, in public services such as the police there will always be more people who need the services on offer than the police can provide services for. Unless it can somehow be controlled, demand will always outstrip supply, and perceived availability of a service will fuel demand for that service. The situation becomes particularly acute when the increases in resourcing are relatively moderate in quantity and relatively short in duration, as is often the case when different issues compete for political and police priority. On the other hand, however, if demand for services can be controlled, for example through the exercise of discretion in decision-making, then the pressures of inadequate resourcing may be eased. In addition to a lack of organisational resources, Lipsky suggests that street-level bureaucrats also experience deficiencies in personal resources, in terms of inexperience, lack of training, and so on. Lipsky points out that part of this inadequacy is attributable to the nature of the job in the sense that some jobs simply cannot be done properly regardless of personal or organisational resources. This point may have particular relevance to the policing of hate crime because there are some important issues that exist in this area over which the individual and the organisation have less potential for control. For example, it has been suggested that hate crime is best viewed as an ongoing process. Viewing hate crime in this way can help to further explain the negative views of the police that many members of minority groups hold. The underlying rationale is that there is a fundamental mismatch between the nature of hate crime and the requirements of the criminal justice system. With reference to racist victimisation, Bowling (1999) argues that the problem is that while hate crime is best viewed as an ongoing process, the police and wider criminal justice system necessarily respond to incidents. Bowling suggests that racist victimisation does not occur in a moment, but rather it is ongoing, dynamic and is embedded in time, space and place, and must always be kept in context if it is to be understood and responded to effectively. However, Bowling explains that as the ‘incident’ is transformed from the world of the victim’s experience 162

Law enforcement and hate crime

into an object for policing it is placed in the context of the police organisational and cultural milieu, an environment that is usually antithetical to that of the victim. The net result of the process– incident contradiction, according to Bowling, is that while the police may feel that they have responded appropriately and effectively, the victims are frequently left with feelings of dissatisfaction, fear, and a perception of being under-protected.

Implications This chapter has considered a number of theoretical issues relating to the law and law enforcement in relation to hate crimes. It has been argued that because hate crimes are complex events, they pose significant challenges to law-makers, the police, and also prosecutors. It has been suggested that there are many issues, both internal and external to those individuals and agencies responsible for law enforcement, and over which they have varying degrees of control, that inevitably impact upon their ability to enforce the law and to respond effectively and to provide a service appropriate to the needs of victims and wider communities. The theoretical perspectives discussed in this chapter highlight the potential for practical difficulties in law enforcement in hate crime cases, which it would seem are reflected in the practical problems experienced in the prosecution of such cases, as illustrated by the relatively low rates of conviction for these offences in most countries where such laws exist. At one level the potential for difficulty arises because, arguably, the law in this field reflects a number of the elements that Pound (1917) argued would limit legal effectiveness, not least because of the problems associated with proving the motivation behind an offence, or that the offence was aggravated by some prejudice-based hostility. In addition, Pound (1917) rightly notes the importance of the role of law enforcement agencies, most notably the police, in the process of law enforcement. The role of the police is crucial, as noted by Cotterrell (1992: 56): How the law is put into effect is clearly as important as its content. The nature of the enforcement agencies used, the degree of commitment of enforcement agents to implementation of the law, their morale and – a closely related factor – the amount of resources available to ensure compliance, are all shown to be 163

Hate Crime

extremely significant factors. In addition, the particular strategy of coercion or persuasion employed in regulation is clearly of great importance. The theoretical perspectives relating to the police discussed in this chapter serve to illustrate the importance of how the law is put into effect. The theoretical issues concerning ‘street-level bureaucrats’ demonstrate the powerful position held by those often in lower ranks of law enforcement agencies, and the potential for these employees to determine the outcomes of both law and organisational policy through their use of discretion. With these issues in mind, it is important to note that in some countries, notably in England and Wales, law enforcement is only one aspect of the policing response to hate crimes. This is important given that the literature discussed in this chapter implies that the nature and characteristics of hate crime effectively mean that the clearup rate is a relatively meaningless measure of police performance. It is therefore important to consider measuring ‘success’ in other ways. The ‘success’ of policing in relation to hate crimes might be better measured against the numbers of hate crimes reported by victims, where depending on the circumstances an increase may signify an increase in trust in the police, or where a decrease may perhaps suggest success in deterring hate crimes, or by victim satisfaction with the police response, as indicated by qualitative and quantitative surveys. The latter is of significant importance in assessing the ‘success’ of the police and is of far greater informational value than clear-up rates. It may also be possible to further measure ‘success’ against the educative role achieved by policing activities in the community, or the strength of the message conveyed to the public that hate crimes will not be tolerated and the suffering of victims will be treated seriously. One might also consider the ability of the police to respond proactively to a perceived threat or tension in a community; or the collecting and dissemination of hate-related intelligence; the training of other police officers to respond appropriately to hate crimes; or the ability of the police to liaise with advocacy groups and ‘build bridges’ between minority communities and the police. All of these represent significant functions undertaken by some police services that are not measurable simply by considering clear-up rates, yet the importance of which should not be underestimated, particularly given the difficulties in law enforcement in this field. In other words, enacting and attempting to enforce hate crime legislation is not a 164

Law enforcement and hate crime

panacea to the problem of hate crime. Rather, it is one part of a much bigger picture.

Conclusion The discussion above has highlighted a number of issues relating to the effective investigation of hate crimes. When combined these issues largely determine the effectiveness of criminal law in this field, which in turn, as Pound (1917) suggested, has important implications for trust and confidence in law enforcement agencies and for the invocation of law by the public. Dependence on an organisation that is necessarily influenced by individual and occupational perceptions and beliefs, finite resources, competing priorities and so on, to enforce the law and provide other related services in response to a given problem raises a host of issues. There are a plethora of issues relating to the role of the police in responding to hate crimes, but which can, arguably, be reduced to two key questions: do the police have the ability to respond effectively to hate crimes, and do the police have the desire to respond effectively to hate crimes? The answers to these two questions largely determine the extent to which legal and organisational goals in terms of responding to the problem, which itself is determined by the social context that gave rise to its construction, are met. The ability and desire of the police to respond adequately to any problem is crucial because, as Cotterrell (1992) suggests, to continue to function effectively the police must protect the social and political bases of their authority, and to do this they must demonstrate an adequate degree of success in the tasks allotted to or assumed by them. In terms of hate crime this situation is particularly acute given the social, historical and political context in which the contemporary policing of hate crime takes place. In policing terms, arguably the most important aspect of that context is the depth of trust and confidence that the public have in policing and the police. The police are largely dependent upon the public to invoke the services on offer, which in turn is dependent upon both the ability and the desire of the public to do so. Demonstrating an adequate degree of success in the tasks allotted to or assumed by the police with regard to hate crime is crucial in determining the extent to which services are invoked. The ability and desire of the police to respond to hate crimes are therefore crucial in influencing the ability and desire of the public to engage with the police, and 165

Hate Crime

the relationship is reciprocal. It is this relationship that is central to ‘success’, however so defined, in the policing of hate crime.

References Arnold, T. W. (1935) The Symbols of Government. New York: Harcourt Brace and World. AVP (2007) National Hate Violence Report. New York: Anti-Violence Project. Bell, J. (2002) Policing Hatred: Law Enforcement, Civil Rights and Hate Crime. New York: New York University Press. Berger, M. (1952) Equality by Statute: Legal Controls over Group Discrimination. New York: Columbia University Press. Bolton, K. Jr. and Feagin, J. R. (2004) Black in Blue: African-American Police Officers and Racism. New York: Routledge. Bowling, B. (1999) Violent Racism: Victimisation, Policing and Social Context. New York: Oxford University Press. Bradley, R. (1998) Public Expectations and Perceptions of Policing, Police Research Series, Paper 96. London: Home Office. Burney, E. and Rose, G. (2002) Racist Offences: How is the Law Working?, Home Office Research Study 244. London: Home Office. Chan, J. (1997) Changing Police Culture: Policing in a Multicultural Society. Cambridge: Cambridge University Press. Cotterrell, R. (1992) The Sociology of Law, 2nd edn. London: Butterworths. CPS (2003) Guidance on Prosecuting Cases of Racist and Religious Crime. London: Crown Prosecution Service. Craig, K. M. (2002) ‘Examining Hate-Motivated Aggression: A Review of the Social Psychological Literature on Hate Crimes as a Distinct Form of Aggression’, Aggression and Violent Behaviour, 7: 85–101. Cronin, S. W., McDevitt, J., Farrell, A. and Nolan, J. J. (2007) ‘Bias-Crime Reporting: Organizational Responses to Ambiguity, Uncertainty, and Infrequency in Eight Police Departments’, American Behavioral Scientist, 51(2): 213–31. Dror, Y. (1959) ‘Law and Social Change’, Tulane Law Review, 33: 787–802. Erlich, E. (1936) Fundamental Principles of the Sociology of Law. New York: Arno Press. Franklin, K. (2002) ‘Good Intentions: The Enforcement of Hate Crime PenaltyEnhancement Statutes’, American Behavioral Scientist, 46(1): 154–72. Gordon, P. (1983) White Law: Racism in the Police, Courts and Prisons. London: Pluto. Green, D. P., McFalls, L. H. and Smith, J. K. (2003) ‘Hate Crime: An Emergent Research Agenda’, in B. Perry (ed.) Hate and Bias Crime: A Reader. New York: Routledge. Grimshaw, R. and Jefferson, T. (1987) Interpreting Policework. London: Allen and Unwin. 166

Law enforcement and hate crime

Grossman, J. B. and Grossman, M. H. (eds) (1971) Law and Change in Modern America. California: Goodyear. Haberman, C. (1999) ‘Finding Flaws in the Logic of Bias Laws’, New York Times, 12 March. Hall, N. (2005) Hate Crime. Cullompton: Willan Publishing. Hall, N. (2009) ‘Policing Hate Crime in London and New York City’, unpublished PhD Thesis: University of Portsmouth. Holdaway, S. (1996) The Racialisation of British Policing. Basingstoke: Macmillan. Human Rights Watch (1997) Racist Violence in the United Kingdom. London: Human Rights Watch/Helsinki. Hunte, J. (1966) Nigger Hunting in England? London: West Indian Standing Conference. Jacobs, J. B. and Potter, K. (1998) Hate Crimes: Criminal Law and Identity Politics. New York: Oxford University Press. John, G. (2003) Race for Justice: A Review of CPS Decision Making for Possible Racial Bias at Each Stage of the Prosecution Process. London:, CPS. Johnson, M. (2003) Street Justice: A History of Police Violence in New York City. Boston, MA: Beacon Press. Kahn-Freund, O. (1969) ‘Industrial Relations and the Law: Retrospect and Prospect’, British Journal of Industrial Relations, 7: 301–16. Landau, S. (1981) ‘Juveniles and the Police’, British Journal of Criminology, 21(1), 27–46. Lester, A. and Bindman, G. (1972) Race and Law. Harmondsworth: Penguin. Levin, B. (2002) ‘From Slavery to Hate Crime Laws: The Emergence of Race and Status-Based Protection in American Criminal Law’, Journal of Social Issues, 58(2): 227–45. Lipsky, M. (1980) Street-level Bureaucracy. New York: Russell Sage Foundation. Macpherson, W. (1999) The Stephen Lawrence Inquiry. Cm 4262. London: The Stationery Office. Perry, B. (2001) In the Name of Hate: Understanding Hate Crimes. New York: Routledge. Pound, R. (1917) ‘The Limits of Effective Legal Action’, International Journal of Legal Ethics, 27: 150–67. Reiner, R. (1992) The Politics of The Police, 2nd edn. Hemel Hempstead: Harvester Wheatsheaf. Reiner, R. (1997) ‘Policing and the Police’, in M. Maguire, R. Morgan and R. Reiner (eds) The Oxford Handbook of Criminology, 2nd edn. New York: Oxford University Press. Scarman, L. (1981) The Brixton Disorders: 10–12 April 1981: Report Of An Inquiry: Presented to Parliament by The Secretary of State for the Home Department, November. London: HMSO. Skolnick, J. (1966) Justice Without Trial. New York: Wiley. Streissguth, T. (2003) Hate Crimes. New York: Facts on File. 167

Hate Crime

Victim Support (2006) Crime and Prejudice. London: Victim Support. Waddington, P. A. J. (1999) Policing Citizens. London: UCL Press. Willis, C. F. (1983) The Use, Effectiveness and Impact of Police Stop and Search Powers, Home Office Research and Planning Unit Paper No. 15. London:. Home Office.

168

Chapter 8

From hate to ‘Prevent’: community safety and counter-terrorism Derek McGhee

Introduction The ideas for this chapter came to me during a particularly interesting exchange at a conference in 2008. On this occasion the discussion had moved from a rather abstract discourse on the relationship between counter-terrorism and human rights to focus more specifically on the ‘policing’ challenges associated with extremism and radicalisation. It was during this discussion that I initially began to talk and think about concepts such as vulnerability in the context of policing. This led to some musing on my part on the diversity of vulnerable, ‘atrisk’ and targeted communities and individuals that are of concern to police. When the editor of this book asked me to contribute a chapter that would reconcile my research on the policing of lesbian, gay, bisexual and transgender (LGBT) communities1 and hate crime in general with my recent research on the impact of British counterterrorism strategies on Muslim communities, I was delighted to take up this opportunity to explore – and ultimately compare and contrast – the policing of these two very different communities. Therefore, in this chapter I explore the relationship between two different types of ‘community policing’: the policing of LGBT communities and the policing of Muslim communities.2 I will suggest in this chapter that building trust between these communities and the police is central in both instances. I also explore the implications of the different definitions of vulnerability that organise the interactions of police forces/services and these communities. For example, LGBT communities are presented in police discourses as being vulnerable to 169

Hate Crime

hate, and although Muslim communities are also presented in police discourses as being vulnerable to hate, the primary contemporary focus of policing with regard to Muslim communities is their vulnerability to extremism. This chapter explores the implications of these two very different discourses of vulnerability for the policing of these ‘communities’. The chapter begins by examining the emergence of a discourse of homophobic hate crime in the context of the emergence of joined-up, multi-agency community safety in the early years of the New Labour government. It then moves on to explore the recent investment in preventative community relations policing under the banner of ‘Prevent’, in the context of the government’s counter-terrorism strategy called CONTEST.

Community safety and ‘hard-to-reach’ groups It was in the passing of the Local Government Act of 1999 that much of this joined-up thinking in relation to communities and local government came into being. However, one year before (in the first year of the Labour administration), a parallel development in relation to the mobilisation of communities in the sphere of law and order was implemented in the Crime and Disorder Act of 1998. Even though this legislation is better known for the shake-up of the youth justice system, it was here that ‘community safety partnerships’, as first suggested in the Morgan Report (the Home Office’s report of the Standing Committee on Crime Prevention Through the Partnership Approach), were introduced as a statutory requirement. The emergence of such partnerships and initiatives should be put in context. It was with the publication of the Morgan Report in 1991 that ‘police-driven’ initiatives associated with crime prevention in general were seen as being limited in scope. It was suggested in this report that the term ‘crime prevention’ should be replaced with ‘community safety’, which is a wider interpretation that could encourage greater participation from all sections of the community in the fight against crime (Home Office 1991: 3). The Morgan Report suggested that the local authority was the ‘natural focus for co-ordinating, in collaboration with the police, the broad range of activities directed at improving community safety’ (Home Office 1991: 4). When these recommendations were eventually drafted in the Crime and Disorder Bill they took the form of statutory partnerships that required local authorities, the police, and other bodies to come 170

From hate to ‘Prevent’: community safety and counter-terrorism

together to develop strategies for tackling crime and disorder (DETR 1998: para. 1.15). This ‘required community consultation process’ in the implementation of community safety strategies and community representation in local partnerships has been described by Crawford as an attempt to implement a ‘holistic, problem-oriented approach’ to crime (Crawford 2001: 57–9). Thus, it was with the introduction of the Crime and Disorder Act, and also the Local Government Act and the programmes and reports that have followed in their wake, that the government attempted to enjoin local authorities and local police forces to the local communities that they serve in the name of ensuring democratic legitimacy and increasing community safety and community confidence in the police (and government). In this context all citizens were to be active citizens who are attached, live in or identify with ‘active’ participating communities. Now, these developments included both opportunities and challenges for ‘minority’ communities. The main challenge is that some ethnic and religious minority groups, as well as sexual minority groups such as LGBT communities, might have good reason not to trust police and other service providers. According to Jones and Newburn (2001: 7), these ‘potentially disaffected and vulnerable social groups’ were designated by the Home Office as ‘hard to reach’; at the same time, the views of these so-called hard-to-reach groups were essential to the ‘inclusive’ ambitions of the Crime and Disorder audits and strategies. According to the Home Office’s Guidance, referring specifically to gay and lesbian groups: It is absolutely central to the success of the partnerships that they should be seen as credible and inclusive by all sections of the community. It is likely that the Home Secretary will use his powers under Section 5(3) of the Crime and Disorder Act to require the police and local authorities to invite the full participation of gay and lesbian groups in the work of the new partnerships. This should do much to ensure that issues of concern to these groups are not overlooked when the audit is conducted and the strategy developed. (Home Office 1998: 2.44, emphasis in the original). Jones and Newburn suggested, however, that the label ‘hard to reach’, used to depict social groups such as lesbian and gay communities, should be changed to reflect the context of the type of disengagement associated with communities so described. For these authors, there are many examples of different social groups and communities who 171

Hate Crime

for diverse reasons have a difficult relationship with police. However, they stipulate that few of these groups are ‘hard to reach’ in any fundamental physical sense; in many cases ‘hard to reach’ actually means ‘hard to engage with on a positive level’ (Jones and Newburn 2001: 13). This change of label from ‘hard to reach’ to ‘hard to engage with on a positive level’ is rather useful for the purposes of this chapter. The reason for this is that the partial citizenship afforded to the gay community in the privacy provisions of the sex offences legislation in the 1950s and 1960s (Moran 1996), and the subsequent policing of this community in line with this legislation, are fundamental components in the designation of the lesbian and gay community as being hard to reach. By replacing ‘hard to reach’ with ‘hard to engage with on a positive level’, and by appreciating the context of police–community relationships – especially the perceptions of members of sexual and ethnic minority communities of being over-policed and under-protected – then it becomes clearer to police and other statutory partners why some groups, including members of LGBT communities, do not trust them. The Crown Prosecution Service (CPS) describe LGBT communities’ lack of trust in police as an ‘entirely understandable reluctance’ on the part of members of LGBT communities due to ‘the way in which members of the LGBT communities have historically been treated by individuals within the criminal justice agencies’ (CPS 2002: 7). In their updated guidance published in 2007, the CPS describe this problem as being a matter of perceptions of how LGBT people might be treated by police when reporting a homophobic incident (2007: 3). LGBT communities’ lack of confidence with regard to the criminal justice system, according to the CPS, stems from the belief that LGBT people who have witnessed or been a victim of a homophobic incident will be treated disrespectfully by the police because of their sexual orientation or gender identity (2007: 3). For ACPO this belief or perception by LGBT victims of a homophobic incident (or LGBT witnesses to a homophobic incident) is associated with their fear of experiencing ‘secondary victimisation’ by police (ACPO 2005: 30). The main way in which police forces have been attempting to break down the barriers between themselves and the lesbian and gay community (and the wider LGBT communities) is through taking homophobic (and transphobic) hate crime seriously. Evidence of LGBT ‘trust-building’ and homophobic hate crime initiatives have appeared all over the country. These initiatives testify to a new mobilisation of sexual citizenship that introduces a significant modification of 172

From hate to ‘Prevent’: community safety and counter-terrorism

the private, detached homosexual subject promoted by the sex offences legislation in the 1960s. These initiatives can be described as attempting to demarginalise lesbian and gay communities (as well as the wider LGBT communities) through inviting them to enjoy an expanded form of (equal) citizenship hitherto unavailable to them (McGhee 2005). These developments in policing strategies have also been supported in legislation. For example, section 146 of the Criminal Justice Act 2003 requires the courts to consider disability and sexual orientation ‘hostility’ as an aggravating factor when deciding on a sentence for any offence (ACPO 2005: 13). According to Chakraborti and Garland’s analysis of CPS guidance, the 2003 Act conveyed upon the courts the capacity to increase sentences for offences in which hostility based upon sexual orientation (or presumed sexual orientation) was present just before, during or just after the offence or when the offence was motivated (wholly or partly) by hostility towards persons who are of a particular sexual orientation (2009: 59). A further development includes the introduction of section 74 of the Criminal Justice and Immigration Act 2008 (which amends Part 3A of the Public Order Act 1986) to include a new offence of inciting hatred based upon a person’s sexual orientation, thereby achieving similar legal provision for LGB people to that already provided to minority ethnic communities under previous incitement legislation (Chakraborti and Garland 2009: 60).

Taking vulnerability to homophobic hate crime seriously The Association of Chief Police Officers (ACPO) was one of the first organisations in the UK to focus on the problems facing minority groups in society targeted by hatred. Significantly, ACPO’s Guide to Identifying and Combating Hate Crime, published in 2000, included homophobic hate crime alongside racist hate crime as the two most high-profile hate crime problems facing the UK. Both of these forms of ‘hate crimes’, according to ACPO, are associated with high-profile media events: for example, the Stephen Lawrence Inquiry and the nail-bombing of the Admiral Duncan (gay) pub in Soho in London in 1999 (ACPO 2000: 17). Hate crime in ACPO’s original guidance is taken to mean any crime where the perpetrator’s prejudice against an identifiable group of people is a factor in determining who is victimised (ACPO 2000: 10). The centrality of victimisation in relation to racist and homophobic hate crime is evident in ACPO’s definitions. For example, a racist 173

Hate Crime

incident is defined as, ‘any incident which is perceived to be racist by the victim or any other person’ and homophobic incidents are similarly defined as ‘any incident which is perceived to be homophobic by the victim or any other person’ (ACPO 2005: 11): in effect, ‘any incident intended to have an impact on those perceived to be lesbian, gay men, bisexual or transgender’ (ACPO 2000: 13). According to Paul Evans (Director of the Policing Standards Unit), writing in the foreword to ACPO’s updated guidance on hate crime published in 2005, hate crimes are special crimes that require special police attention because of the particular vulnerability of victims and the wider impact of such crimes: Hate crimes where people are targeted because of the nature of their diversity affects not only the primary victim but also the wider family and sometimes communities. It is a serious crime often committed against victims who are particularly vulnerable due to their individual circumstances. (ACPO 2005: 6) According to Moran, the original ACPO definitions (which have not changed significantly in subsequent ACPO guidance) of racist and homophobic hate crimes are significant as they focus on the ‘special circumstances’ and the ‘unique effects’ of these ‘value-added’ crimes, which although impersonal are experienced as being very personal and thus are the source of greater harm (Moran 2000: 12). For example, homophobic hate crimes have been described as having a particularly spatial impact on victims and victim communities in that these incidents have been described as a mechanism for ‘policing the closet’, or the boundaries of ‘privacy’ that work to keep lesbians and gays marginalised in the secret fringes of society (McGhee 2005). Thus, homophobic violence can be viewed as ‘a punishment for stepping outside culturally accepted norms and a warning to all gay and lesbian people to stay in “their place”, the invisibility and self hatred of the closet’ (Herek and Berrill 1992: 3). According to McManus and Rivers, homophobia is not just a fear; it is also a hatred of or prejudice towards people who are (or perceived to be) lesbian, gay or bisexual. It can be subtle or overt, and can take many forms (2001: 3): for instance, verbal abuse and harassment; physical assault and emotional violence; property damage; and other crimes (such as extortion). For McManus and Rivers, the victims of homophobia and homophobic incidents can experience a number of long-term mental health problems as a result of the victimisation or discrimination they face. For example, it has been suggested that the victims of 174

From hate to ‘Prevent’: community safety and counter-terrorism

homophobia in adolescence are more prone to self-harm and suicide than their heterosexual counterparts;3 equally, exposure to violence and harassment increases the prevalence of depression, anxiety and relationship problems among lesbians, gay men and bisexual men and women.4 There is also the possibility that the victims of homophobic incidents may suffer from post-traumatic stress disorder (2001: 3). Moreover, and as noted above, ACPO stress that it is not just individuals who are affected by hate crimes, but also the families and friends of the victim (their wider ‘community’). Thus hate crimes, whether racist or homophobic, are perceived by organisations such as ACPO as a serious threat to individuals, neighbourhoods and communities. The policy solutions to hate crime, other than introducing higher tariff offences – such as, for example, racially aggravated offences in the Crime and Disorder Act 1998 – are also associated with attempting to encourage the communities affected by hate crime to work with police in order to fight it. The following is a five-step plan devised by McManus and Rivers for the development of a generic police–LGBT communities anti-hate crime strategy: (1) prepare the ground of partnerships; (2) prepare a common policy statement; (3) consult and involve lesbian, gay and bisexual communities; (4) audit homophobic crime; (5) develop and implement a strategy for dealing with homophobic crime (McManus and Rivers 2001: 4). It will be suggested below that much of the focus of police–LGBT anti-hate crime partnerships is on the third and fourth steps of this five-step plan: that is, on the consultation with certain individuals who are thought to ‘represent’ the LGBT community, and on the auditing of homophobic hate crime. But what is being audited in homophobic victimisation surveys? According to Moran and Skeggs (2004: 16), homophobic victimisation surveys represent violence in two main ways: (1) as hidden violence; and (2) as unreported violence. They suggest that victimisation surveys are mechanisms for recognising ‘what is not recognised’ in official criminal statistics. Moreover, the conducting of a homophobic victimisation survey is often perceived as a reflexive act associated with institutions within the criminal justice system, who have often failed to take these instances of violence seriously to begin the process of doing so. This absence of data on homophobic victimisation is evident, according to Moran and Skeggs, in the absence of police reports of these instances of violence and is related to the lack of police investigation into the instances and perpetrators of this violence, as well as the dearth of prosecutions and convictions in relation to the perpetration of homophobic violence (2004: 16). 175

Hate Crime

However, homophobic victimisation surveys do not exclusively focus on homophobic violence and the experience of homophobic victimisation in local LGBT communities. Homophobic violence and victimisation is only one side of the problem to be audited. McManus and Rivers inform us that lesbian, gay, bisexual and transgender populations are at a greater risk of being victims of crime; they are also disadvantaged when they experience crime and find it more difficult to access the range of services available to many other victims of crime (2001: 2). Service provision, or more accurately the obstacles obstructing vulnerable communities accessing the range of services that should be open to them, is another important aspect of homophobic victimisation surveys or audits of homophobic crime. The presence or absence of adequate services and policies to deal with specific crimes is a significant part of ACPO’s classification of victims of homophobic and racist crime as hate crime victims. The controversial exclusion of domestic violence from ACPO’s framework, according to Moran, is particularly revealing of this classification process: This violence is not ‘hate crime’ [in the ACPO guide], in part because of the existence of well-established services and policies relating to violence against women. We must assume that the argument here is that there is no need to add value to women as victims, as their needs are already well met. (Moran 2000: 12) The presence of an institutionalised prejudice in service provider organisations is also one of the prerequisites for ACPO’s classification of hate crime. Lack of trust in service providers results in a further layer of vulnerability which can impact on individuals and whole communities because hate crime offenders ‘can often keep the same victim or group of victims locked in isolation and fear by keeping the physical extent of each attack at a level where it is unlikely to be reported’ (ACPO 2000: 21). Thus, the reporting behaviour and limited take-up of services with regard to the victims of homophobic crime is seen as another layer of ‘added-value’ in relation to these ‘special victims’. Part of the problem in relation to getting police–LGBT antihate crime partnerships established, according to McManus and Rivers, is the perception among members of LGBT communities that the police as an organisation is homophobic. These authors have produced a definition of institutional homophobia modified from the definition of institutional racism found in the report of the Stephen Lawrence Inquiry: 176

From hate to ‘Prevent’: community safety and counter-terrorism

The collective failure of an organisation to provide an appropriate and professional service to people because of their actual or perceived sexual orientation. It can be seen or detected in processes, attitudes and behaviour that amount to discrimination, through unwitting prejudice, ignorance, thoughtlessness and homophobic stereotyping, all of which disadvantages LGBT people. (McManus and Rivers 2001: 3) As we can see, LGBT communities are presented in police discourses as experiencing two different types of vulnerability: that is, vulnerability with regard to being targeted by hate criminals, and vulnerability in terms of not trusting the police as an organisation to support, protect and take the experience of hate crime seriously. Much of the efforts expended with regard to improving the policing of LGBT communities is therefore a matter of building trust through a range of ‘outreach’ mechanisms (for example, police–LGBT community liaison, multi-agency strategies) and crime reporting strategies (for example, anonymised reporting initiatives, police–LGBT community ‘drop-in’ surgeries). In the next section my attention shifts to examine the context of the emergence of preventative, ‘extremism-focused’ community policing strategies under the government’s counterterrorism strategy called CONTEST, and the very different ways the discourse of ‘vulnerability’ is deployed with regard to those referred to here as ‘Muslim communities’.

‘Prevent’, CONTEST and the policing of Muslim communities The UK has had a strategy for countering international terrorism in place since 2003 called CONTEST. CONTEST has four principal strands: ‘Prevent, Pursue, Protect and Prepare’ (HM Government 2006: 1). According to Blick et al. (2006: 31) there are two interconnected aims in this strategy: ‘first, gaining vital intelligence that will expose the terrorists; and secondly, engaging with the Muslim communities to obtain that intelligence’. The CONTEST strategy has all the hallmarks of a well-rounded counter-terrorism strategy. Unfortunately, according to Blick et al., the Blair administration gave prominence to ‘the hard end of the strategy’ such as legislation, measures against ‘preachers of hate’ and the pursuit and disruption of potential terrorist activities (2006: 31–2). Even the government has admitted that ‘Prevent’ was the least developed of the four CONTEST strands, as priority was given to protecting the country from ‘the immediate threat to life, 177

Hate Crime

rather than to understanding the factors driving radicalisation’ (HM Government 2009: 82). This relative neglect of the ‘Prevent’ aspects of CONTEST was to be addressed within the first few months of Gordon Brown taking up office as Prime Minister. However, much of the groundwork for the ‘Prevent’ programme was under way after the new Department for Communities and Local Government (DCLG) was introduced and the Home Office had been restructured in 2006. The Preventing Violent Extremism Pathfinder Fund (PVEPF) was launched by the DCLG in October 2006. Initially £5 million of funding was made available for distribution among 70 priority local authorities. The aim of the PVEPF programme was to fund projects that were to act as catalyst for the development of ‘bottom-up, community-based programmes to tackle violent extremism’ (DCLG 2007b: 4). This section explores the Blair administration’s ambivalence towards ‘Prevent’, and Gordon Brown and (former Home Secretary) Jacqui Smith’s championing of ‘Prevent’ in order to explore what ‘engagement with’ Muslim communities actually means in our current context. This will allow comparisons to be drawn between the Muslim-focused ‘Prevent’ community policing strategies and the policing strategies employed for engaging LGBT communities examined above. As noted above, the post-9/11 counter-terrorism strategy in the UK has developed in the main at the expense of an appreciation of the wider contexts of ‘radicalisation’ and extremism; more than that, there was a ‘lazy parlance’ in political discourses that does not adequately differentiate between ‘extremism’ and ‘radicalisation’ (DEMOS 2006: 41–2). At the same time, the Blair government’s subordination of alternative discourses of extremism (including the potential link between British foreign policy and extremism) has hampered the development of ‘partnership approaches’ on ‘terrorism’ with Muslim communities in the UK (McGhee 2008). It is suggested here that the major obstacle to the Blair government creating an effective partnership with Muslim communities in the UK is the government’s contradictory approach to Muslim communities as both suspect communities harbouring terrorists and as the communities they most need work with against extremism (see also Chakraborti 2007). Following the United Nations’ Global Counter-Terrorism Strategy Plan of Action 2006 and the Council of Europe’s Convention on the Prevention of Terrorism 2007, which both advocate ‘communityrelations’ approaches to preventing tensions through promoting dialogue and public awareness-raising (Davis 2007: 9), the new government under Gordon Brown took up and ran with the DCLG 178

From hate to ‘Prevent’: community safety and counter-terrorism

and the Countering International Terrorism strategy. As noted above, the rebalancing of CONTEST was to include a greater investment in prevention. For former Home Secretary Jacqui Smith, prevention was central in the fight against extremism. Smith stated at the First International Conference on Radicalisation and Political Violence that ‘stopping people becoming or supporting terrorists – that is the major long-term challenge we face’ (Smith 2007: 2). On the back of this endorsement the PVEPF was introduced from April 2007. This funding was made available to enable ‘priority’ local authorities to take forward programmes of activities to tackle violent extremism at a local level (DCLG 2007a: 4). The DCLG define the priority local authority areas to be targeted for the PVEPF funding as those local authorities with a sizeable Muslim community, with populations of 5 per cent or more Muslims, or those areas with a significant Muslim community concentrated in a few wards (DCLG 2007a: 6). In a similar vein to multi-agency LGBT community–police–local authority trust-building and consultation strategies (see above), the PVEPF projects were established in order to encourage local authorities to better understand and engage in dialogue with their communities; to forge partnerships with police, community and faith groups; and to work more effectively with mosques and ‘institutions of education’ (DCLG 2007b: 4). The PVEPF has been created as part of the DCLG’s strategy to deliver its priority remit. This is to enable local communities to challenge the ideas of extremists in order to prevent violent extremism at the local level and to deliver ‘local solutions’ to ‘local problems’ (DCLG 2007a: 4). Rather than shying away from or refusing to engage with what have been described as the felt grievances5 among some Muslims in Britain (Smith 2007: 15), the PVEPF projects are more in line with the ‘battle of ideas’ approach advocated by Gordon Brown (see speeches made by Brown when Chancellor of the Exchequer in McGhee 2008: 70–1) which were set out in the Department’s Preventing Violent Extremism – Winning Hearts and Minds report (DCLG 2007c). The projects funded by the PVEPF are varied and wide-ranging, depending on the needs of particular local areas. According to the DCLG, the funded projects range from promoting the contributions that Muslims make in local communities, to building the capacity of local communities to tackle violent extremism in their area, to protecting specific groups of vulnerable individuals from being targeted by violent extremists (DCLG 2007c: 8). As with my conclusions with regard to progressive policing strategies developed to build trust between LGBT communities and 179

Hate Crime

the police, on paper the PVEPF programmes seem to be a giant step forward. However, it is at this point where we need to be cautious. The trust-building and engagement between police and LGBT communities is a much more straightforward strategy. The police and CPS are bending over backwards to ensure that the victims and witnesses of homophobic incidents are treated with the utmost confidentiality, respect and discretion. Even when the LGBT witness or victim of a homophobic incident is himself or herself the perpetrator of a crime, the CPS has stipulated that ‘it is more important to prosecute the perpetrator of the more serious crime’ (CPS 2007: 12). The ‘hate crime’ in the form of the homophobic incident is therefore the ‘trump’ crime that will be prosecuted and not the less serious (for example, public sex or recreational drug) offences that the LGBT victims or witnesses to a homophobic incident were perpetrating at the time of the incident. Can similar levels of discretion exist when it comes to extremism? This is an issue I will come back to below. When it comes to the ‘Prevent’ strategy a range of questions must be explored. What does ‘engagement’, ‘capacity’ building and ‘building resilience against extremism in local communities’ (DCLG 2007c: 4) actually mean in practice? How does ‘Prevent’ compete with the other components of CONTEST on the ground? Suffice to say, the ‘Prevent’ strategy does not begin and end with DCLG-funded projects but is also part of a wider policing strategy. Just looking at HMIC reports produced in Hampshire, the county in which I live and work, Hampshire Constabulary and Hampshire Police Authority are very clear in one of their strategic priorities: that is, to protect communities from terrorism and domestic extremism (Hampshire Constabulary 2007: 6). They are also clear about how this strategic priority is to be achieved in their policing plan; that is, through developing closer ‘community engagement’ (Hampshire Police Authority and Hampshire Constabulary 2007: 10). What this entails is first, the ‘identification of vulnerable communities’, and second, engagement with communities so identified through forging ‘partnerships’ with them; these partnerships are described as requiring ‘close co-operation, consistency and compatibility when responding to real or perceived terrorist threats’ (Hampshire Police Authority Community Affairs Committee 2007: 5). In these reports there is (in line with the interdependent aims of CONTEST, see above) an explicit reference to the relationship between the identification of and engagement with ‘vulnerable’ or ‘priority’ communities and the accumulation of intelligence. For example, it was stated in the Local Policing Plan that ‘by forging these closer links it 180

From hate to ‘Prevent’: community safety and counter-terrorism

will be possible to capture accurate and up-to-date intelligence that can be used to build a picture of the issues affecting our community’ (Hampshire Police Authority and Hampshire Constabulary 2007: 10). This raises a number of questions. How will the police use and act upon this intelligence? How will this impact on the priority or targeted communities in question and their potential future relationship with police and other agencies? That is, does the emphasis on intelligence here detract from the other work that police forces (and for that matter local authorities and other multi-agency partners) currently do in and with local Muslim communities with regard to engaging with and empowering them? The government is aware of the precarious balancing act that government departments, security agencies and the police have to perform in order to address ‘Prevent objectives’. In this context these departments, agencies and services are attempting to support vulnerable individuals and enhance the resilience of communities, and to implement operations to disrupt propagandists and their networks; at the same time, they are attempting to build ‘further trust and confidence between police and Muslim communities’ (HM Government 2009: 99). However, there are even more precarious balancing acts going on here. Spalek and Imtoual’s (2007) examination of the multiple and complex roles that the police are expected to perform in the name of CONTEST and ‘Prevent’ are highly valuable for exploring some of these questions. For example, Spalek and Imtoual suggest that police forces often find themselves caught inbetween ‘hard’ and ‘soft’ engagement approaches, where the former are devoted to intelligence gathering through internal community surveillance and the use of informants and the latter are devoted to trust-building, community policing approaches (Spalek and Imtoual 2007: 189). While the policing initiatives associated with taking homophobic hate crimes seriously, examined above, are invariably ‘soft’ engagement approaches, this cannot be said for the policing initiatives associated with preventing violent extremism. Furthermore, the tension between hard and soft engagement approaches haunts Jacqui Smith’s promotion of ‘Prevent’ in that the former Home Secretary sees the ‘Prevent’ strategy as comprising both PVEPF projects and ‘community’ policing initiatives (which are not mutually exclusive). The tension between ‘intelligence gathering’, ‘prevention’ and ‘engagement’ is most apparent when Smith talked about finding new ‘ways of directly supporting vulnerable people – by intervening with individuals when families, communities and networks are concerned about their behaviour’ (Smith 2007: 12). Smith 181

Hate Crime

suggested that ‘support for vulnerable individuals is best provided by their own communities’, and that rather than arresting ‘our way out of the problem’, ‘we need to think about the most effective response – more about rehabilitation, where that will work, and less about the criminal justice system’ (Smith 2007: 13–14). The government, in their updated strategy for countering international terrorism published in March 2009, attempted to unpack the circumstances and potential causes associated with an individual’s vulnerability to extremism: Vulnerability is not simply the result of actual or perceived grievances. It may be the result of family or peer pressure, the absence of positive mentors and role models, a crisis of identity, links to criminality including other forms of violence, exposure to traumatic events (here or overseas) or changing circumstances (e.g. a new environment following migration or asylum). (HM Government 2009: 89) It is in the context of the government’s attempts to understand different types of vulnerabilities to extremism that they have suggested two types of ‘support’. One set of priority programmes will support ‘those who are believed to be vulnerable to radicalisation’ through programmes providing peer mentoring, diversionary activities and leadership programmes; these programmes will equip participants with ‘the knowledge and skills to challenge extremist narratives’ (2009: 89). However, alongside what they describe as these ‘early support’ programmes the government have devised ‘more intensive interventions’ which will focus on people who have already become drawn into violent extremist networks (2009: 89). Whereas Jacqui Smith mentioned that rehabilitation programmes normally used for illegal drug users could be modified and developed for these type of ‘extremists’ (Smith 2007: 12), the government has suggested that the more intensive interventions would build on existing multi-agency support mechanisms at a local level to support vulnerable adults and early intervention work with at-risk young people as part of the Every Child Matters strategy (2009: 89).6 ‘Local community’ derived knowledge and/or intelligence is central to the ‘Prevent’ strategy. Alongside the conflicting roles that the police are expected to play under ‘Prevent’ with regard to gathering and acting upon intelligence gleaned from contradictory (and potentially counter-productive) combinations of hard and soft engagements with priority communities, the PVEPF projects are also to be ‘knowledge 182

From hate to ‘Prevent’: community safety and counter-terrorism

systems’ generating intelligence and measurable effects. The government does not shy away from the relationship between local intelligence, ‘Prevent’ and the other strands of CONTEST in their strategy for countering international terrorism published in 2009. In this document they stipulate that ‘Prevent’-related intelligence requirements will require police, security and intelligence agencies to build on their existing programmes. This will involve the sharing of information between these organisations and local partners in order to ensure ‘the more comprehensive assessment of areas at risk from terrorism and radicalisation and enable authorities to better target “Prevent” related interventions’ (HM Government 2009: 92). The rationale for investing in the PVEPF projects is to further particular strategic objectives,7 ‘to measure success’ and ‘to enable the aggregation and comparison’ of ‘key local issues’ across priority local authority areas (DCLG 2007a: 7–8). In this light, as well as being set up to build the resilience of priority areas to combat and challenge extremism through capacity building, awareness training, youth work and the like, the PVEPF projects are also to be evidence-driven knowledge systems that are to connect central government to local ‘priority’ communities (and the ‘at-risk’ or ‘vulnerable’ individuals in their midst) through what Valverde and Mopas describe as ‘targeted governance’ (2004: 245) and what Rose describes as ‘preventative interventions’ (2007: 71). This signifies a shift away from universal, or in this case national, social programmes – for example, as found in Improving Opportunity, Strengthening Society, the combined ‘community cohesion’ and ‘race inequality’ strategy introduced by the Home Office in 2005 (see Home Office 2005; McGhee 2008: 90–5) – to ‘targeted’ specific ‘priority’ local authority-focused programmes. These evidence-based, intelligence-driven PVEPF programmes could transform the designated priority local authorities or areas into calculable spaces (Miller 1997: 256), which in theory can be acted upon by central government from a distance (Rose 1994: 364). Valverde and Mopas describe this as ‘the dream of targeted governance’ with its misplaced optimism in believing that ‘good information’ can and will be collected and will in turn enable ‘managers’, government departments or police forces ‘to target their organisation’s resources efficiently and with maximum benefit’ (Valverde and Mopas 2004: 246). The precariousness of such dreams are captured by the DCLG’s assumptions that in order to ‘measure success’ in the form of ‘demonstrable changes in attitudes among Muslims and wider communities’, it is essential that current local attitudes are mapped 183

Hate Crime

‘to provide a baseline for future progress’ (DCLG 2007a: 7). This is a tall order for any evaluation team commissioned to monitor the success of a PVEPF project; attempting to establish an attitudinal baseline for a number of Muslim and non-Muslim communities in an entire local authority area in order to measure the direct impact of a PVEPF project on their ‘attitudes’ is quite a challenge in itself. Designing surveys that contain what the DCLG describe as ‘a common core’ to enable aggregation and comparison across local authorities to identify best practice and to identify key local issues introduce, in combination, further challenges. In this light the PVEPF programme is reduced to a calculative regime whose successes or failures are measured less by the quality of the engagement facilitated by the projects and more by the potential for aggregating and comparing data in the context of evidence-based resource distribution apparati. Thus, both ‘Prevent’ policing work and the PVEPF programmes have the potential to become, or to be viewed as, intelligence-driven strategies that connect central government to ‘local’ communities (with high concentrations of Muslims) through particular intelligence or ‘knowledge economies’. Although there is evidence of a re-engagement with Muslim communities under Brown and Smith’s PVEPF programme, this too is a slightly different government–‘Muslim community’ engagement. This is because it bypasses national-level Muslim organisations to focus on a wider range of Muslim groups in particular designated ‘at-risk’ areas through the net-widening partnership approach introduced in the PVEPF programme. One of the remits of the PVEPF projects is to ‘develop a more detailed understanding of the range of Muslim and other organisations and key individuals with whom they can work’ (DCLG 2007a: 8). In fact the DCLG suggested that PVEPF resources should ‘be used to broaden the range of contacts and networks, particularly those engaging with young people and women for whom opportunities to express views and participate can be limited’ (DCLG 2007a: 8). It is on this basis, according to the DCLG, that Local Forums on Extremism and Islamophobia should be built (DCLG 2007a: 8).8 There are considerable parallels here with the strategies recommended for reaching deeper into LGBT communities. For example, rather than just being motivated to find those ‘Muslim groups’ who are willing to tackle extremism, the PVEPF programme encourages a deeper reach into Muslim communities in order to engage with the ‘hard-to-reach’ or traditionally marginalised groups of women and young people. However, it is here where we again 184

From hate to ‘Prevent’: community safety and counter-terrorism

run up against a tension in the ‘Prevent’ strategy. Is this ‘reaching deep’ into Muslim communities simply about increasing the representativeness and enhancing the active participation of women and young people? What is the relationship between the latter and the intelligence-driven focus on ‘at-risk’ or ‘vulnerable’ individuals? Engaging young people and women in this light increases their potential for participating and being represented in local projects and forums; however, it also draws these ‘hard-to-reach’ groups into potential intelligence networks through increasing the potential for internal community surveillance into the marginalised, disaffected ‘hard-to-reach’ parts of these communities. Thus, just as Ruth Kelly (former Local Government and Communities Minister) set out to transform the government–‘Muslim community’ consultation landscape post-7/7 (see McGhee 2008: 78) in response to the view that Muslim leaders had not taken a strong enough stand against extremism, the PVEPF projects have reinstated ‘government through community’ in that central government is to be linked to individuals (for example, ‘at-risk’ or ‘vulnerable’ individuals) through ‘local’ public, private, third-sector, police and ‘Muslim community’ partnerships under ‘Prevent’. This will no doubt have a transformative effect on all those involved, including front-line police officers grappling with their conflicting roles; ‘Muslim leaders’ representing their communities, as well as the ever-wider community groups and all those agencies included in multi-agency partnerships in local authority areas, as they are expected to transform themselves into ‘calculating individuals’ in ‘calculable spaces’ subjected to ‘calculative regimes’ (Miller 1997) set up to measure ‘community’ resilience against extremism. At the same time, the PVEPF projects could be conceived as the Labour government (initiated under Blair and implemented under Brown) forcing responsibility for countering extremism onto Muslim communities through a process of devolving responsibility downwards onto communities (O’Malley 2004: 73). This responsibility is being articulated less in the accusatory mode of ‘Muslims must do more to fight extremism’ and more in terms of the responsibility of Muslims and Muslim communities to perform their duty of being ‘front-line’ ‘vigilant watchers’ in their mosques, and among their neighbours and families. This ‘vigilant or watchful visuality’ (Amoore 2007: 215) is to be taken into their intimate relationships as parents, siblings, cousins, friends and neighbours as they play their part in pre-empting ‘future events’ through scrutinising those in their midst for what the former Home Secretaries John Reid and Jacqui Smith described as the 185

Hate Crime

‘telltale signs’ of potential extremism, or the evidence of vulnerability of ‘vulnerable individuals’ to extremism. There is one further tension that should be briefly examined here. That is the government’s attempts to rally support for their wider counter-terrorism strategy (including ‘Prevent’) through their particular articulation of extremism as being akin to a form of hatred, prejudice and discrimination. The government is attempting to co-opt into their counter-terrorism strategy the ‘intolerance of intolerance’ discourse associated with the hatred and prejudice that since 2000 has become increasingly ‘mainstreamed’ in service provision and linked to protections in law associated with the incitement, motivation and aggravation of hatred targeted at minority communities. That is, the government is attempting to promote the intolerance of extremism alongside the now familiar intolerance in Britain with regard to other forms of hatred, prejudice and discrimination through insisting that ‘the duty on all of us – government, citizens and communities’ is to ‘challenge those who dismiss our shared values’ (HM Government 2009: 97). By so doing the government is attempting to ally their ‘Prevent’ strategy with what they describe as their established ‘tradition of building strong, empowered and resilient communities’ through ‘tackling all forms of hate crime, and promoting equal opportunities’ (HM Government 2009: 87). In such a way, the government can seek to rally support for ‘Prevent’ through stipulating who the target of these strategies are, namely, ‘Those who, for whatever reason or cause, reject the rights to which we are committed, scorn the institutions and values of our parliamentary democracy, dismiss the rule of law and promote intolerance and discrimination on the basis of race, faith, ethnicity, gender or sexuality’ (HM Government 2009: 87). Through this discourse, extremist like ‘hate criminals’ are relegated to what Beck would describe as the ‘counter-modern residuum of our dynamic, complex and diverse contemporary societies’ (1997: 11), who for the sake of ‘our’ society must be identified, rehabilitated or incarcerated for ‘our’ protection.

Conclusion Perhaps the major difference with regard to the policing of these two different communities is that of protection. One could say that the strategy of taking homophobic hate crimes seriously is motivated by the recognition that LGBT communities, like other minority communities, including Muslim communities, have not 186

From hate to ‘Prevent’: community safety and counter-terrorism

been adequately protected from targeted crimes motivated by or aggravated by hatred and hostility. However, when it comes to the policing of Muslim communities under ‘Prevent’ one could say that the strategies are motivated by ‘public security’ rather than increasing the protection of Muslim communities per se. Meanwhile, the policing strategies associated with taking homophobic incidents seriously and Prevent can be described as strategies dedicated to improving the communication and interaction between police and these communities. There are a number of key differences. One difference, as noted above, is the lack of ‘hard’ policing strategies with regard to LGBT communities and the often uncomfortable combination of hard and soft engagement approaches in the policing of counter-terrorism. Another difference is that the strategy of taking homophobic incidents and homophobic victims seriously is a direct ‘victim-centred’ strategy; on the other hand, the ‘Prevent’ strategy is rather more indirect in that it targets the parents and neighbours of individuals at risk of extremism. Furthermore, the LGBT or homophobic incident strategy is dedicated to positively engaging with and supporting victims. The ‘Prevent’ strategy seems to be dedicated ultimately to protecting ‘the general public’ through devising networks of surveillance in the name of early intervention. At the same time, I would suggest that the policing of homophobic hate crime can be described as a strategy of ‘protective inclusionism’ with implications for summoning LGBT communities to greater civic participation and even ‘active citizenship’ in neighbourhood policing and police–LGBT liaison groups. However, in contrast I believe that the policing strategies under ‘Prevent’ could lead to divergent outcomes in different areas of the country. Some of these projects might be highly successful in improving and strengthening community relations between the police and other agencies; however, this might not be the case everywhere. I can imagine that there must be a range of responses, at local authority level, when notification is received that PVEPF funding is to be made available to that local authority. I am sure all sources of funding made available to cashstrapped local authorities and their partner agencies and community group beneficiaries is always very welcome. However, the PVEPF funding comes with several, often ambivalent strings, potential consequences and risks attached to it, especially when it comes to concerns over the impact of news of impending PVEPF project funding on Muslim community groups. Presenting a well-balanced account of the necessity of this intervention, and the funded projects associated with it, is perhaps not always an easy task given that the 187

Hate Crime

communities to be targeted perceive themselves as being viewed with considerable ambivalence. For example, Azad Ali, Chair of the Muslim Safety Forum,9 in 2006 made reference to the ‘strongly held belief amongst grassroots Muslim communities that the government and law enforcement agencies operate on the basis that Muslims are terrorists and or harbour them’ (cited in McGhee 2008: 79). When one considers the definition of a ‘priority area’ under ‘Prevent’ there does seem to be something to Ali’s accusations with regard to the complex Islamophobias associated with Muslim-focused concerns over extremism in contemporary Britain. These areas are designated priority areas because they have a relatively high Muslim population. In the context of this police–‘Muslim community’ PR problem I regard the ‘Prevent’ strategy (including ‘Prevent’ policing strategies and the PVEPF programmes) as potentially harbouring ‘misplaced optimism’ (see also Valverde and Mopas 2004), that will at best only partially fulfil the remits associated with: engagement, capacity building, intelligence and resource distribution in response to aggregated and comparable data. At worst the ‘Prevent’ strategy will be limited by mutual suspicion across the ‘Muslim–non-Muslim’ divide; alternatively it could cause tensions in local communities who refuse to engage, build their capacity and become partners in what they could potentially view as enforced internal community surveillance under a perceived ‘colonial governance’ associated with what Miller describes as ‘a type of mastery made possible by those at a centre having a particular type of information about events and persons distant from them’ (1997: 243). ‘Prevent’ is, after all, part of CONTEST, and as Blick et al. suggest, CONTEST has two interdependent aims: ‘first, gaining vital intelligence that will expose the terrorists; and secondly, engaging with the Muslim communities to obtain that intelligence’ (2006: 31). As such, ‘Prevent’ should be viewed just as integrated into these aims as ‘Protect’, ‘Pursue’ and ‘Prepare’. Whereas the policing strategies explored here with regard to LGBT communities can be described as a strategy dedicated to reassuring and engaging these communities and building trust, the ‘Prevent’ strategy, in contrast, can be described as a strategy dedicated to reassuring the ‘general public’ and one that attempts to get Muslim communities ‘on board’ against extremism in the context of deepseated mistrust of the latter. In this regard the ‘Prevent’ strategy could potentially reinforce the general public’s view of Muslims as a homogenous ‘suspect community’ consisting of a range of diverse

188

From hate to ‘Prevent’: community safety and counter-terrorism

local ‘vulnerable communities’ who present a problematic ‘absence of resilience’ to ‘and in some cases tacit support’ for extremist activities (HM Government 2009: 83). Perhaps an unintended consequence of the various initiatives under ‘Prevent’, and statements such as these, could be that the general public’s views of Muslim communities could be further tinged with fear and suspicion in a climate in which post-9/11 Islamophobia is yet to dissipate.

Notes 1 The single LGBT ‘banner’ employed by academic researchers, police and other agencies is now very common. In this chapter I will employ this ‘umbrella’ banner, but a note of caution must be included. By using this banner I am not suggesting that the experiences of all these very different communities which include, gay men, lesbian women, transgender people, bisexuals (which could include individuals categorised as or who self-define as heterosexual men who have sex with men) are similar enough to group them under one banner: these communities are themselves internally diverse in terms of class, ethnicity and geography and can suffer different levels and forms of hate crime victimisation and also variable experiences of ‘policing’ (Chakraborti and Garland 2009: 70). 2 The author is, of course, aware that Muslim LGBT individuals could be the target of complex prejudices and hate crime activities motivated by ethnicity, religion, sexuality and gender. This chapter will not examine these intersectional experiences of hatred. 3 See Canadian research conducted by Bagley and Tremblay (1997). 4 See American and British research conducted by Hershberger and D’Augelli (1997) and Rivers (2000) respectively. 5 Jacqui Smith’s list of ‘Muslim’ grievances included: grievances about foreign policy, grievances about socio-economic deprivation, and grievances based on perceptions/misperceptions of police and law enforcement activity (2007: 15) 6 The government reported that they have put in place what they call the Channel Programme which will be co-ordinated by police and local authorities with the aims of identifying those at risk from violent extremism and provide help to them, primarily through communitybased interventions (HM Government 2009: 90). 7 According to the DCLG the strategic objectives for the programme (including national, regional and local dimension) are to develop a community in which Muslims in our communities identify themselves as a welcome part of a wider British society and are accepted as such by the wider community; reject violent extremist ideology and actively condemn

189

Hate Crime

violent extremism; isolate violent extremist activity, and support and cooperate with the police and security services; and develop their own capacity to deal with problems where they arise and support diversionary activity for those at risk (DCLG 2007a: 7). 8 One could suggest that this process of increasing the representativeness and reaching deep into Muslim communities is the end point of the recommendations made previously by the then Home Office Faith Communities unit in their 2004 report. 9 The Muslim Safety Forum consists of ‘Muslim community’ representatives, Metropolitan Police and Association of Chief Police Officers. Representatives from other police forces (other than the Metropolitan Police Service) attend by request or invitation. This group holds monthly meetings in New Scotland Yard, which are described as ‘national meetings, albeit mainly London based’ (HM Government 2006: 22).

References Amoore, L. (2007) ‘Vigilant Visualities: The Watchful Politics of the War on Terror’, Security and Dialogue, 38(2): 215–32. ACPO (2000) Guide to Identifying and Combating Hate Crime. London: Association of Chief Police Officers. ACPO (2005) Hate Crime: Delivering a Quality Service – Good Practice and Tactical Guidance. London: Home Office Police Standards Unit. Bagley, P. and Tremblay, A. (1997) ‘Suicidal Behaviours in Homosexual and Bisexual Males’, Crisis, 18(1): 24–34. Beck, U. (1997) The Reinvention of Politics: Rethinking Modernity in the Global Social Order. Cambridge: Polity Press. Blick, A., Choudhury, T. and Weir, S. (2006) The Rules of the Game: Terrorism, Community and Human Rights, a Report by the Democratic Unit, Human Rights Centre, University of Essex for the Joseph Rowntree Trust. Available at www.jrrt.org. Chakraborti, N. (2007) ‘Policing Muslim Communities’, in M Rowe (ed.) Policing Beyond Macpherson: Issues in Policing, Race and Society. Cullompton: Willan Publishing (pp. 107–27). Chakraborti, N. and Garland, J. (2009) Hate Crime: Impact, Causes and Responses. London: Sage. CPS (2002) Guidance on Prosecuting Cases of Homophobic Crime. London: Crown Prosecution Service. CPS (2007) Guidance on Prosecuting Homophobic and Transphobic Crime. London: Crown Prosecution Service. Crawford, A. (2001) ‘Joined-up but Fragmented: Contradiction, Ambiguity and Ambivalence at the Heart of New Labour’s “Third Way” ’, in R. Mathews and J. Pitts (eds) Crime, Disorder and Community Safety. London: Routledge (pp. 54–80). 190

From hate to ‘Prevent’: community safety and counter-terrorism

Davis, T. (2007) ‘Do Terrorists Have Human Rights?’, Keynote Address, First International Conference on Radicalisation and Political Violence, King’s College, London. Available at www.icsr.info/files/ICSR%20Remarks%20b y%20Terry%20Davis.pdf. DCLG (2007a) Preventing Violent Extremism Pathfinder Fund. London: Department for Communities and Local Government. Available at www. communities.gov.uk/documents/communities/pdf/320330. DCLG (2007b) Preventing Violent Extremism Pathfinder Fund 2007/08 Case Studies. London: Department for Communities and Local Government. Available at www.communities.gov.uk/documents/communities/pdf/324967. DCLG (2007c) Preventing Violent Extremism – Winning Hearts and Minds. London: Department for Communities and Local Government. Available at www.communities.gov.uk/documents/communities/pdf/320752. DEMOS (2006) Bringing it Home: Community-based Approaches to CounterTerrorism. Available at www.demos.co.uk. DETR (1998) Modernising Local Government: Local Democracy and Community Leadership. London: Department of the Environment, Transport and the Regions. Hampshire Constabulary (2007) HMIC Inspection Report Hampshire Constabulary, October 2007. Available at http://inspectorates.homeoffice.gov.uk/hmic/ Inspections/phase_1_reports_20071/hampshire_2007?view=Binary. Hampshire Police Authority Community Affairs Committee (2007) Safer Neighbourhoods – Counter-Terrorism and Domestic Extremism – Report of the Chief Constable. Available at www3.hants.gov.uk/item_6_safer_ neighbourhoods__counter_terrorism_and_domestic_extremism.doc. Hampshire Police Authority and Hampshire Constabulary (2007) Local Policing Plan 2007/2008. Available at www.hantspa.org/lpp_final_290607_ june_version.doc. Herek, G. M. and Berrill, K. T. (1992) ‘Introduction’, in G. M. Herek and K. T. Berrill (eds) Hate Crimes: Confronting Violence Against Lesbians and Gay Men. London: Sage (pp. 1–10). Hershberger, D. and D’Augelli, R. (1997) ‘Predictors of Suicide Attempts among Gay, Lesbian and Bisexual Youth’, Journal of Adolescent Research, 12(4): 477–97. HM Government (2006) Countering International Terrorism: the United Kingdom’s Strategy, Cm 6888. Norwich: The Stationery Office. Available at www. intelligence.gov.uk/upload/assets/www.intelligence.gov.uk/countering.pdf. HM Government (2009) The United Kingdom’s Strategy for Countering International Terrorism, Cm 7547. Norwich: The Stationery Office. Available at http://security.homeoffice.gov.uk/news-publications/publication-search/ general/HO_Contest_strategy.pdf?view=Binary. Home Office (1991) Safer Communities: The Local Delivery of Crime Prevention through the Partnership Approach (Morgan Report). London: Home Office. Home Office (1998) Guidance on Statutory Crime and Disorder Partnerships: Crime and Disorder Act 1998. London: Home Office. 191

Hate Crime

Home Office (2000) Setting the Boundaries: Reforming the Law on Sexual Offences. London: Home Office. Home Office (2005) Improving Opportunities, Strengthening Society: The Government’s Strategy to Increase Race Equality and Community Cohesion. London: Home Office. Available at www.homeoffice.gov.uk/documents/ improving-opportunity-strat?view=Binary Jones, T. and Newburn, T. (2001) Widening Access: Improving Police Relations with Hard to Reach Communities, Police Research Series Paper 138. London: Home Office Research, Development and Statistic Directorate. McGhee, D. (2005) Intolerant Britain? Hate, Citizenship and Difference. Maidenhead: Open University Press and McGraw-Hill. McGhee, D. (2008) The End of Multiculturalism? Terrorism, Integration and Human Rights. Maidenhead: Open University Press and McGraw-Hill Education. McManus, J. and Rivers, D. (2001) Without Prejudice: A Guide for Community Safety Partnerships on Responding to the Needs of Lesbians, Gays and Bisexuals. Available at www.nacro.org.uk. Miller, P. (1997) ‘Accounting and Objectivity: The Invention of Calculating Selves and Calculable Spaces’, in A. Megill (ed.) Rethinking Objectivity. Durham, NC and London: Duke University Press (pp. 239–64). Moran, L. (1996) ‘The Homosexualization of Human Rights’, in C. Gearty and A. Tomkins (eds) Understanding Human Rights. London: Mansell (pp. 313–35). Moran L. (2000) ‘Victim Surveys and Beyond: Researching Violence Against Lesbians, Gay Men, Bisexuals and Transgender People’, SCOLAG Journal, September: 10–12. Moran, L. and Skeggs, B. (2004) Sexuality and the Politics of Violence and Safety. London: Routledge. O’Malley, P. (2004) Risk, Uncertainty and Government. Glasshouse Press: London and Sydney. Rivers, I. (2000) ‘Going Against the Grain: Supporting Lesbian, Gay and Bisexual Clients as They “Come Out” ’, British Journal of Guidance and Counselling, 28(4): 503–13. Rose, N. (1994) ‘Expertise and the Government of Conduct’, Studies in Law, Politics and Society, 24: 359–67. Rose, N. (2007) The Politics of Life Itself. Princeton, NJ: Princeton University Press. Smith, J. (2007) ‘Our Shared Values – A Shared Responsibility’, Keynote Address, First International Conference on Radicalisation and Political Violence, King’s College, London. Available at www.icsr.info/files/ICSR %20Remarks%20by%20Jacqui%20Smith.pdf. Spalek, B. and Imtoual, A. (2007) ‘Muslim Communities and CounterTerrorism Responses: “Hard” Approaches to Community Engagement in the UK and Australia’, Journal of Muslim Minority Affairs, 27(2): 185–202. 192

From hate to ‘Prevent’: community safety and counter-terrorism

Valverde, M. and Mopas, M. (2004) ‘Insecurity and the Dream of Targeted Governance’, in W. Larner and W. Walters (eds) Global Governmentality: Governing International Spaces. London: Routledge (pp. 233–50).

193

Chapter 9

Hate crime victims and hate crime reporting: some impertinent questions Kris Christmann and Kevin Wong

Introduction Much of the academic, practitioner and voluntary sector interest in victims of hate crime has focused upon the impacts of hate crime and the practical and emotional support needs and services for victims. Our own work has been somewhat divergent from this. We were commissioned to identify how hate crime reporting could be improved in an English northern town and made inclusive across different equality groups. We undertook a small-scale study that examined individual decision-making by hate crime victims in whether or not to report incidents, and how the available reporting arrangements and associated publicity materials affected these decisions (Wong and Christmann 2008). Somewhat to our surprise, what appeared to be a critical issue in terms of whether or not hate crime policies were likely to succeed was also a much under-researched area.1 While our own research findings cannot be generalised beyond the study site, it did allow us to test out and consider more thoroughly some of the assumptions implicit in policy developments around hate crime reporting, specifically the policy goal of full reporting. We want to reflect back on these findings and the broader research literature to pose some questions on the adequacy and utility of the current reporting agencies’ approaches and the general policy direction to hate crime victims. We believe that this has merit because the statutory criminal justice agencies and the voluntary sector are grappling with the challenges of adopting hate crime in its broadest sense, and providing a responsive, 194

Hate crime victims and hate crime reporting

effective and victim-centred service across markedly different vulnerable groups. Pertinent questions can be asked about what the current policies on hate crime can be expected to achieve given the nature of victim decision-making on the critical issue of whether to report their victimisation. We will draw out some implications that the legacy of the Lawrence Inquiry has had for strategic thinking and policy-making, and make some tentative suggestions on how these might be improved. We argue something that may be considered heresy among hate crime victimology circles and victim campaigning groups: that the current policy message concerning victim reporting does not reflect reality and risks being discredited. What is required, some ten years post-Macpherson, is more nuanced responses and ones that acknowledge: (1) the distance travelled by criminal justice agencies in the intervening years; (2) that the majority of hate crime is manifested as single incidents of harassment (which may not necessarily constitute crimes); and (3) the unlikelihood of full reporting by the public, which realistically fits where the public are in terms of their expectations. In doing so we do not pretend to have any authoritative answers to these issues, but believe that the questions are worth posing to prompt a debate between efficacy of response versus a largely unchallenged view of hate crime victimology.

Hate crime reporting The current policy imperative is aimed at encouraging the reporting of hate crime, driven by the conviction that reporting rates are notoriously low. There is good reason to believe that this is the case as much of the research literature suggests that victims of hate crime are less likely to report their victimisation to the police when compared to other crimes. The Metropolitan Police Service estimates that the majority of racist and religious hatred crimes, and as much as 90 per cent of homophobic hate crime, go unreported, although crucially this has to be seen within the context of low levels of reporting generally. For volume crime, for instance, the British Crime Survey (BCS) 2008/09 suggests that 41 per cent of crimes2 were reported to the police, whereas 59 per cent were never reported, although the likelihood of reporting a crime varies considerably by the type of offence (Home Office 2009). This last point is well illustrated by the analysis by Goudriaan et al. (2004) of the International Crime Victims Survey, which spans 16 industrialised countries; they found that 195

Hate Crime

assaults and threats were the least frequently reported, with only slightly more than a third being reported. Nationally, in 2006 the police recorded 43,000 racially or religiously motivated hate crimes,3 but a more complete measure comes from the BCS, which indicated a substantially higher figure of 184,000 offences that victims believed were motivated by prejudice. This suggests that over three-quarters (76 per cent) were never reported to the police.4 Bouten et al. (2002) found that only half of all incidents across six crime types in 17 industrialised countries (including the UK) are reported to the police. Needless to say, we do not know the true level of unreported hate crime, as is the case with the ‘dark figure’ of volume crime more generally. In response to such deficits, periodic but patchy local campaigns are launched up and down the country to raise awareness and instruct hate crime victims to the many ways to report incidents.5 The policy logic is that achieving higher rates of reporting to the criminal justice system will better identify the level of presenting need and inform how best to address and control victimisation. But do we know enough already, without sacrificing further resources, in a vain effort to achieve something close to accurate reporting levels across groups vulnerable to hate crime? In our own study we found that the official concern over reporting all hate crimes for service planning requirements was not shared by the overwhelming majority of respondents, and furthermore would not have been feasible to deliver (Wong and Christmann 2008). This regularity in victim decision-making is reflected in the wider criminological literature which consistently demonstrates that the higher the perceived seriousness of a crime the greater the probability that this will be reported to the police (see, for example, Skogan 1976, 1984; Sparks et al. 1977; Fishman 1979; Goudriaan et al. 2004). The underlying assumption has been that victims undertake a cost–benefit calculation (consciously or unconsciously) in deciding whether it is worth the effort involved in contacting the police and making a report (Skogan 1984). However, if the victim sustains physical injury or the loss is high, then the benefits of reporting to the police are deemed that much greater, as opposed to when only minor injury or loss, or no loss, is involved. Similarly, in Australia the analysis of victim survey data suggests that incidents causing physical harm to the victim, or with greater potential for threatening the victim’s safety, are reported at higher rates than less serious ones (Carcach 1997). Without exception, then, crime seriousness has been shown to be the principal determinant of reporting behaviour, and the BCS 196

Hate crime victims and hate crime reporting

findings would also appear to confirm this. Table 9.1 summarises the reasons that victims gave for not reporting the incident to the police (across all BCS crime types and for the two most pertinent crime types as regards the reporting of hate crime incidents: violence and vandalism).6 According to the BCS 2008/09, the reasons for victims not reporting volume crime primarily centre on the victim perceiving the incident as too trivial, no loss being involved or believing that the police would (or could) not do much about the offence (76 per cent of incidents).8 A further 14 per cent of respondents felt that the incident was a private matter to be dealt with by them, and a further 5 per cent cited the inconvenience of reporting. Only 1 per cent stated policerelated reasons such as dislike or fear of the police and having had a previous bad experience with the police or courts. That said, it is interesting to note that Goudriaan et al. (2004) found that perceived police confidence had no effect on the reporting of contact crimes, but it did on the chance that property crime would be reported. As we would expect, these reasons for not reporting crime change markedly when examining violent crime as a separate category, with fewer victims citing the incident as too trivial or fearing police inaction (52 per cent). For BCS respondents, considerably more regarded being a victim of violence as a private matter (34 per cent) than for all BCS crime (possibly indicating instances of intimate partner violence). Table 9.1  British Crime Survey 2008/09: victims’ reasons for not reporting crime to police. Reasons for not reporting crime to Police Trivial/no loss/police would not/   could not do anything Private/dealt with ourselves Inconvenient to report Reported to other authorities Common occurrence Fear of reprisal Dislike or fear of police/previous   bad experience with police   or courts Other7

All BCS crime (%)

All violence Vandalism (%) (%)

76 14 5 4 2 2

52 34 5 6 5 6

87 8 4 1 2 2

1 5

3 6

1 3

Source: Home Office (2009: 41). 197

Hate Crime

Fear of reprisals, dislike of the police, repeat victimisation and reporting to alternative authorities all figure higher for not reporting violence when compared to all BCS crime. The overriding reason for not reporting vandalism was that victims considered it too trivial or they concluded that nothing would be or could be done by the police (some 87 per cent). Similar reasons for the non-reporting of crime have been found in the Australian criminological literature (see, for instance, Carcach 1997). With this combined weight of evidence we would reasonably expect that the severity of an incident holds true for hate crime victims as well. This was the case with our own research, which found a steep decline in victims’ propensity to report as a function of both the incident severity and the frequency of its occurrence (although others such as Goudriaan et al. (2004) have argued for a more inclusive model of crime reporting that considers the role of contextual factors in addition to just the attributes of the crime, which we discuss later). While police recorded hate crime figures are no doubt incomplete, they do consistently show that the bulk of reported hate crimes (67 per cent in 2007/08)9 are cases of harassment, while the minority of cases (a further 12 per cent in 2007/08) involve the victim sustaining an injury. That violence and highly emotive crimes are the exception rather than the rule is further supported by research on perpetrators (Burney and Rose 2002; Ray et al. 2003, 2004). Drawing on this research and their own professional practice, Dixon and Ray (2007) argue that ‘much of what gets labelled “hate crime” is more casual confrontations in which racist (or other hate language) is deployed’, and consequently much of the behaviour is ‘more ambiguous and subject to re-evaluation by the participants who often display adherence to dominant norms of acceptable behaviour’ (2007: 110). In fact ‘mission offenders’, who constitute the greatest potential danger due to their tendency for premeditated and targeted offending, are relatively rare, while the more common ‘thrill seekers’ and ‘reactive offenders’ seldom proceed beyond verbal abuse (Dixon and Court 2003: 151). Gadd et al. (2005) study of racially aggravated offenders in Staffordshire found that racism was rarely the sole motivating factor in their offending behaviour, as was the case with our own interviewing of racially motivated offenders (Wong et al. 2008). This is not to deny the impact that hate offences can have on victims, or that as the typologies of perpetrators automatically dictate levels of risk, prejudices can become more entrenched and perpetrators may gravitate and escalate to ‘mission’ activities. Rather, there can be a 198

Hate crime victims and hate crime reporting

tendency among hate crime scholars to presuppose a ‘one size fits all’ view of victims and among policy-makers a limited understanding of offenders involved in hate crime. What we require is a better policy alignment between the range of experiences and expectations of victims and those of offender motivations. As Dixon and Ray note, the evidence from agencies dealing with racially motivated offending and broader community conflict problems ‘suggests that race is often an issue in violence and antisocial behaviour, but not the issue’ (2007: 110, emphasis in original). Where does this leave the discussion as regards a goal of full reporting? Our own research study (Wong and Christmann 2008) acknowledged that hate crime reporting could be impacted upon by measures such as improving access, improving the service experience and deploying social marketing exercises that actively engage end users. To this end we recommended significantly upgrading the local service to a ‘Client Centred Service Model’. This was an effort to offer more tangible benefits to victims and thereby overcome the expenditure of the transaction costs of notification, be it to the police or another reporting agency. But we conceded that the reporting gains were likely to be only marginal, irrespective of the efforts that agencies made, because victims perceived the crime not to be serious, and therefore by definition it was often not considered worth reporting by our respondents. We also found that providing alternative reporting methods to reporting by telephone or in person, such as via a website,10 and a number of other currently unavailable methods (texting, for example) did not provide any quick solution to under-reporting either. In our study, respondents emphasised the importance of requiring proof that their report had been made and crucially, actioned (Wong and Christmann 2008: 28–9); they therefore proved sceptical of non-traditional reporting methods that did not deliver this.

How did we get here? The Stephen Lawrence Inquiry Report published in 1999 did much to shape the current approach to hate crime reporting. Perhaps for the first time this inquiry acted to expose the nature and challenges in the processing of race hate crime by statutory criminal justice agencies and to stimulate policy change that would increase trust and confidence in policing among ethnic minority communities (Macpherson 1999: 199

Hate Crime

46.31). Prior to the inquiry, racially motivated crimes were only a marginal issue for the police in the context of their overall workloads (Sampson and Phillips 1995), despite the fact that the police were the most significant agency for reporting crime. While there have been long-standing tensions between the police and minority ethnic communities, the inquiry served to highlight how those communities’ lack of confidence in the way that cases were dealt with had led to under-reporting, despite a greater risk of victimisation. The Report drew together some 70 recommendations covering nearly all aspects of policing, and constituted the development of a new strategic approach to racial harassment and hate crime more generally throughout England and Wales (Dixon and Ray 2007). The key recommendations of the Report stipulated a common definition of a racist incident that had wider applicability than racial prejudice alone. There was also an explicit commitment to tackle underreporting through recommending codes of practice to create a more comprehensive system of reporting and recording of racist incidents, and that local criminal justice agencies (in consultation with local communities) should encourage the reporting of racist incidents and crimes, including the ability to report at non-police station locations and the ability to report 24 hours a day. Emphasis was put on the need to allocate sufficient resources to tackle incidents and to support victims: again, partly achieved through the development of common reporting systems and encouraging ‘third party’ reporting centres for groups reluctant to report directly to the police (Wong 2002). Many of these recommendations were adopted by the government of the day and incorporated in the ensuing legislative and policy programmes. The Crime and Disorder Act 1998 and the Race Relations Amendment Act 2000 introduced new statutes on racially aggravated offences and committed community safety units to develop effective multi-agency initiatives to address under-reporting. Additional legislation revised and extended this to include religious hatred, and also required the courts to consider hostility to disability or sexual orientation as an aggravating factoring in sentencing.11 Two government reports assessing the progress made since the Lawrence Inquiry ten years on found that there had been significant improvement in the reporting and recording of racist incidents, although under-reporting was still felt to remain high (Foster et al. 2005), especially for incidents perceived by victims to be ‘less serious’ but which could have a cumulative impact (Docking and Tuffin 2005). The overall message was that more should and could be done to continue to increase reporting through increasing trust 200

Hate crime victims and hate crime reporting

and confidence in the police (2005: 24). We have no quarrel with the sentiment being expressed here, and recognise that these combined legislative and policy changes have resulted in a systematic reconfiguration of the ways in which the criminal justice system now deals with prejudice-related offending, including a greater preparedness to recognise a plurality of hate crime victims. However, we do not believe that the current policy messages will substantially increase the reporting of hate crimes. Consequently the goal of full reporting risks remaining at the status of policy aspiration for the reasons that we have rehearsed. The Macpherson Report reforms were driven by the need to address the glaring deficiencies with how the police responded to racially motivated offending, but moreover, they were and remain a priority partly for symbolic reasons – in reforming the police as an institution and improving the quality of their operational responses. This is what the policies are really about, despite there being only a meagre evidence base for the effectiveness of some of the adopted approaches, such as non-police, third party reporting centres.

Target those most likely to report The point of departure for us was how to affect victim reporting with regard to ‘lower level’ racial or other hate harassment (typically manifested in isolated incidents of name calling/verbal abuse) which is not considered sufficiently serious by victims to report. Moreover, the problems do not stop here, as the victim may not even realise that the incident constitutes a crime (at times requiring a technical interpretation of the law), or alternatively the verbal abuse may not be regarded as sufficiently disturbing because it does not violate the wider community standards thought to justify intervention by community appointed agents, namely the police. What we do know is that if it is perceived as serious, it is more likely to be reported. If low-level incidents recur, this persistence is also more likely to prompt reporting. We found that victims of hate crime make a clear distinction between violent and non-violent crimes and discriminate accordingly (Wong and Christmann 2008: 30–1). It was only with the repeated occurrence of a non-violent, non-serious hate crime that the likelihood of this type of incident being reported increased because the perpetrator would be more identifiable and consequent action likelier to result in effective enforcement action in the minds of respondents. 201

Hate Crime

Figure 9.1 sketches out diagrammatically where we see the most profitable point of intervention, namely those victims who suffer a hate-motivated incident that is sufficiently severe to warrant real consideration of reporting and where there is a realistic potential that this will be carried through. Clearly there are many caveats to note here. Victim characteristics differ widely and therefore they may experience incidents differently. Hate crime also covers a variety of offences across a range of different vulnerable groups. However, the figure is for illustrative purposes only and relates to the two key determinants of reporting: offence severity and persistence. We want to argue that resources should be targeted at the greatest point of impact in encouraging reporting: this ‘ambivalent group’ who may go either way in independently reporting or not (as indicated by the downward arrow in Figure 9.1). We suggest that this can be achieved by focusing social marketing messages at this core audience and their social networks, and by using neighbourhoods as the social unit for hate crime control. In our study we reviewed the social marketing materials such as posters, leaflets and information packs used in the city area. With the exception of some of the information packs, these marketing materials were targeted at an unsegmented market; anyone who may experience any form of hate crime, from violence to verbal abuse. Instead of this broad-brush approach, we propose that materials need to be targeted at individuals who experience crime types that Point of greatest policy impact

Most severe hate offence

Victim considers reporting

Likely to report victimisation

Unlikely to report victimisation

Persistence of hate crime incident

Figure 9.1  Schematic of hate crime reporting. 202

Least severe hate offence

Hate crime victims and hate crime reporting

straddle the boundaries between seriousness and non-seriousness, and those who experience persistent low-level harassment. In addition, messages need to be targeted at individuals who form the social networks of this targeted group of victims. The rationale for focusing on this social network group comes from a consideration of the broader social context of reporting. Describing the results from a 13-year programme on research in the USA on crime victims, Greenburg and Ruback (1992) found that the strongest and most consistent findings concern the role that social influence variables play in the victim’s decision to report or not report the crime to the police. While not looking at hate crime victims in particular, the authors found that crime victims often talk with others before making a decision about reporting, and due to their more emotionally aroused state are more susceptible to the influence of these others. They concluded that the real gatekeeper of the criminal justice system is not the police, or indeed the victim, but rather those people whom the victim consults directly after the crime has occurred. This is the ‘neglected decision-maker’ in the criminal justice system, and one whom information campaigns should be directed at (Greenberg and Ruback 1992: 240–1). All crime is nested within social contexts, and one can move to the neighbourhood level and consider it a social unit for hate crime control. Using intelligence-led policing at the neighbourhood level to try to address low-level hate crime harassment offences would be one potentially profitable approach. Work on identifying vulnerabilities could be undertaken among Neighbourhood Policing Teams (NPTs). Devolving down those more targeted responsibilities would facilitate the identification of individuals and/or geographical areas that are most vulnerable to hate crime. For example, black and minority ethnic staff working in a takeaway, open till late at night next to a pub in an area of high deprivation with a predominantly all-white residential population, are likely to be vulnerable to racial harassment. NPTs already undertake community consultation work, and having the police engage with these more vulnerable communities or their social networks would sell the reporting message to a susceptible audience. An additional, potential outcome of this activity is that it would enhance the reputation of the police among those communities who may have reservations about engaging with them. The role of local authorities in supporting this approach is crucial. Instead of directing their efforts into ‘grandstanding’ (Wong 2009) over the numbers of non-police reporting centres (often within local authority premises) within their areas, their resources would be 203

Hate Crime

better spent working with NPTs and raising the confidence within vulnerable communities to report incidents to the police.

Conclusion We have suggested that the current emphasis on full reporting needs to be revised and assert that agencies need a more intelligent strategy in how they operationalise hate crime generally. Our own findings suggested that broad-brush unsegmented local publicity campaigns are likely to be ineffective. Keeping hate crime as a policing priority needs to be maintained, not least because it has yet to permeate sufficiently high in the ‘hierarchy of police relevance’ (Bowling 1999) across an adequately large number of front-line officers. Chakraborti’s (2009) recent review examining the policing of hate crime questioned whether sufficient progress has been made at the operational level and argued that reforms needed to be sustained. We feel that a more complex narrative is required at the operational and practice level in the form of a tiered approach for hate crime victims, and one that recognises the reality of under-reporting. It is simply unrealistic to expect victims to report low-level harassment motivated by prejudice, no more so than with lower severity volume crime more generally. To concede this much is to accept that there is an unacknowledged deficiency with current policy, but at the same time there is a danger in being seen as backpedalling on the Lawrence agenda, and this risks interrupting police reform and confusing the public. Clearly there is a tension here that needs to be sensitively managed between holding on to realistic policy aims while retaining the direction of policy travel from Lawrence. We want community safety practitioners at the local level to recognise that full reporting is a wasted effort. Rather, they need to refocus their approach and target their limited provision in a more effective way. Having non-police reporting centres (third party reporting) is at best an interim measure while the police upgrade their response to hate crime victims, and at worst a diversion absolving the police from upgrading their response to hate crime victims. It would be more profitable to support police efforts that target and engage those vulnerable communities in an effort to buttress confidence in the police. After all, victims realise that the only agency that has the capability to enforce laws and punish offenders is the police. The veracity of this point should not be underestimated. Our own study found that the predominant minimum agency response expected 204

Hate crime victims and hate crime reporting

across a wide range of hate incidents was the clear expectation to catch and convict the culprit (Wong and Christmann 2008). Perhaps the best way to build confidence in the police is through the effect of cumulative peer recommendations from victims, their families and support networks. The spread of positive rumours and stories recounting how the police responded to hate incidents has the potential to create a virtuous cycle and further growth in the message’s exposure and influence. However, the opposite also holds true: our study found that a previous negative encounter with the police, be it through police disinterest in the report or failing to take any action, acted as a powerful disincentive to report in future. Furthermore, this was the case whether experienced directly or learned secondhand from someone else. Such negative encounters also hardened attitudes, making people resistant to progressive reporting messages (Wong and Christmann 2008:26). While these barriers to reporting need to be addressed in a proactive way, the central issue we have focused on is how to operationally deal with low-level hate crimes in a way that makes the best use of resources. This is one that the police are all too aware of and that the public intuitively realise and accept. We have made some tentative suggestions on some potentially profitable approaches, but remain convinced that significantly increasing reporting for the bulk of hate crime (as manifested as single incidents of harassment) present insurmountable challenges. What is required, then, is an approach that remains aspirational in its policy aim but realistic in terms of operational resources.

Notes 1 As acknowledged in Victim Support’s (2005) literature review of the support needs of victims of hate crime. 2 This refers to comparable volume crime which is a subset of crimes (such as vandalism, burglary, vehicle-related theft, bicycle theft, theft from the person, wounding, robbery, assault with minor injury and assault without injury) rather than all list offences. 3 It should be noted that the term ‘hate crime’ in this context has no legal status in the UK, but rather denotes a series of 11 racially and religiously aggravated offences of violence, harassment or criminal damage. These 11 Home Office list offences comprise: racially or religiously aggravated inflicting GBH; racially or religiously aggravated less serious wounding; racially or religiously aggravated ABH or other injury; racially or religiously aggravated harassment/public fear; 205

Hate Crime

4

5 6 7

8

9 10 11

racially or religiously aggravated harassment; racially or religiously aggravated public fear, alarm or distress; racially or religiously aggravated assault without injury; racially or religiously aggravated criminal damage to a dwelling; racially or religiously aggravated criminal damage to a building other than a dwelling; racially or religiously aggravated criminal damage to a vehicle; and racially or religiously aggravated other criminal damage. Although there are additional reasons that explain this discrepancy between BCS and police recorded crime, namely trends in reporting rates and police recording practice as well as variation within the BCS sample (Home Office 2009: 26). For instance, the ‘Talk to Us’ campaign launched by Camden Council and Camden Police and a range of community and support groups is one of many. The Home Office’s annual reports on Crime in England and Wales do not provide figures for why BCS victims did not report incidents of ‘harassment’. Category includes: something that happens as part of job; partly my/ friend’s/relative’s fault; offender not responsible for actions; thought someone else had reported incident/similar incidents; tried to report but was not able to contact the police/police not interested; other (Home Office 2009: 26). Unfortunately the BCS merges these different reasons due to the perceived similarity for respondents, so we are unable to know whether it is the victim’s subjective feelings concerning the crime or alternatively their assessments of the police response that results in non-reporting – this is especially unhelpful when trying to understand victim decisionmaking. Figures were not available for 2008/09 at time of writing due to a quality assurance inspection by Her Majesty’s Inspectorate of Constabulary of some categories of racially and religiously motivated offences. The Police operate a ‘True Vision’ website which allows victims to report hate incidents online. The Anti-Terrorism, Crime and Security Act 2001, the Racial and Religious Hatred Act 2006, the Criminal Justice Act 2003 and the Criminal Justice (No. 2) Northern Ireland Order 2004.

References Bouten, E., Goudriaan, H. and Nieuwbeerta, P. (2002) ‘Criminal Victimization in Seventeen Industrialised Countries’, in P. Nieuwbeerta (ed.) Crime Victimization in Comparative Perspective: Results from the International Crime Victims Survey, 1989–2000. The Hague: Boom Juridische Uitgevers (pp. 13–28). 206

Hate crime victims and hate crime reporting

Bowling, B. (1999) Violent Racism: Victimisation, Policing and Social Context. Oxford: Oxford University Press. Burney, E. and Rose, G. (2002) Racist Offences – How is the Law Working? The Implementation of the Legislation on Racially Aggravated Offences in the Crime and Disorder Act 1998, Home Office Research Study 244. London: Home Office. Carcach, C. (1997) ‘Reporting Crime to the Police’, Trends and Issues in Crime and Criminal Justice, Australian Institute of Criminology, No. 68, March. Chakraborti, N. (2009) ‘A Glass Half Full? Assessing Progress in the Policing of Hate Crime’, Policing: A Journal of Policy and Practice, 3(2): 121–8. Dixon, L. and Court, D. (2003) ‘Developing Good Practice with Racially Motivated Offenders’, Probation Journal, 50(2): 149–53. Dixon, L. and Ray, L. (2007) ‘Current Issues and Developments in Race Hate Crime’, Probation Journal, 54(2): 109–24. Docking, M. and Tuffin, R. (2005) Racist Incidents: Progress since the Lawrence Inquiry, Home Office Online Report 42/05. London: Home Office. Fishman, G. (1979) ‘Patterns of Victimisation and Notification’, British Journal of Criminology, 19(2): 146–57. Foster, J., Newburn, T. and Souhami, A. (2005) Assessing the Impact of the Stephen Lawrence Inquiry, Home Office Research Study 294. London: Home Office. Gadd, D., Dixon, B. and Jefferson, T. (2005) Why Do They Do It? Racial Harassment in North Staffordshire. Staffordshire: Keele University. Goudriaan H., Lynch, J. P. and Nieuwbeerta P. (2004) ‘Reporting to the Police in Western Nations: A Theoretical Analysis of the Effects of Social Context’, Justice Quarterly, 21(4), December. Greenberg, M. S. and Ruback, R. B. (1992) After the Crime: Victim DecisionMaking, Perspectives in Law and Psychology. New York: Plenum Press. Home Office (2009) Crime in England and Wales 2008/09, Home Office Statistical Bulletin 11/09. London: Home Office. Macpherson, W. (1999) The Stephen Lawrence Inquiry, Cm 4262. London: The Stationery Office. Ray, L., Smith, D. and Wastell, L. (2003) ‘Understanding Racist Violence’, in E. Stanko (ed.) The Meanings of Violence. London: Routledge (pp. 112–29). Ray, L., Smith, D. and Wastell, L. (2004) ‘Shame, Rage and Racist Violence’, British Journal of Criminology, 44(3): 350–68. Sampson, A. and Phillips, C. (1995) Reducing Repeat Racial Victimisation on an East London Estate, Crime Detection and Prevention Series: Paper 67. London: Home Office Police Research Group. Skogan, W. G. (1976) ‘Citizens Reporting of Crimes: Some National Panel Data’, Criminology, 13: 535–49. Skogan, W. G. (1984) ‘Reporting Crimes to the Police: The Status of World Research’, Journal of Research in Crime and Delinquency, 21(2): 113–37. Sparks, R., Genn, H. and Dodd, D. (1977) Surveying Victims. Chichester: John Wiley and Sons. 207

Hate Crime

Victim Support (2005) The Impact of Hate Crime: Needs, Practical and Emotional Support for Victims: A Literature Review. London: Victim Support. Wong, K. (2002) ‘The (Re)emergence of Hate Crime as a Policy Issue’, Criminal Justice Matters, 48: 34–5. Wong, K (2009) ‘Do the Right Thing: The Stephen Lawrence Inquiry Report 10 Years On’, Safer Communities, 8(3): 10–13. Wong, K. and Christmann, K. (2008) ‘The Role of Decision-Making in Reporting Hate Crimes’, Community Safety Journal, 7(2) 19–34. Wong, K., Christmann, K., Wilcox, A., Smithson, H. and Monchuk, L. (2008) ‘The Prevalence of Racially Motivated Offending and the Provision of Targeted Interventions’, unpublished Final Report to the Youth Justice Board.

208

Chapter 10

Racial aggravation or aggravating racism: overcoming the disjunction between legal and subjective realities David Gadd1 Introduction When, how and whether racism motivates offending is rarely easy to discern. The aim of this chapter is to introduce readers to the definitional challenges the notion of ‘racially aggravated crime’ presents policy-makers, practitioners and academics with. Using a case study to illuminate, the chapter illustrates what is to be gained by contrasting social-legal, sociological and psychosocial perspectives on the same phenomenon, and makes a case for legal reform that presents greater opportunities for offenders to come to understand in which ways their motives might be perceived by others to have been racially aggravated. The chapter begins with a short account of the research on offenders’ motives and the legislative remedies the UK has implemented to tackle crimes deemed ‘racially aggravated’.

The advent of racially aggravated crime Often regarded as the first study in the UK to put racial harassment on the government’s agenda, the Home Office’s (1981) report Racist Attacks anticipated the difficulties of defining racial motivation legislators as follows: … a racial incident is one that is in some sense motivated by racial hatred or antipathy. Motives, however, are not open to 209

Hate Crime

direct inspection but have to be inferred from the circumstances of the incident or offence. Inferences of this kind may call for difficult and highly subjective judgments … Ideally, the only reliable source of information on racial motivation would be from the offender. (Home Office 1981: 7) Nearly thirty years later we are little closer to understanding offenders’ ‘motives’ and the interpretive challenges that making sense of ‘highly subjective judgments’ about them entail. This is because the bulk of criminological work on racist crime tends to be engaged only with second-hand accounts of what offenders have said and done; where offenders are interviewed, it is inclined to conflate the motives of particular offenders with the cultural milieu of resentment felt by marginalised white working-class communities. Meanwhile, the criminal justice reforms implemented following the Stephen Lawrence Inquiry have made getting to grips with offender motivation all the more necessary. Police services in the UK now operate with a highly subjectivist definition of a racist incident – any incident that ‘is perceived to be racist by the victim or any other person’ – while the advent of hate crime law has made the courts responsible for not only assessing whether defendants are guilty of the offences of which they are accused but also for coming to some conclusion as to why they also committed them. Since the implementation of the 1998 Crime and Disorder Act, the police in England and Wales have been able to charge offenders with ‘racially aggravated’ offences. Sections 29–32 of the Crime and Disorder Act permitted courts to enhance the sentences of those convicted of racially aggravated offences beyond the tariffs imposed for their nonracially aggravated equivalents. While the legal picture is complicated by the precedents set by case law and by the introduction into the statute book of ‘religiously aggravated offences’, racial aggravation is still regarded as in evidence if (a) at the time of committing the offence, or immediately before or after doing so, the offender demonstrates towards the victim of the offence hostility based on the victim’s membership (or presumed membership) of a racial group; or (b) the offence is motivated (wholly or partly) by hostility towards members of a racial group based on their membership of that group. Consequently, the courts have been able to convict defendants of racially aggravated offences in cases where racism was not regarded by the prosecution as the primary motive: that is, where the offence was only ‘partly’ motivated by hostility towards members of a racial group. In such cases, evidence of a defendant’s use of any ethnic 210

Racial aggravation or aggravating racism

referents at the time of the offence can be sufficient enough proof that the offence was ‘racially aggravated’ (Dixon and Gadd 2006).

How the law ‘works’ The complications this has caused are similar to those experienced in many US jurisdictions. There, the defendant’s ‘bias’ against an outgroup can be evidence enough of the ‘animus’ required to secure a conviction for ‘hate crime’: an outcome some socio-legal scholars have deemed socially divisive and unconstitutional (Jacobs 2002). In Britain, by contrast, criminologists have remarked little on this subject, with the exception of Iganski (2008: 94) who defends the 1998 Crime and Disorder Act as a ‘radical intervention … designed to promote justice by attempting to mould the collective conscience’. Demonstrating that the public are even minimally aware of the law in this regard, let alone influenced by it, is, however, rather difficult, not least because it remains unclear whether the legislative changes have had any impact in terms of diminishing the prevalence of racial harassment. Since the implementation of the Crime and Disorder Act there have been sharp increases in the rates at which racially aggravated offences are reported to the police and prosecuted by the courts. Nevertheless, the majority of racist incidents still do not culminate in anybody being convicted of a racially aggravated offence. While 184,000 racially motivated incidents were experienced by the populations represented in the British Crime Survey in 2006/07 – this in itself probably a gross underestimate, given the survey’s use of an ethnic typology into which it is increasingly difficult to fit the asylum-seeking and refugee populations living in Britain – the police in England and Wales recorded just 42,600 racially or religiously aggravated offences. Moreover, only 38 per cent of these offences recorded by the police were ‘cleared up’ (Ministry of Justice 2008: 15), and not all those cases that were cleared up by the police made it to court. Consequently, in 2006 only 11,500 charges of racially aggravated offending were prosecuted in England and Wales and only 5,166 defendants were convicted for racially aggravated offences (2008: 19). The net effect of this steady attrition makes for a ratio of racist incidents to racially aggravated offenders convicted higher than 35 to 1. When one considers just how many incidents involve multiple perpetrators this attrition rate looks steeper still. 211

Hate Crime

Of course, there are many reasons for this attrition. Many victims do not want to report being harassed. Some perceive their victimisation as too trivial, not really criminal, and/or think the police will not take it seriously. Some are too afraid to come forward, unwilling to give evidence in court, or see racial harassment as too depressingly common for anyone to do anything about. In many cases the perpetrators cannot be identified with sufficient certainty, or the police take the view that there is insufficient evidence that a crime has been committed. But post-1998 there is no avoiding the need to make sense of offenders’ motives, something that has proved a stumbling block that the courts in England and Wales remain poorly equipped to overcome (Newman 2008). As Burney and Rose’s (2002) analysis of how the law on racially aggravated crime is working demonstrated, the courts have struggled to respond in a consistent manner to the high volume of cases brought ‘where the racist element is ancillary to the substantive offence, rather than the cause’ (2002: x). What constitutes the ‘hostility’ needed to prove ‘racial aggravation’ is a source of contention for legal practitioners. In some local magistrates’ courts, for example, there is a current of opinion, most often expressed by stipendiary magistrates and defence solicitors, that the law came down rather hard on people who, in the course of ‘normal working class mayhem’ as one person put it, uttered words which were part of their natural vocabulary. (2002: 20) Meanwhile, other sentencers have taken the view that ‘any reference to the victim’s ethnicity in connection with an offence’ proves ‘the necessary element of racial aggravation’ (2002: 114). For them, ‘anything less condones racism’ (2002: 114). Inevitably, defendants have tended to contest this inference, often proving willing to admit to the basic (non-racially aggravated) charges, but unwilling to plead ‘guilty’ to their racially aggravated equivalents. In these instances, ‘not-guilty’ pleas expose defendants to the possibility of harsher sentencing, trials being handled by the Crown Court which has greater sentencing powers than the magistrates’ courts. Burney and Rose became convinced that some defendants take this risk because they are genuinely upset and indignant at the prospect of a ‘racist’ tag … it is clearly socially unacceptable (except in a few circles) to be branded as a racist. A racially aggravated conviction is seen 212

Racial aggravation or aggravating racism

as a shaming event in a different class to a conviction for a basic offence. (2002: 91)

Elusive motives As the practice-based literature attests, these complications are rarely resolved when convicted ‘racially motivated offenders’ serve out their sentences. In prisons there are few if any interventions that explicitly address the racism of racially motivated offenders, and as the inquiry into the murder of Zahid Mubarek attested, the witting and unwitting racism of prison officers has in some institutions only enhanced the dangerousness hate crime perpetrators pose to minority ethnic people also remanded in custody (Keith 2006). In probation, the picture is more equivocal, but it remains the case that only a minority of services have developed provision that explicitly addresses the racism of racially motivated offenders (Isal 2004; Hankinson 2008). The Probation Inspectorate is not convinced that anything better than ‘limited’ evidence of ‘positive change in attitudes, beliefs and behaviour in relation to racially motivated offending’ can be achieved through this provision (HMIP 2005: 32). The elusive quality of racist motivation presents probation officers with enormous challenges (Ray et al. 2003). Because they know that racism will not be tolerated by the probation service, and because they know that being known as ‘a racist’ is morally stigmatising, perhaps even more so than being known as a violent offender or thief, many probationers are reluctant to talk about issues to do with race with the workers assigned to them. This has meant that racism is often inadequately tackled in intervention work; some probation officers avoid the subject completely, while others perceive only the small minority of offenders who are overtly interested in the far right as ‘real’ racists. Much everyday prejudice goes unremarked upon because practitioners are afraid of compromising a viable working relationship with a typically unwilling client group troubled by complex social and psychological difficulties (Dixon 2002; HMIP 2005: 27; Smith 2006a). Both because of the difficulties entailed in confronting racism overtly and because the number of convicted racially motivated offenders attending each probation office at any given time is always small, the probation service has started to look away from specialist accredited racially motivated offender programmes and towards developing the kind of one-to-one work the majority of offenders 213

Hate Crime

benefit from (Smith 2006a; McGhee 2007). How exactly racism should be tackled in one-to-one work remains open to debate; however, many commentators identify offender needs that cannot be reduced to enhancing cognitive, social and/or employability skills. Lemos (2005: 34), for example, underlines the importance of tackling the offender’s dependence on ‘unconscious racist stereotypes’. He observes that hate crime perpetrators are prone to ‘winding’ themselves up psychologically and are susceptible to a – perhaps quite justifiable – feeling that that no one ever listens to them (2005: 38). Court (2003) argues that practitioners need to develop the skills and confidence required to open up discussions with racist offenders, to get them to disclose earlier incidents, and explore the contradictions between their own accounts and what victims have alleged about them. Because attending to ‘subconscious’, ‘dynamic’ and ‘emotional factors’ is so critical, we need more ‘informed’ interventions ‘likely to uncover and effectively address the complex psychological processes underlying this form of aggravated offending’ (Dixon and Court 2003: 150). Smith (2006b) argues similarly that it is crucial to attend to racism’s ‘appeal’, the ‘unacknowledged shame’ that often lies behind it, and the ‘defences’ of ‘denial’ and ‘repression’ that protect the individual from the painful humiliation such acknowledgement can entail. For him, ‘effective work with racist offenders will need to be based in a recognition that racism often emerges from feelings not of power and superiority … but from a sense of weakness and subordination’ (2006b: 28). McGhee (2007) pushes the point further. Having felt themselves typecast in the course of the prosecution process, many convicted offenders need to hear an acknowledgement that not all of those deemed to be racists are racist most of the time. The majority of perpetrators understand their offending behaviour as an extraordinary departure occurring in the context of a ‘highly individualised “trigger” situation’ (2007: 221) and thus expect any intervention work to begin by recognising this. According to McGhee, practitioners must work with a model of the offender that assumes not an unthinkingly prejudiced personality, but an individual with ‘inconsistent, contradictory and highly ambiguous’ attitudes towards race (2007: 218).

The contradictorily racist subject In many respects, this notion of the contradictorily racist subject was the conceptual foundation upon which Adorno et al.’s (1950) classic 214

Racial aggravation or aggravating racism

analysis of anti-semitism was built. It has also been the starting point for a number of insightful, if too often overlooked, psychoanalytically informed studies of racism since then (Sherwood 1980; Young-Bruehl 1996; Dalal 2002). While few commentators have realised it, the criminological evidence points in a similar direction, at least if one does proper justice to the data criminological researchers have collated. Sibbitt (1997), for example, provided numerous illustrations of the contradictory ways in which racism is interwoven into the social and psychodynamics of community and family life: people worried about their own educational ability, financial security, physical health and mental well-being, projecting their anxieties on to all those deemed ‘foreign’, while at the same time espousing tolerance and exempting their black neighbours from all such criticism. To the extent that they feel they have to keep re-accomplishing their identities through the perpetration of hate crime, the typically young, educationally challenged, economically disadvantaged young white men who are the focus of ‘structured action theory’ are self-evidently beset with insecurities about their own whiteness, masculinity and/or sexual virility (Messerschmidt 1997; Perry 2001). And one does not have to look very far in the work of Ray et al. (2004) to find a depiction of the racially motivated offender – albeit something of a composite of many different individuals – whose personality is layered with selfdeceptions, spliced with grudges, and who is unable to face up to a catalogue of painful home truths: [T]hey saw themselves as weak, disregarded, overlooked, unfairly treated, victimised without being recognised as victims, made to feel small; meanwhile, the other – their Asian victim … was experienced as powerful, in control, laughing, successful, ‘arrogant’. (2004: 355) But because all this work lacks an adequately theorised account of the relationship between the outward hostility articulated in racist attacks and the inner world insecurity, powerlessness and disregard many perpetrators feel, most criminological commentary on this topic struggles to provide an analysis of motivation that is anything other than ‘middle range’. This is a serious shortcoming, because as McGhee argues this space, created by the contradictions in the racialised discourses adopted by offenders, is often the only room available to manoeuvre offenders into a mindset within which they can contemplate the prospect of relinquishing their dependency on prejudice and hate. 215

Hate Crime

The revisionist psychoanalytic work, informed by Kleinian and relational insights, that has cultivated a psychosocial approach to racism freed of the determinism that marred the work of Adorno et al., has much to offer in this regard. Take, for example, the argument, expressed most cogently by Mike Rustin, that ‘race’ is an ‘empty container’ that provides an ideal receptacle for the ‘mechanisms of psychotic thought’. These mechanisms include the paranoid splitting of objects into the loved and hated, the suffusion of thinking processes with intense, unrecognized emotion, confusion between self and object due to the splitting of the self and massive projective identification, and hatred of truth and reality. (Rustin 2000: 187) Here we have the opening to a theory that is exceptionally versatile. It provides an account of how racial difference comes to symbolise the things we as a society collectively choose to disown – as evidenced in the collective cultural racism that blames immigrants for taking our jobs, for example – and the idiosyncratic, sometimes bizarrely sexual, meanings individual racists invest in racism as they attempt to keep out of mind thoughts and feelings infused with ‘intense but unrecognised emotion’ (2000: 187). It provides an account of why racism comes to the fore at times of financial and social crisis, but also why those undergoing personal crises, grappling with loss, or suffering poor mental or physical health might find racism especially alluring. And through the concept of projective identification – the process through which disowned parts of the self are psychically attributed to the attacked other – we have an explanation as to why it is that racists often feel menaced by their victims, subject to malicious accusations by them, and rendered naked in all their shameful failings by the very presence of the person they have attacked (Clarke 2003). Adopting this kind of psychosocial approach can also help understand why confrontational work with racially motivated offenders is likely to be counter-productive. Quite plainly, paranoia fuels racism, so persecuting the racist tends only to compound the problem. Because so much is kept unconscious by splitting and projection, the racist has to learn to accept through psychic and embodied experiences – not just through rational argument – that hostility is conveyed by their words and deeds. In order to come to terms with their own hate for the parts of themselves they unconsciously disown through splitting 216

Racial aggravation or aggravating racism

and projection, the racist must to some extent submit themselves to an intersubjective engagement with someone willing and able to both represent the attacked other and withstand and survive the attacker’s hostility (Benjamin 1998; Gobodo-Madikizela 2003).

The case of Alan The case analysis that follows is elicited from an ESRC research project conducted in north Staffordshire, a region which once hosted substantial pot, pit and steel industries, but is now home to a pre­dominantly white working-class population that has expressed growing levels of support for the British National Party over the past decade, and in which a small, long-established south Asian community has been blamed, criticised and attacked for a host of social problems erroneously attributed to the arrival of asylum seekers and refugees. During this research I – together with Bill Dixon and Tony Jefferson – conducted focus groups with 13 naturally occurring groups of people living in this locality. These included people from a residents’ association, a neighbourhood watch, a working men’s club, a day centre and two anti-racist groups, as well as young offenders, asylum seekers, and white and minority ethnic users of two youth centres. I also interviewed 15 people implicated in acts of racial harassment using the Free Association Narrative Interview Method designed by Hollway and Jefferson (2000). The pen portrait of Alan that follows is derived from the two interviews I conducted with him using this method. Alan was a 39-year-old man on a 12-month probation order for the racially aggravated assault of an Asian taxi driver. This was an event Alan claimed not to be able to recall because he was mentally ill at the time of the alleged offence. Alan was prepared to accept that he might have spat in the face of the taxi driver in the ‘heat of the moment’: ‘I’m terrible for it. I think it’s like the ultimate insult to somebody’. This was why he had pleaded guilty to common assault. However, Alan did not believe that he was guilty of the racially aggravated element of his conviction. He did not consider himself to be racist and neither did his probation officer. (As an aside, neither did I back in 2005 when I first attempted to make sense of this case, there being little in Alan’s interview responses to suggest particularly racist attitudes and much in his words and behaviour to suggest that he was someone who could be incredibly open and caring in the company of strangers.) Alan could not ever imagine using the words he was alleged to have used – ‘Paki bastard’. 217

Hate Crime

Alan’s disavowal of the racial dimension to the assault accorded well with the stories he told depicting his non-racist beliefs. When, as a schoolboy, he worked on the markets, he ‘got on like a house on fire’ with the ‘Indian or Pakistani’ traders. As a teenage football hooligan, he remembered two ‘coloured’ guys in their gang and ‘they never got no racist abuse of us’. He used to wear a ‘Rock Against Racism’ badge and would take issue with younger hooligans who purported to have connections with militant far right groups. Nowadays, he is appalled by post-9/11 Islamophobia. This kind of racism had, in his view, done ‘a lot of damage to the Muslim community … which they don’t deserve’. Alan even claimed to be ‘good friends’ with the other Asian taxi drivers who worked for the ‘same firm’ as the man he had assaulted. He had persuaded one to come with him to a ‘Stoke away game and introduced him’ to all of his ‘friends down the match’. Three of his niece’s children, all of whom he loves ‘to bits’, have a father who is Pakistani (and also a taxi driver). While there had been some family animosity over this niece’s choice of husband, it was, Alan claimed, ‘nothing racial’. The animosity, Alan said, had more to do with ‘how they treat their women’ and the strong suspicion that the father, with whom ‘everybody gets on’, has a second family in Pakistan. As for the taxi driver who had been assaulted, Alan had hired him on two previous occasions before the alleged assault. On both of those two previous occasions the taxi driver, according to Alan, had made sexually derogatory remarks about women. The taxi driver’s offensive remarks on the first occasion were with reference to two young girls walking past. In his first account of this journey, Alan claimed the taxi driver had said: ‘I wouldn’t mind giving it her up the arse’, to which Alan said he reacted by saying: ‘I’m not like that. I don’t enjoy people speaking like that about women.’ In the second interview, however, Alan recalled the taxi driver’s remarks as ‘I bet you like having it up her arse, mate’ to which Alan said he had replied: ‘I’m not like that … I don’t enjoy people speaking like that about women … You don’t know me.’ Alan had ‘had a few to drink’ and was returning from a night out, at the time of the second taxi journey, when the driver allegedly asked him: ‘Where you been mate? You been fucking?’ Alan said he replied: ‘No offence mate, but this is twice you’ve made derogatory offence like remarks about women and now you’re asking me if I been shagging …You’re nothing but a pervert.’ For this reason, Alan thinks that on the third taxi journey, when he spat on the driver, he would have called him a ‘pervert’ not a ‘Paki’: ‘because that’s what I genuinely thought of him’. 218

Racial aggravation or aggravating racism

Asked whether he could recall other times when he felt offended by ‘sexist or sexually derogatory comments’, Alan immediately remembered a boy at school with ‘dark-looking skin’ that ‘put’ him ‘ill at ease’. The same boy, after a history of rape charges and a fouryear jail sentence for ‘a string of attacks’, is now serving a life sentence in prison for ‘something like 16 or 17 rapes’. Alan knew some of the victims and said he would ‘joyfully’ act as the rapist’s executioner. Another instance when questions of sexuality were touched upon by Alan arose in relation to his explanation of the ‘male bonding’ dimension of the football hooliganism he had found so ‘addictive’ since his teens. ‘It was’, he said, ‘always nice if … you was on the coach and you were a bit scared because you were young … and some older lad would say, “Stick with us kid … You’ll be alright.” It’s like a … a male bonding type of thing. I’m not homosexual, or nothing like that.’ Perhaps predictably, this male bonding often involved the bloodiest of violence, Alan’s fights with rival fans and police officers exposing him to some serious injuries, numerous bans from the club, and the ‘short, sharp, shock’ of detention centre in his late teens. It also involved overt sexism not all that different from that expressed by the taxi driver whose insinuations had turned Alan ‘cold’. ‘It’s something different when ten thousand lads are singing to a pretty model parading round the pitch, “Get your tits out for the lads.” That’s a bit of a laugh.’ The bravado this male-bonding entailed contrasted sharply with the pain Alan felt in relation to his own relationships with women: ‘I wear me heart on me sleeve. I throw me whole self in and me whole self out.’ His first relationship coincided with his time in detention centre, his girlfriend’s ultimatum on his release – football or her – persuading Alan to give up going to football matches for a year and a half. This same relationship culminated in an engagement, but when Alan’s fiancée found out that he had ‘two-timed’ her, she broke it off. Alan was devastated, took an ‘overdose’, ‘slit his wrist’, and became ‘really, really down depressed’: ‘everything really went shit up’. As Alan put it, ‘I’ve never been very good at handling relationship, er problems … I’m not very emotionally strong’. Thereafter, alcohol continued to be ‘a big problem’ in Alan’s life, as did cannabis use and the hooliganism he drifted back into, even if he did not always perceive them as such. In his late twenties, Alan married a woman who he met when she began comforting him after he had been knocked unconscious in a pub fight. While 219

Hate Crime

their relationship was violent – after ‘a few weeks she hit me with a glass’ – once they had married Alan ‘packed the hooliganism’ in and ‘kept out of trouble as much as humanly possible’. Having ‘vastly underachieved’ at school, Alan had spent much of his life ‘drifting from job to job’. After he married, however, his employment situation also improved, Alan finding the motivation to go to college, qualify as a painter and decorator, and become self-employed. For a few years, during the ‘boom time’ of the late 1980s/early 1990s he was ‘doing alright’. But then: There was a sort of depression again and … I was struggling … You never knew when your money was coming and I don’t know. I think that was when the depression started hitting hard like … There was nothing right about your life, as I thought, and that carried on with me right through me early years of marriage … This emotional downturn was not, however, simply about the pressures of a working man with a family of three children to feed struggling to bring home a decent wage in the midst of an economic downturn. Having, he claimed, endured much violence from his wife over the years, when Alan discovered her having an affair with their next-door neighbour he punched her on the shoulder, hard enough to put ‘her arm in a sling for a few weeks’, and ‘spat in her face’. After a period of terminal illness, Alan’s father died of kidney failure that same week. ‘Cut up’ with grief, Alan ‘went really down hill’ as he and his ex-wife battled over access to the children. He became ‘uppity all the time … couldn’t rest’. He was ‘not eating properly’, started ‘drinking heavily’ and began ‘trawling’ the pubs and clubs ‘looking for another partnership’. Alan’s mind was filled with ‘overwhelming thoughts’ that he ‘was dying’. He also started hearing voices but kept this to himself. This stopped after he saw a psychiatrist who prescribed medication, but this was only after the assault on the taxi driver, which Alan believes occurred during a ‘psychotic bout’. Now getting on better with his ex-wife, Alan thinks that after all the years of fighting they have ‘got no more hate to show each other’. But Alan fears his mental health problems have not gone away: ‘It terrifies me that I’ve had two bouts of it. Maybe if I have another one I might never come out of it.’ Meanwhile, he has kept his conviction for racially aggravated assault a secret, shared only with those who know him well: his ex-wife, his children, his ‘closest brother’, and his Asian friend who works for the same taxi firm as the victim. 220

Racial aggravation or aggravating racism

Reading racial motivation On the basis of this evidence, there would appear to be at least three ways in which Alan’s story can be read. The first way, let us for the sake of simplicity call it a socio-legal way, of reading this story fits with what Burney and Rose’s (2002) research reveals about the perspectives of people prosecuted for racially aggravated offences. It was also the way preferred by Alan and his probation officer; a straightforward refutation of the racially aggravated element of the charges brought. Racism was neither intended nor in evidence. Alan was not overtly racist and was certainly much less openly prejudiced than many people living in and around his neighbourhood. When the charges were prosecuted it was the victim’s word against Alan’s. And as a ‘fully fledged hooligan’ who could ill afford a decent solicitor, Alan was going to have a tough time convincing a jury that his testimony was the more plausible of the two, especially given his inability to remember what actually happened. From this perspective, Alan was the unfortunate victim of both prejudice and ill-conceived legal reform. A second, more sociological way of reading this story, however, is to interrogate it through the wider cultural meaning frame in which it is embedded. This entails attending to the contrast between the chauvinism exhibited by the football hooligans Alan was involved with and the sexually explicit remarks of the Asian taxi driver he regarded as so unacceptable. For inspiration in this regard, we need not look much further than the seminal works of structured action theorists like Messerschmidt (1997) and Perry (2001). Underachieving at school, unsuccessful in work, and unlucky in love, conceived through the lens of structured action theory, one can construe the assault Alan perpetrated as involving both the ‘exercise of power’ that was otherwise ‘unavailable’ to him and the ‘maintenance of power in relation to subordinate racial, ethnic’ and ‘gender groups’ in a structurally racist, classist and sexist society (Perry 2001: 38). Not unlike the southern US white men involved in lynching, who at the twentieth century’s inception felt so threatened by the ‘interracial sexuality’ they had themselves once inflicted on black slave women, Alan, the ‘fully fledged’ football hooligan positioned himself as the chivalric protector of white women: women whom he saved from the insidious sexual aggression of men unlike him, those who did not show women ‘total respect’, sexual ‘perverts’ like the Asian taxi driver, and men from Pakistan, like his niece’s husband, who treat women as ‘second class citizens’. 221

Hate Crime

Using structured action theory one can read into Alan’s behaviour evidence of difference being constructed in ‘negative relational terms’, the salience of ‘multiple structures of domination’ that routinely privilege white heterosexual masculinity even when its authority is contested (Perry 2001: 47). From this sociological perspective, perpetrators recreate their own masculinity, or whiteness, while punishing victims for their deviant identity performance (2001: 55). Perhaps the taxi driver was being punished by Alan for both ‘transcending normative conceptions’ of what Asian men are entitled to be like and for ‘conforming’ to a certain stereotype of Asian masculinity (2001: 55, emphasis in original). From a third, psychosocial vantage point, however, both of the aforementioned explanations are unsatisfactory, even though they both add crucial explanatory value. The socio-legal explanation fails to satisfy because it still leaves the question of why Alan spat on the taxi driver under-theorised. The sociological explanation falls short of persuading because one of the necessary components of structured action theory was not in evidence. There was no ‘social audience’ that became more convinced of the purity of Alan’s whiteness or the potency of his masculinity in the course of the assault. There were no witnesses at the time of the offence and Alan expressed no sense of pride about perpetrating the assault. It is therefore hard to see that masculinity or whiteness were really accomplished by it. That Alan felt it likely that he had spat on the taxi driver is revelatory in this context, for we know that Alan is not only ‘terrible’ for spitting, but that he also spat on his wife – whom he used to ‘hate’ – when he found her in the arms of the man who lived next door. For Alan then, spitting might well have been a measure of intense feelings of disgust, anger, humiliation, self-doubt and betrayal, and this may be why he considered it ‘the ultimate insult’. Placed in this context, the contradictions in Alan’s account of his first taxi ride are all the more intriguing. In Alan’s first account the taxi driver allegedly said what he would like to do to the two young women they were driving past (‘I wouldn’t mind giving it her up the arse’). In the second account, however, Alan remembered this differently, in terms of what the taxi driver was implying about him (‘I bet you like having it up her arse, mate’). We do not know enough about this taxi driver to be able to understand why he used such crude and sexually explicit language. That said, it is easy to imagine why, when faced with the prospect of driving around a man whom he knew to be a ‘fully fledged hooligan’, a man who had recently taken one of his co-workers to a football match to experience the behaviour 222

Racial aggravation or aggravating racism

of the crowd in all its aggressive and sexual explicitness, and whose presence, especially when drunk or mentally ill, might seem more than a little menacing, an Asian taxi driver might be tempted to break the ice by indulging in the crude chauvinism he suspected Alan also subscribed to. Alan, as we now know, certainly interpreted the taxi driver’s words as about him, and may well have assumed this taxi driver already knew some things about him, knowing, or knowing someone who knew, Alan’s niece’s husband, the Pakistani taxi driver, about whom Alan and his family had their own strong suspicions. Could this be why Alan countered the taxi driver’s crude innuendos with the assertion, ‘You don’t know me’? What was there for the taxi driver to know about Alan? Given that on the third occasion he was picking Alan up from a night of heavy drinking, it seems more than likely that Alan, a recently divorced man who by his own admission is ‘not very emotionally strong’, who had a long history of falling foul of depression and suicidal thoughts whenever his relationships became problematic, was returning from a night of ‘trawling’ the pubs and clubs ‘looking for another partnership’ he had failed to find. This may have been humiliating enough for Alan the taxi driver’s innuendos being interpreted as referring to Alan’s desperation for any kind of sexual relationship with any woman. Moreover, we know that Alan was also sensitive to the homosocial nature of male bonding he had spent a lifetime indulging through his hooligan activities, a sensitivity that may well have become more acute when he was not in a committed heterosexual relationship. He could well have read the taxi driver’s reference to his purported interest in anal sex as implying both desperation and homosexual inclinations, hence his unnecessarily defensive assertion that he ‘was not a homosexual’. Following Rustin (2000: 87), this defensiveness should alert us to the unravelling of the psychotic-like ‘paranoid splitting’, ‘suffusion of thinking processes with intense, unrecognised emotion’, ‘confusion between self and object’, and ‘massive projective identification, and hatred of truth and reality’ that often characterises acts we take to be motivated by racism. Within this framework one can see how the chauvinist conversation-making of an Asian taxi driver was enough to elicit the spiralling hostility of a mentally fragile, if physically daunting, man like Alan: calling to mind unconscious doubts about his own emotional and sexual adequacy as a man; causing him to perceive the taxi driver in racialised terms akin to those he associated with his niece’s (Asian, taxi-driving) husband and a ‘dark-skinned’ classmate whom he knew to have been convicted of numerous sexual 223

Hate Crime

assaults, both of whom he half-knew things about, just as they may have half-known things about him; making him fear that this taxi driver too also knew something secret about him, the dynamics of projective identification evoking a self-disgust that was physically projected out and disowned through spitting perhaps because it felt too threatening to Alan for him to articulate verbally.

Conclusion As criminologists we need to move away from one-dimensional models of the racially motivated offender. Instead we must attempt to understand the contradictory and changing inner worlds such perpetrators occupy and the different ways in which these worlds are understood by those working within criminal justice settings. Working with three models of the racist subject – socio-legal, sociological, psychosocial – what I have tried to show in this chapter is that it is possible to understand how offenders whose racism is hardly evident in the relatively controlled surroundings of a probation office can appear so hateful to their victims in the ‘heat’ of a confrontation. Such transformations help explain why those convicted of racially aggravated offences often feel unjustifiably condemned, without having to deny that race matters in the perpetration of these offences, and without having to foreclose the question of how, why and in what ways race becomes salient in conflictual circumstances. Of course, not all racially aggravated crimes require such a complex explanation as the case of Alan. But I hope that the analysis this chapter offers persuades more sceptical readers that advocating for the benefits of a more dynamic psychosocial approach – an approach that helps us understand why offenders behave as they do – need not necessarily preclude the kinds of critical analyses that highlight the wider ramifications of racist attacks and the kinds of social relationships that are reproduced in their perpetration. More generally, I would argue that if we are serious about using hate crime law to send a message to those deemed racially aggravated offenders we must use the enhanced penalty provisions at the courts’ disposal to ensure that offenders actually get this message: that they are helped to appreciate in what ways their offending might have been perceived as racially motivated by those on the receiving ends of it; that perpetrators are given the space and skills needed to reflect critically on their own words and actions, so that in time they might be able to comprehend how the feelings that they felt towards 224

Racial aggravation or aggravating racism

their victims became racialised, whether before, during or after the offence. For practitioners, what a more psychosocial approach helps illuminate is that getting perpetrators into a frame of mind from which they can appreciate such arguments – whether expressed socio-legally or sociologically – requires that we pay attention to the social and psychic dynamics that get too readily reduced to the term ‘racial motivation’. Getting a perspective on the contradictions in what perpetrators remember, making sense of the chains of associations that inform their use of ethnic referents, and being able to comprehend how these qualities produce effects in the minds of those who feel persecuted, will be crucial if we are to truly understand how hate motivates crime. Helping perpetrators reach a similar level of understanding about their own actions is probably the only way in which we are ever likely to reduce their risk of offending. Letting them know that the criminal justice system is prepared to help them reach this level of understanding is probably a much better message to send out than the exclusively condemnatory one currently used to justify the enhanced sentences to which those convicted of racially aggravated crimes are routinely subject.

Note 1 The author would like to thank Bill Dixon and Tony Jefferson for their reflections on the case study contained in this chapter. A longer and more complex analysis of the case study presented here appears in the British Journal of Criminology, Volume 49, Number 6.

References Adorno, T. W., Frenkel-Brunswik, E., Levinson, D. J. and Sanford, R. N. (1950) The Authoritarian Personality. New York: Harper. Benjamin, J. (1998) Shadow of the Other. London: Routledge. Burney, E. and Rose, G. (2002) Racist Offences – How is the Law Working? London: Home Office Research, Development and Statistics Directorate. Clarke, S. (2003) Social Theory, Psychoanalysis and Racism. Houndmills: Palgrave. Court, D. (2003) ‘Direct Work with Racially Motivated Offenders’, Probation Journal, 50(1): 52–8. Dalal, F. (2002) Race, Colour and the Process of Racialization. London: Routledge. Dixon, B. and Gadd, D. (2006) ‘Getting the Message? “New” Labour and the Criminalization of “Hate” ’, Criminology and Criminal Justice, 6(2): 309–28. 225

Hate Crime

Dixon, L. (2002) ‘Tackling Racist Offending: A Generalized or Targeted Approach’, Probation Journal, 49(3): 205–16. Dixon, L. and Court, D. (2003) ‘Developing Good Practice with Racially Motivated Offenders’, Probation Journal, 50(2): 149–53. Gobodo-Madikizela, P. (2003) A Human Being Died That Night. Claremont: David Philip. Hankinson, I. (2008) ‘Working with Racially Motivated Offenders: Practice Issues’, Probation Journal, 55(1): 119–20. Her Majesty’s Inspectorate of Probation (2005) I’m Not a Racist But … An Effective Supervision Inspection Thematic Report. London: Home Office. Hollway, W. and Jefferson, T. (2000) Doing Qualitative Research Differently. London: Sage. Home Office (1981) Racial Attacks. London: Home Office. Iganski, P. (2008) ‘Hate Crime’ and the City. Bristol: Policy Press. Isal, S. (2004) Preventing Racist Violence: Interim Findings, Runnymede Trust Working Paper. London: Runnymede Trust. Jacobs, J. (2002) ‘Hate Crime: Criminal Law and Identity Politics’, Theoretical Criminology, 6(4): 481–4. Keith, B. (2006) Report of the Zahid Mubarek Inquiry. London: Her Majesty’s Stationery Office. Lemos, G. (2005) The Search for Tolerance. York: Joseph Rowntree Foundation. McDevitt, J., Levin, J. and Bennett, S. (2003) ‘Hate Crime Offenders: An Expanded Typology’, in B. Perry (ed.) Hate and Bias Crime. New York: Routledge (pp. 109–16). McGhee, D. (2007) ‘The Challenge of Working with Racially Motivated Offenders: An Exercise in Ambivalence?’, Probation Journal, 54(3): 213–26. Messerschmidt, J. (1997) Crime as Structured Action. Thousand Oaks, CA: Sage. Ministry of Justice (2008) Statistics on Race and the Criminal Justice System – 2006/7. London: Ministry of Justice. Newman, C. (2008) ‘Divisional Court. Racially Aggravated Public Order Offence: Sufficiency of Partial Racial Hostility’, Journal of Criminal Law, 72(4): 265–67. Perry, B. (2001) In the Name of Hate. New York: Routledge. Ray, L., Smith, D. and Wastell, L. (2003) ‘Racist Violence from a Probation Service Perspective’, in R. M. Lee and E. A. Stanko (eds) Researching Violence. London: Routledge (pp. 217–31). Ray, L., Smith, D. and Wastell, L. (2004) ‘Shame, Rage and Racist Violence’, British Journal of Criminology, 44(3): 350–68. Rustin, M. (2000) ‘Psychoanalysis, Racism and Anti-Racism’, in P. Du Gay, J. Evans and P. Redman (eds) Identity: A Reader. London: Sage (pp. 183– 201). Sherwood, R. (1980) The Psychodynamics of Race. Brighton: Harvester. 226

Racial aggravation or aggravating racism

Sibbitt, R. (1997) The Perpetrators of Racial Harassment and Racial Violence. London: Home Office. Smith, D. (2006a) ‘Racially Motivated Offenders and the Probation Service’, in S. Lewis, P. Raynor, D. Smith and A. Wardak (eds) Race and Probation. Cullompton: Willan Publishing (pp. 25–40). Smith, D. (2006b) ‘What Might Work with Racially Motivated Offenders?’, in S. Lewis, P. Raynor, D. Smith and A. Wardak (eds) Race and Probation. Cullompton: Willan Publishing (pp. 200–16). Young-Bruehl, E. (1996) The Anatomy of Prejudice. Cambridge, MA: Harvard University.

227

Chapter 11

Healing harms and engendering tolerance: the promise of restorative justice for hate crime Mark Walters and Carolyn Hoyle

Introduction Hate crime scholars have spent the best part of twenty years investigating the prevalence of hate crime, its aetiological determinants, the harms it causes, and how – within a democratic and diverse society – it could be diminished or eradicated (Herek and Berrill 1992; McDevitt and Levin 1993; Jacobs and Potter 1998; Lawrence 1999; Perry 2001; Iganski 2008). Yet within the ‘hate debate’ there has been little attention paid to the potential efficacy of restorative justice. This chapter explores whether restorative justice practices (henceforth RJ) might have the potential to help to repair the harms caused by acts of hatred. It draws upon theoretical and empirical research on both hate crime and RJ and proffers some tentative observations from the early stage of a three-year empirical study into RJ and hate crime being carried out by the first author. Before considering the potential of RJ in cases of hate crime, we briefly present the current criminal justice responses to the majority of hate crimes in the UK. We then move on to describe the harms suffered by hate victims, before considering justifications for using RJ. Finally we explore the potential of RJ to challenge prejudice and bigotry and to repair harms to victims, while considering if the vulnerabilities and power imbalances in hate crime cases should preclude RJ conferencing. The chapter then draws towards the conclusion that the promise of RJ can only be realised and the pitfalls avoided with effective facilitation.

228

Healing harms and engendering tolerance

The retributive response to hate crime In 1998 the government established new aggravated forms of assault, criminal damage, public order offences and harassment where the offender demonstrates hostility towards the victim based on his/her (perceived or actual) race, ethnicity or nationality (ss. 28–32, Crime and Disorder Act 1998). These new offences considerably enhance the punishment of offenders where hostility is proven. More recently, the Act has been amended to include religiously aggravated offences, while the Criminal Justice Act 2003 has also introduced provisions for increasing the sentence of an offender where there is evidence that he/she demonstrated ‘hostility’ towards the victim based on race, religion, disability and sexual orientation (ss. 145 and 146). Retribution provides a clear rationale for hate crime laws, with proportionality the main justification: if hate crimes cause greater harm to victims and minority communities (as research suggests; see below), and if they frustrate aspirations for harmonious multicultural societies, then punishments should be enhanced in order to reflect their seriousness (Lawrence 1999; Levin 1999; Walters 2005, 2006). Moreover, expressive denunciation – the censuring of ‘hate offenders’ – is called for in order to deter future hate crimes (Iganski 1999, 2001, 2002; Lawrence 1999; Walters 2005, 2006). Expressive denunciation may not have any direct individual deterrent effect (Jacobs and Potter 1998), but may contribute towards evolving attitudes to racism, homophobia and other pernicious prejudices. In other words, hate crime laws may help to create a social climate that rejects public displays of identity prejudice, over time potentially bringing about a reduction in hate crime (Walters 2006: 75). There is clearly some weight to this argument. At a theoretical level the criminalisation of hate-motivated perpetrators and the enhancement of their punishments seem to be persuasive; the claims for censure are intuitive, and it is only fair that the punishment fits the crime. However, retribution offers little to victims beyond the satisfaction that the penalty is tough, and – it seems to us – fails to address the aetiology of hate, despite government assertions to the contrary (Home Office 2005: 50). Retribution alone is unlikely to repair the harms caused by hate crime or provide for an effective challenge to prejudice (Jacobs and Potter 1998; Moran et al. 2004). Furthermore, punishment enhancements might serve to uphold victims’ emotional attachments to ‘hate, anger, malice and revenge’, rather than diminish them (Moran et al. 2004: 42). Thus, we believe

229

Hate Crime

that hate crime legislation, in the absence of other measures, cannot deliver on its promise of justice for victims and their communities.

Hate harms While legal scholars looked to legislation to tackle hate crime, criminological attention turned to the experiences and needs of victims and minority communities (Herek and Berrill 1992; Herek et al. 1997; Levin 1999; Iganski 2008). Violence and intimidation motivated by prejudice has been characterised as a form of ‘terror’ as it threatens the personal safety of individuals with certain identity characteristics (Weinstein 1992). Terror spreads within identity groups among others who fear the same fate (Lawrence 1999: 43–4). Their vulnerability seems to exacerbate their experiences of victimisation. Research has shown that a significant proportion of hate crime victims experience heightened levels of emotional harm such as fear, anxiety, depression and feelings of insecurity as a result of being victimised because of who they are or are perceived to be (Herek et al. 1997; Lawrence 1999; Iganski 2001, 2008; McDevitt et al. 2001). Iganski has demonstrated the elevated psychological damage to hate victims through the analysis of data from three British Crime Surveys dating from 2002 to 2005. He found statistically significant higher levels of victims who had been involved in racially motivated incidents (compared to non-hate-related incidents) reporting feelings of shock, fear, depression, anxiety, panic attacks, loss of confidence, vulnerability, insomnia and crying (Iganksi 2008: 12, 13, 82 and 83).1 Furthermore, the duration of psychological harm is often extended for victims of hate crime, with some hate victims suffering from periods of depression, stress and anger for up to five years (Herek et al. 1997). In contrast, psychological harms suffered by victims of crimes not motivated by hate showed considerable improvements within two years (Herek et al. 1997). Clearly hate can aggravate harms caused to victims.

Restorative justice and hate crimes In light of research demonstrating that hate crimes have deleterious consequences for victims and minority communities, a criminal justice response seeking only to punish the offender is inadequate. An alternative or additional approach is called for to reduce the 230

Healing harms and engendering tolerance

heightened levels of emotional harms experienced by victims, and to address the underlying causes of prejudice and hatred. To this end, academics have begun to theorise about the potential of RJ (see, for example, Shenk 2001; Gavrielides 2007). For some considerable time responses to crime in western common law jurisdictions were retributive. The term restorative justice emerged in the latter part of the twentieth century to describe both emerging practices based on the principles of victim–offender mediation, and a new theoretical approach that puts the damaged relationship between the victim, the offender and the wider community at the heart of the criminal justice response. Restorativists in the US and UK developed the ideas and practices of mediation and began to articulate the need for a new way of responding to crime which had at its heart a balanced approach addressing the needs of victims, offenders and the wider communities from which they came. Restitution and reparation – to victims directly and to communities more generally – became the key words for the restorative movement. Since then many advocates of RJ have taken conventional retributive justice as their point of departure, either rejecting outright its utility, or suggesting that alone it fails adequately to respond to the needs of victims, offenders and communities. They argue instead that justice is best achieved by attempting to repair the harms caused by crime (Strang 2002). Healing those affected by crime involves renewing damaged interpersonal relationships or establishing legitimate connections between divided communities or people who have previously seen themselves as estranged. Perpetrators have obligations to restore and repair the harms caused to all ‘stakeholders’, including the victim, other affected community members and even the offender him or herself (Zehr 1990). Thus, rather than stigmatising offenders, RJ is primarily concerned with the engagement of those affected by crime in a dialogic process which aims to achieve reparation – be it emotional, material or to relationships (Considine 1995). Of course, in cases where there has been a dispute, but no clear infraction of the criminal law, the responsibility for cessation of hostilities and for repairing the relationship is shared between them in a process that resembles civil mediation. The two most frequently discussed examples of restorative justice are victim–offender mediations (VOM) and family group conferences (henceforth conferences).2 These practices typically involve a face-toface meeting between the victim and offender (or those involved in conflict where disputants cannot be categorised as such) in a safe environment to discuss the incident, the harms it has caused and 231

Hate Crime

how these harms should be repaired. Conferences, unlike VOMs, typically include supporters of the disputants and other concerned community members and sometimes representatives of the state, such as police officers, social workers or housing officers. RJ meetings can also be facilitated through indirect or ‘shuttle’ mediation; victims and offenders discuss their case individually with an RJ facilitator, who then feeds back to the other party. The first author is currently conducting research on restorative mediation for hate crimes, with the Hate Crimes Project – set up by the Southwark Mediation Centre in 2000 – being one key site for the study. The Hate Crimes Project aims to resolve hate-motivated incidents and ongoing disputes through restorative dialogue and reparation agreements. Between 2007 and 2009, 144 hate incidents were referred to the project for shuttle or direct mediation.3 So far, we have observed ten direct and indirect mediations. Incidents have included verbal abuse, harassment and serious assault involving a perceived (or actual) motivation (or partial motivation) based on sexual orientation, race and/or disablism. We also observed an indirect RJ meeting facilitated by the Oxford Youth Offending Team in a case of racially aggravated harassment. While the study continues, and many more cases are required to make definitive claims for RJ, qualitative analysis of the data derived from these cases allows for some tentative observations.

Engendering tolerance of ‘difference’ through restorative dialogue The dialogic process that characterises RJ meetings typically begins by focusing on the perpetrator’s accountability for the crime/incident. RJ generally presupposes that the offender has acknowledged responsibility for an offence. In other words, it is not concerned with fact-finding but with an appropriate response to an admitted offence. Its realm is that of sentencing, not that of the criminal trial. But the meeting goes beyond an admission of guilt to probing questions about why the offender committed the offence – what were the motivations or triggers – and aims to challenge any post-hoc justifications. The facilitator then encourages those harmed by the incident to describe its impact, at the time and since, with a view to exploring the relationship between actions and consequences, and encouraging the perpetrator to take full responsibility for those harms. RJ meetings should be safe, empowering and inclusive and 232

Healing harms and engendering tolerance

enable all stakeholders to reveal fully how the incident has affected them, with no one silenced by domination (Braithwaite 2003: 157). This structured dialogic process helps to humanise victims while allowing offenders to witness first-hand the impact of their actions (Shenk 2001; Sapir 2007). So what promise does such an approach hold for hate crime cases? We are fairly confident that RJ has the potential to heal victims and communities, and believe that it can challenge prejudices held about minority communities if, and only if, it is facilitated well. We are unsure that one-off RJ meetings can fully realise the potential to change offenders’ attitudes, although we remain optimistic that in some cases they will change their behaviours and just might alter their mindset. Before exploring how this could happen, and what factors would thwart these aims, it might be useful to say something about the nature of hate crime. The term ‘hate’ in hate crime refers not only to feelings of hatred but also to the prejudice or bigotry that offenders use to victimise certain people. Within the body of literature on the causes of prejudice motivating hate crime,4 many academics have focused their attention on social psychological analyses and macro-sociological theories,5 with those that have concentrated on fear relating to cultural and/or identity ‘difference’ being of particular relevance here (Young 1999: 100ff.; Perry 2001; Hudson 2008). Cultural or identity difference refers simply to belonging to, and sharing characteristics with, an identity group that differs from the dominant group in society (Perry 2001). Identity groups may coalesce around ethnicity, religion, sexual orientation, or other such characteristics. In many respects, cultural and social differences are celebrated by those who embrace the notion of a multicultural society. In fact, grouping by identity can help individuals to gain ontological security by asserting their own ways of life and holding tight to an assertion of naturalness (Young 1999: 100–1). Yet essentialism in pursuit of ontological security and social equality is a double-edged sword. Indeed, Young (1999; 104–5) argues that essentialising ‘others’ can have the unintended consequence of encouraging stereotyping that allows dominant identity groups to project their own social insecurities onto others and can lead to demonisation of minority groups. This further paves the way for the formation of cultures of prejudice within local communities and society more generally (Sibbitt 1997; Ray and Smith 2002; Ray et al. 2003). This fear of cultural difference, as Perry (2001) argues, results in intimidation and violence aimed at suppressing minority identity groups. Hate crimes 233

Hate Crime

can thus become everyday occurrences for many individuals who are identified as a threat to the dominant norms and values (Ray et al. 2003; Iganski 2008). Unlike other sentences in the criminal process, RJ explicitly aims to explore the underlying causes and consequences of a crime. In hate crime cases the reparation of harms is likely to be contingent on the process effectively exposing and challenging bigotry and fallacies around identity difference. This should not entail an attempt by facilitators to prescribe appropriate attitudes, but to create the conditions whereby all participants feel able to challenge prejudice and correct ignorance, and, where necessary, to impede harassment or denigration of one or more of the participants during the conference. Effective communication between participants is crucial if victims are to receive emotional reparation and all participants are to gain greater insight into others’ cultural backgrounds, attitudes and beliefs. In successful RJ meetings offenders understand the hurt and suffering experienced by victims, and as a result recognise the immorality of their behaviour and feel empathy and remorse (Harris et al. 2004: 201). In some cases victim forgiveness may follow, although forgiveness should never be the goal. Empathy is central to RJ theory; when confronted with victims’ suffering offenders are more likely to experience understanding and compassion than when simply told by a third party about the harms caused. Empathy can act as a catalyst for attitudinal change, which can benefit both parties, but also the wider communities to which they return (Dzur and Olson 2004). However, there is often an inevitable ‘empathic divide’ between victims and defendants, especially when they come from different social backgrounds (Smith 2006).6 People find it difficult to understand the lives of others from different social or cultural groups. In hate cases prejudice and abhorrence present further barriers to understanding. The route to empathy in hate crime cases therefore lies not just with exploration of the physical or psychological impacts of the crime on that particular victim, but with attempts to humanise them and their communities, whether that be gay, disabled or British minority ethnic individuals. There is likely to be little progress without breaking down the empathic divide. In hate cases offenders will come to the meeting holding negative attitudes towards victims based on ignorance or fear, attitudes that might until then have gone unchallenged (McConnell and Swain 2000; see generally Perry 2001). When confronted by other participants, offenders might attempt to rationalise their offences – and their

234

Healing harms and engendering tolerance

behaviours more generally – by seeking refuge in ‘techniques of neutralisation’, which if left unchallenged can undermine the restorative process (Sykes and Matza 1957). These techniques can involve offenders denying or underplaying the seriousness of their offence or their own culpability. Offenders might refuse to accept full responsibility by blaming others for their behaviour; or admit that they have broken the law but not that they have caused any real harm; or – and of particular relevance here – assert that the victim is undeserving or illegitimate, or less worthy of respect (Byers et al. 1999), even less human. History has shown us that ‘othering’, by recourse to dehumanising language and images, precedes that most extreme of hate crimes, genocide; witness the denunciation of Jews in Nazi Germany as ‘vermin’, and Tutsis in pre-genocide Rwanda as ‘cockroaches’. Language and the images it conveys are particularly potent resources in attempts to justify the apparently unjustifiable. RJ processes can only succeed if they challenge crude stereotypes and prejudicial assumptions. Techniques of neutralisation can be managed within a restorative process where other participants feel able to challenge them. A skilful facilitator can help by scrutinising the offender’s account and asking questions that encourage participants to reflect and comment on the offender’s version of events. Restorative processes are well placed for challenging techniques of neutralisation when they confront offenders with the results of their actions; it is harder to pretend that the victim was not harmed or deserved to be offended against when that very person is present and articulating their feelings. A defendant’s racist or homophobic attitudes may not necessarily be deeply entrenched but could instead be the result of unthinking adoption of what is perceived at that time to be local consensus on immigration, for example (Sibbit 1997; Ray and Smith 2002). It may be easy – without evidence to the contrary – to hold ignorant beliefs about ‘outsiders’, where it is argued, albeit erroneously, that such people threaten the status quo, but attitudes can change when confronted with an individual who seems, in many ways, not dissimilar (Smith 2006; see also Sapir 2007). Indeed, the RJ meeting may present the offender with his or her first opportunity to talk to a ‘gay’ or a ‘black’ person in a safe environment, and, importantly, in the absence of provocation from others, and this may be enough to demonstrate that the prejudice is without merit (McConnell and Swain 2000).7

235

Hate Crime

Repairing hate harms The second half of this chapter draws on three of our case studies to show that the potential of RJ to challenge prejudice and heal victims can only be realised by effective facilitation.8 The first case illustrates how RJ can help to break down stereotypes and facilitate communication among people divided by ignorance and intolerance.

Case study 11.1: Homophobic abuse Mr J, a young gay man of mixed race, perceived himself to be the victim of homophobic harassment from his elderly neighbours, Mr and Mrs M (a black Caribbean man and a white British woman). They were accused of giving Mr J ‘dirty looks’, and pointing at him in the street. One evening Mr J confronted Mr and Mrs M but was met with ‘a barrage of abuse’, with Mr and Mrs M both repeatedly shouting ‘People like you are sick in the head’ [our emphasis]. They also called him ‘a thief, a drug addict and a psycho’. Mr J perceived these words to be homophobic. When later approached by their housing officer, Mr M said that he would hit Mr J with a bat if he came back. This resulted in an intervention by the neighbourhood policing team and a referral to the Hate Crimes Project. Both parties took part in indirect mediation meetings before participating in a face-to-face meeting. During indirect mediation Mr J explained that the incident had made him feel ‘really worried’ that he may be physically assaulted and that he wanted someone to ‘make them understand’ how he felt and for them to stop harassing him. The mediator raised the issue of ‘identity’, and Mr J said: ‘People have misconceptions about gay people, especially elderly Caribbean people. There is a misconception that gay people are drug addicts and steal from people to feed their habit.’ The mediator interjected, saying that ‘not all Caribbean people’ feel like this, and Mr J agreed. At direct mediation both parties initially appeared defensive and tensions were running high. The mediator explained what the process was about, and set the parameters for discussion: ‘There will be no Jerry Springer … do not tell people what to do … Looking to the future, what do we want to do to make the future different? … You will hear different views today, be respectful of those different views … please ask questions in a 236

Healing harms and engendering tolerance

polite and respectful manner. It is not always easy to talk … My role as mediator is not to point the finger or to judge or blame … This is a place of respect for communication.’ She then summarised both parties’ account of the incident and invited them to talk about what had happened. This revealed significant disagreement in perspectives and the meeting soon descended into what appeared to be an intractable dispute. However, the mediator kept control and used various techniques to find common ground. Finally, there was a breakthrough and the tone of the meeting began to change as the mediator reiterated points of agreement. Mr J then apologised, saying, ‘I wanted to speak with you. If I was rude I didn’t intend to be.’ Mr M responded by holding out his hand and smiling and they shook hands. The issue of homophobia was raised by Mr J: ‘I thought you may have had different ideas about gay people … a lot of people think gay people are party animals on drugs, etcetera ... I wanted to prove I was a nice person, a quiet person …’ Mr M replied: ‘I never said anything personal about you … we have a gay nephew, we would never turn him away.’ A lengthy conversation ensued, with Mr J explaining his experience of homophobia and previous abuse from other neighbours, and Mr and Mrs M saying that they understood that everyone was an individual and it was not right to stereotype people. They agreed that they had reached a better understanding of each other; as Mrs M put it, ‘He’s found out how we feel and we’ve found out how he feels.’ By the end of the meeting all parties appeared to be happy, an agreement was made and signed and they then thanked the mediator for helping them to resolve the incident. This RJ intervention illustrates how the process, if managed sensitively and according to RJ principles, can defuse conflict and expose ignorance and bigotry to critical scrutiny. Once prejudices are explored in a safe environment there is scope for effective challenge and subsequent modification of perspectives. In this case both parties benefited from the discussion, and there was ‘victim’ healing and ‘offender’ edification. The Hate Crimes Project mediator contacted the parties again several weeks after the meeting and both stated that no further incidents had occurred. Of course, by challenging stereotypes restorative processes might change defendants for the better, but may do little for victims. If RJ is to be more than rehabilitation in a different guise, it needs to 237

Hate Crime

repair the harms endured by hate victims, many of whom experience heightened levels of emotional harm, disempowerment, and even low self-esteem through marginalisation by the dominant groups in society. RJ can put hate victims in the centre of the justice process by providing them with a voice to express their experiences (McConnell and Swain 2000). Our observation of an indirect restorative meeting carried out by the Youth Offending Team RJ practitioner in Oxford provides a good example.

Case study 11.2: Racial harassment (indirect mediation) An 18-year-old woman was the victim of racial harassment that had left her constantly anxious, ‘too frightened to go out’, and unable to sleep. During an indirect RJ meeting her mother indicated that her daughter had felt ‘ashamed’ about what had happened and reluctant to talk about it, and described the ‘traumatic’ impact on the whole family and her hopes that her daughter could regain her feeling of ‘safety’. Both mother and daughter appeared to experience relief in describing to the YOT practitioner the impact of their experiences. The practitioner promised to feedback to them any progress made by the offender and arranged for the victim to attend a victim support course aimed at improving self-esteem and individual security, which (according to feedback) made a great difference in helping her to feel safe again. RJ promises many potential benefits to both victims and offenders. It can reduce levels of fear, anger, anxiety and insecurity among victims of violent and non violent crimes (Daly 2001: 78; Strang 2002). Research has shown that these benefits are most likely to be realised when victims attend sessions with their offenders. However, they can be realised to some extent even when victims do not meet their offenders by the effective facilitation of indirect communication between victims and offenders. Best practice leaves such victims less angry and less afraid (Hoyle 2002: 115), as this case example shows. By speaking out about how the crime has impacted upon their lives victims can be rescued from the shadows of silence, isolation and despair (Kay 2008). Not only can they express their emotions, they can also ask questions about the offender’s motivations (McConnell and Swain 2000). In turn, the offender’s story can help them to realise that they were not personally to blame for their own victimisation and they can obtain assurances from offenders – either directly or 238

Healing harms and engendering tolerance

mediated through the facilitator – that the offence will not happen again, reducing their fears of repeat victimisation. This allows them to move on with their lives and gain what social psychologists call ‘a sense of closure’ about the incident. RJ meetings typically result in a reparation agreement. This may require an apology only, but it can comprise financial restitution or the repair of damaged property. Some material reparation agreements have no direct connection with the offence, but simply demonstrate a willingness to give back to the community; others are more personally associated with the offence or victim. Some hate crimes involve abusive graffiti directed at the victim or a wider identity group, with racist or homophobic graffiti being the most common. A reparation agreement to clean up graffiti could provide material and symbolic reparation. An offer and acceptance of an apology demonstrates a commitment by both parties to draw a line under the incident and allow those involved to move on with their lives. However, it is important that apologies are perceived as being genuine, otherwise they may have adverse effects on the victim’s healing process (Daly 2001). A genuine apology is an expression of remorse and contrition and can alleviate victims’ feelings of resentment and frustration (Strang 2002: 56), raising the prospect of reconciliation between the parties. Research by Theo Gavrielides has documented various RJ case studies that illustrate these benefits (2007: 203; see also Gavrielides et al. 2008; and further Rice 2003 who explores her own journey of hate victimisation and restorative justice).

Vulnerability and power imbalances in hate crime conferencing The more that victims are affected by the crimes against them the more likely they are to want to meet the offender (Mattinson and Mirrlees-Black 2000: 43). Nonetheless, some will not wish to explore in front of the offender the emotions evoked by recalling the incident and may be more afraid of their offender because they know that they were targeted because of what they represent (Gavrielides 2007). Low victim participation in RJ initiatives is not uncommon, although there is a good deal of jurisdictional variation. Victims exclude themselves for a variety of reasons (Hoyle 2002) and practitioners should make informed judgements, first on whether a RJ meeting is appropriate for all parties, and second, how far they should encourage the victim to get involved. This is particularly important when the victim is deemed to be vulnerable. 239

Hate Crime

Critics of RJ express concerns that it can result in secondary victimisation, particularly when there is a significant power differential between the parties (Hoyle 2007). Restorative processes might perpetuate power imbalances if offenders are allowed to manipulate the informal process to diminish guilt, trivialise the violence or shift the blame to the victim. There are, for example, pitfalls in using restorative justice for domestic violence; not least of these concern the power and control some (mostly male) perpetrators have over their victims. Under such conditions victims’ participation could be disempowering, with little chance of emotional healing if the process fails to challenge entrenched gendered roles (Coker 2002; Stubbs 2007). Homophobia, racism, disablism and other forms of prejudice mean that hate victims can feel similarly vulnerable, marginalised and disempowered by the dominant white, heterogenous culture (Smith 2006). The social distance between offender and victim may mean that neither party can identify with each other’s cultural backgrounds. They may live in different neighbourhoods, hold different beliefs and will typically socialise in different groups (Smith 2006). This social distance has led some to question the potential efficacy of RJ for those without a shared community (Kelly 2002) if, for example, minority groups are viewed with scepticism or as deserving of discrimination. As Gavrielides (2007: 199) points out, it would be idealistic to assume that years of learned racism or homophobia could be transformed into caring and empathetic attitudes during one RJ meeting. Offenders may not be able to empathise with their victims (Kelly 2002; Smith 2006; Stubbs 2007), and victims’ revelations might be perceived as provocative and elicit a hostile response from the offender (Smith 2006). In such circumstances ineffective facilitation by someone who does not fully understand the cultural tensions could fail to prevent repeat victimisation (Smith 2006). Like victims of domestic violence, hate victims in such situations might feel compelled to accept disingenuous apologies in an attempt to appease the offender and reduce the likelihood of reprisals and this, of course, is not the path to healing.

The dangers of ineffective facilitation Cultural difference, power imbalances, and suspicion of ‘others’ will undoubtedly impact on the success of face-to-face communication and may make it unlikely that the parties will find common 240

Healing harms and engendering tolerance

ground (Gavrielides 2007: 198). Differences in communication style, mannerisms and tone can shape the progress of a conference (Smith 2006: 170). The way that a participant tells a story and subsequently the way that story is interpreted can depend on racial, ethnic, religious or sexual orientation variables (Smith 2006). The challenge for RJ facilitators will be to create a safe environment for victims and offenders to divulge sensitive information and expose their fears or prejudices to the other participants, and to encourage participants to challenge prejudice and praise pro-social sentiments. This requires careful planning; the dynamics of a successful restorative intervention are dependent on the facilitator having identified and invited the most appropriate participants to the meeting and being able and prepared to encourage challenging and yet respectful interaction between all parties. Offenders should be discouraged from bringing supporters who share their prejudices and might condone unacceptable behaviour or use the meeting to further intimidate the victim and his or her supporters. Facilitators can ‘vet’ potential supporters prior to a meeting, or they could exclude any co-offenders, holding separate meetings for them (McConnell and Swain 2000). Of course, while they are safe options, these measures challenge one of the key aims of restorative justice: to engage communities. In some cases facilitators manage to locate appropriate pro-social supporters for offenders, whether grandparents, teachers, social workers or local football coaches. However, while these people can effectively challenge the offender’s prejudices and then provide avenues for rehabilitation, they may not so easily be able to provide reintegration into the most crucial communities for offenders – most typically kinship or friendship groups. Hence, it may sometimes be necessary to include those people closest to the offender, even when they appear to be a negative influence. And in such circumstances it is appropriate for facilitators to challenge some of the norms and values of the wider community in encouraging offenders to reflect on and account for their behaviour, hence improving the prospects for the whole group including the offender and reducing the risk of repeat victimisation. Clearly, facilitators will need extensive training on all areas of prejudice in order effectively to negotiate the minefield of restorative conferencing for hate crime (McConnell and Swain 2000). Social distance can be reduced when individuals are encouraged to communicate with one another through structured dialogue facilitated by a trained and experienced mediator and with the support of people close to the main parties (Smith 2006). Facilitators 241

Hate Crime

need to be aware of and responsive to language, whether that means employing translators, or encouraging people to use their own dialects and idioms (Smith 2006) to reduce power imbalances and social distance. Furthermore, reflections on the social context in which the crime occurred must be encouraged. This does not mean condoning techniques of neutralisation, but exploring the social and cultural factors behind the unacceptable behaviours (Smith 2006: 183).9 Our third case example shows that without effective facilitation restorative justice will never realise its potential to reduce social distance, to challenge prejudice and hatred and to heal the harms caused to victims.

Case study 11.3: Racial harassment A black couple (Mr and Mrs V), originally from South America, complained several times to the police and housing association that their neighbour, Mr H, was constantly harassing them by hammering on their walls, and had at one point shouted ‘Fucking nigger! Nigger! Nigger!’ at Mrs V and made monkey sounds at her. A conference was attended by Mr H, Mr and Mrs V, two housing officers, a housing manager, and the Hate Crimes Project mediator. The housing manager (Mr J) facilitated the meeting and began by stating that proof of racial abuse would put Mr H in breach of his tenancy agreement and the criminal law. Mr J then asked Mr and Mrs V, followed by Mr H, to describe what had happened. As Mrs V’s English was limited, her husband frequently translated what she wanted to say. Throughout the meeting Mr H vehemently denied the allegations of racial abuse, and made a counter complaint about the noise coming from Mr and Mrs V’s flat to explain his ‘hammering on their walls’. Mr J asked the participants to resolve the situation, but no consensus could be found at this point. Miss E (HCP mediator) tried to explore whether racism was behind the alleged abuse by asking Mr H:, ‘If your neighbour had been white would the allegation have been made?’ Mr H responded: ‘They don’t speak English, it’s always another language … It could be. Who knows?’ He repeated this several times during the meeting. Miss E followed this up by asking, ‘If they had been speaking English would it be different?’, which Mr H could not answer. Still looking for a resolution, Mr J asked: 242

Healing harms and engendering tolerance

‘What is triggering [Mr H’s] behaviour? How can we move on?’ Mr H replied that he would stop shouting and banging on the walls, and then proposed – referring to his relationship with his neighbours – ‘Let’s try and start again.’ An agreement was drawn up that Mr H would not hammer on Mr and Mrs V’s wall or use racist language towards them. Mr and Mrs V agreed that they would not make any unreasonable noises. This trivial agreement was the product of the superficial treatment of a complex and sensitive matter and poor facilitation. The meeting failed fully to reveal and therefore challenge the racism behind the abuse. Not surprisingly, Mr and Mrs V appeared still to be very distressed at the end of the meeting and were clearly unhappy about the outcome.10 Cultural and identity differences were at the heart of this case. Mr and Mrs V’s foreignness was central to the racism behind the dispute and the communication difficulties within the conference. The facilitator failed to provide a safe space within which participants could explore their own and others’ behaviours and did not probe fully the apparent racism. He was unable therefore to find common ground between the participants or create the conditions for empathy. Part of this failure was due to communication difficulties, which might have been helped by a translator, but this was exacerbated by the failure to lay down ground rules and follow restorative principles. The housing manager was not best placed to facilitate as he was not impartial, but instead adopted an investigatory and prosecutorial role. This stemmed in part from his power to evict either of the parties from their home, a power that clearly dissuaded Mr H from providing an honest account of his behaviour. Furthermore, he was not trained. The HCP mediator, as a neutral party working for a third sector organisation, would have been a more appropriate choice. Given the failure to hold Mr H accountable for his behaviour, it should be of no surprise that the reparation agreement suggested that this was simply a neighbour dispute, rather than an incident of racial abuse.

Conclusion Hate crimes can have dramatic emotional and physical effects on victims, their families and minority communities more generally. The government’s response to this has been to enhance the punishment 243

Hate Crime

of offenders who demonstrate hostility based on the victim’s race, religion, sexual orientation or disability. Yet increasingly punitive responses do little to repair the harms experienced by victims – beyond perhaps an initial visceral satisfaction that the offender is being hurt – and fail effectively to challenge the prejudices of individual offenders. Hence, this chapter has explored the potential efficacy of restorative justice practices in responding to hate crime, either as a diversionary measure (for example, community mediation) or in addition to criminal sanctions (such as RJ meetings carried out by youth offending teams or the probation service). The case studies described above suggest that RJ meetings, if facilitated carefully, may help victims to move on from their experiences – potentially reducing feelings of fear, anger and insecurity. Given the often exacerbated harms caused by crimes motivated by discrimination and prejudice, the ability of RJ to provide reparation and emotional healing makes it a particularly helpful tool in efforts to assuage the pains of hate crime. Of course, it is not only victims who can benefit from participation in restorative processes. Given effective and sensitive facilitation, and the inclusion of relevant community members in meetings, offenders too can be helped to better understand their prejudices and appreciate the consequences of their actions, as well as learning more about other cultures and identities, with a view to reducing prejudice and consequent hate crime. The effective facilitation of structured dialogue between victims and offenders, and their respective communities, can allow for the emergence of new relationships between participants based on respect and mutual understanding. In this sense, RJ may provide a propitious mechanism for members of our increasingly multicultural communities to find resolution to hate-motivated conflicts. Projects such as that implemented by Southwark Mediation Centre may well emerge as exemplars for restorative processes in fractured communities. Within the criminal justice system, the police, probation services and youth offending teams may find that using specialised restorative justice practitioners who can deal specifically with cases involving prejudice, domestic violence and/or sexual assault will enable them to more effectively tackle the root cause and effect of such offences.11 Indeed, there is now growing support for using RJ in difficult cases where power imbalances exist, such as domestic, sexual or racial violence (Hudson 2002; Hoyle 2007). However, there is still a need for thorough empirical investigation to better understand RJ’s potential in hate crime and other ‘difficult’ cases and how best facilitators can negotiate the difficulties raised by bringing together 244

Healing harms and engendering tolerance

those who would otherwise be antagonistic towards each other. From our research findings so far we believe that if victims and offenders are properly prepared for restorative processes and meetings are facilitated effectively by people with adequate training and resources, conferences can promote empathy, provide harm reparation and potentially promote acceptance of cultural difference.

Notes 1 See also research by the National Institute Against Prejudice and Violence (1986, reported in Levin 1999). 2 The RJ label is confusingly applied to a variety of often disparate practices, including the provision of victim support services and various court-imposed sanctions (Dignan 2005: 2), but only the genuinely dialogic processes, such as conferencing and mediation are, as far as we are concerned, restorative justice. 3 Referrals are from statutory and third sector organisations (e.g. police, housing association, anti-social behavioural unit, victim support, and self-referrals). 4 A review of this literature is beyond the scope of this chapter (see generally Green et al. 2001). 5 Though some have looked at psychological traits, economic and political causes (see Green et al. 2001). 6 The empathic divide refers to the inability to understand the lives of, and empathise with, ‘others’ from different sociocultural backgrounds, and is particularly apparent across the racial divide (Haney 2005: 189–210). 7 See case study examples of RJ in hate crime cases in Umbreit et al. (2002) and Gavrielides et al. (2008). 8 The first author observed and made detailed notes on these RJ meetings. 9 The Hate Crimes Project at Southwark Mediation is a positive illustration of where these forms of good practice take place. 10 Although recently Mr and Mrs V said they were pleased that the harassment has stopped. 11 The first author is currently working with Sussex Police to implement a pilot scheme using RJ as a response to low-level hate crime offences.

References Braithwaite, J. (2003) ‘Restorative Justice and Social Justice’, in E. McLaughlin, R. Fergusson, G. Hughes and L. Westmarland (eds) Restorative Justice Critical Issues. London: Sage. 245

Hate Crime

Byers, B., Crider, B.W. and Biggers, G.K. (1999) ‘Bias Crime Motivation: A Study of Hate Crime Offender Neutralization Techniques Used Against the Amish’, Journal of Contemporary Criminal Justice, 15(1): 78–96. Coker, D. (2002) ‘Transformative Justice: Anti-Subordination Processes in Cases of Domestic Violence’, in H. Strang and J. Braithwaite (eds) Restorative Justice and Family Violence. Cambridge: Cambridge University Press. Considine, J. (1995) Restorative Justice: Healing the Effects of Crime. Lyttelton, NZ: Ploughshares Books. Daly, K. (2001) ‘Conferencing in Australia and New Zealand: Variations, Research Findings, and Prospects’, in A. Morris and G. Maxwell (eds) Restorative Justice for Juveniles: Conferencing, Mediation and Circles. Oxford: Hart. Dignan, J. (2005) Understanding Victims and Restorative Justice. Maidenhead: Open University Press. Dzur, A. and Olson, S. (2004) ‘The Value of Community Participation in Restorative Justice’, Journal of Social Philosophy, 35(1): 91-–107. Gavrielides, T. (2007) Restorative Justice Theory and Practice: Addressing the Discrepancy. New York: Criminal Justice Press. Gavrielides, T., Parle, L., Sall, A., Liberatore, G., Mavadia, C. and Gloria, A. (2008) Restoring Relationships Project: Addressing Hate Crime through Restorative Justice and Cross-Sector Partnerships: A London Study. London: ROTA. Green, D., McFall, L. and Smith, J. (2001) ‘Hate Crime: An Emergent Research Agenda’, Annual Review of Sociology, 27: 479–504. Haney, G. (2005) Death by Design: Capital Punishment as a Social Psychological System. New York: Oxford University Press. Harris, N., Walgrave, L. and Braithwaite, B. (2004) ‘Emotional Dynamics in Restorative Justice’, Theoretical Criminology, 8(2): 191–210. Herek, G. and Berrill, K. (1992) Hate Crimes: Confronting Violence Against Lesbians and Gay Men. Thousand Oaks, CA: Sage. Herek, G., Cogan, J., Gillis, J. and Glunt, E. (1997) ‘Hate Crime Victimization Among Lesbian, Gay, and Bisexual Adults: Prevalence, Psychological Correlates, and Methodological Issues’, Journal of Interpersonal Violence, 12(2): 195–215. Home Office (2005) Improving Opportunity, Strengthening Society. London: Home Office. Hoyle, C. (2002) ‘Securing Restorative Justice for the “Non-Participating” Victim’, in C. Hoyle and R. Young (eds) New Visions of Crime Victims. Oxford: Hart Publishing. Hoyle, C. (2007) ‘Feminism, Victimology and Domestic Violence’, in S. Walklate (ed.) Handbook of Victims and Victimology. Cullompton: Willan Publishing. Hudson, B. (2002) ‘Restorative Justice and Gendered Violence? Diversion or Effective Justice?’, British Journal of Criminology, 42: 616–34. 246

Healing harms and engendering tolerance

Hudson, B. (2008) ‘Difference, Diversity and Criminology: The Cosmopolitan Vision’, Theoretical Criminology, 12(3): 275–92. Iganski, P. (1999) ‘Why Make Hate a Crime?’, Critical Social Policy, 19(3): 386–95. Iganski, P. (2001) ‘Hate Crimes Hurt More’, American Behavioural Scientist, 45(4): 626–38. Iganski, P. (2002) ‘Hate Crime Hurts More But Should They be Punished More Harshly?’, in P. Iganski (ed.) The Hate Debate: Should Hate be Punished as a Crime? London: Profile Books. Iganski, P. (2008) Hate Crime and the City. Bristol: Policy Press. Jacobs, J. B. and Potter, K. (1998) Hate Crimes. New York: Oxford University Press. Kay, J. W. (2008) ‘Murder Victims’ Families for Reconciliation: Story Telling for Healing, as Witness and in Public Policy’, in D. Sullivan and L. Tifft (eds) Handbook of Restorative Justice: A Global Perspective. Abingdon: Routledge. Kelly, T. L. (2002) ‘Is Restorative Justice Appropriate in Cases of Hate Crime’, conference paper, Western Pacific Association of Criminal Justice Educators Conference, Lake Tahoc, Nevada. Lawrence, F. M. (1999) Punishing Hate: Bias Crimes under American Law. London: Harvard University Press. Levin, B. (1999) ‘Hate Crime: Worse by Definition’, Journal of Contemporary Justice, 15(1): 6–21. Mattinson J. and Mirrlees-Black, C. (2000) Attitudes to Crime and Criminal Justice Findings from the 1998 British Crime Survey, Home Office Research Study 200. London: Home Office. McConnell, S. and Swain, J. (2000) ‘Victim–Offender Mediation with Adolescents who Commit Hate Crimes’, paper presented at the Annual Conference of the American Psychological Association, Washington, DC, 4–8 August. McDevitt, J. and Levin, J. (1993) Hate Crimes: The Rising Tide of Bigotry and Bloodshed. Boulder, CO: Westview Press. McDevitt, J., Balboni, J., Garcia, L. and Gu, J. (2001) ‘Consequences for Victims: A Comparison of Bias- and Non-bias-Motivated Assaults’, American Behavioural Scientist, 45(4): 697–713. Moran, L., Skeggs, B., with Tyrer, P. and Corteen, K. (2004) Sexuality and the Politics of Violence and Safety. London: Routledge. Perry, B. (2001) In the Name of Hate: Understanding Hate Crimes. New York: Routledge. Ray, L. and Smith, D. (2002) ‘Hate Crime, Violence and Cultures of Racism’, in P. Iganski (ed.) The Hate Debate: Should Hate be Punished as a Crime? London: Profile Books. Ray, L., Smith, D. and Wastell, L. (2003) ‘Understanding Racist Violence’, in E. Stanko (ed.) The Meanings of Violence. London: Routledge. Rice, S. (2003) ‘Restorative Justice: A Victim’s Personal Exploration Towards 247

Hate Crime

Acceptance’, paper presented at the Sixth International Conference on Restorative Justice, 1–4 June. Sapir, B. (2007) ‘ Healing a Fractured Community: The Use of Community Sentencing Circles in Response to Hate Crimes’, Cardozo Journal of Conflict Resolution, 9: 227-235. Shenk, A. H. (2001) ‘Victim-Offender Mediation: The Road to Repairing Hate Crime Injustice’, Ohio State Journal on Dispute Resolution, 17: 185–217. Sibbitt, R. (1997) The Perpetrators of Racist Violence and Harassment. London: Home Office. Smith, K. (2006) ‘Dissolving the Divide: Cross Racial Communication in the Restorative Justice Process’, Dalhousie Journal of Legal Studies, 15: 168–203. Strang, H. (2002) Repair or Revenge: Victims and Restorative Justice. Oxford: Oxford University Press. Stubbs, J. (2007) ‘Domestic Violence and Critical Questions for Restorative Justice’, Criminology and Criminal Justice, 7: 169–87. Sykes, G. and Matza, D. (1957) ‘Techniques of Neutralization: A Theory of Delinquency’, American Sociological Review, 22: 664–70. Umbreit, M., Coates, R. and Vos, B. (2002) Community Peacemaking Project: Responding to Hate Crimes, Hate Incidents, Intolerance, and Violence through Restorative Justice Dialogue. Minnesota: Center for Restorative Justice. Walters, M. (2005) ‘Hate Crimes in Australia: Introducing Punishment Enhancers’, Criminal Law Journal, 29(4): 201–16. Walters, M. (2006) ‘Changing the Criminal Law to Combat Racially Motivated Violence’, University Technology Sydney Law Review, 8: 66–86. Weinstein, J. (1992) ‘First Amendment Challenges to Hate Crime Legislation: Where’s the Speech?’, Criminal Justice Ethics, 11(2): 6–20. Young, J. (1999) The Exclusive Society: Social Exclusion, Crime and Difference in Late Modernity. London: Sage. Zehr, H. (1990) Changing Lenses: A New Focus for Crime and Justice. Scottsdale: Herald Press.

248

Index

Added to a page number ‘t’ denotes a table and ‘n’ denotes notes. ability, to enforce law 152, 165 Aboriginal communities 24 Abortion Act (1967) 91 accountability 232 acquaintances, victimisation by 110t, 126, 129t Action Plan (cross-government) 58 active citizenship 171, 187 age of consent 95n hate crime policy 61, 62, 68, 69, 70–1, 72, 75 of offenders 134t Age Concern 68 aggravated assault 128, 129t aggravated bodily harm 112t, 116, 117t, 119 Agreeing to Agree: A Proponent and Opponent of Hate Crime Laws Reach for a Common Ground 27 Ali, Azad 188 alternative subcultures 41, 47, 51, 54n

‘Alternatives Have Rights Too’ website 48, 54n ambiguity of motivation 138–9, 156 American Civil Liberties Union 27 animus-based violence 20, 22, 211 anti-gay violence 21–2, 24 anti-Islamic protests 2 anti-Muslim violence 25 anti-semitic violence frequency 106t lack of studies devoted to 23 locations 107t offences 108t, 109 verbal-textual hostility 114t, 115, 119 victim-offender relationship 109, 110t anti-semitism 116, 214 Anti-Terrorism, Crime and Security Act (2001) 59, 102, 106 anti-white violence 24 apologies 239 ‘arousing fear’ 83 arrests 129t, 130 Asian communities 2, 20, 24 assaults 51, 128 see also aggravated assault; common assault 249

Hate Crime

Association of Chief Police Officers (ACPO) 4, 59, 69, 86, 95n, 172, 173 assumptive world theory 22 attitudes inability of law to control 152 mapping of local 183–4 racist/homophobic 235 towards victims 234 Austin’s speech act theory 101–2, 113–14, 119, 120n authoritative illocutions 101 Baynham, Ian 2 Belfast (Good Friday) Agreement (1998) 80, 82 beliefs, inability of law to control 152 bias motivation 25, 125, 139, 154 bigotry 1, 2, 233, 234 black male offenders 126 black and minority ethnic see BME communities Blair administration 177, 178 blame 5, 235 BME communities hate crimes against defensive 133 government concern for reporting 106 lack of studies devoted to 23 location 107 offenders 126, 129t verbal-textual hostility 113, 115 police stereotypes 158 see also Aboriginal communities; Asian communities; Jews; Latinos; Muslim communities Boston Police Department 131, 133, 139 boundaries, policing of 43, 53 British Crime Surveys 195, 196–7, 206n British National Party (BNP) 2 Brown, Gordon 92, 178, 179 250

bullying 1, 7, 29 Bureau of Justice Statistics (US) 126 business colleagues, victimisation by 110t business environments, hate crime location 107–8 bystanders 138 calculative regimes 185 campaign groups disability and gay rights 63–4 hate crime policy exclusion of age and gender 72–3 influence of 66–8 use of ‘hate crime’ term 60–1, 75 Canada 25, 26 capacity-building (Prevent strategy) 180 case studies 21 Catholics, Northern Ireland 80, 90, 91 Channel Programme 189n Church of Ireland 95n Civil Partnerships Act (2004) 82, 90 clarity of proof 152 clear-cut hatred 61, 70 Client Centred Service Model 199 colonial governance 188 colonisation, and racist violence 24 Commission for Racial Equality 63 commitment to bias 134, 140 common assault 112t, 116, 117t, 119 communication, in RJ 234, 241 communities educative role, and policing success 164 organising resistance to racist groups 29 responsibility for responding to hate crime 6 in rural settings 19 studies on effects of hate crime 22 victimisation of ‘invisible’ 6

Index

see also BME communities; LGBT communities; vulnerable communities community cohesion 3, 64, 150, 183 Community Disorders Unit (Boston Police Dept) 131, 139 community engagement 178, 179, 180–1, 184, 185, 187 ‘community relations’ approaches 178 community safety 170–3 community safety practitioners 204 community-based responses 31 community-based theoretical models 18 community-derived intelligence 182 ‘competition for suffering’ 63 comprehensive social change 151 confidence, in police 165, 181, 197, 205 confrontational work 216 consent, age of 95n content analysis 20 CONTEST 177–86, 187, 188–9 contradictorily racist subject 214–17 Convention on the Prevention of Terrorism 2007 (Council of Europe) 178 Copeland, David 2, 12n, 42 cost-benefit calculation, crime reporting 196 Council of Europe Convention on the Prevention of Terrorism (2007) 178 council workers, inactivity 1–2 counter-terrorism 106, 177–86 crime(s) ‘hierarchy of police relevance’ 156–7, 204 see also hate crimes Crime and Disorder Act (1998) 27, 59, 65, 82, 102, 106, 170, 171, 175, 200, 210, 211, 229 crime prevention 170 Crime and Security Act (2001) 106

criminal damage 112t, 116, 117t, 119 criminal justice gatekeepers 203 and hate crime interventions 130 policy 58 responses 62, 229–30 student perceptions 20 RJ practitioners within 244 understanding of victimhood 74 Criminal Justice Act (2003) 59, 83, 102, 106, 173, 229 Criminal Justice and Immigration Act (2008) 95n, 173 Criminal Justice Inspectorate Northern Ireland 86, 87, 93 Criminal Justice (No. 2) (Northern Ireland) Act (2004) 83 Criminal Justice (Northern Ireland) Order (2004) 83 Criminal Law (Amendment) Act (1885) 81 criminal liability 153 criminalisation 81, 104t, 106, 154, 229 Critical Discourse Analysis (CDA) 99, 102, 111–18 Cronulla Beach ‘race riots’ 21 Crown Prosecution Service (CPS) 59, 60, 64, 154, 172, 180 culpability 132, 135–8, 141 cultural difference 47, 89, 233, 240 cultural geography 19 cultural racism 216 cultural studies 19 culture of hate 21 cyberhate 6, 26, 28–9 Data Protection Act (1998) 100 defensive hate crimes 133, 134t, 136, 139, 142 defensive offenders 137t defensiveness 223 demarginalisation 173 demonisation 1, 104t, 114t, 115, 117t 251

Hate Crime

denunciation 229 Department for Communities and Local Government (DCLG) 178–9, 183–4, 189n desire, to enforce law 152, 165 deterrence 134, 135, 143, 229 dialogic process, restorative justice 231, 232 difference engendering tolerance of 232–5 Goths and tolerance of 43, 44 and hate crime 29, 41, 47, 52, 53, 89 see also cultural difference; ‘dilemma of difference’; ‘doing difference’; identity difference; racial difference ‘dilemma of difference’ 25 Disability Advisory Group (MPS) 68 disability campaigners 63–4 disability hate crimes 59, 75 marginalised in ‘hate debate’ 7 prosecution 59 research 25 responses to 1–2, 62 sentencing 173 victim hierarchy 62 Disability Rights Commission 63 ‘disability targeted violence’ 61 disablism 240 discrimination, by police 157, 161 disputes 156 disturbances 156 Dobrowski, Jody 40, 51, 78 ‘doing difference’ 89 domestic violence 69, 70 Dudgeon, Jeffrey 81 educative role, policing success 164 elder abuse 69, 72 empathetic divide 234, 245n empathy 234 empirical approaches 19 encouragement, in attacks 127, 143 252

engagement of offenders in RJ 231 of public with police 165–6 see also community engagement England and Wales hate crime increase in recorded racist 161 legislation 149, 153 politicisation 159 prosecutions of racist 211 towards LGBT communities 78 law enforcement 164 police exercise of discretion 157 reluctance to enforce hate crime legislation 158 as white, male-dominated organisations 158 English Defence League 2 environmental factors, law enforcement 28 Equal Opportunities Commission 63 equality, Northern Ireland 81, 82, 94 Equality Commission survey (2009) 89–90 Equality and Human Rights Commission 61, 63 essentialism 233 ethnic minorities see BME communities Europe 8–9, 26, 27 Evans, Paul 174 Every Child Matters 182 evidence-driven knowledge systems 183 exclusionary practices 19, 89 exercitives 101 expatriation 104t, 114t, 115, 117t experimental studies 19–20 extremism 178, 180, 182, 184, 185, 186, 189 faith-based hate crimes frequency 106t locations 107t

Index

offences 108t offender characteristics 129t unreporting of 195 verbal-textual hostility 111, 114t, 115, 116, 119 victim hierarchy 62 victim-offender relationships 110t see also religiously aggravated crimes family group conferences 231–2 family members, victimisation by 110 fear of difference 47 of persons 83 of police 197 of reprisals 197t, 198 of victimisation 50 Federal Bureau of Investigation (FBI) 124, 125, 132 fellow travellers 135, 137t, 138, 143 female offenders 126 First Amendment (US Constitution) 27, 101 First International Conference on Radicalisation and Political Violence 179 focus group approaches 20 forensic linguistics 99, 102, 118, 119, 120 forgiveness, of homosexuality 90 Fourteenth Amendment (US Constitution) 27 free speech 27, 28, 100–1, 118 friends, victimisation by 110t, 126 gatekeepers 28, 155, 203 gay men estranged relationship with police 106 see also anti-gay violence; homosexuality; lesbian and gay community gay rights groups 63–4 Gellman, S.B. 27

gender 21, 61, 62, 63, 68, 69, 70, 71, 72, 75 ‘geography of hate’ 19 Gibbs, Paul 49 Global Counter-Terrorism Strategy Plan of Action 2006 (UN) 178 ‘good crimes’ 156 Good Friday Agreement (1998) 80, 82 goth subculture 43–5, 54n goth teenagers 48 goth victimisation 41 campaign for recognition as hate crime 47–8 comparison with other hate crimes 49–51 inclusion in hate crime legislation 52–3 nature and frequency of 48–9 see also Lancaster, Sophie ‘government through community’ 185 grievous bodily harm 112t, 116, 117t, 119 group classification 65 group subcultural identity 44 Guide to Identifying and Combating Hate Crime (ACPO) 173–5 guilt, about sexuality 91 harassment 1, 7, 29 as defensive hate crime 133 reported 198 verbal-textual hostility 112t, 117t, 119 see also racial harassment ‘hard’ engagement 181, 187 ‘hard to reach’ groups, community safety 170–3, 185 harm see hate harms Harris, Brendan 46 hate 3, 21, 159, 233 hate crime(s) 1 ACPO definition 4, 95n ACPO manual 59, 66, 69 253

Hate Crime

classification issue 6 consequences see hate harms as contested area of study and policy 3–4 cross-government strategy 75 greater awareness of 8 issues demanding enquiry 6–7 as a legal concept 124 need for fresh thinking about 7–9 offenders see offenders as policing priority 204 politicisation 159 recording 28, 83–6, 111–14, 157, 161 reporting see reporting responses to policing 1, 28, 62 in policy and scholarly domains 2–3 punitive/retributive 31, 229–30 responsibility for shaping 5–6 uncertainty surrounding effectiveness of 9 as a social issue 150 terminology operational usefulness 60–1 uncertainty surrounding 8–9 use of 41–3 verbal see hate speech victims see victims see also anti-semitic violence; disability hate crimes; faithbased hate crimes; goth victimisation; homophobic hate crimes; racist hate crimes; sexual orientation hate crimes Hate Crime in America: The Debate Continues 26 hate crime legislation 58, 81 balkanisation 63 categories 149–50 effect of social change on 151 enforcement see law enforcement inclusion of goth victimisation in 47–8, 52–3 254

limits of effective 151–3 literature 26–30 Northern Ireland 82–3 police reluctance to enforce 158 policy career 105, 120n racially aggravated offences 211–13 rationale behind 150 Scotland 59 United States 3, 27, 149, 153 victim groups included in 60 see also individual acts hate crime policy 58–76 campaign groups and influence 66–8 challenging assumptions 69–72 development, research findings 59–61 encouragement of communities to work with police 175 examination of future challenges 59 explaining exclusions 72–3 hierarchies within 61–4, 62 post-Macpherson measures 2–3 victim silos and diverse identities 64–6 hate crime scholarship 3, 17–32 organised hate groups 26 promoting 30–2 responding to hate crime 26–30 theorisation and approaches 18–22 victim groups 22–5 Hate Crimes 32 Hate Crimes: The Rising Tide of Bigotry and Bloodshed 131 Hate Crimes Project 232, 244, 245n The Hate Debate: Should Hate be Punished as a Crime? 26–7 hate groups 26 hate harms 52, 174–5, 230 capacity to respond to 119 reparation 231, 234, 236–9

Index

verbal-textual hostility 101–2, 116, 118 hate speech 22, 27, 83 see also verbal-textual hostility study hatred 28, 70, 71, 83, 92, 173, 186 hegemonic identity 42 hegemonic masculinity 88 Help the Aged 68 Herbert, Ryan 46 heroes 136, 137t, 138, 140, 143 heterosexual intercourse, age of consent 95n hierarchies of subject positions through naming 118 of victims 60, 61–4, 74 ‘hierarchy of police relevance’ 156–7, 204 HIV/AIDS, homophobia and 88 Hodkinson, Mark 40, 47, 54n Home Grown Hate: Gender and Organized Racism 26 homeless people 7, 21, 66 homophobia 79, 174, 240 goth victimisation 49–50 Northern Ireland 87–93 shaping of anti-gay violence 21 homophobic hate crimes 2 ACPO definition 173–4 frequency 106t harm reparation 236–8 hate crime policy 64, 65, 68, 75 high-profile media events 173 locations 107t Northern Ireland against LGBT communities 78, 79 legislation 83 prosecutions 87 recording 83–6 offences 108t, 109 policing of 187 verbal-textual hostility 111, 113, 114t, 119

victim hierarchies 62 victim-offender relationships 109, 110t victimisation surveys 175–6 victims 174–5 homophobic and transphobic hate crime 64, 172 homosexual intercourse, age of consent 95n homosexuality discourses of 87–8 Iris Robinson comment 78, 79, 83, 92 Northern Ireland and the law 81–7 opposition to 89–90 political condemnation 92–3 hostility, racially aggravated offences 212, 215 How to Do Things with Words 101 Hulme, Joseph and Danny 46 identity criminalisation of 106 punishment of 42–3, 222 reassertion of hegemonic 42 identity difference 233, 234 identity groups 62, 64–6, 230, 233 identity politics 151 identity theories 19 ill will 59 illocutionary speech acts 101–2, 113–14, 119, 120n Improving Opportunity, Strengthening Society 183 in terrorem effect 6, 119 incitement to hatred 83, 92, 173 inconvenience, and non-reporting 197t incremental social change 151 information sharing 183 infrequency, recognition of hate crimes 156 insiders 19 Institute for Conflict Research 85 255

Hate Crime

institutional homophobia 176 institutional prejudice 157–8, 161, 176 intelligence 182, 183 intelligence-gathering 181 intelligence-led policing 203 intensive interventions 182 intent 83, 153, 154 interdisciplinary hate studies 31–2 International Crime Victims Survey 195–6 interpellation 104t, 114, 115, 116, 117t intersubjective relationships 103 interview techniques 20–1 ‘intolerance of intolerance’ 186 invisible communities 6, 24 Islamophobia 65, 184, 188, 189 Jews 20, 25, 115, 126 see also anti-semitic violence Johnson, Alun 58 Johnston, Sgt Detective 139–40 Journal of Hate Studies 31 juvenile offenders 31, 128, 129t, 133 juvenile victims 129t, 175 Kelly, Ruth 185 knowledge systems (PVEPF) 182–3 Lancaster, Sophie 40–1, 43, 45–7, 51, 53 language and justification 235 responsiveness to, in RJ 242 see also forensic linguistics; sociolinguistics Latinos 24, 25, 126 law enforcement 149–66 ability and desire 152, 165 agencies 126, 152, 188 practical difficulties 163 problem of motive 153–5 recording of hate crime 28 256

student perceptions’ of hate crime 20 see also police; policing Lawrence, Frederick 27 Lawrence v. Texas 21 leaders 135, 136–8, 143 Lesbian Advocacy Services Initiative (LASI) 85–6, 89 Lesbian and Gay Anti-Violence Project 113 lesbian and gay community as hard-to-reach 171–2 police attempts to break down barriers with 172–3 verbal-textual hostility 115 victimisation by family members 110 see also LGBT communities lesbian women 85–6 Lewington, Neil 2, 12n LGBT communities definition 12n hate crimes against effects of 21–2 heavy concentration of 24 research 24 Northern Ireland see Northern Ireland and police lack of trust in 172 presentation in discourses 169–70, 177 trust-building 180 research 19 see also gay men; police-LGBT communities anti-hate strategy liberal political theory 100 libertarian values 27 local authorities 5, 170, 179, 203–4 ‘local community’ knowledge 182 Local Government Act (1999) 170, 171 Local Policing Plan 180–1 location of crimes 107–8, 134t locutionary acts 101

Index

low-level bullying 7 McDaid, Kevin 93 Macpherson Report 42, 201 see also Stephen Lawrence Inquiry McVeigh, Timothy 2, 12n, 143n male goths, homophobic abuse 49–50 malice 59 Mallett, Daniel 46 Maltby, Robert 40–1, 45, 46, 47, 51, 53 Martin, Brent 40 Mediated Discourse Analysis (MDA) 102 mediation 231–2 Members of the Legislative Assembly (MLAs) 81 mental ill health 174–5, 230 message crimes 42, 47 Metropolitan Police Service (MPS) 68, 82, 99, 100, 161, 195 minority groups community safety initiatives 171 and police belief in unwillingness to investigate 161 lack of trust in 171 prejudice 157–8 violence against 47, 89, 233 see also BME communities; goths; LGBT communities mission hate crimes 134t, 136, 142 mission offenders 134, 135, 137t, 142, 198 moral conservatism 89–91, 92 moral panics 21 Morgan, Jo 66, 67 Morgan Report (1991) 170 Morley, David 78 Mosher 54n motivation(s) ambiguity and complexity of 138–9, 156 and culpability 135–6, 137t, 141

data limitations 139 identifying primary 140 law enforcement and problem of 153–5 multiple 139, 140, 156 police reluctance in decisionmaking 158–9 and prosecution 141 and typologies 131–2 see also racial motivation Mubarek, Zahid 213 multi-agency support 182 multiculturalism 21, 233 multiple motivations 139, 140, 156 multiple offenders 127, 129t, 134t, 135 Muslim communities counter-terrorism and disciplinary practice against 106 in police discourses 170 policing 177–86, 187 suspect communities 178, 189 threats of death and violence 115 see also anti-Islamic protests; antiMuslim violence; Islamophobia Muslim Safety Forum 188, 190n naming the addressee 104t, 114, 115, 116, 117t, 118, 119 National Hate Crime Statistics 124–5 National Incident-Based Report System (NIBRS), using to understand offenders 125–32 ‘neglected decision-maker’ 203 Neighbourhood Policing Teams (NPTs) 203 neighbours, victimisation by 110t, 126–7 net-widening partnership 184 neutralisation, techniques of 235 ‘non-guilty’ pleas 212 non-police reporting centres 204 257

Hate Crime

non-traditional interventions 29–30 Northern Ireland homophobia 87–93 homosexuality and the law in 81–7 LGBT communities homophobic hate crime against 78, 79 legislative protection 83 recognition as equal citizens 94 silence surrounding attacks on 93 stakeholders in social and political initiatives 82 the ‘new’ 80–1 Northern Ireland Life and Times (NILT) survey 90 Northern Ireland Office (NIO) 81 occupational culture, policing 157, 161 offenders 4–5, 124–43 benefit from RJ 244 motivation see motivation(s) punishment 243–4 research 21, 30–1, 124–5 typology 132–6 1993 original 131–2 critique of 138–41 evolution of original 141–3 expansion of culpability roles 136–8 using NIBRS to understand 125–32 and victims ‘empathetic divide’ 234, 245n social distance between 240, 241 see also juvenile offenders; racially aggravated offenders; victimoffender relationships Office for Democratic Institutions and Human Rights (ODIHR) 149 258

Office of the First Minister and Deputy First Minister (OFMDFM) 81 On Liberty 100 ‘one size fits all’ solutions 5 ‘one size fits all’ view of victims 199 one-to-one work, with offenders 213–14 online petitions 47–8, 92 organisational resources, policing 161–2 organised violence 1, 2, 8–9, 19, 26 Organization for Security and Cooperation in Europe (OCSE) 149 ‘other accepted crime’, verbaltextual hostility 112t, 116, 117t, 118 other/othering 1, 4, 5, 22, 47, 118, 233, 235, 240 outgroups 52, 53 outsiders 19, 47, 235 over-policing 161, 187 Paisley, Ian (Jr) 93 Paisley, Ian (Rev) 81 paranoid splitting 216, 223 Parkes, James 2 partnerships against terrorism and extremism 178, 179, 180, 184 community safety 170–1 pathologisation 104t pathways to crime 142 patrol officers, decision-making 155–6 Patten Report 80 perceptions, scenarios and 20 perlocutionary speech acts 101, 102 perpetrators see offenders persistence of crime, and reporting 201 personal resources, policing 162 physical injury 128, 129t, 132, 196, 198 physical violence see assault

Index

Pilkington, Fiona 1–2, 7 police attempts to break down barriers with lesbian and gay communities 172–3 confidence in 165, 181, 197, 205 data 139 as decision-makers 155–9, 160 discretion 155, 157, 160 dislike of, and non-reporting 197t, 198 engagement approaches 181 estranged relationships with 106 investigations 160, 161 law enforcement 152, 163–4 as policy-makers 159–63 prediction and anticipation of violence 142 problem of identifying motive 154 recorded hate crime figures 198 requirement to provide proof and intent 154 responses to hate crime 1, 28, 62 subject definition of racist incident 210 trust-building 169, 179–80 trust/lack of 106, 165, 171, 172, 181 Police Service of Northern Ireland (PSNI) 79, 80, 83, 86, 87, 92, 94 police-LGBT communities anti-hate strategy 175, 187 policing of boundaries 43, 53 modern-day 5 Muslim communities 177–86, 187 Northern Ireland 80 success of 164 vulnerability in context of 169–70, 177 see also intelligence-led policing; over-policing policy and religious doctrine 92

see also hate crime policy policy career, hate crime legislation 105, 120n policy-makers, police as 159–63 politicisation, hate crime 159 ‘politics of hate’ 159 post 9/11 hate crime scholarship 17–32 UK counter-terrorism strategy 178 power imbalances, RJ conferencing 239–40 power relations, maintenance of 42 prejudice 1, 2, 50, 125, 230, 233 see also homophobia; institutional prejudice; societal prejudice Prevent strategy (CONTEST) 177–8, 180, 181, 182, 183, 184, 185, 186, 187, 188–9 prevent-related intelligence 183 preventative interventions 183 Preventing Violent Extremism Pathfinder Fund (PVEPF) 178, 179, 180, 182–3, 184, 185, 187 Preventing Violent Extremism – Winning Hearts and Minds 179 Prevention of Terrorism Act (2005) 106 priority areas (Prevent) 188 prison officers, racism 213 privacy, and non-reporting 197 privilege, victimisation and maintenance of 42, 53 Probation Inspectorate 213 profanity/profane naming 104t, 114, 115, 116, 117t, 118, 119 Programme for Government documents 81 projection/projective identification 18, 216 proof 83, 152, 154 propaganda theory 18 proportionality 229 Prosecuting Crimes Against Older People (CPS) 74 259

Hate Crime

prosecution(s) difficulty of successful 154 disability hate crime 59 lack of literature on 28 motivation and culpability 141, 142 Northern Ireland gay men 81 homophobic hate crimes 87 no requirement to prove intent 83 racially aggravated offences 211 see also Crown Prosecution Service; Public Prosecution Service protective inclusionism 187 Protestantism 90–1, 95n psychological damage 174–5, 230 psychological studies 18, 19 psychosocial approach, to racism 215–17, 225 psychosocial perspective, racial motivation 222–4 public trust and confidence in police 165 willingness to engage with police 165–6 public buildings, hate crime within 107 Public Order Act (1986) 95n, 102, 173 Public Order (Northern Ireland) Order (1987) 83, 92 Public Prosecution Service (PPS) 87, 94 public security 187 public spaces, hate crime in 107 punishment homophobic violence as 174 of identity performance 42–3, 222 of offenders 243–4 punitive responses 31 punks, assaults on 51 qualitative research 20–1 260

quantitative research 19–20 race 21, 23, 199, 216 ‘race inequality’ strategy 183 Race for Justice project 74–5 Race Relations (Amendment) Act (2000) 200 ‘race riots’, Cronulla Beach 21 racial difference 216 racial harassment harm reparation 238–9 ineffective facilitation 242–3 non-reporting 212 racial motivation 198, 209–10, 213–14 reading 221–4 racial violence 18, 19, 23, 24, 27 Racial Violence in the United States 18–19 racially aggravated offences 83 advent of 209–11 law 211–13 legislation 59, 200 motivation 213–14 reporting 196 racially aggravated offenders 198 case of Alan 217–20 confrontational work with 216 contradictorily racist subject 214–17 one-to-one work with 213–14 racism 6, 94, 198, 212, 213–14, 215–17, 240 Racist Attacks (Home Office) 209 racist hate crimes ACPO definition 173–4 as an ongoing process 162 frequency 105–6 in ‘hierarchy of police relevance’ 156–7 high-profile media events 173 increase in recorded 161 locations 107t offences 108t offender characteristics 129t

Index

rejection of provision to encompass religion 65 reporting 195 reporting of 106 verbal-textual hostility 114t, 115, 116 victim hierarchies 62 victim-offender relationships 110t racist incidents 157, 200, 209, 210 Racist Victimisation: International Reflections and Perspectives 27 radicalisation 178, 182, 183 rank-and-file officers, perception of racist incidents 157 rationalisations 235 reactive offenders 198 recording, of hate crimes 28, 83–6, 111–14, 157, 161 rehabilitation programmes 182 Reid, John 185 relatives, victimisation by 126 religious doctrine, policy and 92 religiously aggravated crimes 59, 83, 196, 210 see also faith-based hate crimes reparation 231, 234, 236–9 repeat victimisation, and nonreporting 198 reporting hate crime 106, 194, 195–9 encouraging 201–4 legacy of Lawrence Inquiry 199–201 racially aggravated offences 211 verbal-textual hostility 111–14 ‘required community consultation process’ 171 residential areas, hate crime location 107t resilience, against extremism 180, 181, 185, 189 resources, policing 161–2 Responding to Hate Crime (US Dept of Justice) 132

responsibility 232, 235 restorative justice 228–45 dangers of ineffective facilitation 240–3 engendering tolerance of difference 232–5 facilitators 241–2 and hate crimes 230–2 meetings 232–3, 234, 239, 244 referrals 245n research 30, 228 vulnerability and power imbalances in conferencing 239–40 retaliatory hate crimes 133–4, 134t, 136, 142 retaliatory offenders 137t retributive response 229–30 revisionist psychoanalytic work 215–16 revolutionary social change 151 right-wing violence 2, 8–9 ripple effects 22 Robinson, Iris 78, 79, 83, 92, 93 Royal Ulster Constabulary (RUC) 80, 81 ‘rubbish crimes’ 156 Rural Racism 19 Russell, Anthony (QC) 40–1, 46 Rustin, Mike 216 safe environment, in RJ process 237, 241 Save Ulster from Sodomy 81 scenarios, and perceptions 20 scholarly interest, in hate crime 2, 3 school-based anti-hate programming 29 Scotland hate crime legislation 59 hate crime policy 68 Scotland, Baroness 74 Scottish Executive 59, 68, 70, 76n secondary victimisation 240 sectarianism 80, 94 261

Hate Crime

self-perception, and harm 118 sentence enhancement 64 sentencing guidelines 54, 83, 173 serious injury 129t seriousness of crime, and reporting 196–7, 201 service provision, and vulnerable communities 176–7 sex workers, exclusion from hate crime policy 66 sexual citizenship 172–3 Sexual Offences (Northern Ireland) Order (2007) 95n sexual orientation hate crimes growing interest in 24–5 legislation 59, 82 offender characteristics 129t sentencing 83, 173 sexualisation 104t, 114t, 115, 116–17, 117t Shepard, Matthew 21, 22, 82 ‘shuttle’ mediation 232 single offenders 127 single speech acts 113 single victims 127 slippage, between policy and practice 160 Smith, Jacqui 178, 179, 181, 182, 185, 189n social change 151 social cohesion 150 social conservatism 47 social distance 240, 241 social influence, decision to report 203 social marketing materials, encouragement of reporting 202–3 social psychology 88 social relations, law and 152–3 social theory 18–19 societal prejudice 158 socio-legal reading, racial motivation 221 sociolinguistics 102 262

sociological perspective, racial motivation 221–2 ‘soft’ engagement 181, 187 Southern Poverty Law Center 142 Southwark Mediation Centre 232, 244, 245n spectators 143 speech see free speech; hate speech speech act theory 101–2, 113–14, 119, 120n splitting 216, 223 spontaneity of incidents 127 Stamp Out Prejudice, Hatred and Intolerance Everywhere campaign 48–9 Stephen Lawrence Inquiry 3, 58, 105, 106, 161, 173, 176, 195, 199–201 stereotypes 158, 214, 235, 236 stereotyping 233 Stonewall 67, 85 Stormont Assembly 81 strain theory 18 ‘stranger danger’ 42, 47, 70 strangers, victimisation by 126, 129t street-level bureaucrats 157–8, 159–63, 164 structured action theory 215, 221, 222 student perceptions 20 subcultures see alternative subcultures; youth subcultures subjectivity 118 subordination 118 support, for vulnerable people 181–2 supporters, in RJ 241 surrogate role, of hate speech 119 surveillance 185, 187 suspect community, Muslims as a 178, 189 suspicion of others 240 symbolic violence 120 sympathisers 143

Index

Tackling Hate Crime in Europe conference 74 techniques of neutralisation 235 terrorisation 104t, 114t, 115–16, 117t, 119 terrorism 183 Terrorism Act (2006) 106 thrill crimes 133, 134t, 136, 139, 140, 142 thrill offenders 137t, 142, 198 training curriculum (FBI) 132 transference 18 transgender/transsexuals, violence against 24–5 transphobic hate crime 64, 65, 172 transport, hate crime location 107t trivial victimization, non-reporting 197, 212 Troubles, the 80 trust 106, 165, 171, 172, 176, 181 trust-building 169, 172, 179–80 UN Global Counter-Terrorism Strategy Plan of Action (2006) 178 uncertainty, recognition of hate crimes 156 unconscious racist stereotypes 214 under-reporting 86, 200 Uniform Crime Reporting programme (FBI) 125 United Kingdom see England and Wales; Scotland United States constitutional amendments 27, 101 cyberhate 26 hate crime legislation 3, 27, 149, 153 police 157, 158 politicisation of hate crime 159 unwilling participants 135–6, 137t, 138, 143 vandalism 197, 198

verbal bullying/victimisation 29, 86 verbal-textual hostility study 99–121 approach 100–3 coding 105t context 105–11 source 103–4 text 111–18 themes 104t verdictives 101, 113 ‘vicarious traumatisation’ effect 21–2 victim hierarchies 60, 61–4, 74 victim interchangeability 69 victim reports 164 victim services 30 victim silos 64–6 victim–offender history 134t victim–offender mediation 231–2 victim–offender relationships 5, 69–70, 109–11, 126, 129t victimisation Aboriginal communities 24 generalisations 4 highlighting of different forms of 60 see also goth victimisation; repeat victimisation; secondary victimisation victimisation surveys, homophobic 175–6 victims 4 crime reporting see reporting distinction between violent and non-violent crime 201 estranged relationship with police 106 hate crime legislation 60 hate crime policy 61–8 hate crime scholarship 22–5, 30 of homophobic incidents 180 interest in 194 and offenders ‘empathetic divide’ 234, 245n social distance between 240, 241 263

Hate Crime

‘one size fits all’ view 199 single 127 ‘vigilant or watchful visuality’ 185 Violence Against Women unit (Scotland) 68, 70 violent crime, victims’ distinction between non-violent and 201 vulnerability in context of policing 169–70, 177 distinction between hatred and 70, 71 in hate crime conferencing 239–40 to extremism 182, 186 to hate crime 230 victimisation and perceived 127 vulnerable communities bigotry against 1 experimental studies 20 identification of 180–1 inclusion of ‘disabled’ category 25 lack of knowledge of experiences of 7 obstacles to service provision 176–7 vulnerable individuals, support for 182

264

Wales see England and Wales Walker, Anthony 40 weapons, use of 127, 129t, 134t Weinberg, Dr George 88 Where Do We Go From Here? Future Directions in Hate Crime Scholarship 17 white male/female offenders 126 white supremacy 21, 26, 116, 121n Wolfenden Report 81 women see female offenders; lesbian women women’s groups absence in hate crime debate 73 involvement in hate movement 26 Working Group on Hate Crime (Scotland) 68, 70–1, 76n xeno-racism 6 yobishness 156 young people see juvenile offenders; juvenile victims youth subcultures attacks on members 50–1 see also goth subculture