Blood, Threats And Fears: The Hidden Worlds Of Hate Crime Victims 3030319962, 9783030319960, 9783030319977

This book offers unparalleled insight into the ways in which hate crime affects individuals and communities across the w

488 130 2MB

English Pages 178 Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Blood, Threats And Fears: The Hidden Worlds Of Hate Crime Victims
 3030319962,  9783030319960,  9783030319977

Table of contents :
Acknowledgements......Page 6
Praise for Blood, Threats and Fears......Page 8
Contents......Page 9
Part I Setting the Scene......Page 11
Abstract......Page 12
References......Page 16
Abstract......Page 19
Conceptualising Hate Crime......Page 20
Hypervisible: Recognising ‘Difference’ and ‘Vulnerability’......Page 24
Rendered Invisible: Disempowered and Disengaged......Page 26
Conclusion......Page 28
References......Page 29
Abstract......Page 32
The Domestic Context......Page 33
The International Landscape......Page 36
Signs of Failure......Page 39
References......Page 44
Part II Undertaking Research with Hate Crime Victims......Page 48
Abstract......Page 49
Connecting with Hidden Voices......Page 50
Connecting with Our Sample......Page 60
Connecting with Our Data......Page 63
Conclusion......Page 68
References......Page 69
Abstract......Page 73
Acknowledging Positionality......Page 74
Acknowledging Resistance......Page 77
Acknowledging Resilience......Page 79
References......Page 82
Part III Revealing Hidden Problems......Page 85
Abstract......Page 86
Targets of Hate Crime......Page 87
Forms of Hate Crime......Page 94
Conclusion......Page 100
References......Page 101
Abstract......Page 102
Perpetrators of Hate Crime......Page 103
Drivers of Hate Crime......Page 109
Conclusion......Page 114
Abstract......Page 115
Physical Harms......Page 116
Emotional Harms......Page 118
Wider Harms......Page 123
Coping Strategies......Page 127
References......Page 130
Abstract......Page 132
Barriers to Reporting......Page 133
Barriers to Justice......Page 140
Barriers to Support......Page 144
Conclusion......Page 148
References......Page 149
Part IV Transforming Responses......Page 150
Abstract......Page 151
Defining Hate Crime......Page 152
Theorising Hate Crime......Page 154
Researching Hate Crime......Page 157
References......Page 160
Abstract......Page 162
Improving Front-Line Responses......Page 163
Improving Justice Outcomes......Page 167
Improving Support Provision......Page 169
Conclusion......Page 172
References......Page 173
Index......Page 174

Citation preview

PALGRAVE HATE STUDIES

Blood, Threats and Fears The Hidden Worlds of Hate Crime Victims Stevie-Jade Hardy Neil Chakraborti

Palgrave Hate Studies Series Editors Neil Chakraborti School of Criminology University of Leicester Leicester, UK Barbara Perry Faculty of Social Science and Humanities University of Ontario Oshawa, ON, Canada

This series builds on recent developments in the broad and interdisciplinary field of hate studies. Palgrave Hate Studies aims to bring together in one series the very best scholars who are conducting hate studies research around the world. Reflecting the range and depth of research and scholarship in this burgeoning area, the series welcomes contributions from established hate studies researchers who have helped to shape the field, as well as new scholars who are building on this tradition and breaking new ground within and outside the existing canon of hate studies research. Editorial Advisory Board Tore Bjorgo (Norwegian Institute of International Affairs) Jon Garland (University of Surrey) Nathan Hall (University of Portsmouth) Gail Mason (University of Sydney) Jack McDevitt (Northeastern University) Scott Poynting (The University of Auckland) Mark Walters (University of Sussex) More information about this series at http://www.palgrave.com/gp/series/14695

Stevie-Jade Hardy · Neil Chakraborti

Blood, Threats and Fears The Hidden Worlds of Hate Crime Victims

Stevie-Jade Hardy School of Criminology University of Leicester Leicester, UK

Neil Chakraborti School of Criminology University of Leicester Leicester, UK

Palgrave Hate Studies ISBN 978-3-030-31996-0 ISBN 978-3-030-31997-7  (eBook) https://doi.org/10.1007/978-3-030-31997-7 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Palgrave Pivot imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Acknowledgements

As we write these acknowledgements we do so with a palpable sense of relief and with enormous gratitude to those who made it possible. We had wanted to write this book for some years prior to now but a combination of things kept getting in the way: new roles, new research projects, new family members, new challenges, new commitments … and new excuses. Fortunately, we managed to get our heads together, signed a contract and here we are writing our acknowledgements. Finally. Hurrah! This book simply couldn’t have been written without the consent, support and inspiration provided by our research participants. We speak at length within this book about the challenges associated with undertaking sensitive fieldwork of this nature but these pale into irrelevance when set against the kinds of challenges that many thousands of hate crime victims have to contend with on a daily basis. We are enormously grateful to our research participants for sharing their stories with us and for allowing us into their lives. We hope that you, the reader, can take as much inspiration as we did from their narratives. Thanks to those who funded the research that underpins this book— namely the Economic and Social Research Council, the Equality and Human Rights Commission, and the Offices for the Police and Crime Commissioner in Hertfordshire and the West Midlands. Thanks also to the large number hate crime practitioners, policy-makers, scholars, researchers and activists whom we have been lucky enough to work with over recent years. You have all shaped our thinking enormously, as have v

vi  

ACKNOWLEDGEMENTS

our colleagues Chris Allen and Fay Sadro at the Centre for Hate Studies, the fantastic team at the School of Criminology, University of Leicester and the brilliant students whom we teach (and learn huge amounts from) every year on our Hate Crime module. Finally, a quick mention to our families for their limitless love and support. Writing about horrible themes during horrible times can take its toll, and it’s our families who keep us smiling.

Praise

for

Blood, Threats and Fears

“This is a wonderful book.” —I. M. Smart, Professor, Political Science, Harvard University, USA

vii

Contents

Part I  Setting the Scene 1

Increasing Problems, Increasing Indifference 3

2

Visible yet Invisible: Challenges Facing Hate Crime Victims 11

3

Relevant yet Irrelevant: Challenges Associated with Hate Crime Policy 25

Part II  Undertaking Research with Hate Crime Victims 4

The Process of Engagement: Hard to Reach or Easy to Ignore? 43

5

Lessons from the Field 67

ix

x 

CONTENTS

Part III  Revealing Hidden Problems 6

Everyday Hate 81

7

Everyday Contexts 97

8

Invisible Harms 111

9

Invisible Victims 129

Part IV  Transforming Responses 10 Implications for Scholarship 149 11 Implications for Policy 161 Index 173

PART I

Setting the Scene

CHAPTER 1

Increasing Problems, Increasing Indifference

Abstract  This chapter highlights the originality, significance and need for this book by referring to the rising levels of hate and extremism both online and offline, the physical, emotional and community-level harms associated with hate crimes, and to the growing scepticism towards the concept of hate crime and ignorance of the harms associated with it. Keywords  Hate crime

· Extremism · Victims · Perpetrators

Hate crime constitutes one of the biggest global challenges of our time and blights the lives of millions of people across the world. The term ‘hate crime’ has been used within the domains of policy and scholarship as a way of distinguishing those forms of violence and micro-aggressions which are directed towards people on the basis of their identity, ‘difference’ or perceived vulnerability. The coining of a collective descriptor for these forms of victimisation has acted as a catalyst for improved awareness, understanding and responses amongst a range of different actors including lawmakers, non-governmental organisations, activists and professionals within and beyond the criminal justice sector. Importantly, it has facilitated increased prioritisation across disciplines, across communities and across borders. The need for such prioritisation has become all the more urgent amidst a context of rising levels of hate and extremism. As stated by © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_1

3

4  S.-J. HARDY AND N. CHAKRABORTI

Commissioner Věra Jourová at the Launch of the EU High Level Group on combating racism, xenophobia and other forms of intolerance in 2016: Over recent years, racism, xenophobia and other forms of intolerance have been growing and spreading across Europe at very high speed … there is also an exponential spread of hate speech on online fora, including social media and chats.

Within the United Kingdom (UK), 103,379 hate crimes were recorded by the police in England and Wales in 2018/19, which was not only an increase of 10 per cent compared to the previous year but it was also a continuation of an upward trend since 2012/13, with recorded hate crime having more than doubled in that time frame (Home Office 2019). While this rise is likely to be the result of a culmination of factors—including increased reporting and improved recording—‘trigger’ events of local, national and international significance have influenced the prevalence and severity of hate-fuelled violence and micro-aggressions (Littler and Feldman 2015; Hanes and Machin 2014; King and Sutton 2014). For instance, the EU referendum result of June 2016 led to an upsurge in reports of hate crime with more than 14,000 recorded by police forces in England and Wales between July and September 2016, which amounted to record levels of hate crime for three-quarters of police forces (Home Office 2018). ‘Trigger’ events such as the EU referendum also lead to a proliferation of hate online (Awan and Zempi 2017; Burch 2017; Williams and Burnap 2015), with social media ‘acting as a force-amplifier for cyberhate as it can open up a potential space for the rapid galvanising and spread of hostile beliefs, via the spread of rumours through online contagion’ (Williams and Pearson 2016: 7). Equally worrying spikes in hate crime have been observed further afield, including escalating tensions during and after the 2016 presidential campaign within the United States of America (USA) (Human Rights Watch 2019; Southern Poverty Law Centre 2016). In the ten days following the election of Donald Trump, 867 hate crimes were reported to the Southern Poverty Law Centre (SPLC), many of which were targeted towards ethnic and religious minority groups, thereby emulating the hostile and degrading discourse espoused by Trump during the election campaign (ibid., 2016). Populist parties have also experienced success across Europe, with othering, xenophobia and

1  INCREASING PROBLEMS, INCREASING INDIFFERENCE 

5

scapegoating permeating the political landscape in countries such as Austria, Denmark, France, Germany, Greece, Holland, Hungary, Italy, Norway, Poland, Slovenia and Sweden (Human Rights Watch 2019; Dearden 2017; Lazaridis et al. 2016). Similar to the presidential election in the USA and the EU Referendum in the UK, these narratives have fermented a fertile environment for the commission of hate crime, as exemplified by substantial increases in anti-Semitic and anti-Muslim hate crime in France, a rise in targeted violence towards refugees and refugee shelters in Germany and Greece, and a sharp increase in racist hate crime towards immigrants, minority ethnic and Roma communities in Italy and Hungary (Human Rights Watch 2019). The growth of hate and extremism in online and offline environments paints a worrying picture, not least because of the considerable damage they are known to cause. On an individual level, hate crime has been found to have a more harmful effect upon a victim’s emotional and physical well-being when compared to non-hate motivated crimes (Paterson et al. 2019; Vedeler et al. 2019; Iganski and Lagou 2015). According to the 2015/16 and 2017/18 Crime Survey for England and Wales, hate crime victims are more likely than victims of crime overall to cite being emotionally affected by the incident (89 per cent and 77 per cent respectively) and more likely to be ‘very much’ affected (36 per cent and 13 per cent, respectively) (Home Office 2018). Additionally, hate crime victims are more than twice as likely to have experienced a loss of confidence, difficulty sleeping, anxiety, panic attacks and depression when compared with victims of crime overall (Home Office 2018). The implications of suffering mental ill-health are far-reaching, often resulting in a decline in physical health, educational achievement and work productivity. It is also now widely acknowledged that the impacts of hate crime extend well beyond the actual victim, transmitting a sense of apprehension and vulnerability amongst family and community members (Walters et al. 2019; Bell and Perry 2015; Perry and Alvi 2012). As found in research by Paterson et al. (2019), knowing someone whom you share an identity or lifestyle characteristic with and who has been subjected to hate crime can also lead to indirect impacts such as an escalation in feelings of fear, anger and shame, and changes to everyday practices. Both directly and indirectly, hate-fuelled violence and micro-aggressions have the propensity to reinforce long-standing social divisions, a situation which is compounded further by those who seek to exploit these

6  S.-J. HARDY AND N. CHAKRABORTI

fractures to advance a particular ideology or to legitimise hateful and extremist views (Greater Manchester Preventing Hateful Extremism and Promoting Social Cohesion Commission 2018; Casey 2016). The originality, significance and need for this book become all the more evident at a time when scepticism towards the concept of hate crime and ignorance of the harms associated with it are becoming ever more palpable. As illustrated by the quotation below, attempts to devalue, disparage and deny the pervasiveness of hate crime not only reinforce the sense of isolation and marginalisation felt by many hate crime victims but also seek to silence their voices and to invalidate their experiences. Britain is in the grip of an epidemic, apparently. An epidemic of hate. Barely a day passes without some policeman or journalist telling us about the wave of criminal bigotry that is sweeping through the country … what the BBC calls an ‘epidemic’ is a product of the authorities redefining racism and prejudice to such an extent that almost any unpleasant encounter between people of different backgrounds can now be recorded as ‘hatred’ … According to one leftie online magazine, Britain now evokes ‘nightmares of 1930s Germany’. But this doesn’t square with the reality of our country today, and you shouldn’t believe it. The hate-crime epidemic is a self-sustaining myth — a libel against the nation. (O’Neill 2016) To be honest the term “hate crime” was cooked up by the extreme Left as a whip to crack over white British heterosexual Christians, and it’s just another way of gagging free speech. If you call someone a twat, a bastard or a wanker it’s no big deal, people can brush it off, so why is it different if you call someone a nigger, a poof or a whore? Message received by the authors via Facebook

The prevailing shallow, repetitive cycle which sees leading figures simply condemn hate crimes as ‘utterly unacceptable’ and as ‘having no place in our society’, to coin recent phrases used by the then Home Secretary Amber Rudd, does little to stem this growing tide of cynicism (Home Office 2016: 5). After having spent more than fifteen years investigating this phenomenon and hearing from thousands of hate crime victims, many of whom are living in despair, scared to do their weekly shop, to drop their children at school or to catch a bus, we felt compelled to write this book to show that hate crime is a very-real, repetitive

1  INCREASING PROBLEMS, INCREASING INDIFFERENCE 

7

and damaging problem. Without urgent action, hate crime victims will continue to encounter violence and micro-aggressions; to be met with incredulity and apathy; to reject opportunities to report their experiences; to become increasingly detached from support structures; and to have little faith in criminal justice responses. This book draws from a series of empirical research studies which have generated testimonies from more than 2000 victims of hate crime who come from different backgrounds and who are based within different parts of the UK. These testimonies offer unparalleled insights which transform our understanding of the ways in which victims experience and respond to hate incidents; the physical, emotional and community-level harms associated with hate crimes; and the implications for justice in the context of punitive, restorative, rehabilitative and educative interventions. In addition to generating new knowledge on victim experiences and expectations, the book is designed to shape innovation in hate crime research methodology and policy formation. We document the sensitivities, subjectivities and practicalities associated with undertaking complex research of this nature. In doing so, we offer an authentic account of the very necessary—and sometimes unconventional—steps which are fundamental to the process of engaging with ‘hard-to-reach’ groups and undertaking harrowing fieldwork. We also give specific focus to the policy implications of our research in highlighting ways to generate improved responses to hate crime from within the spheres of criminal justice, health, social care and education.

References Awan, I., & Zempi, I. (2017). ‘I will blow your face off’: Virtual and physical world anti-Muslim hate crime. British Journal of Criminology, 57(2), 362–380. Bell, J. G., & Perry, B. (2015). Outside looking in: The community impacts of anti-lesbian, gay, and bisexual hate crime. Journal of Homosexuality, 62(1), 98–120. Burch, L. (2017). ‘You are a parasite on the productive classes’: Online disablist hate speech in austere times. Disability & Society, 33(3), 392–415. Casey, L. (2016). The Casey Review: A review into opportunity and integration. London: Department for Communities and Local Government. Dearden, L. (2017). Attacks on refugee homes double in Austria as accommodation firebombed and sprayed with Nazi graffiti. http://www.independent.co.uk/ news/world/europe/refugee-crisis-austria-migrants-asylum-seekers-homesattacksfirebombings-doubles-accomodation-a7661831.html.

8  S.-J. HARDY AND N. CHAKRABORTI Greater Manchester Preventing Hateful Extremism and Promoting Social Cohesion Commission. (2018). A shared future. Manchester: Greater Manchester Preventing Hateful Extremism and Promoting Social Cohesion Commission. https://www.greatermanchester-ca.gov.uk/media/1170/preventing-hateful-extremism-and-promoting-social-cohesion-report.pdf. Hanes, E., & Machin, S. (2014). Hate crime in the wake of terror attacks: Evidence from 7/7 and 9/11. Journal of Contemporary Criminal Justice, 30(3), 247–267. Home Office. (2016). Action against hate: The UK Government’s plan for tackling hate crime. London: Home Office. Home Office. (2018). Hate crime, England and Wales, 2017/18. London: Home Office. Home Office. (2019). Hate crime, England and Wales, 2018/19. London: Home Office. Human Rights Watch. (2019). World report 2019. New York: Human Rights Watch. Iganski, P., & Lagou, S. (2015). The personal injuries of ‘hate crime’. In N. Hall, A. Corb, P. Giannasi, & J. G. D. Grieve (Eds.), The Routledge international handbook on hate crime (pp. 34–46). London: Routledge. King, R. D., & Sutton, G. M. (2014). High times for hate crimes: Explaining the temporal clustering of hate motivated offending. Criminology, 51, 871–894. Lazaridis, G., Campani, G., & Benveniste, A. (2016). The rise of the far right in Europe: Populist shifts and ‘othering’. London: Palgrave. Littler, M., & Feldman, M. (2015). Tell MAMA reporting 2014/2015: Annual monitoring, cumulative extremism, and policy implications. Teesside: Teesside University Press. O’Neill, B. (2016). Britain’s real hate crime scandal. https://www.spectator. co.uk/2016/08/the-real-hate-crime-scandal/. Paterson, J. L., Brown, R., & Walters, M. A. (2019). Feeling for and as a group member: Understanding LGBT victimization via group-based empathy and intergroup emotions. British Journal of Social Psychology, 58(1), 211–224. Perry, B., & Alvi, S. (2012). ‘We are all vulnerable’: The in terrorem effects of hate crimes. International Review of Victimology, 18(1), 57–71. Southern Poverty Law Centre. (2016). The Trump effect: The impact of the 2016 presidential election on our nation’s schools. https://www.splcenter.org/20161128/ trump-effectimpact-2016-presidential-election-our-nations-schools. Vedeler, J. S., Olsen, T., & Eriksen, J. (2019). Hate speech harms: A social justice discussion of disabled Norwegians’ experiences. Disability & Society, 34(3), 368–383. Walters, M. A., Paterson, J., McDonnell, L., & Brown, R. (2019). Group identity, empathy and shared suffering: Understanding the ‘community’ impacts of anti-LGBT and Islamophobic hate crimes. International Review of Victimology. https://journals.sagepub.com/doi/full/10.1177/ 0269758019833284.

1  INCREASING PROBLEMS, INCREASING INDIFFERENCE 

9

Williams, M., & Burnap, P. (2015). Cyberhate on social media in the aftermath of Woolwich: A case study in computational criminology and big data. British Journal of Criminology, 56(2), 211–238. Williams, M., & Pearson, O. (2016). Hate crime and bullying in the age of social media. https://orca-mwe.cf.ac.uk/88865/1/Cyber-Hate-and-Bullying-PostConference-Report_English_pdf.pdf.

CHAPTER 2

Visible yet Invisible: Challenges Facing Hate Crime Victims

Abstract  Within this chapter, we illustrate the prima facie contradictory positions in which hate crime victims often find themselves in by being on the one hand all too visible to perpetrators on the basis of their ‘difference’, and yet largely invisible to professionals, mainstream society and policy formation because of the marginal positions which they are seen to occupy. We begin by reflecting upon how the term ‘hate crime’ has been—or perhaps more pertinently how it should be—defined. This discussion is designed to steer readers through some of the main challenges and ambiguities relating to defining ‘hate crime’ and by doing so offers conceptual clarity for the purposes of the chapters which follow. Keywords  Hate crime Prejudice · Identity

· Victims · Difference · Targeted hostility ·

Within this chapter, we illustrate the prima facie contradictory positions in which hate crime victims often find themselves in by being on the one hand all too visible to perpetrators on the basis of their ‘difference’, and yet largely invisible to professionals, mainstream society and policy formation because of the marginal positions which they are seen to occupy. We begin by reflecting upon how the term ‘hate crime’ has been— or perhaps more pertinently how it should be—defined. This discussion is designed to steer readers through some of the main challenges © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_2

11

12  S.-J. HARDY AND N. CHAKRABORTI

and ambiguities relating to defining ‘hate crime’ and by doing so offers conceptual clarity for the purposes of the chapters which follow. It emphasises the need to recognise and to integrate concepts such as ‘difference’ and ‘otherness’ within established frameworks in order to reconfigure a more inclusive concept. As this chapter highlights, visible markers of difference and patterns of economic, structural, political and cultural inequality contribute to certain groups being especially vulnerable to hate crime victimisation. Despite this enhanced risk of victimisation, this chapter demonstrates that by virtue of the multiple and intersecting layers of discrimination encountered by minority groups within the context of education, employment, housing and health, they are often detached from or denied access to policy-making, civic engagement and services. This disempowered position has significant implications for the visibility of hate crime victims within and beyond the criminal justice system and the perceived deservingness of their victim status.

Conceptualising Hate Crime ‘Hate’ is an emotive and conceptually ambiguous label which can mean different things to different people, and this influences the offences which are classified as forms of ‘hate crime’ and for the groups of victims who are offered protection under its umbrella framework. As Chakraborti and Garland (2012: 501) explain, hate crime ‘emerges from a complex network of events, structures and underlying processes, and, as such, will be constructed according to different actors’ perceptions, whether they are scholars, law enforcers or victims’. An array of definitions can be found within the wider body of academic and policy literature, but almost all are consistent in referring to a broader range of factors than hate alone to describe the motivations which underpin hate crime victimisation (see, inter alia, Hall 2013; Gerstenfeld 2013; Petrosino 2003; Jacobs and Potter 1998). In particular, terms such as ‘targeted hostility’, ‘prejudice’, ‘bias’ and ‘intolerance’ have all been used interchangeably by scholars as a way of highlighting that the presence of ‘hate’ itself is not central to the commission of a hate crime. In the absence of a universal definition, it is Barbara Perry’s (2001) work in this area which has been especially influential upon contemporary interpretations of hate crime. According to Perry (2001: 10):

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

13

Hate crime … involves acts of violence and intimidation, usually directed towards already stigmatised and marginalised groups. As such, it is a mechanism of power and oppression, intended to reaffirm the precarious hierarchies that characterise a given social order. It attempts to re-create simultaneously the threatened (real or imagined) hegemony of the perpetrator’s group and the ‘appropriate’ subordinate identity of the victim’s group. It is a means of marking both the Self and the Other in such a way as to re-establish their ‘proper’ relative positions, as given and reproduced by broader ideologies and patterns of social and political inequality.

There are a number of noteworthy elements to this framework which have shaped the way in which hate crimes—and responses to victims— have come to be conceived of with regard to scholarship and policy. Significantly, it highlights the complexity of hate crime by emphasising the relationship between structural hierarchies, institutionalised prejudices and acts of violence and micro-aggressions. It gives primacy to the idea that perpetration is qualitatively different when it is motivated by bigotry and directed towards already disempowered, marginalised and stigmatised populations. As such, hate crime cannot be divorced from the power dynamics present within modern societies which reinforce the ‘othering’ of those who are different. Indeed, the process of ‘doing difference’ is a central tenet of Perry’s framework which conceives of hate crime as being rooted in the ideological structures of societal oppression that govern normative conceptions of identity. Within such a process, hate crime emerges as a response to the threats posed by ‘others’ when they attempt to step out of their ‘proper’ subordinate position within the structural order: a mechanism through which hate crime is used to sustain both the hegemonic identity of the perpetrator and the boundaries between dominant and subordinate groups by reminding the victim of their place. Moreover, the focus on group membership underlines the symbolic relevance of hate crime in this context. For Perry (2001), hate crimes are acts of violence and intimidation directed not solely towards the individual victim but towards the collective wider community whom the victim is perceived to represent. As such, they symbolise the ‘natural’ relations of superiority and inferiority within the confines of structural norms and are designed to transmit a message to the victim’s community that they are ‘different’ and that they ‘don’t belong’. As ‘message crimes’, hate crime—whether it is perpetrated in online or offline spaces—reaches ‘into the community to create fear, hostility and suspicion’ and in so doing reaffirm ‘the hegemony

14  S.-J. HARDY AND N. CHAKRABORTI

of the perpetrator’s and the “appropriate” subordinate identity of the victim’s group’ (Perry 2001: 10). Within this framework, the victims themselves are interchangeable as they have been chosen on the basis of their generic subordinate identity rather than any individual characteristics and are, therefore, likely to be strangers with whom the perpetrator has had little or no contact (Perry 2001: 29). Perry’s drive to conceptualise hate crime within the broader psychological and sociopolitical contexts that condition hostile reactions to the ‘other’ has been of considerable value. Crucially, the framework described above recognises that hate crimes are not static events which occur in a cultural or social vacuum, but instead are part of a process of repeated or systematic victimisation involving context, structure and agency (see also Bowling 1993). However, despite the strengths of Perry’s definition and its influence upon the field of hate studies, it can restrict our understanding of how hate crime can play out in the ‘realworld’ by suggesting that victims and perpetrators will almost invariably be strangers to each other. Evidence from our research contests this assumption by showing that a substantial proportion of hate crimes are actually committed by perpetrators known to the victim in some capacity, such as a neighbour, a colleague, a classmate or a carer to name just some examples. As discussed in greater detail within the next section, conceiving of hate crime purely as a mechanism of oppression or subordination overplays what for some perpetrators will be an act borne from boredom, jealousy, convenience or unfamiliarity with ‘difference’. Noteworthy developments in establishing a common understanding of hate crime have emerged outside of the academic domain. One such attempt has come from the Office for Democratic Institutions and Human Rights (ODIHR), whose guidance for OSCE member states describes hate crimes as ‘criminal acts committed with a bias motive’ (2009: 16). In line with the academic definitions, the underlying motive does not have to manifest itself as ‘hate’ for the offence to be thought of as a hate crime. Rather, bias in this sense refers to online and offline acts where the victim is targeted deliberately because of a ‘protected characteristic … shared by a group, such as “race”, language, religion, ethnicity, nationality, or any other similar common factor’ (2009: 16). Importantly, ODIHR’s guidance does not seek to specify which protected characteristics should form the basis of a member state’s hate crime policy, aside from making reference to aspects of identity which are ‘fundamental to a person’s sense of self’ and to the relevance of ‘current

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

15

social problems as well as potential historical oppression and discrimination’ (2009: 38). As expanded upon within the next chapter, this broad, pan-national framework for understanding hate crime was developed in response to an increased awareness of hate crime and the pressing need for states, statutory and non-governmental organisations across Europe to respond more effectively to these forms of victimisation. Within the UK, the College of Policing is the professional body responsible for setting standards in professional development across English and Welsh police forces and their interpretation of hate crime is especially relevant as it illustrates the broad conceptualisation and application of hate crime. The College of Policing (2014: 4) defines hate crime as ‘any criminal offence which is perceived, by the victim or any other person, to be motivated by a hostility or prejudice’ and stipulate that the offence must be recorded as a ‘hate crime’ if the incident was targeted towards a person on the basis of their actual or perceived disability, race, religion, sexual orientation and transgender identity. Importantly, the guidelines provided by the College of Policing include a requirement for the police to record and investigate all hate incidents in the same way as hate crimes even if they lack the requisite elements to be classified as a notifiable offence later in the criminal justice process. This means that at the recording stage, any hate incident, whether a prima facie ‘crime’ or not, is to be recorded if it is perceived by the victim or any other person (such as a witness, a family member or a carer) as being motivated by hostility or prejudice. Conceiving of hate crime in such a way enables the police to acknowledge and respond to the micro-aggressions which many victims are routinely subjected to, including harassment, online abuse or other forms of intimidatory behaviour in addition to more violent expressions of hate. As noted previously, there is no one universally accepted definition of hate crime and the interpretations highlighted in preceding pages illustrate the elasticity of the concept. Of note is that the presence of ‘hate’ is not central to the commission of a hate crime, and the label itself can convey a misleading impression. Equally, using group identity politics to define who is—and who is not—a victim of hate crime can sometimes add to this confusion, as such an approach creates division between communities of identity rather than highlighting the shared nature of their victimisation. For this reason, increasing numbers of scholars have called for a re-evaluation of the way in which hate crime has come to be stringently associated with particular identities and particular forms

16  S.-J. HARDY AND N. CHAKRABORTI

of victimisation (see, inter alia, Campbell 2014; Garland and Hodkinson 2014a; Chakraborti and Garland 2012).

Hypervisible: Recognising ‘Difference’ and ‘Vulnerability’ Conventional conceptualisations of hate crime have tended to regard the phenomenon in a broadly similar way: that hate crimes ‘hurt more’ as they are an attack on the victim’s identity (Iganski 2001); that they are ‘message crimes’ as they can affect the victim’s wider identity community too (Paterson et al. 2019); and that they are perpetrated by members of dominant social groups against those from historically marginalised and disadvantaged communities who have suffered from discrimination and victimisation for decades, even centuries (Perry 2001). However, in recent years there has been a growing call for both the academic and policy spheres to widen the parameters of the hate debate in recognition that some victims can be subjected to hate not exclusively because of their membership of a particular identity group but also because they are seen as somehow ‘different’ or especially vulnerable in the eyes of the perpetrator. The selection of a ‘soft’ or convenient target because they are obviously ‘different’ (for instance, through markers of language, skin colour, culture, dress or disability) or because they appear vulnerable (because of their age, isolation and physical presence, to name just a handful of factors) may often have little to do with any conscious intent to suppress the ‘other’ or to communicate a message of hate to the victim’s wider community. Visibility—or more pertinently, being visibly ‘different’— in a public space can exacerbate the perceived threat posed by a specific group. Moran et al. (2003: 7) developed the notion of ‘regimes of placement’ to highlight how historical and institutionalised practices have established and continue to reinforce what is considered legitimate and illegitimate visual representations within a given space (Pile and Thrift 1995). Valentine’s (2010: 531) research on expressions of prejudice from majority groups is illustrative of this point, highlighting that ‘various spatial metaphors including “invasion”, “taking over”, “being out of place” or “not knowing their place” were used by respondents to justify the cultural threat allegedly posed by “difference”’. A growing evidence base adds further weight to the significance of markers of ‘difference’ to the process of victim selection within

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

17

socio-spatial contexts. For instance, substantial proportions of lesbian, gay and bisexual people avoid holding hands in public with a same-sex partner to reduce their risk of victimisation (Government Equalities Office 2018); trans people adjust the way that they dress out of fear of harassment (Stonewall 2018b); male goths are targeted due to their ‘effeminate’ and distinctive appearance and dress sense (Garland and Hodkinson 2014b); anyone ostensibly ‘looking Asian’, for instance through wearing a turban or veil, sporting a beard or simply by virtue of being ‘dark-skinned’, is susceptible to the risk of physical assault or verbal abuse (Awan and Zempi 2017); disabled people are now at a greater risk of being subjected to hate crime as a result of growing independence and a withdrawal of services which has led to an increased presence within housing estates, communities and city centres (Hall and Bates 2019; Power and Hall 2018); and being ‘different’ in some conservative, monolithic and traditional rural environments can be dangerous when communities are intolerant of non-conformity and outsiders (Burnett 2012a, b, c; Neal 2009). In these instances, vulnerability to hate crime victimisation is exacerbated through social conditions, prevailing norms and people’s reactions to ‘difference’. Intersectionality is another dynamic central to the process of hate crime victimisation which is often overlooked by scholars and policy-makers alike. Conceiving of hate crimes simply as offences directed towards discrete strands of a person’s identity fails to give adequate recognition to the interplay of identities with one another and with other personal, social and situational characteristics. Hate crime victims are not homogeneous groups of people with uniform characteristics and experiences, and any attempts to investigate lived experiences should acknowledge the differences within and intersections between a range of identity characteristics. For instance, hate crimes can often be triggered and exacerbated by socio-economic conditions, and some victims will invariably be better placed than others to avoid persecution by virtue of living at a greater distance from prejudiced neighbours or in less overtly hostile environments (Hardy 2017; Walters and Hoyle 2012). According to Cops and Pleysier (2011: 59), vulnerability refers to ‘the perception of exposure to danger, a loss of control over the situation and a perceived inadequate capacity to resist the direct and indirect consequences of victimisation’. Typically vulnerable groups or individuals include those who are socially and economically disadvantaged, or subject to othering, vilification and discrimination (Nyamathi 1998).

18  S.-J. HARDY AND N. CHAKRABORTI

This point is particularly pertinent at a time when ‘other’ identities are under greater political and public scrutiny than perhaps ever before as a result of unstable, fractious social, political and economic contexts. For example as part of the UK Government’s austerity agenda, substantial cuts in welfare spending were implemented and as a result of this overhaul, many disabled claimants found themselves stigmatised as ‘benefit scroungers’ by sections of the media and some politicians (Dodd 2016; Ralph et al. 2016). Similarly, Burnett (2017) observes that a hostile political environment did not simply emerge after the EU Referendum was announced by David Cameron in 2016; instead, the propagation of fear, intolerance and ill-treatment of minority ethnic and faith communities had been present within the political sphere for a much longer time frame. Such narratives position minority groups like asylum seekers, disabled people, immigrants and Muslims as being responsible for society’s ill, increasing their vulnerability to victimisation. Under these social, political and economic conditions, minority groups who are visibly ‘different’ have greater ‘exposure to danger’; a situation which they have no control over or are able to resist (Cops and Pleysier 2011: 59).

Rendered Invisible: Disempowered and Disengaged This hypervisible status juxtaposes the peripheral position many minority groups occupy within society as a result of the multiple and intersecting layers of discrimination that they face. As illustrated by the most recent iteration of the ‘Is Britain Fairer?’ report, the ‘picture is still bleak for the living standards of Britain’s most at-risk and “forgotten” groups of people’ (EHRC 2018: 5). Socio-economic disadvantage is known to adversely affect participation and achievement in nearly every aspect of life, including education, employment and health, which is why the ever-growing number of people in poverty—particularly those from minority groups including disabled people and certain minority ethnic communities—is of great concern (Joyce and Xu 2019; EHRC 2018). For instance, within the context of ethnicity, evidence suggests that minority ethnic communities are more likely to encounter barriers when accessing health care and to become homeless (ibid., 2018). Despite improvements for those who are Indian and Chinese within the realm of work, young Muslims face obstacles when entering the labour market and yet are more likely than ever to prosper in education and to go on to university when compared to other groups (EHRC 2018; Social

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

19

Mobility Commission 2017). Additionally, Black African, Bangladeshi and Pakistani people are at a higher risk of experiencing deprivation, while Gypsy, Roma and Travellers encounter multiple disadvantages across different areas of life (EHRC 2018). In nearly all areas of public life, disabled people continue to face inequality and discrimination and have borne the brunt of UK Government’s austerity measures (Papworth Trust 2018; United Nations 2018; Runswick-Cole and Goodley 2015). Disabled people are now more likely to be sent to a special school rather than a mainstream one; to have no qualifications; to be unemployed or to be paid less; to have poorer health outcomes; and to experience difficulties in finding suitable housing (EHRC 2018; Papworth Trust 2018). Similarly worrying trends are seen in relation to lesbian, gay, bisexual and trans people with research indicating that one in five of those searching for work, one in six who visited a café, restaurant, bar or nightclub, and one in ten looking for a house or flat to rent or buy were discriminated against on the basis of their sexual orientation and/or gender identity in the last year (Stonewall 2018a). As Camino and Zeldin (2002: 213) explain, ‘inclusive participation is a primary component of civil society’ as ‘engagement is both a right and a responsibility of citizenship’ which ‘benefits both the individual and the collective’. Fraser (2003: 29) explains that ‘parity of participation’ is predicated upon two mutually inclusive conditions: recognition and redistribution of societal resources. In essence, the groups who lack recognition within society have been adversely affected by the institutionalised patterns which promote and sustain the dominant identity, their values and customs. These hegemonic ideals are embodied through a state’s laws, policies and practices, creating an environment whereby ‘difference’ is seen as incompatible and inferior. For instance, disabled people face difficulties in finding employment—and thus participating within society—as their potential to contribute and to add value is questioned because of the institutional practices and deep-seated beliefs which have misrecognised disabled people as being less worthy, weak and burdensome (Ralph et al. 2016; Quarmby 2011). Additionally, inclusive participation hinges on fair redistribution of a state’s material resources ‘to ensure participants’ independence’ (Fraser 2003: 36). As illustrated in the preceding paragraphs, certain minority groups encounter mis- or nonrecognition through discrimination and unconscious bias which leads to being denied access to quality housing, health care and services,

20  S.-J. HARDY AND N. CHAKRABORTI

and to opportunities to flourish within the context of education and employment. Correspondingly, these groups receive distributive injustice through economic marginalisation and deprivation (Fraser 2003). Without inclusive participation within society not only are these groups rendered invisible—along with their ideas, experiences and needs—but also the status quo which prioritises the ruling class in every respect, prevails. This situation is further compounded by the wider economic climate which has seen the rescinding of outreach and engagement work within state agencies and non-government organisations as part of cuts to frontline resources (Chakraborti 2018). Research reveals that front-line professionals have less time and fewer resources to meaningfully engage with those who are considered ‘hard to reach’—a term commonly used but contested due to the inference that it is the groups themselves who are to blame because they are difficult to access or unwilling to engage (Hardy and Chakraborti 2017; Burrall and Carr-West 2009; Friemuth and Mettger 1990). Both civic and community engagement contribute to a person’s social capital by encouraging the development of new social networks and interactions between different people, which leads to increased knowledge, awareness and bonds of trust (Putnam 2000). This lack of capital is especially relevant within the context of hate crime because it could explain why many hate crime victims—particularly those from disenfranchised communities—are unaware that there are policies and laws in place to facilitate justice or that support provisions exist to aid in repairing the harm caused.

Conclusion This chapter has illustrated that the concept of hate crime is more complex and multi-layered than many might imagine. We have seen that the term ‘hate’ is a problematic, ambiguous and in many cases an inaccurate descriptor of the offences with which it is commonly associated. We have highlighted that hate crime is rooted within a wider, structural process whereby expressions of hate, prejudice and hostility are used to marginalise ‘difference’ and to sustain hegemonic boundaries. An increasingly extensive body of literature has shown that attacks against the ‘other’ can feed off economic instability, political scaremongering and media stereotyping to the point where violence becomes a mechanism used to reinforce power dynamics between dominant and subordinate groups and to

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

21

create cultures of fear within alien communities. However, we have also challenged the idea that a hate crime victim group must be conceived of as a minority that has been historically marginalised and disadvantaged and is targeted in order to maintain the privileged position of society’s powerful groups. Instead, a vulnerability-based approach was proposed which acknowledges the heightened level of risk posed to certain groups or individuals that can arise through a complex interplay of different factors, including prejudice, hostility, unfamiliarity, discomfort, opportunism, economic disadvantage or political scaremongering.

References Awan, I., & Zempi, I. (2017). ‘I will blow your face off’: Virtual and physical world anti-Muslim hate crime. British Journal of Criminology, 57(2), 362–380. Bowling, B. (1993). Racial harassment and the process of victimization. British Journal of Criminology, 33(2), 231–250. Burnett, J. (2012a). The new geographies of racism: Peterborough. London: Institute of Race Relations. Burnett, J. (2012b). The new geographies of racism: Stoke on Trent. London: Institute of Race Relations. Burnett, J. (2012c). The new geographies of racism: Plymouth. London: Institute of Race Relations. Burnett, J. (2017). Racial violence and the Brexit state. Race & Class, 58(4), 85–97. Burrall, S., & Carr-West, J. (2009). Citizen power in recession? The case for public engagement in local government. London: Local Government Information Unit. Camino, L., & Zeldin, S. (2002). From periphery to center: Pathways for youth civic engagement in the day-to-day life of communities. Applied Developmental Science, 6(4), 213–220. Campbell, R. (2014). Not getting away with it: Thinking sex work and hate crime in Merseyside. In N. Chakraborti & J. Garland (Eds.), Responding to hate crime: The case for connecting policy and research (pp. 55–70). Bristol: The Policy Press. Chakraborti, N. (2018). Responding to hate crime: Escalating problems, continued failings. Criminology & Criminal Justice, 18(4), 387–404. Chakraborti, N., & Garland, J. (2012). Reconceptualising hate crime victimization through the lens of vulnerability and ‘difference’. Theoretical Criminology, 16(4), 499–514. College of Policing. (2014). Hate crime operational guidance. Coventry: College of Policing.

22  S.-J. HARDY AND N. CHAKRABORTI Cops, D., & Pleysier, S. (2011). ‘Doing gender’ in fear of crime. British Journal of Criminology, 51(1), 58–74. Dodd, S. (2016). Orientating disability studies to disablist austerity: Applying Fraser’s insights. Disability & Society, 31(2), 149–165. Equality and Human Rights Commission (EHRC). (2018). Is Britain fairer? The state of equality and human rights 2018. London: EHRC. Fraser, N. (2003). Social Justice in the age of identity politics: Redistribution, recognition, and participation. In N. Fraser & A. Honneth (Eds.), Redistribution or recognition? A political-philosophical exchange (pp. 7–109). London: Verso. Friemuth, V. S., & Mettger, W. (1990). Is there a hard-to-reach audience? Public Health Reports, 105(3), 232–238. Garland, J., & Hodkinson, P. (2014a). Alternative subcultures and hate crime. In N. Hall, A. Corb, P. Giannasi, & J. G. D. Grieve (Eds.), The Routledge international handbook on hate crime (pp. 226–236). London: Routledge. Garland, J., & Hodkinson, P. (2014b). “F**king freak! What the hell do you think you look like?” Experiences of targeted victimisation among goths and developing notions of hate crime. British Journal of Criminology, 54(4), 613–631. Gerstenfeld, P. B. (2013). Hate crimes: Causes, controls and controversies (3rd ed.). London: Sage. Government Equalities Office. (2018). National LGBT survey: Research report. London: GEO. Hall, N. (2013). Hate crime (2nd ed.). London: Routledge. Hall, E., & Bates, E. (2019). Hatescape? A relational geography of disability hate crime, exclusion and belonging in the city. Geoforum, 101, 100–110. Hardy, S. (2017). Everyday multiculturalism and hidden hate. London: Palgrave. Hardy, S., & Chakraborti, N. (2017). Hate crime: Identifying and dismantling barriers to justice. Leicester: University of Leicester. Iganski, P. (2001). Hate crimes hurt more. American Behavioural Scientist, 45(4), 626–638. Jacobs, J., & Potter, K. (1998). Hate crimes: Criminal law and identity politics. Oxford: Oxford University Press. Joyce, R., & Xu, X. (2019). Inequalities in the twenty-first century: Introducing the IFS Deaton Review. London: The Institute for Fiscal Studies. Moran, L., Skeggs, B., Corteen, K., & Tyrer, P. (2003). The formation of fear in gay space: The ‘straight’ story. Capital and Class, 80, 173–199. Neal, S. (2009). Rural identities: Ethnicity and community in the contemporary English countryside. Farnham: Ashgate. Nyamathi, A. (1998). Vulnerable populations: A continuing nursing focus. Nursing Research, 47(2), 65–66. Office for Democratic Institutions and Human Rights (ODIHR). (2009). Hate crime laws: A practical guide. Warsaw: OSCE Office for Democratic Institutions and Human Rights.

2  VISIBLE YET INVISIBLE: CHALLENGES FACING HATE CRIME VICTIMS 

23

Papworth Trust. (2018). Facts and figures 2018: Disability in the United Kingdom. Cambridge: Papworth Trust. Paterson, J. L., Brown, R., & Walters, M. A. (2019). Feeling for and as a group member: Understanding LGBT victimization via group-based empathy and intergroup emotions. British Journal of Social Psychology, 58(1), 211–224. Perry, B. (2001). In the name of hate: Understanding hate crimes. London: Routledge. Petrosino, C. (2003). Connecting the past to the future: Hate crime in America. In B. Perry (Ed.), Hate and bias crime: A reader (pp. 9–26). London: Routledge. Pile, S., & Thrift, N. (1995). Mapping the subject: Geographies of cultural transformation. London: Routledge. Power, A., & Hall, E. (2018). Placing care in times of austerity. Social and Cultural Geography, 19(3), 303–313. Putnam, R. (2000). Bowling alone: The collapse and revival of American community. New York: Simon and Schuster. Quarmby, K. (2011). Scapegoat: Why we are failing disabled people. London: Portobello Books. Ralph, S., Capewell, C., & Bonnett, E. (2016). Disability hate crime: Persecuted for difference. British Journal of Special Education, 43(3), 215–232. Runswick-Cole, K., & Goodley, D. (2015). Disability, austerity and cruel optimism in big society: Resistance and “the disability commons”. Canadian Journal of Disability Studies, 4(2), 162–186. Social Mobility Commission. (2017). The social mobility challenges faced by young Muslims. London: Social Mobility Commission. Stonewall. (2018a). LGBT in Britain: Hate crime and discrimination. London: Stonewall. Stonewall. (2018b). LGBT in Britain: Trans report. London: Stonewall. United Nations. (2018). Realization of the sustainable development goals by, for and with persons with disabilities. New York: United Nations. Valentine, G. (2010). Prejudice: Rethinking geographies of oppression. Social & Cultural Geography, 11(6), 519–537. Walters, M., & Hoyle, C. (2012). Exploring the everyday world of hate victimization through community mediation. International Review of Victimology, 18(1), 7–24.

CHAPTER 3

Relevant yet Irrelevant: Challenges Associated with Hate Crime Policy

Abstract  This chapter describes the way in which hate crime policy has become increasingly relevant to many Western states and highlights the value of these developments in terms of facilitating access to justice, embodying the values of a state and deterring prejudicial and hostile sentiment. It begins by considering how hate crime has ‘translated’ across national boundaries, comparing the implementation of different policy frameworks designed to combat hate crime. It is argued that while there have been some successes in dealing with hate crime in certain individual countries, in others much work still needs to be done in developing effective policies and practices which offer meaningful protection from hate crime. Keywords  Hate crime Policing

· Victims · Policy · Criminal justice system ·

Hate crime is recognised and embraced within many countries around the world through legislation and policy. This chapter describes the way in which hate crime policy has become increasingly relevant to many Western states and highlights the value of these developments in terms of facilitating access to justice, embodying the values of a state and deterring prejudicial and hostile sentiment. It begins by considering how hate crime has ‘translated’ across national boundaries, comparing the © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_3

25

26  S.-J. HARDY AND N. CHAKRABORTI

implementation of different policy frameworks designed to combat hate crime. It is argued that while there have been some successes in dealing with hate crime in certain indi-vidual countries, in others much work still needs to be done in developing effective policies and practices which offer meaningful protection from hate crime. Within the final section, we show that established policy frameworks are still alarmingly irrelevant to the majority of hate crime victims. Illustrated through a set of specific problems with policy- and operational-level responses to hate crime, we demonstrate that progress has been only partially effective in addressing victims’ needs and lived experiences, and that the damage caused by hate crime can often be reinforced—and not alleviated—by the continued failures of policy-makers and professionals to respond and engage effectively.

Signs of Progress The Domestic Context In the UK, the concept of hate crime is now well-established in both practitioner and academic circles. As a number of scholars have noted (Hall 2013; Ray and Smith 2001), popular usage of the term in the UK gained currency in the wake of much-publicised events that took place towards the end of the last and the start of the current century. In particular, the racist murder of black teenager Stephen Lawrence in 1993— and the resultant publication of the Macpherson Report in 1999—drew widespread attention to the problems posed by crimes targeted towards members of minority groups, as did the racist and homophobic nail bomb attacks of 1999 instigated by the neo-Nazi David Copeland in the diverse London communities of Brixton, Brick Lane and Soho. While the Macpherson Report acted as a catalyst for debate in government, policy and scholarly circles, the remit soon broadened from the policing of race and ethnicity to a much more inclusive discussion of how a range of minority groups were viewed and treated by the criminal justice system. Much of this focused upon how the police, Crown Prosecution Service (CPS) and courts were dealing with the victims of targeted violence and harassment. Thus, gradually, the term ‘hate crime’ was adopted to describe the unique form of abuse suffered by disabled people, lesbian, gay and bisexual people, religious communities and trans people, in addition to minority ethnic groups.

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

27

Policy guidance on hate crime was first produced in 2000 by the Association of Chief Police Officers (ACPO), with subsequent iterations published in 2005 and again in 2014 by the College of Policing who are now responsible for setting standards in professional development across English and Welsh police forces. From its earliest conceptualisation, the operational definition—outlined in the previous chapter—gave primacy to the perception of the victim or any other person as being the defining factor in determining a hate crime and not the discretion of the investigating police officer (College of Policing 2014). Importantly, the victim is not required to provide corroborating evidence or justification to support their belief, and ‘police officers or staff should not directly challenge this perception’ (College of Policing 2014: 5). In theory at least, framing the policy in this way is designed to improve levels of trust and confidence within historically marginalised communities and to increase the numbers of victims coming forward to report hate crime. The operational guidance also stipulates that the police are required to record an offence as a ‘hate crime’ if the incident was motivated by hostility towards any of five monitored strands of identity: namely disability, race, religion, sexual orientation and transgender status. Thus, it is not just any form of prejudice or hostility which can form the basis of a hate crime, but rather prejudice or hostility against particular groups of people or based upon particular grounds. Limiting the application of hate crime to certain groups and not others is inherently problematic as this requires difficult decisions to be made with regard to who should be deserving of ‘special protection’, an issue discussed in greater depth in the next section. However, there is a degree of flexibility in how the boundaries of hate crime are framed within UK policy, with official guidance stating that the five monitored strands ‘are the minimum categories that police officers and staff are expected to record’ (College of Policing 2014: 7). This enables police forces to record other forms of targeted hostility as hate crime in addition to those five strands and has been developed as a result of tragic cases and an emerging body of research highlighting the targeting of ‘other’ identities and groups who have not routinely been considered as hate crime victims. In response, a number of individual forces now monitor other offences—including violence against the homeless, alternative subcultures and sex workers, and more recently misogynistic harassment—as additional strands of hate crime within their local areas (Mullany and Trickett 2018; Campbell 2014; Garland and Hodkinson 2014b; Lancaster 2014).

28  S.-J. HARDY AND N. CHAKRABORTI

Mirroring policy developments, a number of laws have been introduced which adhere to the principle that crimes motivated by hostility or prejudice towards the five monitored strands of identity should be treated differently from ‘ordinary’ crimes. While it is beyond the scope of this chapter to examine in depth the complex contours of the domestic legal landscape (see Walters et al. 2017), for the purposes of conveying legislative progress we briefly outline the salient points. Through the introduction of various pieces of legislation, the courts are able to increase the sentence for any offence in which the aggravated element— or in other words, where there is sufficient evidence that the crime was motivated by hostility—is proven. These laws, and specifically the enhanced sentencing framework which accompanies these laws, have significant value in terms of their capacity to express collective condemnation of prejudice; to send a declaratory message to offenders; to convey a message of support to victims and stigmatised communities; to build confidence in the criminal justice system within some of the more disaffected and vulnerable members of society; and to acknowledge the additional harm caused by hate offences (Walters et al. 2017; ODIHR 2009). These developments and the publication of successive UK Government action plans are indicative of the increasing prioritisation of hate crime by the state over the course of the last decade (HM Government 2012, 2018; Home Office 2016). The most recent action plan, Action Against Hate, was published in the wake of the spike in recorded hate crimes following the EU referendum and refers to the importance of protecting ‘the shared values that underpin the British way of life’ (Home Office 2016: 5) through a series of actions and priority areas which seek to identify best practice across all strands of hate crime. Simultaneously, the UK Government has taken steps to improve victim support, formally adopting the EU Directive on minimum standards of victim support in 2012. Alongside strengthening the Code of Practice for Victims’ of Crime (the ‘Victims Code’) in 2013—and again in 2015—to incorporate these directives, there are now requirements for those responsible for delivering support to victims to ensure that victims receive the information, services and support that they are entitled to. The most recent iteration of the Victims Code sets out a number of additional responsibilities providers must comply with, which includes enhanced entitlements for people affected by the most serious and persistent crimes, including hate crime (Ministry of Justice 2015). These additional provisions include access to support within one day (rather than

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

29

five days for other types of crimes) and implementing special measures during criminal court proceedings, thus improving the support for hate crime victims within and beyond the criminal justice system. The International Landscape The USA has been at the international forefront of developing hate crime legislation and policy, at both the federal and state levels. At the federal level, the first piece of hate crime legislation was the Civil Rights Act 1968, which contained a provision for the prosecution of offences in which the victim was targeted on the basis of their race, colour, religion or national origin under certain specific circumstances. Subsequently, the Hate Crime Statistics Act 1990 required the collation of national hate crime statistics on the basis of bias against a person’s race, ethnicity, religion, disability or sexual orientation. This legislative framework, however, did not include a provision for penalty enhancement— which was subsequently rectified through the Hate Crimes Sentencing Enhancement Act 1994—and only permitted protection for those targeted on the basis of their race and religion. The introduction of the Matthew Shepard and James Byrd Jr. Hate Crimes Prevention Act 2009 heralded a significant step forward in terms of recognising the plight of other minority groups and in facilitating greater access to justice through increased investment in the investigation and prosecution of hate crime. The first piece of state legislation was enacted by California in 1978, and since this time many states have created laws protecting a wide range of identity groups and characteristics. Canada, Australia and New Zealand have also developed their own approaches to enshrining the concept of hate crime within a legal framework through the adoption of different models including the penalty enhancement model, the sentence aggravation model and the substantive offence model (Mason 2010). Canada, like the USA, developed some of the earliest policy and legal interventions in the context of hate crime. For instance, Sections 318–320 of the Canadian Criminal Code, which cover ‘hate propaganda’ and incitement to hatred, were adopted by the Canadian Parliament as early as 1970. However, there are issues with how this legislation is framed: a police officer needs to obtain written consent from the attorney general before he or she can proceed with charges for these offences, causing severe delays in many cases and thereby resulting in prosecutions for hate crime being disappointingly

30  S.-J. HARDY AND N. CHAKRABORTI

low (Corb 2014). More recently, amendments to the Canadian Human Rights Act and the Criminal Code in 2017 extended protection against hate speech to cover gender identity or expression and clearly articulated that evidence that an offence was motivated by bias, prejudice or hate based on gender identity or expression constitutes an aggravating circumstance which a court must take into consideration when it imposes a sentence (Government of Canada 2019). There has also been notable progress at a provincial level with Ontario Government establishing an Anti-Racism Directorate in 2016 who are responsible for delivering the government anti-racism initiatives, and administering the newly implemented Anti-Racism Act which is designed to tackle systemic racism within the country (Ontario 2019). Interestingly, it has only been within the last twenty years or so that the idea of ‘hate crime’ has gained currency within the European Union. Indeed, the Office for Democratic Institutions and Human Rights (ODIHR)—an intergovernmental institution which deals with the ‘human dimension’ of security (ODIHR 2017)—notes that the term ‘hate crime’ was only used officially for the first time by the Organization for Security and Co-operation in Europe (OSCE) in 2003, although some individual states had been employing it since the early 1990s (ODIHR 2013). Since that time, the concept has struggled to establish itself in many of the nation states which are part of the EU despite the publication of policy guidance and frameworks. In an attempt to improve recognition of and responses to hate crime across the 57 countries in the OSCE, ODIHR developed a definition which frames hate crime as having two key aspects: first, the incident must constitute a criminal offence, and second, the victim of the offence must have been deliberately targeted ‘because of [their] ethnicity, “race”, religion or other status’ (ODIHR 2013: 6). Despite this broad, pan-national framework, inconsistencies between countries in terms of their understandings and policy priorities remain. This led the EU to establish a Framework Decision (2008/913/JHA) on combating racism and xenophobia, which was designed to ‘harmonise’ relevant legislation and create a common approach to tackle hate crime in those areas across the EU. As well as containing provisions on genocide, crimes against humanity and war crimes, the Decision required member states to criminalise incitement to hatred directed at people ‘defined by reference to race, colour, religion, descent or national or ethnic origin’ (FRA 2012: 25). It also directed member states to ensure that ‘racist and xenophobic motivation

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

31

is considered an aggravating circumstance’ which should be taken into account when sentencing in such cases is passed (2012: 25). Furthermore, in late 2012 the EU adopted a Directive (2012/29/ EU) which contained provisions that recognised the specific needs of victims of hate crime and the support they require. Amongst other things, the Directive instructed member states to ensure that employees throughout the criminal justice system were trained in recognising and dealing with hate crime cases. Article 22, meanwhile, obliged member states to identify, at the individual level, any measures that needed to be taken to ‘protect’ people from being victimised, based upon their ‘personal characteristics’ such as ‘age, gender and gender identity or expression, ethnicity, race, religion, sexual orientation, health, disability, residence status, communication difficulties, relationship to or dependence on the offender and previous experience of crime’ (ODIHR 2013: 39). In response to this Directive, the European Union Agency for Fundamental Rights (FRA) established a working party consisting of 28 EU member states, the European Commission, the Council of Europe’s Commission against Racism and Intolerance (ECRI) and ODIHR, to produce a compendium of practices and protocols aimed at improving reporting rates and recording practices across the EU (FRA 2018). Additionally, ODIHR have created a suite of ‘tools for change’ (ODIHR 2016: 2) to support member states in operationalising the provisions outlined in the Directive. This includes a reporting website which provides annual data on hate crime and developments reported by participating States; training for law enforcement and for prosecutors; a guide on data collection and monitoring; and the Tolerance and NonDiscrimination Information System which provides information on international standards and instruments (ODIHR 2016). The issue of online hate has been relatively peripheral to these debates and Directives, and therefore a welcome development came from the European Commission who introduced a Code of Conduct which tasks social media platforms such as Facebook, Microsoft, Twitter and YouTube with greater responsibility in combatting hate speech and a set of operational measures to do so (European Commission 2019). It is evident that there has been considerable growth in both the deployment and the currency of the hate crime concept and that many countries have taken meaningful action to develop legislation and enforcement policies which are designed to protect individual freedoms, to prioritise victims and to bring hate crime offenders to justice.

32  S.-J. HARDY AND N. CHAKRABORTI

However, while there has been undoubted progress, the extent to which the implementation of new policies and laws has translated to real and sustained improvements in the support offered to victims remains in question.

Signs of Failure One of the main issues with policy and legislative responses to hate crime is that there remains a lack of consistency regarding which groups are actually defined as hate crime victim groups and thus afforded protection through these frameworks. This debate extends not just to which minority groups are included but whether majority groups can also be incorporated (Chakraborti and Garland 2012). This quandary is reflected in the framing of laws within EU member states, with some protecting all groups and others just minorities (ODIHR 2018). In the context of the UK, for instance, anyone from any background can potentially be the victim of a hate crime: consequently, those belonging to minority and to majority communities are protected by law against acts of targeted hostility (College of Policing 2014). In Sweden, however, a member of a majority community cannot be deemed to be the victim of a hate crime if targeted by someone from a minority community, while similarly a minority community member cannot be treated as a hate crime victim if attacked by somebody from another minority background (Klingspor 2008). The most monitored type of hate crime by OSCE’s 57 member states is that which is motivated by racism or xenophobia, with approximately half of member states officially recording this data (ODIHR 2018). A third monitor sexual orientation or gender identity hate crime (19 states), around a quarter anti-Muslim (16 states) and antisemitic hate crime (15 states), and just a fifth disability hate crime (ODIHR 2018). Although the Charter of Fundamental Rights of the European Union obliges member states to combat racist, xenophobic, religiously motivated, disablist, homophobic and transphobic hate crimes, it appears that this obligation is not felt as keenly by as many states as it should be (Whine 2019). That some states do not demonstrate the expected commitment to tackle hate crime is especially concerning in the context of recent EU-wide austerity climate, in which many minority groups across the EU, including Roma, Somalis, sub-Saharan Africans, Jewish, disabled, and lesbian, gay, bisexual and trans people, have been ‘scapegoated’

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

33

by some sections of the press and political classes, resulting in a rise in the number of hate crimes (FRA 2013, 2018). It is these inconsistencies that form the basis of a now-familiar criticism of conventional hate crime policy (see, inter alia, Chakraborti and Garland 2012; Mason-Bish 2010; Jacobs and Potter 1998) but one which has not been adequately resolved. It is often said that hate crime policy creates and reinforces hierarchies of identity, where some victims are deemed worthy of inclusion within hate crime frameworks and others invariably miss out. As Mason-Bish (2010: 62) notes: … 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.

Activists and campaigners have undoubtedly played a key role in revealing the hate-fuelled violence and micro-aggressions experienced by certain victims groups, and in stimulating the debate and momentum necessary to influence lawmaking and policy enforcement (Lancaster 2014). This is a process that has been crucial to the maturation of hate crime as an issue of international significance, but there is a downside to this process too: namely that the parameters of what we cover under the hate crime ‘umbrella’ can be contingent upon the ability of campaign groups to lobby for recognition under this umbrella. Whether because of greater resources, a more powerful voice, public support for their cause or a more established history of stigma and discrimination, campaigners working to support certain strands of hate crime victim will invariably be able to lobby policy-makers more effectively than other potential claim-makers (Chakraborti 2015). This explains why there are certain ‘others’ who can find themselves marginal to or excluded from hate crime policy and scholarship despite being targeted because of characteristics fundamental to perceptions of their sense of self. An ever-growing body of research has raised concerns about the lack of recognition and support for a range of victims who are regular targets of violent and intimidatory behaviour, including the homeless (Allison and Klein 2019; Wachholz 2009), members of alternative subcultures (Garland and Hodkinson 2014a), sex workers (Campbell 2014), people with mental ill-health (Chakraborti and Garland 2012) and women

34  S.-J. HARDY AND N. CHAKRABORTI

(Mullany and Trickett 2018; Mason-Bish 2014). These groups have much in common with the more familiar groups of hate crime victims in that they too are often singled out as targets of hostility specifically because of their ‘difference’. However, lacking either the support of lobby groups or political representation and typically seen as ‘undesirables’, criminogenic or less worthy than other more ‘legitimate’ or historically oppressed victim groups, they are commonly excluded from conventional hate crime frameworks (Chakraborti 2015). For these marginalised victims, the process of inclusion and exclusion is much more than simply a thorny conceptual challenge; it is a fundamental human rights and equality issue which has life-changing consequences in the context of experiences of targeted violence that go unnoticed and unchallenged in the absence of policy recognition. At present, these experiences tend to fall between the cracks of existing conceptual and policy frameworks. In many ways, policy frameworks and operational responses are contingent on having reliable data which evidences a given problem, but within the context of hate crime under-reporting is a long-standing and pervasive issue. Figures collated by ODIHR show substantial disparity between OSCE member states: for instance, the number of hate crimes recorded by police in England and Wales in 2017—95,552—is far greater than the corresponding numbers provided by the USA (8437), Germany (7913), Canada (2073), France (1505), Spain (1419) and Italy (1048) to cite just some examples (ODIHR 2018). Evidently, these figures are not an accurate measure by which to gauge the true scale of hate crime in each respective country; they are more a reflection of how hate crimes are defined, recorded, reported and statistically collated by different states, than of any genuine disparity in levels of hate crime. They also highlight how much variation there is between the way in which hate crime policy has been developed and prioritised in different parts of the world. Within the UK, while the high number of officially recorded incidents may indicate that victims of hate crime are more aware of reporting procedures and have more confidence in them (and in the police and wider criminal justice system), there are still issues with under-reporting. The Crime Survey for England and Wales (CSEW)—which provides an alternative measure of hate crime victimisation—estimated that 184,000 hate crimes took place within the same time frame, a figure double that recorded by the police service (Home Office 2018). Moreover, the ‘real’ figure of hate crimes taking place is likely to be higher still,

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

35

as many cases of hate crime are simply not recognised as such by criminal justice agencies, non-governmental organisations or by victims themselves (HMICFRS 2018; Thorneycroft and Asquith 2015; McDevitt and Williamson 2003). The CSEW also illustrates that rates of underreporting vary significantly between different strands of hate crime: recent figures suggest that one in two racist hate crimes are reported to the police, while this rate drops to one in three for homophobic hate crimes, one in four for religiously motivated hate crimes and one in seven for disability hate crimes (Home Office 2018). The importance of reporting in the context of generating effective prioritisation and service delivery is well documented, both within previous academic research on hate crime (Pezzella et al. 2019; Pickles 2019; Chakraborti 2018) and within key sources of policy guidance (Home Office 2016; College of Policing 2014; HM Government 2012). Indeed, the most recent Hate Crime Action Plan issued by the UK Government pledges to ‘ensure that everyone has the opportunity and information needed to report hate crime’ (point 89) while in its foreword the secretary of state for communities and local government states that ‘if we report every incident of hate crime, we can drive it from our streets’ (Home Office 2016: 6). And yet, despite recognition of the importance of reporting, it appears that little has been done to understand the factors which underpin a victim’s reluctance or unwillingness to report hate crime and to dismantle these barriers. Similar failings can be seen within the context of police responses to hate crime. Although hate crime has received increased levels of strategic prioritisation and has been the focus of numerous policing manuals, guidance documents and strategies, hate crime victims are less likely to be satisfied with the police response in terms of both fairness and effectiveness of the service provided when compared to victims of non-hate motivated crimes. Based on combined 2015/16 to 2017/18 surveys, just 51 per cent of hate crime victims within England and Wales were found to be very or fairly satisfied with the handling of their case, compared to 69 per cent of general crime victims (Home Office 2018). Following a recent inspection of how the police in England and Wales investigate hate crime, Her Majesty’s Inspectorate of Constabularies and Fire and Rescue Service (HMICFRS 2018: 6–14) concluded that ‘progress has been too slow’, with victims often encountering ‘an inconsistent response from control room staff’, and that there were ‘problems with the accuracy’ of data with ‘most forces doing too little to put this

36  S.-J. HARDY AND N. CHAKRABORTI

right’. Of particular concern was the finding that robust risk assessments were not widely used, and that risk management plans were rarely developed or followed, despite hate crime victims being at an increased risk of repeat victimisation (HMICFRS 2018; Home Office 2018). In this respect, it is unsurprising that there has been a decline both in referrals from the police to the CPS and in successful hate crime prosecutions (CPS 2018). While the UK has one of the strongest legislative frameworks for responding to hate crimes, the number of convictions remains stubbornly low despite an increase in police recorded hate crime (CPS 2018). A series of Criminal Justice Joint Inspection reports have highlighted particular failings within the context of disability hate crime, attributing low conviction levels to inadequate investigation and collection of evidence by the police service, to a poor understanding of disability and of existing policy definitions of hate crime, and to a disregard of victim needs and their rights to reasonable adjustments (CJJI 2018). Indeed, a significant number (28 per cent) of non-convictions are as a result of victim disengagement with the process, emphasising the importance of victims receiving effective and meaningful support throughout the criminal justice process (CPS 2018). Both the CJJI and HMICFRS reports highlight issues with the support provided to hate crime victims in terms of inconsistent referral processes and compliance with the Victims’ Code of Practice (CJJI 2018; HMICFRS 2018). In the recent cross-government Victim Strategy (2018), which sets out a framework for dealing with victims through the criminal justice system, it was noted that current practices fall short of expected standards, particularly in the coordination of support services, information sharing and the promptness of the support offered to facilitate effective recovery (HM Government 2018). Non-compliance with the Victims Code can be reported to the police or to the CPS, but it is not legally enforceable. There is limited evidence to suggest that violations of the Victims Code are reported, and this is likely to be explained through a lack of awareness of victims’ rights and entitlements within the general public, especially within marginalised and disempowered communities (Waxman 2019). Despite well-intentioned measures being implemented, hate crime victims continue to encounter a defective service with insensitive responses from front-line professionals, long periods with limited updates, short notices of court appearances and a lack of choice in terms of an alternative outcome as opposed to a punitive approach (Walters and Brown 2016; Wedlock and Tapley 2016).

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

37

Conclusion This chapter has illustrated a number of positive developments globally: we have seen the emergence of new policies, laws and interventions which have facilitated some degree of change in cultural attitudes towards the prejudice suffered by a range of minority groups, and wider prioritisation across governments, criminal justice agencies and other service providers. However, in the context of prevailing economic, political and social conditions which act as enabling factors for the continued demonisation of ‘marginal’ communities, and with levels of hate crime escalating to increasingly alarming levels both within the UK and beyond, it is all the more important to both acknowledge and address the shortcomings identified within this chapter. The latter part of this chapter identified a series of failings within existing policy-level responses to hate crime, including restrictive policy which excludes certain identity strands and groups, under-reporting, and inadequate recording and investigating procedures, all of which undermine their credibility and effectiveness in the eyes of victims. These failings not only exacerbate the harms associated with the original hate incident and increase the likelihood that victims and wider communities will not come forward in the future but also results in lawmakers and professionals overlooking or not fully comprehending the lived realities of hate crime victimisation.

References Allison, K., & Klein, B. R. (2019). Pursuing hegemonic masculinity through violence: An examination of anti-homeless bias homicides. Journal of Interpersonal Violence. https://doi.org/10.1177/0886260518821459. Campbell, R. (2014). Not getting away with it: Thinking sex work and hate crime in Merseyside. In N. Chakraborti & J. Garland (Eds.), Responding to hate crime: The case for connecting policy and research (pp. 55–70). Bristol: The Policy Press. CJJI (Criminal Justice Joint Inspection). (2018). Joint inspection of the handling of cases involving disability hate crime. London: CJJI. Chakraborti, N. (2015). Re-thinking hate crime: Fresh challenges for policy and practice. Journal of Interpersonal Violence, 30(10), 1738–1754. Chakraborti, N. (2018) Responding to hate crime: Escalating problems, continued failings. Criminology & Criminal Justice, 18(4), 387–404. Chakraborti, N., & Garland, J. (2012). Reconceptualising hate crime victimization through the lens of vulnerability and ‘difference’. Theoretical Criminology, 16(4), 499–514.

38  S.-J. HARDY AND N. CHAKRABORTI College of Policing. (2014). Hate crime operational guidance. Coventry: College of Policing. Corb, A. (2014). Hate and hate crime in Canada. In N. Hall, A. Corb, P. Giannasi, & J. G. D. Grieve (Eds.), The Routledge international handbook on hate crime (pp. 163–173). London: Routledge. CPS (Crown Prosecution Service). (2018). Hate crime annual report 2017–18. London: CPS. European Commission. (2019). Measures to address online hatred. https:// ec.europa.eu/info/policies/justice-and-fundamental-rights/combatting-discrimination/racism-and-xenophobia/combating-racism-and-xenophobia_ en#targeted-measures-to-address-online-hatred. FRA (European Union Agency for Fundamental Rights). (2012). Making hate crime visible in the European Union: Acknowledging victims’ rights. Vienna: FRA. FRA (European Union Agency for Fundamental Rights). (2013). FRA brief: Crimes motivated by hatred and prejudice in the EU. Vienna: FRA. FRA (European Union Agency for Fundamental Rights). (2018). About the compendium.  https://fra.europa.eu/en/theme/hate-crime/compendiumpractices/about-compendium. Garland, J., & Hodkinson, P. (2014a). “F**king freak! What the hell do you think you look like?” Experiences of targeted victimisation among goths and developing notions of hate crime. British Journal of Criminology, 54(4), 613–631. Garland, J., & Hodkinson, P. (2014b). Alternative subcultures and hate crime. In N. Hall, A. Corb, P. Giannasi, & J. G. D. Grieve (Eds.), The Routledge international handbook on hate crime (pp. 226–236). London: Routledge. Government of Canada. (2019). An Act to amend the Canadian Human Rights Act and the Criminal Code. https://laws-lois.justice.gc.ca/eng/annualstatutes/2017_13/FullText.html. Hall, N. (2013). Hate crime (2nd ed.). London: Routledge. Her Majesty’s Inspectorate of Constabulary and Fire & Rescue Services (HMICFRS). (2018). Understanding the difference: The initial police response to hate crime. London: HMICFRS. HM Government. (2012). Challenge it, report it, stop it: The government’s plan to tackle hate crime. London: HM Government. HM Government. (2018). Action against hate: The UK Government’s plan for tackling hate crime—Two years on. London: Home Office. Home Office. (2016). Action against hate: The UK Government’s plan for tackling hate crime. London: Home Office. Home Office. (2018). Hate crime, England and Wales, 2017/18. London: Home Office.

3  RELEVANT YET IRRELEVANT: CHALLENGES ASSOCIATED … 

39

Jacobs, J., & Potter, K. (1998). Hate crimes: Criminal law and identity politics. Oxford: Oxford University Press. Klingspor, K. (2008). The challenges of collecting statistical data in the field of hate crime: The case of Sweden. In J. Goodey & K. Aromaa (Eds.), Hate crime: Papers from the 2006 and 2007 Stockholm Criminology Symposiums (pp. 40–55). Helsinki: European Institute for Crime Prevention and Control. Lancaster, S. (2014). Reshaping hate crime policy and practice: Lessons from a grassroots campaign. In N. Chakraborti & J. Garland (Eds.), Responding to hate crime: The case for connecting policy and research (pp. 39–54). Bristol: The Policy Press. Mason, G. (2010). A picture of bias crime in New South Wales. Cosmopolitan Civil Societies: An Interdisciplinary Journal, 11(1), 47–66. Mason-Bish, H. (2010). Future challenges for hate crime policy. In N. Chakraborti (Ed.), Hate crime: Concepts, policy, future directions (pp. 58–77). Devon: Willan. Mason-Bish, H. (2014). We need to talk about women: Examining the place of gender in hate crime policy. In N. Chakraborti & J. Garland (Eds.), Responding to hate crime: The case for connecting policy and research (pp. 169– 182). Bristol: The Policy Press. McDevitt, J., & Williamson, J. (2003). Hate crimes directed at gay, lesbian, bisexual and transgendered victims. In H. Wilhelm & J. Hagan (Eds.), International handbook of violence research (pp. 801–815). London: Springer. Ministry of Justice. (2015). Code of practice for victims of crime. London: Ministry of Justice. Mullany, L., & Trickett, L. (2018). Misogyny hate crime evaluation report. Nottingham: Office of Nottinghamshire Police and Crime Commissioner. ODIHR (Office for Democratic Institutions and Human Rights). (2009). Preventing and responding to hate crimes: A resource guide for NGOs in the OSCE region. Warsaw: OSCE. ODIHR (Office for Democratic Institutions and Human Rights) (Office for Democratic Institutions and Human Rights). (2013). Hate crimes in the OSCE region: Incidents and responses—Annual report for 2012. Warsaw: OSCE. ODIHR (Office for Democratic Institutions and Human Rights). (2016). ODIHR’s efforts to counter hate crime. Warsaw: OSCE. ODIHR (Office for Democratic Institutions and Human Rights). (2017). Factsheet of the OSCE Office for Democratic Institutions and Human Rights. Warsaw: OSCE. ODIHR (Office for Democratic Institutions and Human Rights). (2018). Hate crime reporting. http://hatecrime.osce.org. Ontario. (2019). Anti-racism directorate. https://www.ontario.ca/page/antiracism-directorate.

40  S.-J. HARDY AND N. CHAKRABORTI Pezzella, F. S., Fetzer, M. D., & Keller, T. (2019). The dark figure of hate crime underreporting. American Behavioral Scientist. https://doi.org/10. 1177/0002764218823844. Pickles, J. (2019). Policing hate and bridging communities: A qualitative evaluation of relations between LGBT+ people and the police within the North East of England. Policing and Society, 1–19. Ray, L., & Smith, D. (2001). Racist offenders and the politics of ‘hate crime’. Law and Critique, 12(3), 203–221. Thorneycroft, R., & Asquith, N. (2015). The dark figure of disablist violence. Howard Journal of Criminal Justice, 54, 489–507. Wachholz, S. (2009). Pathways through hate: Exploring the victimization of the homeless. In B. Perry (Ed.), Hate crimes: The victims of hate crime (pp. 199–222). Westport, CT: Praeger. Walters, M. A., & Brown, R. (2016). Preventing hate crime: Emerging practices and recommendations for the improved management of criminal justice interventions. Brighton: University of Sussex. Walters, M. A., Wiedlitzka, S., & Owusu-Bempah, A. (2017). Hate crime and the legal process: Options for law reform. Brighton: University of Sussex. Waxman, C. (2019). Review of compliance with the Victims’ Code of Practice (VCOP): Findings, recommendations and next steps. London: Victims’ Commissioner. Wedlock, E., & Tapley, J. D. (2016). What works in supporting victims of crime: A rapid evidence assessment. London: Victims’ Commissioner. Whine, M. (2019). Cooperation between criminal justice agencies and civil society in combating hate crime. Crime, Law and Social Change, 71(3), 275–289.

PART II

Undertaking Research with Hate Crime Victims

CHAPTER 4

The Process of Engagement: Hard to Reach or Easy to Ignore?

Abstract  This chapter begins by highlighting the overarching aims of our research before discussing the methodologies employed within each study. In particular, it focuses on the complexities and challenges associated with researching diversity, and more specifically with attempts to engage with, and to collect data from, groups and communities considered ‘hard to reach’. This chapter draws from the methodological dilemmas encountered during fieldwork to illustrate that it is only through developing a more nuanced, less formulaic approach to conducting research that social scientists will achieve more inclusive constructions of ‘difference’, more complete accounts of victimisation and a more comprehensive understanding of the specific issues and problems facing different communities. Keywords  Hard to reach · Diverse communities Victims · Qualitative · Survey

· Engagement ·

This book is based on a body of transformative research, the details of which are provided within this chapter. It begins by highlighting the overarching aims of the research before discussing the methodologies employed. In particular, it focuses on the complexities and challenges associated with researching diversity, and more specifically with attempts to engage with, and to collect data from, groups and communities considered ‘hard to reach’. Despite a growth in research involving victim © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_4

43

44  S.-J. HARDY AND N. CHAKRABORTI

groups traditionally described as ‘hard to reach’, we suggest that there are still methodological barriers which prevent researchers from fully engaging with diversity and understanding experiences of victimisation. As preceding chapters have highlighted, there are many groups of ‘Others’ who are targeted because of a specific identity characteristic, and who experience similar forms of victimisation and suffer the same level of physical and emotional impacts as the more familiar, more established victim groups, and yet they continue to occupy a peripheral position within research frameworks. Both methodological and practical challenges have resulted in hate crime victims’ experiences being homogenised—or worse still, ignored—through essentialised conceptualisations of identity. This chapter draws from the methodological dilemmas encountered during fieldwork to illustrate that it is only through developing a more nuanced, less formulaic approach to conducting research that social scientists will achieve more inclusive constructions of ‘difference’, more complete accounts of victimisation and a more comprehensive understanding of the specific issues and problems facing different communities.

Connecting with Hidden Voices This book is grounded within the findings from a series of studies which were based in different geographical locations and which were conducted over the course of seven years. Each of these four studies was commissioned by a different funding body but they shared the common objectives of seeking to uncover lived experiences of hate crime; to understand the physical and emotional harms suffered by victims, their families and wider communities; and to identify ways of improving the quality of support offered to victims. The four studies can be described in the following terms: a two-year piece of research funded by the Economic and Social Research Council which ran from 2012 to 2014 (henceforth referred to as Study One); a four-month study in 2015 for the Equality and Human Rights Commission (Study Two); a four-month study in 2016 commissioned by the OPCC in Hertfordshire (Study Three); and a six-month study in 2017–2018 on behalf of the Office for the Police and Crime Commissioner (OPCC) in the West Midlands (Study Four). Collectively through these studies, we heard from 2073 victims of hate crime (a detailed description of the sample is provided within the next section).

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

45

From the outset, we chose to follow a different path to that taken within conventional studies of hate crime by adopting a new and deliberately inclusive definition of hate crime in order to capture the experiences of anyone, from any background, who felt that they had been targeted on the basis of their identity, ‘difference’ or perceived vulnerability. In doing so, our intention was to give individuals a sense of agency in deciding whether they had been victims of hate crime, and on what grounds, as opposed to interpreting experiences of hate crime through the narrower prism of official definitions or recorded data. To do justice to this intention, however, required us to access and forge connections with the full range of diverse groups and communities who might conceivably be victimised because of their identity, ‘difference’ or perceived vulnerability. Such groups and communities are typically described as ‘hard to reach’, a term which has been widely debated (see, inter alia, Devota et al. 2016; Friemuth and Mettger 1990). “Hard-to-reach” audiences have been called obstinate (2), recalcitrant (3), chronically uninformed (4), disadvantaged (5), have-nots (6), illiterate (7), malfunctional (8), and information poor (5). These labels reflect communicators’ frustration in trying to reach people unlike themselves. (Friemuth and Mettger 1990: 232) [The term hard to reach] will sometimes be used to refer to minority groups, such as immigrants, LGBT people, or the homeless; it can be used to refer to “hidden populations,” groups of people who do not wish to be found or contacted, such as illegal drug users or gang members; at other times it may refer to broader segments of the population, such as the elderly, or young people, or people with disabilities. (Rabnett 2009 cited in Froonjian and Garnett 2013: 832)

Many of the populations referred to by Rabnett (2009) are groups who are especially vulnerable to targeted violence and micro-aggressions, and accordingly ones whom we wanted to access and engage with as part of these studies. What follows is an authentic account of the practical and personal challenges that we faced while undertaking these studies—challenges which are often unacknowledged within the context of research publications (Farrant 2014; Jewkes 2012)—along with the steps that we took to overcome these barriers. Issues relating to positionality, resistance and resilience are covered within the next chapter.

46  S.-J. HARDY AND N. CHAKRABORTI

One of the most significant limitations with conventional engagement approaches within the social sciences is an over-reliance on gatekeepers based within mainstream organisations and on community leaders. While this can be a necessary and appropriate method of engagement in some contexts (Zempi 2016; Shirazi and Mishra 2010), research can be constrained through an over-reliance on gatekeepers and community leaders as a means of representing the opinions and experiences of a specific group. Gatekeepers and self-appointed community leaders are commonly out of touch with, or oblivious to, the concerns of those whom they purportedly represent. As Garland et al. (2006: 428) note, ‘community leaders are often middle-aged (or elderly) males who have no real comprehension of the viewpoints of younger community members or those of females, whose voices may therefore remain unheard by those who are supposedly researching and consulting them’. Community leaders and gatekeepers occupy a position of power in controlling whether researchers are permitted access to a specific group or community and can often deny access without contact or explanation. These are problems that we regularly encountered during the initial engagement phase of each study. Typically, we would establish a steering group or advisory board comprised of agency representatives who were tasked with providing a list of ‘official’ organisations, community groups and associated leaders who they believed would facilitate the most appropriate points of contact. We embarked upon the standard process of engagement familiar to many qualitative researchers which included sending emails, making phone calls and arranging meetings with gatekeepers. However, this approach met with limited success as emails and phone-calls were frequently unanswered, and the majority of meetings led merely to suggestions of further meetings with a ‘more relevant person’ within that organisation or within a different organisation entirely. As the following field diary entry—written by the lead researcher— denotes, we also encountered gatekeepers and community leaders who offered a range of reasons as to why research exploring hate crime victimisation was irrelevant to their service users or community: Today a faith community leader told me that there would be no point in engaging with his community as “no-one experiences this kind of victimisation”. This can now be added to the frustratingly long list of community groups and organisations who have denied us access, such as the asylum seekers and refugee organisation who suggested that being a victim of ‘low level’ hate

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

47

crime was of little significance to their clientele; a mental health organisation who suggested that their service users were too vulnerable to speak to us; and a disability organisation who stated that our presence as researchers would be a hindrance as we lacked the necessary skills to engage with that specific group.

In contrast, there were many organisations that permitted access to service users known to be ‘easy’ to access and ‘willing’ to participate. Reliance on these convenience samples can make the process of engagement and data collection more straightforward (Berg 1999), but at the same time reinforces reliance on familiar, over-researched groups and the marginalisation of ‘hidden’ communities (Devota et al. 2016; Clark 2008). As a way of overcoming this issue, we found that framing interactions with a mainstream organisation—such as a local authority, criminal justice agency, housing association or a well-established charity—as a personalised exchange with tangible benefits for both parties maximises the potential for a positive outcome (see also Benoit et al. 2005). As highlighted in Chapter 2, public and third sector agencies have been forced to make significant cuts to their services and staff over recent years as part of government austerity measures. One of the areas hardest hit has been community engagement, and many front-line professionals have less time and fewer resources to meaningfully engage with diverse communities, as illustrated by the following quotations taken from Study Four: It seems to have got worse since the restructure … You can’t do meaningful engagement. And it’s a shame because there’s real value in it. We did try and work with other departments, agencies and voluntary organisations, and we’d try and get involved with local groups. And that’s how you make people living in the community aware of all these services, it’s important. Without that, I am worried as to where some of this work is, if anywhere, and what’s going to happen as a result of that. Community enforcement officer who works for local government I see a lot of people now who are being made redundant … and that team which dealt with these difficult to engage groups became absorbed. In terms of community engagement, yeah, there’s still stuff happening but it’s not at the same level that it was. Community safety officer who works for local government

48  S.-J. HARDY AND N. CHAKRABORTI I’m expected to go and engage with these groups which I don’t have the time to do and I don’t know how to do it … If I go to a mosque, how am I meant to act? Police officer

The dearth of resource to donate to community engagement, and the ensuing loss of knowledge and awareness-raising, gives researchers an opportunity to remunerate gatekeepers with valuable insights into the perceptions, experiences and support needs of minority communities (Benoit et al. 2005; Denner et al. 1999). As part of our research, we went beyond communicating the benefits of our research for staff members, service users and organisations at an abstract level, and instead provided tangible and tailored outputs for those who facilitated access. This included committing to return to present our research findings, to offering evidence-based recommendations that would inform future practice, or to providing briefing reports which had direct relevance to a particular service provider. Additionally, we factored in time to engage with relevant professionals who have extensive experience of accessing our target populations before attempting to broker access ourselves. Similar to the principle of conducting research with rather than on a particular group or community, we found that inclusive and meaningful engagement approaches are designed with and informed by key stakeholders. This can range from stakeholders acting as ‘experts’ as part of an expert choice sampling approach, or actively participating through community-based research projects or as co-investigators of a study (Mohebbi et al. 2018; Blakeslee et al. 2013; Benoit et al. 2005). Again, in isolation this ‘top-down’ model to participant recruitment is unlikely to yield opportunities to access and engage with ‘hidden’ minority groups but an approach which communicates with gatekeepers in a personalised way and which produces real and valued benefits for stakeholders is more likely to generate sustained support. Another drawback of conventional approaches to engaging with potential participants from minority communities is the use of inaccessible recruitment material and the methods used to disseminate it. In order to garner interest in a particular study, researchers often rely on fairly standard methods to recruit potential participants including posters, leaflets and press releases. A problem with this approach is that the discourse commonly adopted within policy and academic spheres and

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

49

used within recruitment materials is often inaccessible and fails to resonate with the wider public and their experiences, as evidenced by the following field diary extract: I met with Joyce today who had set up a small community group to help African women in the local community. I asked her if she had ever been a victim of hate crime, to which without hesitation she replied “No”. This led to an awkward pause. I decided to fill this by detailing what forms hate crime can take and what aspects of identity someone could be targeted for. She looked up at me, raised her arms and replied “Oh I get that every day”. Joyce had never heard of the term ‘hate crime’ but had experienced racially motivated verbal abuse and harassment as part of her daily life and had never reported it to anyone.

In the early stages of participant recruitment, we found that the term ‘hate crime’, which we had initially—and somewhat naively—assumed would be a commonly known concept, was one which the vast majority of our participants were unfamiliar with. By using the official terminology of ‘hate crime’ and labels such as ‘victim’, we were inadvertently creating a barrier between ourselves and these participants as they seldom recognised their experiences as hate crimes or saw themselves as victims. We sought to address this issue through the use of more accessible, engaging and personalised recruitment materials as part of a nuanced engagement process (Images 4.1 and 4.2). While consideration was given to ensuring that recruitment materials were accessible for those with learning difficulties and for whom English is not a first language, there were still accessibility issues which we had failed to consider. For instance, our recruitment materials were not appropriate for those with physical disabilities who had visual impairments, and we had not taken differences in dialect into account when translating materials into languages such as Arabic. A further challenge associated with recruitment materials relates to how and where content is publicised. Often insufficient thought is given to whether the spaces in which physical copies of recruitment materials are displayed are those in which marginalised and disempowered groups frequent. The same is true of the deployment of mainstream channels of local and national media as a recruitment strategy. While we invested considerable time into taking part in TV and radio interviews as a way of promoting our research, there was an overreliance on orthodox channels of communication which had limited success in enabling us to reach and connect with a diverse audience.

50  S.-J. HARDY AND N. CHAKRABORTI

Images 4.1 and 4.2  Participant recruitment materials used within Study One

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

51

The way in which we addressed some of these challenges was to gather in-depth information on the demography, geography and infrastructure within a given research site, which enabled us to design and implement an engagement approach that was resource-efficient and evidence-based, and that facilitated meaningful opportunities to engage with diversity in its broadest sense (Mohebbi et al. 2018). The intelligence gathered in relation to the demography and geography was used to inform the ways in which we communicated our research in terms of discourse, format and dissemination mechanisms. For instance, the knowledge that we gained about the infrastructure in each area in terms of what services are offered and to who, was used to tailor communication such as email correspondence and phone calls so that it reached the right individual and had direct relevance to their role, services users or organisational purpose. Alongside using the intelligence to produce materials which were accessible—in terms of translating content into languages and dialects of local relevance, and into formats which are appropriate for those with learning and/or physical disabilities—the information that we collected on each research site informed our decision-making in relation to where and how we disseminated the recruitment material which included community and neighbourhood centres, libraries, schools, colleges and universities, health centres and other public spaces. In addition to devoting time to communicating with established and minority radio stations, magazines, and community and neighbourhood newsletters, we also built an active and engaging social media presence via dedicated project websites, Facebook (and Facebook advertising in particular) and Twitter. The websites provided visitors with detailed information about hate crime, sources of support, video updates on each study from the research team and an interactive form entitled ‘Tell Us Your Story’ which enabled participants to share their stories anonymously. This proved to be an effective way of empowering victims who would otherwise be ‘hard to reach’—or reluctant to engage through conventional approaches—to instead share their experiences. The following comment, taken from the ‘Tell Us Your Story’ section of our project website for Study One, highlights the value of using this approach: When I discovered the project and saw the posters across campus I always wondered whether hate crime/abuse in terms of weight issues (obesity or eating disorders, both ends of the scale) would be of concern to your research. I am glad

52  S.-J. HARDY AND N. CHAKRABORTI I can contribute in this way, and I was glad to see the term ‘body shape’ on the header of your website … it is an issue that I think a lot of people seem to ignore or deem as ok.

Of the social media formats, we found that Twitter was the most effective tool in enabling us to connect with a much wider audience and to engage in dialogue with multiple groups of people (see Burns [2010] for a wider discussion of the use of new technologies in engaging with actual and potential research participants). In addition to posting regular bite-size tweets which kept followers up to date with each study and upcoming events and which promoted the link to the online surveys, our use of Twitter extended to developing and delivering a series of ‘Everyday Prejudice’ weeks, inspired by the wider success of the international ‘Everyday Sexism’ project (Bates 2014). Designed to expose the forms of prejudice and hostility that people face in their everyday life as a routine feature of their ‘difference’, these week-long micro-projects received an overwhelmingly positive response from our Twitter followers, with people tweeting in their own experiences of ‘everyday prejudice’, re-tweeting the stories of others and taking part in discussions on the themes raised. As such, the creative use of websites and social media facilitated an increased awareness of, and increased participation within each study as it provided an opportunity to those who we could not physically reach, or who may be reluctant to engage through the more traditional methodological approaches, to become involved in the research (Martinez et al. 2014; Hesse-Biber and Burke Johnson 2013; Bhutta 2012). Although there are invariably significant challenges associated with gaining access to particular groups and communities, researchers have the agency to be more flexible and creative when seeking to engage with such groups. Venturing away from the more formulaic approaches to engagement enables researchers to connect with and capture the lived reality of diverse communities and to recognise that the ‘hard to reach’ are, in fact, eminently reachable. The ‘grassroots’ component of our engagement model relied not only on an initial intelligence gathering phase but also from the additional insights gained from engaging with statutory and charitable organisations. This component was the most time-intensive because it involved identifying, familiarising ourselves with and regularly returning to unconventional access points, referred to as ‘community immersion’ by Mohebbi et al. (2018: 38). As illustrated by the following field diary extract, while undertaking street-based

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

53

reconnaissance in each research site, we came across ‘hidden’ community and social spaces such as international supermarkets, cafes and restaurants, charity shops, community centres, places of worship, pubs and clubs, taxi ranks and homeless shelters. Today I was walking down a residential street when I came across a red door. For the most part it looked like any other terrace house on the street, apart from a bright door with a small sign above it which read ‘Bangladeshi Centre’. This community centre is completely missable; it’s not come to our attention via internet searches or through engagement with our advisory board or gatekeepers.

We found that key ingredients for successful ‘grassroots’ engagement included good interpersonal skills, high levels of patience and a thick skin. Typically, after we had been granted access to a particular group by a gatekeeper or once we had located a new, ‘hidden’ community venue, we would need to interact with potential participants in a much more informal way in order to develop rapport and trust with those groups and communities who are typically disconnected from research, activism and policy-making. Consequently, as part of each of the studies outlined above we spent prolonged periods of time becoming active participants within these different social spaces. This involved taking part in Bhangra classes, various arts and crafts activities, question and answer sessions in English for Speakers of Other Languages (ESOL) lessons, playing chess, pool and bingo, poetry readings, exercise classes, charity walks and a range of other unfamiliar activities. A similar strategy was utilised by Vershinina and Rodionova (2011) who found that the use of communal focus points like Ukrainian food shops, bars or restaurants was highly effective in locating and informally engaging with illegal Ukrainians in London. Tyler (2004: 294) also spoke of the importance of engaging people in Coalville through informal conversations at a grassroots level: I also spent many days just hanging around Coalville’s town centre. I talked to people sitting in the library, various pubs in the town and the Coalville Café, a popular café for local people to drop in for a coffee, a cigarette and a chat with friends. This casual networking enabled me to initiate short conversations with people about their views on Asian, black and Chinese settlers to the area and also made me privy to passing racist comments that I recorded in my fieldwork diary.

54  S.-J. HARDY AND N. CHAKRABORTI

Accessing the voices of those who lack social, cultural and economic capital through official channels or conventional approaches is impractical, and often impossible. However, the adoption of a more immersive grassroots approach enables researchers to become a familiar face, to show their personality and to gain invaluable insights into the lived realities of hate crime for those who are amongst the most vulnerable and disempowered within society.

Connecting with Our Sample Within each of the four studies, we utilised a combination of sampling techniques, including purposive, expert choice, and snowball sampling, which have been found to be especially fruitful when engaging with ‘hard to reach’ audiences or when investigating difficult or sensitive topics (Benoit et al. 2005). As mentioned above, Study One was funded by the Economic and Social Research Council and took place from 2012 to 2014. Through the use of a multi-strand sampling approach, we heard from 1106 victims of hate crime (aged 16 and over) who shared their experiences and perceptions through completion of an online or hard copy survey which had been translated into eight different languages including Arabic, Bengali, Gujarati, Mandarin, Polish, Punjabi, Somali and Urdu, and 374 victims took part in in-depth, semi-structured interviews (a detailed account of the data collection methods is provided in the next section). Study Two was conducted over a four-month period in 2015 and was funded by the Equality and Human Rights Commission. This study focused specifically on the experiences of Lesbian, Gay, Bisexual and Transsexual (LGBT) victims of hate crime based in Leicester and Leicestershire. The team administered a short survey and conducted in-depth face-to-face qualitative interviews to explore experiences and expectations with a total of 50 people who identified as LGB and/or T, of whom 44 identified as being a hate crime victim. Study Three was commissioned by the OPCC in Hertfordshire and took place over a four-month time frame in 2016. We sought to capture the perceptions and experiences of 1652 actual and at-risk victims of hate crime through responses to an online and hard copy survey and in-depth, semi-structured interviews. The term ‘at risk’ is used to define those participants who belong to or share an identity characteristic with a certain group who are especially vulnerable to hate crime victimisation including, for instance, asylum seekers and refugees, minority ethnic

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

55

and/or faith communities, people with physical disabilities and/or learning difficulties and/or mental ill-health, and the trans population. This enabled us to collect and to analyse a broader range of perceptions and expectations relating to decision-making within the context of reporting and accessing support. The same sampling approach and data collection methods were mirrored within Study Four referred to within this chapter, which was conducted over a six-month time frame in 2017 on behalf of the OPCC in the West Midlands. A total of 373 people took part in the West Midlands-based study through participation in an online or hard copy survey or through an interview, of which 142 had experienced hate crime on at least one occasion (Table 4.1). Overall, we heard from a cumulative sample of 2073 research participants, whose profile varied considerably in terms of age, gender identity, ethnicity, religion, sexual orientation and disability. The diversity of the sample is especially apparent within the context of ethnicity, with significant numbers of participants identifying as being Asian British (n = 204), Asian ‘Other’ (n = 42), Black African (n = 113), Black British (n = 73), Black Caribbean (n = 19), Chinese (n = 42), Indian (n = 150), Pakistani (n = 59), White British (n = 1038), White European (n = 66), White ‘Other’ (n = 65) and of Mixed Ethnic Heritage (n = 54). Additionally, 19.5 per cent (n = 405) of participants identified as having some form of disability including mild to severe learning difficulties, long-term health conditions such as muscular sclerosis and HIV, physical disabilities such visual impairments, issues with mobility and impaired speech, and mental health conditions such as anxiety, depression and schizophrenia. Additionally, 55.2 per cent (n = 1145) of participants were female and 43.3 per cent (n = 897) were male. Collectively, the four studies included significant numbers of participants from communities and groups which are often considered ‘hard to reach’. For instance, the research team engaged with members of recently arrived migrant groups such as the Roma, Polish, Somalian, Congolese and Iranian communities (n = 197); asylum seekers and refugees (n = 93); those who identified with Buddhism, Hinduism, Islam, Jainism, Judaism, Sikhism, Paganism, Spiritualism (n = 514); members of the trans ‘community’ (n = 85); lesbian, gay, bisexual, pansexual and asexual people (n = 336); and those who were under 18 years old (n = 124) and over 65 years old (n = 85). The participants within these studies also lived in very different geographical environments, including rural spaces, multicultural settings and economically advantaged and

56  S.-J. HARDY AND N. CHAKRABORTI Table 4.1  Number of participants who had experienced a hate crime and method of data collection Funding body

Study

Number of hate crime victims

Number of hate crime victims who participated through a survey

Economic and Social Research Council Equality and Human Rights Commission Office for the Police and Crime Commissioner (OPCC) for Hertfordshire Office for the Police and Crime Commissioner (OPCC) in the West Midlands Total

Study One

1421

1106

374

Study Two

44

44

44

Study Three

466

446

48

Study Four

142

130

17

2073a

1726

483

a136

Number of hate crime victims who participated through an interview

participants completed a survey and took part in an interview

disadvantaged areas. For example, two of the studies were conducted within the city of Leicester and the surrounding county of Leicestershire. Leicester is one of the most diverse cities within the UK, with just 45 per cent of the city’s almost 330,000 residents identifying as White British within the 2011 Census and yet this differs starkly to the demography of the population within the surrounding county, of whom 89 per cent categorised themselves as White British (Office for National Statistics 2012). This diversity in terms of ethnicity and religion was also apparent within the West Midlands—the site of Study Four—which has been described as a county ‘of contrasts’ (Medland 2011: 1), consisting of seven metropolitan boroughs many of which have a strong industrial heritage. In terms of social and economic diversity, the West Midlands also includes areas of high deprivation with 43 per cent of children in Birmingham specifically growing up in poverty, as well as affluent areas (End Child Poverty 2018).

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

57

Connecting with Our Data Throughout these studies, we adopted a pragmatic paradigm in order to design a methodological approach which was most suited to the exploratory nature of this research and to the participants with whom we were seeking to engage (Creswell 2003; Darlington and Scott 2002). Adopting an exclusively quantitative or qualitative approach to investigate this complex phenomenon would have failed to capture an adequate breadth of experiences or sufficient depth in relation to the lived realities of hate crime (Tashakkori and Teddlie 2003). Often termed a ‘what works’ approach to answering research questions (Onwuegbuzie and Johnson 2006), a mixed-methods pragmatic approach enabled us to collect data in a simultaneous manner and to transition from a subjective stance which permitted participants to construct and to communicate their own reality to a more objective one when analysing the data (Creswell and Plano Clark 2011; Teddlie and Tashakkori 2009). In each of the four studies, participants were asked to share their perceptions and experiences through a survey. As part of Study One, we invested considerable time into the design and content of the survey to ensure that we collected the necessary contextual data on participants’ experiences, along with detailed demographic data. As Garland et al. (2006) noted, there is a tendency for researchers to make generic assumptions about those who fall under ‘catch-all’ identity categories such as ‘Black’, ‘Asian’ or ‘BME’, ‘LGBT’ (lesbian, gay, bisexual and transgender) and ‘disabled’. Such crude analytical frameworks have contributed to a failure by researchers to engage with diversity in its fullest sense and to underplay and even ignore discrete experiences, specificities and intersections in victimisation. Ahn Lin’s (2009) research on hate crimes against Asian Americans is illustrative of this point, highlighting that the experiences of Asian Indians, Cambodians, Chinese, Filipinos, Hmong, Japanese, Koreans, Samoans, Thai and Vietnamese are homogenised by the label ‘Asian Americans’. In recognition of this failure, we adopted a more inclusive and comprehensive approach within the survey. For instance, the question pertaining to a participant’s ethnicity had 29 variables, and the question relating to religious identity included ten variables with the additional option of an open-text box. We were also conscious of designing a questionnaire which minimised the potential for respondent dropout. The majority of survey dropouts occur either before or immediately after the first question, which

58  S.-J. HARDY AND N. CHAKRABORTI

is why the first set of questions were designed to ease respondents into the subject matter by encouraging them to describe how long they had lived in Leicester and to what extent they agreed or disagreed that residents in their local area respected ‘difference’ and got on well together. Additionally, the questionnaire was created to be visually appealing (Dillman et al. 2009), and the online version included a progress bar which has been found to aid completion. Along with having short questions which were written in accessible language, we also separated the questionnaire into distinct sections, with more sensitive questions about personal experiences of victimisation being placed towards the end. The questionnaire included the following sections: • You and your local area • Your thoughts and feelings about hate crime • Your experiences of hate crime in Leicester • Your most recent experience of hate crime • Organisational responses to hate crime • About you. Finally, we undertook cognitive analysis of the pilot questionnaire in both hard copy and online format with ten respondents who came from different backgrounds. Through these in-depth interviews, we were provided with invaluable insight into respondents’ thought processes, including whether they had interpreted the questions and concepts in the way in which we had envisaged and how they had dealt with difficulties and ambiguities when they arose (Collins 2015). We identified a number of issues through the cognitive analysis, including the realisation that certain terms such as ‘harassment’ were not ones with which respondents were familiar or able to associate particular behaviours or acts with. As a result of this process, we made revisions to the questionnaire to maximise its usability and robustness. The questionnaire used within the Study One was used as a template to design the surveys distributed within the subsequent studies.1 As this initial survey was comprehensive in terms of collecting data on hate crime victims’ experiences (including questions relating to when the incident happened, the form it took, the location, how many perpetrators were involved and their demographics, and the steps taken to feel safer) and from a large sample (n = 1106), the focus of the questionnaires administered in Study Three and Four shifted from the nature of hate

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

59

crime victimisation to experiences and expectations of reporting pathways and support provision. Additionally, revisiting the questionnaires provided us with the opportunity to ask about issues which were particularly pertinent at that time, such as the result of the EU referendum and its impact upon perceptions of safety and belonging.2 During the early stages of Study One, datasets were downloaded and analysed on a weekly and then monthly basis in order to keep track not only of the way in which respondents were responding to the questionnaire, but also to see whether there were any significant gaps in overall responses which would warrant any revisions to our engagement approach. In total, 808 questionnaires were completed in hard copy and 298 were completed online, which further highlights the importance of adopting a more hands-on, ‘grassroots’ approach to engagement with those groups who occupy a marginal position within society and who are unlikely to respond via digital technologies. Within each of the studies, after the survey was closed both the hard copy and online versions were integrated into SPSS and subjected to validation checks to assess logic, consistency and reasonableness. Significance testing was undertaken using SPSS, producing both frequency and contingency tables which enabled us to identify prominent findings within the overall sample, along with the commonalities and differences between various groups of hate crime victims. In order to complement the survey findings, we also conducted in-depth semi-structured interviews within each study, resulting in a collective sample of 483 testimonies. This qualitative approach facilitated the exploration of a complex phenomenon by generating ‘thick’ description of each participant’s social world and their experiences within it (Yeo et al. 2014). This method of data collection is reliant upon a researcher’s ability to foster a meaningful collaboration with the participant, as explored more fully within the next chapter. In-depth interviews have the potential to empower participants by enabling them to articulate their own story in their own way, and by conveying a message that the participant’s experiences are valued (Rubin and Rubin 1995). Lincoln (1995: 6) describes qualitative research as involving a commitment to promoting ‘social justice, community, diversity, civic discourse and care’. This was particularly pertinent in the context of our research where many of the participants occupied a peripheral position within society and had never shared their experiences with a criminal justice agency, with any other relevant organisation or through academic research.

60  S.-J. HARDY AND N. CHAKRABORTI

For the most part, interviews were conducted in naturalistic settings such as cafes, community centres, libraries, offices and homes as we wanted to accommodate participants’ preferences while ensuring the safety of the research team. The majority of interviews were conducted face to face with a single participant, but in some instances interviews were conducted with multiple victims, or in the presence of family members, friends, a carer or translator, as appropriate, to facilitate participation. Similarly, where participants requested an alternative mode of discourse such as a telephone interview or through asynchronous email exchange, we were able to facilitate this. Within each study, we developed an interview schedule which set out key topics to address during the interview, while maintaining a flexible approach in order to probe significant themes which arose over the course of each conversation.3 All interviews were digitally recorded, transcribed and stored in a secure file store in accordance with relevant information security policies and the Data Protection Act, and which only the research team had access to. To ensure that we distinguished between two interrelated but key stages of analysis—data management and interpretation (Ritchie and Lewis 2014)—we adopted a robust, inductive approach to analysing the large volume of qualitative data generated within each study. Our analytic journey was iterative, involving thematic content analysis when batches of interview transcripts were available and modifying the interview schedule depending on emergent themes. The lead researcher would undertake open coding by reading each transcript, summarising key points and themes within the margins and highlighting ‘dross’ (Burnard et al. 2008). These codes were then brought together to compile a list of relevant analytical categories, and we then worked through each transcript, appropriately highlighting portions of data which had relevance to the overarching thematic categories. Finally, relevant sections of data from each transcript were transferred to a separate document, enabling us to identify and disseminate significant, common and divergent themes. During the initial open-coding process, members of the research team independently reviewed a sample of analysed transcripts to verify the emergent themes and to reduce the element of bias (Spencer et al. 2014; Barbour 2001). In order to facilitate triangulation of our research data, we made use of a field diary. Primarily, the field diary was used to document relevant contextual information which complemented both the quantitative data collected through the survey and the qualitative data generated from

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

61

in-depth interviews (Phillippi and Lauderdale 2017; Creswell 2013). During the course of each study, we produced rich descriptions of our observations before, during or after an interview, and documented our experiences and feelings while in the field. This data collection method has been termed auto-ethnography and has increasingly been recognised as an important source of data which encourages researchers to become active participants within their research. Auto-ethnographical observations provide additional insight into the context and impact of victimisation, as well as the relationship between the researcher and the researched. As Jewkes (2012: 63) explains: Criminology has largely resisted the notion that qualitative inquiry has auto-ethnographic dimensions and remained quiet on the subject of the emotional investment required of ethnographic fieldworkers studying stigmatized and/or vulnerable ‘others’ in settings where differential indices of power, authority, vulnerability and despair are felt more keenly than most.

Often personal experiences, and more specifically the researcher’s emotional endeavour while in the field, are underappreciated and fail to feature within the ‘final product’ (Jewkes 2012; Farrant 2014). Although social scientists have become more aware that research topics are often influenced by the researcher’s personal interests and are therefore inherently personal (Liebling 1999), academics often fail to document the emotional investment required when conducting a study (Jewkes 2012; Fleetwood 2009). Sections of the field diary were transcribed, and relevant nonverbal content was incorporated into the interview transcript to provide further contextual information (Sandelowski 1994). Below is one example of an extract which provided additional contextual data for an interview: The interview recording is powerful but it fails to capture some of the things I’ve observed during the course of getting to know George. Take, for instance, the physical environment George lives in and the safety measures he incorporates into his everyday life to reduce the risk of victimisation. He has to have multiple locks on his front door; he keeps the curtains drawn at all times; he is too scared to turn on the lights even if it means he has to sit alone in darkness; and he rarely ventures out of his flat. That’s the reality of what life is like for George.

62  S.-J. HARDY AND N. CHAKRABORTI

Our personal experiences, feelings and reflections were stored in a separate document which was subjected to thematic content analysis, and these entries are drawn upon within the next chapter to bring to life key issues such as positionality, subjectivity, resilience and resistance.

Conclusion This chapter has outlined some of the dilemmas facing researchers who wish to locate, engage with and understand marginalised or ‘hidden’ communities. In particular, it has highlighted that conventional approaches are hampered by methodological and practical shortcomings which prevent researchers from fully engaging with diverse populations. We discussed the process of conducting a series of studies focusing on the experiences of ‘hidden’ populations in order to explore some of the limitations inherent within our engagement approaches. In particular, we reflected upon the barriers which can arise when adopting a ‘top-down’ approach relying on gatekeepers, community leaders and official organisations to access potential participants. With this in mind, we embraced a rather different model of engagement which incorporated a more sophisticated and intelligence-led approach to accessing marginalised and diverse communities involving more innovative modes of grassroots engagement. Communal hubs of social interaction such as international supermarkets, cafes and restaurants, places of worship and taxi ranks (to name just some examples) were invaluable access points, while our involvement in a wide range of activities such as arts and crafts, sports groups, dance classes and faith events proved fruitful methods of engagement with ‘hidden’ communities. The chapter moved on to provide a comprehensive account of the sample of participants involved in our research and the methods of data collection, management and analysis that we employed. Throughout all stages of the research process—whether designing our data collection tools, administering the questionnaires, conducting in-depth interviews, documenting field notes or analysing the data—every effort was taken to maximise rigour and robustness. Employing a pragmatic paradigm which incorporated a mixed-methods approach enabled a triangulation of data, thereby yielding the level of rich data required to fully understand the multifaceted phenomenon of hate crime victimisation.

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

63

Notes 1.  The original questionnaire can be accessed via the UK Data Service https://doi.org/10.5255/ukda-sn-851570. 2. Part III of this book—which focuses on the lived experiences of hate crime victims—draws from quantitative data taken largely from Study One, but where relevant the findings from the other studies are drawn upon and this is explicitly stated. 3. As an example, the interview schedule that we used during the Study One and which was a template for subsequent studies can be accessed via the UK Data Service.

References Ahn Lin, H. (2009). Race, bigotry and hate crime: Asian Americans and the construction of difference. In B. Perry (Ed.), Hate crimes volume 3: The victims of hate crime (pp. 65–84). Westport, CT: Praeger. Barbour, R. S. (2001). Checklists for improving rigour in qualitative research: A case of the tail wagging the dog? British Medical Journal, 322, 1115–1117. Bates, L. (2014). Everyday sexism. London, UK: Simon & Schuster. Benoit, C., Jansson, M., Millar, A., & Phillips, R. (2005). Community-academic research on hard-to-reach populations: Benefits and challenges. Qualitative Health Research, 15(2), 263–282. Berg, J. (1999). Gaining access to underresearched populations in women’s health care research. Health Care for Women International, 20, 237–243. Bhutta, C. B. (2012). Not by the book: Facebook as a sampling frame. Sociological Methods & Research, 41(1), 57–88. Blakeslee, J. E., Del Quest, A., Powers, J., Powers, L. E., Geenen, S., Nelson, M., et al. (2013). Reaching everyone: Promoting the inclusion of youth with disabilities in evaluating foster care outcomes. Children and Youth Services Review, 35(11), 1801–1808. Burnard, P., Gill, P., Stewart, K., Treasure, E., & Chadwick, B. (2008). Analysing and presenting qualitative data. British Dental Journal, 204(8), 429–432. Burns, E. (2010). Developing email interview practices in qualitative research. Sociological Research Online, 15 (4). http://www.socresonline.org.uk/15/4/8. html. Clark, T. (2008). ‘We’re over-researched here!’: Exploring accounts of research fatigue within qualitative research engagements. Sociology, 42(5), 953–970. Collins, D. (2015). Cognitive interviewing practice. London: Sage. Creswell, J. W. (2003). Research design: Qualitative, quantitative, and mixed methods approaches (2nd ed.). London: Sage.

64  S.-J. HARDY AND N. CHAKRABORTI Creswell, J. W. (2013). Qualitative inquiry & research design: Choosing among five approaches (3rd ed.). London: Sage. Creswell, J. W., & Plano Clark, V. L. (2011). Designing and conducting mixed methods research. London: Sage. Darlington, Y., & Scott, D. (2002). Qualitative research in practice: Stories from the field. Buckingham: Open University Press. Denner, J., Cooper, C., Lopez, E., & Dunbar, N. (1999). Beyond “giving science away”: How university- community partnerships inform youth programs, research, and policy. Social Policy Report: Society for Research in Child Development, 13(1), 1–17. Devota, K., Woodhall-Melnik, J., Pedersen, C., Wendaferew, A., Dowbor, T. P., Guilcher, S. J. T., et al. (2016). Enriching qualitative research by engaging peer interviewers: A case study. Qualitative Research, 16(6), 661–680. Dillman, D. A., Phelps, G., Tortora, R., Swift, K., Kohrell, J., Berck, J., et al. (2009). Response rate and measurement differences in mixed-mode surveys using mail, telephone, interactive voice response (IVR) and the Internet. Social Science Research, 38, 1–18. End Child Poverty. (2018). More than half of children now living in poverty in some parts of the UK. Available at https://www.endchildpoverty.org.uk/ more-than-half-of-children-now-living-in-poverty-in-some-parts-of-the-uk/. Accessed 22 October 2019. Farrant, F. (2014). Unconcealment: What happens when we tell stories. Qualitative Inquiry, 20(4), 461–470. Fleetwood, J. (2009). Emotional work: Ethnographic fieldwork in prisons in Ecuador. Special Issue: Critical Issues in Researching in Hidden Communities, pp. 28–50. Friemuth, V. S., & Mettger, W. (1990). Is there a hard-to-reach audience? Public Health Reports, 105(3), 232–233. Froonjian, J., & Garnett, J. L. (2013). Reaching the hard to reach: Drawing lessons from research and practice. International Journal of Public Administration, 36(12), 831–839. Garland, J., Spalek, B., & Chakraborti, N. (2006). Hearing lost voices: Issues in researching “hidden” minority ethnic communities. British Journal of Criminology, 46(3), 423–437. Hesse-Biber, S., & Burke Johnson, R. (2013). Coming at things differently: Future directions of possible engagement with mixed methods research. Journal of Mixed Methods Research, 7(2), 103–109. Jewkes, Y. (2012). Autoethnography and emotion as intellectual resources: Doing prison research differently. Qualitative Inquiry, 18(1), 63–75. Liebling, A. (1999). Doing research in prison: Breaking the silence? Theoretical Criminology, 3(2), 147–173.

4  THE PROCESS OF ENGAGEMENT: HARD TO REACH OR EASY TO IGNORE? 

65

Lincoln, Y. S. (1995). Emerging criteria for quality in qualitative and interpretive research. Qualitative Inquiry, 1(3), 275–289. Martinez, O., Wu, E., Shultz, A. Z., Capote, J., López Rios, J., Sandfort, T., et al. (2014). Still a hard-to-reach population? Using social media to recruit Latino gay couples for an HIV intervention adaptation study. Journal of Medical Internet Research, 16(4), 1–14. Medland, A. (2011). Portrait of the West Midlands. Office for National Statistics. London. Mohebbi, M., Linders, A., & Chifos, C. (2018). Community immersion, trust-building, and recruitment among hard to reach populations: A case study of Muslim women in Detroit metro area. Qualitative Sociology Review, 14(3), 24–44. Office for National Statistics. (2012). Ethnicity and national identity in England and Wales: 2011. Available at https://www.ons.gov.uk/peoplepopulationandcommunity/culturalidentity/ethnicity/articles/ethnicityandnationalidentityinenglandandwales/2012-12-11. Accessed 22 October 2019. Onwuegbuzie, A. J., & Johnson, R. B. (2006). The validity issues in mixed research. Research in the Schools, 13(1), 48–63. Phillippi, J., & Lauderdale, J. (2017). A guide to field notes for qualitative research: Context and conversation. Qualitative Health Research, 28(3), 381–388. Ritchie, J., & Lewis, J. (2014). Qualitative research practice: A guide for social science students and researchers. London: Sage. Rubin, H. J., & Rubin, I. S. (1995). Qualitative interviewing: The art of hearing data. London: Sage. Sandelowski, M. (1994). Focus on qualitative methods: Notes on transcription. Research in Nursing & Health, 17(4), 311–314. Shirazi, F., & Mishra, S. (2010). Young Muslim women on the face veil (niqab): A tool of resistance in Europe but rejected in the United States. International Journal of Cultural Studies, 13(1), 43–62. Spencer, L., Ritchie, J., & O’Connor, W. (2014). Analysis: principles and processes. In J. Ritchie & J. Lewis (Eds.), Qualitative research practice: A guide for social science students and researchers (pp. 199–218). London: Sage. Tashakkori, A., & Teddlie, C. (2003). Handbook of mixed methods in social and behavioural research. London: Sage. Teddlie, C., & Tashakkori, A. (2009). Foundations of mixed methods research: Integrating quantitative and qualitative approaches in the social and behavioural sciences. London: Sage. Tyler, K. (2004). Reflexivity, tradition and racism in a former mining town. Ethnic and Racial Studies, 27(2), 290–309. Vershinina, N., & Rodionova, Y. (2011). Methodological issues in studying hidden populations operating in informal economy. International Journal of Sociology and Social Policy, 31(11/12), 697–716.

66  S.-J. HARDY AND N. CHAKRABORTI Yeo, A., Legard, R., Keegan, J., & Ward, K. (2014). In-depth interviews. In J. Ritchie & J. Lewis (Eds.), Qualitative research practice: A guide for social science students and researchers (pp. 177–210). London: Sage. Zempi, I. (2016). Negotiating constructions of insider and outsider status in research with Veiled Muslim women victims of islamophobic hate crime. Sociological Research Online, 21(4), 1–8.

CHAPTER 5

Lessons from the Field

Abstract  This chapter draws upon the observations, experiences and reflections documented within a field diary to offer insight into the reality of conducting research of this nature. It reveals the kinds of challenges, dilemmas and harrowing encounters that we, like many scholars who conduct research on sensitive topics or with ‘hard to reach’ groups, have encountered while in the field. In particular, this chapter draws attention to the relevance of researcher positionality and subjectivity, of resistance and research fatigue, and of empathy and resilience in the context of hate crime fieldwork. Keywords  Researcher subjectivity Barriers to engagement

· Resilience · Hard to reach ·

The previous chapter provided a detailed account of the methodological approaches employed within each of the four studies that we completed over a five-year period. It outlined the overarching aims of these studies and described the multilevel engagement approach that we employed to maximise opportunities to connect with a diverse sample which included those who often find themselves excluded from mainstream society, policy-making and academic research. It also explained why we utilised a triad of methods—including surveys, in-depth interviews and field notes—to collect data on hate crime victimisation, and how the data generated through these methods was stored and analysed. © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_5

67

68  S.-J. HARDY AND N. CHAKRABORTI

This chapter builds upon those points by drawing upon the observations, encounters and reflections documented within a field diary to offer insight into the reality of conducting research of this nature. There are growing calls for researchers to acknowledge the subjective experience much more prominently within their analysis and presentation of findings. As Farrant (2014: 461) suggests, the ‘story of “doing” research’— which details the ‘concerns, fears, thrills, and frustrations involved in the research process’—should be embraced as it helps us to understand the complexity of human behaviour and the importance of ‘contaminated’ research relationships. In particular, this chapter draws attention to the relevance of researcher positionality and subjectivity, of resistance and research fatigue, and of empathy and resilience in the context of hate crime fieldwork.

Acknowledging Positionality It is widely acknowledged that the demographics, appearance and behaviour of a researcher can significantly influence the research process, particularly in relation to access and engagement with ‘hidden’ minority groups (see, inter alia, Zempi 2016; Savvides et al. 2014; Zubair et al. 2012; Corbin-Dwyer and Buckle 2009). Researchers often occupy a position of power, ‘insulated from the realities they are studying’ which can lead to ‘multiple blind spots’ (Devotta et al. 2016: 663; Denzin 2003). The issue of whether researchers from a minority group are best placed to understand the lived experiences of minority communities has been widely discussed by social scientists. One line of argument posits that ‘insiders’ (i.e. those from within the ascribed status group) have a greater ability to relate to their research participants’ cultural practices and norms. Phillips and Earle (2010: 361) explain that the ‘insider’ position assumes that ‘only black people can possess a genuine understanding of the social realities of black feelings, values, behaviours and cultural experiences’. In Tyler’s (2004: 294) study, she found that her whiteness ‘was automatically understood in these contexts as a sign’ that she would conform to the expressions of racism in the ex-mining town of Coalville, and therefore, engagement with a relatively ‘difficult’ community was easier due to ‘insider’ status. While the advantages of being an ‘insider’ have been discussed in depth, what actually qualifies someone as an ‘insider’ is still unknown. Mohebbi et al. (2018: 26) suggests:

5  LESSONS FROM THE FIELD 

69

At the most basic level, scholars often treat major social statuses as insider/outsider categories; that is, scholars who claim membership in the status groups they study – gender groups, ethnic groups, racial groups – are insiders (Zinn, 2001; Young, 2008). But, as a fairly sizeable literature demonstrates, there are numerous other group belongings, experiences, and/or characteristics that can serve insider purposes, including religion and religiosity (Widdicombe 2015; Ahmed 2016), motherhood (Brown and de Casanova, 2009), experiences of domestic abuse (Malpass, Sales, and Feder, 2016), length of residence (Crow, Allan, and Summers, 2001), shared profession (Teusner, 2016), friendship groups (Appleton, 2011; Jenkins, 2013), fashion modeling (Mears 2013), and criminal conviction (Earle, 2014; Newbold et al., 2014).

As highlighted by Garland et al. (2006), researchers have multiple strands of identity and discrete personal experiences, characteristics and histories which will all impact on their capacity to engage with diverse populations. In her work exploring the experiences of Muslim women who wear the Hijab, Spalek (2002) emphasised her position as a woman in order to overcome potential barriers in establishing affiliation and trust with participants who had a different religious and ethnic identity. Similarly, Zempi and Chakraborti (2014) suggest that being able to convey religious and cultural ignorance as an ‘outsider’ can in fact be advantageous because it can empower participants by putting them in a position of authority as educators. The research teams involved in the four studies outlined in the previous chapter had a diverse composition in terms of gender, ethnicity, sexuality, age and regional background, and we all encountered challenges when trying to engage with different populations. In particular, the lead researcher faced a range of difficulties and mixed receptions as a young White British female which were noted in the following field diary accounts: I had my first meeting with a religious community leader today. It didn’t go particularly well, especially when I asked about ways to engage with his community and he said, “I don’t mean to be racist or sexist …” (I had an inkling he might be about to say something both racist and sexist) “… but you’re not going to be successful being white and a woman”.

And:

70  S.-J. HARDY AND N. CHAKRABORTI I visited the Pakistan Community Centre today and met with the Director. I had to work incredibly hard for him to take me seriously. After 20 minutes he asked me what I thought about the recent video on YouTube which mocked Mohammed. I knew that this was the question; the question that if I answered it correctly – or at least correctly from his perspective – it would permit us access to his service users.

However, she actually found that her gender and regional accent were advantageous during the fieldwork phase, transcending other ‘differences’ when it came to engagement with both female and male participants (Labaree 2002). Equally changing her appearance—particularly her dress—and taking part in the grass-roots activities outlined in the previous chapter assisted in breaking down the perceived power imbalance between herself and certain groups, including asylum seekers and refugees, the homeless, Gypsies and Travellers and young people (Okely 2007). Similarly, her ‘outsider’ status was perceived to make her ‘safer’ to talk to when accessing particularly close-knit communities, such as minority faith groups (Zempi 2016; Spalek 2005). Attempting to identity-match, the researcher with a particular community was unachievable within the context of our research because we sought to connect with anyone, from any background, who felt that they had been targeted on the basis of their identity, ‘difference’ or perceived vulnerability. Additionally, there are benefits to occupying an ‘outsider’ position in terms of being less likely to overlook emergent themes during the course of interviews or the data analysis process as a result of shared experiences (Kitzinger and Wilkinson 1997). The objectivity which comes from being detached from a given lived reality can also reduce the likelihood of participants not disclosing certain information due to assumed knowledge (Zempi 2016; Turnbull 2000). In essence, we occupied what Labaree (2002: 102) describes as the ‘space between’, embodying and displaying both ‘insider’ and ‘outsider’ characteristics which enabled us to ‘operate in a fluid space somewhere between the two’ (Zempi 2016: n.p.). As a research team, we actively embedded critical reflexivity into the research journey to ensure that we were consciously aware of, considered and revisited issues relating to positionality throughout each study (Savvides et al. 2014; Hellawell 2006).

5  LESSONS FROM THE FIELD 

71

Acknowledging Resistance Policy-makers, practitioners and academics have become somewhat adept at attributing a lack of engagement with certain groups to a perceived failing on the part of those groups not to engage, as opposed to acknowledging their own failings in not investing the time, resource and effort to develop the necessary bonds of trust or channels of communication required for this engagement. Throughout the fieldwork phase, we were told on countless occasions by potential participants that community engagement—whether this be led by a researcher, policy-maker or professional—is tokenistic as these interactions are often one-dimensional or a one-off occurrence. Researchers in particular can encounter additional criticism for appearing to be ‘tourists’ who dipin to an environment, extract distressing accounts of victimisation and then disappear (Höglund 2011; Tomlinson et al. 2006). Such an engagement strategy not only tarnishes the reputation of academics and calls into question their commitment to a particular topic or community, but also has the potential to hamper future researchers in being able to negotiate access and recruit participants. Within each of the four studies, we experienced resistance to our research from gatekeepers and key stakeholders. As Benoit et al. (2005: 270) observes, community organisations ‘tend to be wary about occupants of the university “ivory tower”’, not only in relation to their intentions but also in their ability to engage with their service users. This is an issue that we encountered during the commission of our research, and the following field diary extract depicts one such instance: I attended the breakfast session at a homeless shelter this morning and the reception from both staff and services users was unwelcoming to say the least. At one point I went over to get a tea and the staff member who was serving loudly stated, “I’m guessing that you’ll want a fresh pot making and not one of the crap ones that these lot are forced to drink”. I felt the room silence and I wanted the ground to open up. All I could do was respond with “No, happy to drink anything!”

As discussed in the previous chapter, there is no easy solution to overcoming scepticism, suspicion and unfriendliness as part of the research process. In this instance, the lead researcher returned to the group week after week, sitting on different tables, playing card games and chess,

72  S.-J. HARDY AND N. CHAKRABORTI

sharing insights into her own life, helping services users with administrative tasks and staff members by cleaning mugs and plates, in an attempt to establish the necessary bonds of trust and rapport. Cynicism from gatekeepers can extend to the perceived usability of research outputs and the potential impact that a study’s findings will have on ‘the real world’ (Benoit et al. 2005). We were confronted with this issue early on when an experienced criminal justice professional on the advisory board for Study One announced to the research team and his fellow advisory board members that: “I won’t be reading the report, it’ll be too long, too boring and I probably won’t understand it!” While this was hugely frustrating given the direct relevance of the findings to his organisation, it reinforced the importance of producing accessible outputs with tangible recommendations in different formats, including executive summaries, briefing reports, manifestos and online training modules. Over the course of our empirical work, we have also been presented with a series of challenges in relation to resistance from potential participants. A reluctance or unwillingness to engage with researchers has been associated with a wide range of factors, including language barriers, practical issues with getting to or taking part in a particular study such as inadequate transportation or inaccessible data collection tools, past negative experiences with statutory or charitable organisations, and involvement in illicit activities (Devotta et al. 2016; Froonjian and Garnett 2013). While we were able to overcome many of these challenges by adopting a flexible, multifaceted methodological approach, we still faced resistance in some instances as a result of research fatigue: I walked in and I could feel that all eyes were on me. I sat down near the buffet and observed as the group caught up with each other, feeling like a voyeur. I knew immediately that getting out my Dictaphone and introducing myself as a researcher for a hate crime project would ruin the mood and any hope that I had of engaging with them. I walked over to a small group of trans women and the first thing they said to me was, “You’re a researcher aren’t you?” The group went on to explain that they only get the opportunity to meet, to be themselves and to feel comfortable once a month and that this space was often tainted by the amount of researchers who visited to dissect their lives.

Research fatigue is often a consequence of feeling over-researched and under-rewarded (Armitage 2008; Clark 2008), as highlighted by a participant as part of Brugge and Missaghian’s (2006: 493) research:

5  LESSONS FROM THE FIELD 

73

The researcher has the luxury of studying the community as an object of science, whereas the young Indian, who knows the nuances of tribal life, receives nothing in the way of compensation or recognition for his knowledge, and instead must continue to do jobs, often manual labor, that have considerably less prestige. If knowledge of the Indian community is so valuable, how can non-Indians receive so much compensation for their small knowledge and Indians receive so little for their extensive knowledge?

While this sentiment was apparent within the context of certain communities, such as trans people and Muslim women, many of the groups that we accessed through grass-roots access points had never been involved in academic research, let alone a study focusing on their experiences of hate crime victimisation. Although a significant proportion of those whom we engaged with expressed some level of wariness about their participation, we found that approaching interactions in an empathetic and authentic manner served to allay initial concerns. It also became apparent that many participants had attempted to report their experiences to the police or through another relevant organisation on previous occasions but had encountered inadequate and dismissive responses. These participants often observed that the process of sharing their story through an in-depth interview with an empathetic and receptive researcher had helped to make them feel validated and valued. As part of these studies, we were also able to inform participants—many of whom were unaware that hate crime policy existed and that recording practices are predicated on their own perception of the incident being motivated by hostility or prejudice towards a particular feature of their identity—about their rights and about the services available to support them.

Acknowledging Resilience The final theme which emerged prominently from our analysis of field diary entries relates to resilience both in terms of our participants and the research team. Most researchers are well versed in the importance of ensuring that participants are unharmed during the research process, and throughout each of the studies the well-being of our participants was at the forefront our minds, influencing how we approached the data collection, analysis, presentation and dissemination. And yet, despite adhering to ethical guidelines, creating a safe environment and providing

74  S.-J. HARDY AND N. CHAKRABORTI

after-care through support information leaflets, there were many times that these steps did not feel sufficient enough to safeguard those whose experiences of hate crime were ongoing: I conducted a difficult interview with Jenny today. The joviality that surrounded us in the café starkly contrasted with the harrowing account that she gave. She told me that every time she leaves her front door she’s called “tranny”, “faggot”, “queer” and when she gets on the bus, she’s called “tranny”, “faggot”, “queer”. She told me that she can’t get a job and that her family have disowned her because she’s trans. And at the end of the interview, she said “Is this going to make a difference?” and I felt such a weight of responsibility. Today I received a call from a police officer who knew that I had interviewed Jenny a couple of months ago. She wanted to let me know that Jenny had killed herself.

The sense of guilt, frustration and sadness felt by researchers during and after conducting an in-depth interview on a sensitive theme is rarely acknowledged. One of the biggest dilemmas that we needed to navigate came as a result of needing to establish a meaningful relationship with participants in order to facilitate an environment in which they felt sufficiently comfortable and confident to disclose traumatic events and at the same time needing to maintain a sense of professional distance and objectivity in the light of our role as researchers. Hearing distressing accounts from someone that you share a characteristic with or that you have established a relationship with can stay with you for days, months and even years. I can’t stop thinking about David who I interviewed yesterday. I met him in the car park, he was smiling and we instantly ‘hit it off’. We sat before the interview drinking a coffee and it felt as if we had known each other for ages. This is probably why I felt so under prepared for what he had been through. He was gay and had been befriended by a group of men after he’d broken up with his partner. Whilst asleep after a night out, one of these ‘friends’ raped him and subsequently gave him HIV. Whilst he was coming to terms with his status, his sister died. Since this time he has faced micro-aggressions, verbal abuse and harassment from his own family, friends and strangers. He got upset during the interview and I asked on multiple occasions whether he’d like to stop and he said, “No, I need to tell you what I’ve been through.”

5  LESSONS FROM THE FIELD 

75

And: I had one of the most horrific interviews I’ve ever done today. The family did not speak English so Adil offered to translate. I couldn’t believe what I heard [Read the field diary extract on page ** for a description of the family’s experience]. At the end of the interview the father shook my hand and said something in Kurdish. I was still holding the man’s hand as I turned to Adil who said “he says thank you for being a White face that’s been friendly and listened”. I could feel the tears in my eyes and I struggled to form any words. I got into my car and just cried. After a few minutes I drove around the corner to see the whole family standing outside the building, smiling at me and waving goodbye. I could feel the tears coming again.

As Lee-Treweek and Linkogle (2000: 1) highlight, ‘the issue of their own or their co-researchers’ safety and welfare needs is often thought through in a cursory manner or in an ad hoc contingent fashion once in the field’. Our experience of conducting research in this field over the course of the last two decades meant that we were acutely aware of the importance of building resilience within the research team and avoiding emotional and physical exhaustion. From the outset of each study, we embedded a formalised approach involving weekly supervisions to debrief and to discuss difficult situations, and opportunities for members of the research team to access professional counselling support. Managing the relationships that you build with participants is particularly crucial when engaging with people who are especially isolated and disempowered. Knowing the most appropriate way to ‘finish’ what Farrant (2014: 461) refers to as a ‘contaminated’ relationship is difficult, particularly when participants continue to reach out for support through phone calls, emails and visits: A few months ago I met Sarah; she has mental ill-health and felt that she was being targeted by her neighbours because she lived on her own and is perceived to be an easy target. I sat with Sarah for over an hour on the first meeting to conduct the interview. She was upset afterwards so I made her a tea and chatted about everyday things. Since this time Sarah has returned every week or two to drop in and talk. She asked me yesterday if I could go with her to a mediation session involving her neighbours. It was such a difficult conversation to have as I know that she has nobody else to turn to but I can’t provide the support that she needs.

76  S.-J. HARDY AND N. CHAKRABORTI

Encounters such as these undoubtedly exacerbate existing feelings of guilt and frustration and reinforce the weight of responsibility upon us to ensure that the ordeal that victims subject themselves to in revisiting traumatic experiences has been worthwhile. As a research team, we have often reflected upon how our interactions within the field have affected us but we are of the view that this is not necessarily a wholly negative experience. These encounters, and the emotions that they engender, drive us to do everything that we can to achieve social change; and to ensure that those in positions of power who are often oblivious to these lived realities are required to confront them, are held accountable for them and are provided with the means to deliver meaningful action to safeguard against them.

Conclusion This chapter has revealed the kinds of challenges, dilemmas and harrowing encounters that we, like many scholars who conduct research on sensitive topics or with ‘hard to reach’ groups, have encountered while in the field. While some of these relate to positionality, in terms of negotiating our changeable positions as ‘insiders’ and ‘outsiders’, we also faced ongoing difficulties in establishing, managing and closing relationships with gatekeepers and participants alike. We also discussed the importance of resilience while undertaking harrowing fieldwork and acknowledged the feelings of guilt, sadness and frustration which often permeate all aspects of the research journey. By offering this authentic account, we have sought to convey the originality, complexity and robustness of the methodological approach that we employed within each study which gave us unparalleled insights into the ‘hidden’ worlds of more than 2000 hate crime victims.

References Armitage, J. S. (2008). Persona non grata: Dilemmas of being an outsider researching immigration reform activism. Qualitative Research, 8(2), 155–177. Benoit, C., Jansson, M., Millar, A., & Phillips, R. (2005). Community-academic research on hard-to-reach populations: Benefits and challenges. Qualitative Health Research, 15(2), 263–282.

5  LESSONS FROM THE FIELD 

77

Brugge, D., & Missaghian, M. (2006). Protecting the Navajo People through tribal regulation of research. Science and Engineering Ethics, 12, 491–507. Clark, T. (2008). ‘We’re over-researched here!’: Exploring accounts of research fatigue within qualitative research engagements. Sociology, 42(5), 953–970. Corbin-Dwyer, S., & Buckle, J. (2009). The space between: On being an insideroutsider in qualitative research. International Journal of Qualitative Methods, 8(1), 54–63. Denzin, N. (2003). The cinematic society and the reflexive interview. In J. Gubrium & J. Holstein (Eds.), Postmodern interviewing (pp. 141–156). London: Sage. Devotta, K., Woodhall-Melnik, J., Pedersen, C., Wendaferew, A., Dowbor, T. P., Guilcher, S. J. T., et al. (2016). Enriching qualitative research by engaging peer interviewers: A case study. Qualitative Research, 16(6), 661–680. Farrant, F. (2014). Unconcealment: What happens when we tell stories. Qualitative Inquiry, 20(4), 461–470. Froonjian, J., & Garnett, J. L. (2013). Reaching the hard to reach: Drawing lessons from research and practice. International Journal of Public Administration, 36(12), 831–839. Garland, J., Spalek, B., & Chakraborti, N. (2006). Hearing lost voices: Issues in researching “hidden” minority ethnic communities. British Journal of Criminology, 46(3), 423–437. Hellawell, D. (2006). Inside-out: Analysis of the insider-outsider concept as a heuristic device to develop reflexivity in students doing qualitative research. Teaching in Higher Education, 11(4), 483–494. Höglund, K. (2011). Comparative field research in war-torn societies. London: Routledge. Kitzinger, C., & Wilkinson, S. (1997). Validating women’s experience? Dilemmas in feminist research. Feminism and Psychology, 7, 566–574. Labaree, R. V. (2002). The risk of ‘going observationalist’: Negotiating the hidden dilemmas of being an insider participant observer. Qualitative Research, 2(1), 97–122. Lee-Treweek, G., & Linkogle, S. (2000). Danger in the field: Risk and ethics in social research. London: Routledge. Mohebbi, M., Linders, A., & Chifos, C. (2018). Community immersion, trust-building, and recruitment among hard to reach populations: A case study of Muslim women in Detroit metro area. Qualitative Sociology Review, 14(3), 24–44. Okely, J. (2007). Fieldwork embodied. In C. Shilling (Ed.), Embodying sociology: Retrospect, progress and prospects (pp. 65–79). Oxford: Blackwell. Phillips, C., & Earle, R. (2010). Reading difference differently? Identity, epistemology and prison ethnography. British Journal of Criminology, 50(2), 360–378.

78  S.-J. HARDY AND N. CHAKRABORTI Savvides, N., Al-Youssef, J., Colin, M., & Garrido, C. (2014). Journeys into inner/outer space: Reflections on the methodological challenges of negotiating insider/outsider status in international educational research. Research in Comparative and International Education, 9(4), 412–425. Spalek, B. (2002). Islam, crime and criminal justice. Cullompton: Willan. Spalek, B. (2005). A critical reflection on researching Black Muslim women’s lives post September 11th. International Journal of Social Research Methodology, 8(5), 405–418. Tomlinson, M., Swartz, L., & Landman, M. (2006). Insiders and outsiders: Levels of collaboration in research partnerships across resource divides. Infant Mental Health Journal, 27(6), 532–543. Turnbull, A. (2000). Collaboration and censorship in the oral history interview. International Journal of Social Research Methodology, 3(1), 15–34. Tyler, K. (2004). Reflexivity, tradition and racism in a former mining town. Ethnic and Racial Studies, 27(2), 290–309. Zempi, I. (2016). Negotiating constructions of insider and outsider status in research with Veiled Muslim women victims of islamophobic hate crime. Sociological Research Online, 21(4), 1–8. Zempi, I., & Chakraborti, N. (2014). Islamophobia, victimisation and the veil. London: Palgrave Macmillan. Zubair, M., Martin, W., & Victor, C. (2012). Embodying gender, age, ethnicity and power in ‘the field’: Reflections on dress and the presentation of the self in research with older Pakistani Muslims. Sociological Research Online, 17(3). http://www.socresonline.org.uk/17/3/21.html.

PART III

Revealing Hidden Problems

CHAPTER 6

Everyday Hate

Abstract  The purpose of this chapter is to enhance understanding of who is affected by hate crime. We showcase the experiences of the more familiar groups of victims, while also demonstrating that there are many groups of ‘Others’ who are targeted on the basis of their identity, and who experience similar forms of victimisation as the more established groups of hate crime victims, and yet occupy a marginal position within research frameworks. It also captures the lived reality of hate crime victimisation which is often characterised by repeated episodes of abuse, harassment and threats, and which frequently culminates in a violent episode. Keywords  Hate crime Repeat victimisation

· Victims · Violence · Harassment ·

The objective of this book is to open the reader’s eyes to the hidden worlds of hate crime victims. We are able to reveal what is often rendered invisible because within our research we adopted a deliberately broad definition of hate crime to capture the experiences of anyone from any background who felt that they had been targeted because of their identity or a particular feature of ‘difference’. By using this more inclusive framework, we were able not only to connect with a diverse range of victim groups, but also to capture the discrete experiences, specificities and intersections in their victimisation. The purpose of this chapter is to © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_6

81

82  S.-J. HARDY AND N. CHAKRABORTI

enhance understanding of who is affected by hate crime. We showcase the experiences of the more familiar groups of victims, while also demonstrating that there are many groups of ‘Others’ who are targeted on the basis of their identity, and who experience similar forms of victimisation as the more established groups of hate crime victims, and yet occupy a marginal position within research frameworks. It is worth highlighting from the outset that for many of the participants involved within our research hate crime is part of a repetitive, persistent and damaging process. For three in five respondents to the survey, their most recent experience of hate crime had taken place within the past year (59 per cent), and for a quarter within the last month (24 per cent). This chapter challenges existing assumptions of what forms hate crime takes by shining a light on the more ‘everyday’ expressions of hostility which plague the lives of victims. As will become evident through this chapter, these experiences are rarely ‘one-off’ incidents and instead form part of a broader continuum of prejudice which is encountered by minority groups on a day-to-day basis. As highlighted within the previous chapter, the research sample comprised a more diverse range of victim groups than is often the case in conventional hate crime studies. This means that within this chapter—and those to follow—we are able to illustrate the similarities and specificities in different groups’ experiences of hate crime, thereby promoting a more nuanced understanding of the nature and scale of the more ‘hidden’ forms of victimisation.

Targets of Hate Crime As highlighted in Chapter 3, many countries across the world have introduced policy guidance and legislative frameworks in order to respond to cases of hate crime. However, the scope of such guidance and lawmaking varies from country to country which creates and reinforces hierarchies of identity where some victims are deemed worthy of inclusion within hate crime frameworks, and others invariably miss out. For instance, of the Office for Security and Cooperation in Europe’s 57 participating states from Europe, Central Asia and America, 23 states record racist or xenophobic hate crime, 19 record sexual orientation or gender identity hate crime; 16 record anti-Muslim hate crime; 15 record antiSemitic hate crime; and 11 record disablist hate crime (ODIHR 2018). Within the context of England and Wales, criminal justice agencies

6  EVERYDAY HATE 

83

are required to monitor five strands of hate crime: namely, hostility directed towards a person’s disability, race, religion, sexual orientation or transgender status. In order to create a more inclusive framework with the capacity to recognise experiences which are often intersectional, under-reported and invisible within the context of recorded hate crime figures, we used a survey instrument which gave explicit acknowledgement to identity characteristics which often slip under the radar such as age, gender and asylum seeker or refugees status, and which provided an ‘Other’ option. This enabled respondents—when asked about which identity or lifestyle characteristic they felt that they had been targeted on the basis of—to self-select multiple variables from a list of 14. The survey findings highlight that the five identity strands monitored by criminal justice agencies within England and Wales featured highly in why victims felt they were targeted. Race was the most commonly cited variable by respondents (33 per cent), and this was followed by 19 per cent of respondents who selected religion; 12 per cent their sexual orientation; 5 per cent learning disability; 5 per cent physical disability; and 2 per cent transgender status. By adopting a more inclusive approach, our research goes beyond the restrictive confines of previous studies to instead highlight the hostility suffered by some of the more unfamiliar victim groups surfaced which are all too often subsumed within broad categories. For instance, within the overall sample who had experienced racist hate crime, substantial numbers of respondents belonged to new and emerging communities, such as Black African (n = 86) (which included significant numbers of Congolese, Somalian and Zimbabwean), as well as to communities with a more established presence such as Indian (n = 61), White British (n = 38) and Chinese (n = 29). The diversity of the groups affected by racist hate crime, and the nuances in their experiences, is often masked because of the adoption of broad and homogenised labels such as ‘BAME’ or ‘Asian’, ‘Black’, ‘White’ and ‘Other’. In relation to religiously motivated hate crime, there were notable patterns which emerged, with Muslim respondents being more likely than others to cite their religion as the reason for being targeted (76 per cent compared with 29 per cent overall), as were Asian respondents (52 per cent compared with 29 per cent overall) and women (33 per cent compared with 25 per cent for men). This finding gives an indication of the ways in which different identity characteristics can intersect with one another to increase the vulnerability of certain groups

84  S.-J. HARDY AND N. CHAKRABORTI

to victimisation, as will be discussed in greater detail in due course. Interestingly, there were victims who had a certain ‘obvious’ identity characteristic but that this feature was not the reason that they were targeted. For example, the survey sample included 77 asylum seekers and refugees and yet only 14 felt that they had been targeted because of their asylum seeker or refugee status, and instead cited their race (68 per cent), their gender (40 per cent), their dress and appearance (22 per cent), and their religion (19 per cent) as being the object of hostility. It could be argued that while discrimination is encountered by asylum seekers and refugees within the context of education, employment and housing, the asylum seeker and refugee ‘status’ is somewhat invisible within the context of everyday life. This may also explain why of the 173 people who identified as being lesbian, gay and bisexual (LGB), only 111 cited this as underpinning their experience of hate crime. Of the 62 respondents who identified as being LGB but who had not experienced homophobic or biphobic hate crime, 31 per cent felt that they had been targeted on the basis of their dress and appearance, 19 per cent their race, 13 per cent their subcultural status and 11 per cent their gender. Again, these characteristics are visible markers of ‘difference’, whereas an individual’s sexual orientation can be concealed as demonstrated by the 68 per cent of the respondents to this survey who admitted to hiding this aspect of their identity in order to avoid victimisation. The findings from the survey challenge assumptions inherent to conventional hate crime theories which see such incidents as linked exclusively to the victim’s membership of a particular group, and instead reinforces the importance of ‘difference’ and of the visibility of specific identity markers. Dress and appearance emerged as the second most commonly cited characteristic (27 per cent) underpinning experiences of hate crime victimisation. Dress and appearance includes visual indicators of identity traits which can contribute to people being more easily identifiable by virtue of a specific ethnicity, religion, sex and sexual orientation, or a learning difficulty or physical disability, as evidenced by the following quotations: I was actually quite shocked, you know. At the time I had just recently started growing my beard. I didn’t actually have a beard before. So straight away that connection, oh my God, had I not grown the beard would he have still said that? Rashid, targeted on the basis of his religion1

6  EVERYDAY HATE 

85

When I looked less convincing it did happen a lot more. And it was just sort of random, verbal abuse, and it would normally be where I was walking through town, walking through the centre of town. Liz, targeted on the basis of her transgender status The year before last, he was indecently assaulted in the street, waiting for a bus to go to the drop-in. And he is very visibly a person with needs, so quite easily targeted, befriended. Carol (a care assistant) describing the experiences of Frank, targeted on the basis of his learning difficulty When you carry a white stick you’re very vulnerable to people who are looking out for susceptible people to take from, to rob, and you do find the white stick’s a giveaway … It does make you feel vulnerable and it’s the same when you live on your own. See if you hear a noise outside you can’t look out the window and see what’s there. Doris, targeted on the basis of her visual impairment I’m not a very feminine girl. Although I’m a straight girl, I have short hair and wear jeans a lot. And people always think that I’m a lesbian. They shout “Oh, you old dyke”. Rachel, targeted on the basis of her perceived sexual orientation I’ve only just started covering my hair, and people do look upon you differently. You do get picked on and called names more. But they don’t realise who’s Pakistani, who’s Indian and who’s Hindu or Muslim or whatever, they just target you on one thing. Like you’re Muslim, and that’s it. Ameera, targeted on the basis of her race and religion

The relevance of this aspect of visible ‘difference’ was also evident in the context of people who wear unconventional modes of dress or have a strikingly alternative appearance—such as those who belong to a subculture including goths, emos and moshers—as well as those who have physical imperfections such as birthmarks and scars. Of note is that dress and appearance was rarely the only variable selected by respondents, but rather one factor alongside various other contributory variables. Of the total sample, 50 per cent selected more than one identity or lifestyle characteristic with respondents indicating that they had been targeted on the basis of their race and their religion; their mental ill-health and their physical or learning disability; or their subcultural status and their dress

86  S.-J. HARDY AND N. CHAKRABORTI

and appearance. The multifaceted nature of hate crime victimisation also emerged within interviews with participants who had encountered hostility on the basis of multiple aspects of their identity or lifestyle. I’ve experienced some abuse for my race, but more for my religion and my sexuality. I’ve been called a terrorist, and just generally saying that we [Muslims] are a problem … I got called Paki in the workplace, I was really, really angry about it. Then I started a new job, no one made any racist comments but they made it about my sexuality. It just replaced race. Mo, targeted on the basis of his religion and sexual orientation It’s very difficult to actually distinguish whether [I’ve been targeted] because I’m disabled or whether it’s because of my ethnicity. Rahul, targeted on the basis of his physical disabilities and race If you actually express your gender identity in ways that don’t conform to perceived societal norms … because everyone has their own idea about what a Trans person ought to look like as well, so if you actually combine that with say a goth identity or a metal identity then you really are going to attract a lot of unwanted attention. Jenny, targeted on the basis of her transgender status and alternative appearance He’s received prejudice for his sexual orientation, he’s been attacked a few times through his disability. When he was at school he went through an awful lot. My son’s had 21 operations in his life to keep him alive … he has been bullied in his workplace [because of this sexuality] for the last four and a half years. Sharon, describing her son’s experiences of being targeted on the basis of his sexual orientation and physical disabilities I’m a young Asian female and I think there’s a lot of negative stuff that can come from that … Racism is something that’s probably going to follow me throughout my life unfortunately. I’m actually in a relationship with a White man and we have experienced racism and passers-by making comments to us, particularly from the Asian community … I’ve also experienced classism because I’m working-class and from the north … I get a lot of abuse because of my accent. Leela, targeted on the basis of her race and socio-economic background

6  EVERYDAY HATE 

87

When respondents were asked about which features of their identity made them concerned about being targeted again in the future, twothirds referred to more than one identity or lifestyle characteristic (64 per cent). This could be indicative of the prejudice and hostility which is woven into the fabric of everyday life and which contributes to minority groups feeling especially conscious about a particular aspect of their identity or ‘difference’. Importantly, those respondents who reported being targeted on the basis of more than one variable were more likely to say that their experience had had a significant impact upon their quality of life. The findings from this body of research challenge conventional ideas regarding which victim groups or identity and lifestyle characteristics should be considered relevant to hate crime policy. 17 per cent of respondents felt that they had been targeted on the basis of their gender; 14 per cent their age; 7 per cent their social status; 6 per cent their mental ill-health; 5 per cent their alternative or subcultural status; 3 per cent their Gypsy/Traveller roots; and 2 per cent their asylum seeker or refugee status. The experiences of some of the more marginalised groups of hate crime victims—such as asylum seekers and refugees, English and European Roma Gypsies, the homeless and those who have mental illhealth or are economically disadvantaged—and of those targeted on the basis of having unfamiliar (in the context of hate crime policy) visual identity markers—such as age, gender or alternative subcultural status— bore all the hallmarks of the more recognised hate crime victim groups. I was morbidly obese for most of my life … I never had any physical abuse; it’s verbal and just blanking you, ignoring you. People pointed at me in shopping centres and people would say mean things, clearly within earshot, so I could hear it. And it makes you feel like you’re not worth the same as other people. Caroline, targeted on the basis of her body shape In 2009 I was stabbed opposite a police station on a main road. It took the police a few days to catch up with the person that did the offence and …. when they caught up with him, he explained to them that he just didn’t like my appearance, the fact that I had pictures tattooed on my body. Dennis, targeted on the basis of his alternative appearance Well it’s very fearful when they attack your home because you’re always that fearful that they’re going to burn it down or something in the night; we’ve had that said to us before. Jo, targeted on the basis of being an English Gypsy

88  S.-J. HARDY AND N. CHAKRABORTI Most recently a gentleman was targeted because he was just old and vulnerable. This young lady used to target this elderly gentleman and would pretend to befriend him, go in, make a cup of tea, sit there. And then when he fell asleep, she would steal finances from him out of his wallet, or steal pieces of items that she could sell. Ian (a housing association officer) describing the experiences of a victim of elder abuse I’ve had verbal comments but I can deal with verbal; it’s when it becomes physical. If you can spit at me what else can you do … they assume just because you sell the Big Issue you’re a smackhead, you live on the streets, you’re a druggie, you’re this and you’re that. Dan, targeted on the basis of being homeless

A small but substantial proportion (13 per cent) of the survey sample ticked ‘Other’ when asked why they felt that they had been targeted. Within this category, many aspects of ‘difference’ were identified as contributing to their victimisation, including: • a distinctive or strong accent; • a lack of religion; • body shape or weight; • education status; • homelessness; • immigration status; and • social status. The same proportion of respondents (13 per cent) stated that they did not know which aspect of their identity of identity or lifestyle had motivated their attacker. While being unable to select a particular identity marker, the notion of vulnerability featured heavily within both the survey and interview data with participants referring to ‘being too nice’, being ‘in a vulnerable situation’ or being ‘an easy target’. These experiences of victimisation would not have surfaced without the adoption of a more inclusive research framework which recognises that victims can be subjected to hostility not exclusively because of their membership of a particular identity group but also because of their perceived vulnerability or difference—and by inference a sense of inferiority—in the eyes of the perpetrator.

6  EVERYDAY HATE 

89

So far this chapter has highlighted the diverse profile of hate crime victims and the intersectional nature of victimisation. It is apparent that a substantial proportion of victims are targeted on the basis of multiple aspects of their identity, and that dress and appearance plays a significant role in increasing a person’s vulnerability to victimisation. It was also revealed that hate crime poses a very real and continuing threat for both the more visible and established groups of hate crime victims, as well as those groups who are often peripheral to empirical and policy frameworks.

Forms of Hate Crime Often it is the extreme manifestations of hate crime which attract media, political and academic attention, while the experiences and the cumulative harms of the more ‘ordinary’, everyday forms of hostility, abuse and harassment are overlooked. Within our research, we wanted to examine all types of victimisation and not just those which would constitute a criminal offence. This broader conceptual lens enabled victims to share— often for the first time—their experiences of being persistently ignored, stared at and spoken to in a belittling and derogatory manner: It was the manner in which she spoke and how she looked at me, I felt so degraded. She looked up to me with this certain look in her eyes of pure dislike … that was the most racist experience I have ever had and she never said a racist word. Shaf, targeted on the basis of his race I have only recently lost weight and I notice that people are treating me different now. And especially with losing the weight, the more I lose the more respectful people treat me, and it makes me really angry, because it just controls your life if you are always afraid of going out. Caroline, targeted on the basis of her body shape They [young people] knock on the windows and on the door…they’re shouting or talk swear words … and she [neighbour] all the time puts rubbish in my garden. Fatima, targeted on the basis of her refugee status

90  S.-J. HARDY AND N. CHAKRABORTI If I go with anyone in my wheelchair they always speak to the person pushing it. Even though I find it difficult standing up, I end up standing up and saying “I’m here, talk to me”. James, targeted on the basis of his physical disabilities We’re not just being targeted verbally and being given the cold shoulder, dirty looks and fingers. We’re also getting something that has moved slightly further and is a bit more dangerous. I’ve had people when I walk past them spit at me and spit at the floor. You can see that it’s a show of disgust. Ameera, targeted on the basis of her race and religion

To an outsider, these incidents in isolation may seem relatively trivial but for many of the victims within this research these experiences were a routine feature of their daily lives and had a profound cumulative set of impacts upon their emotional well-being and physical health, as discussed more fully in Chapter 8. Within the survey respondents were asked what form their victimisation took and how often they were targeted. Verbal abuse emerged as the form of hate crime most likely to have been experienced by respondents (87 per cent had been targeted in this way) and it was also the form that victims suffered the most frequently (16 per cent selected regularly and 32 per cent occasionally). This revelation should not be underappreciated; nearly half of the total sample experience being called abusive, derogatory and upsetting names related to their identity or lifestyle on a frequent basis. Within the research sample, there were some notable variations in terms of the types and frequencies of hate crime experienced by different groups of hate crime victims. For instance, the survey findings revealed that those targeted on the basis of their subcultural status (51 per cent), their physical disability (37 per cent), their transgender status (36 per cent) and their mental ill-health (31 per cent) encounter verbal abuse more regularly than other groups of hate crime victims. The regularity of these ‘everyday’ experiences was also evident from our own interactions with participants during fieldwork. Prior to one interview, for instance, a transgender participant was verbally abused and harassed multiple times during an eight-minute walk to meet the researcher. We also witnessed hostile looks, disparaging remarks and intimidatory behaviour directed towards participants on countless occasions. While quantitative data provides a snapshot of the scale of hostility, abuse and harassment directed towards certain groups of hate crime

6  EVERYDAY HATE 

91

victims, it often fails to capture the lived reality of experiencing an inherently personal and damaging form of victimisation. In-depth interviews, however, provide a more realistic and detailed account of hate crime victimisation, illustrating the devastating regularity in which many victims are exposed to hate crime. I more or less expect something to happen every day. I’m getting quite used to it, and you know I’m so used to the verbal abuse. Cath, targeted on the basis of her transgender status In terms of verbal abuse, loads and loads. Like F’ing old dyke … you got very used to it. Nicola, targeted on the basis of her sexual orientation It’s mainly racist like being called ‘You black bastard’ or, ‘You paki bastard’, and things like that. Rish, targeted on the basis of his race [They shout] Are you a man, are you a woman? What are you? Grow a pair; tranny; gender bender; queer - I’m not going to say the actual word - but you mother F’ing queer. Paula, targeted on the basis of her transgender status I have had a number of experiences where I just get yelled abuse being directed at me (fat bitch variety) as I’m walking, cycling or out in evenings. This abuse is usually started having had no interactions with the person until that moment. Sue, targeted on the basis of her body shape

Our research findings suggest that the second most commonly experienced form of hate crime is harassment which includes threatening and intimidatory behaviour in both online and offline settings. Seven in ten respondents reported that they had been harassed in offline environments (70 per cent), and of this sample 34 per cent experienced it repeatedly. Nearly every respondent who had been targeted on the basis of their learning difficulty (93 per cent), their mental ill-health (93 per cent) and their physical disability (85 per cent) had been on the receiving end of threatening and intimidatory behaviour, indicating variations in the types of victimisation experienced by different groups of hate crime victims. In relation to online harassment, just over a quarter of survey

92  S.-J. HARDY AND N. CHAKRABORTI

respondents admitted to being targeted in this way (27 per cent), with those who had encountered hostility because of their subcultural status (40 per cent), their gender (38 per cent), their mental ill-health (37 per cent) and their transgender status (36 per cent) being most likely to experience this mode of victimisation. The survey also captured the more extreme manifestations of hostility, revealing that 44 per cent of respondents had experienced property crime—which includes instances of burglary, theft, or damage to a victim’s house or car—with 11 per cent experiencing this frequently. In terms of this form of hate crime, those targeted on the basis of being Hindu (62 per cent), or because of their Gypsy/Traveller status (61 per cent) or asylum seeker or refugee status (52 per cent) reported higher levels of property crime. To some extent, the patterns in this type of victimisation can be explained by the symbolic value of property in relation to these groups of hate crime victims. For instance, the hostility expressed towards Gypsies and Travellers is underpinned by an intolerance of their alternative way of life, and thus a caravan is a visible signifier of a perceived ‘difference’. Property for these groups, whether it be a caravan or a place of worship, is interwoven with a particular identity or lifestyle, and therefore it becomes a viable outlet for hostility especially when considered alongside the finding that survey respondents were least likely to know who was responsible for committing property crime compared to other forms of victimisation. It is likely that the reason why asylum seekers and refugees are more vulnerable to property crime differs from that explaining the rates of victimisation affecting Gypsies and Travellers and those identifying as Hindu. Those seeking asylum as well as those who have received their right to remain are more likely to reside in insecure housing located in socially and economically deprived areas, which is already fertile environment for anti-social behaviour and criminality. The role that situational factors play in increasing the vulnerability of certain victim groups is explored in greater detail within the next chapter. Of the total sample of respondents to the survey, 32 per cent had been physically attacked—sometimes with weapons—on at least one occasion, with 8 per cent being targeted in this way repeatedly. These violent expressions of hostility appeared to affect those who had been targeted on the basis of their subcultural status (70 per cent), their transgender status (59 per cent), their learning difficulty (55 per cent) or their mental illhealth (54 per cent) more acutely. It is worth highlighting that the sample

6  EVERYDAY HATE 

93

of respondents, who cited their subcultural status as underpinning their experiences of hate crime, were predominantly male and between the ages of 18 and 34 years old which is a subsection of society who are more likely to experience violent crime generally. In terms of sexual assault and rape, 10 per cent of the overall survey sample reported experiencing this form of victimisation which they felt was motivated by hostility towards a particular feature of their identity or lifestyle. Discrete patterns emerged as with the other forms of victimisation, with those targeted on the basis of their gender (36 per cent), their mental ill-health (33 per cent) and their sexual orientation (25 per cent) experiencing this form of victimisation to a greater extent when compared to other groups of hate crime victims. While this data provides a worrying picture of the frequency with which victims are being subjected to violent expressions of hostility, the interviews revealed the sheer diversity and brutality of these experiences. I’ve been spat on, kicked, punched, thrown up against a wall. Keith, targeted on the basis of his learning difficulties I was smashed in the face and completely disorientated and blinded, and was unable to make any response. I staggered out of the woods, fortunately without this guy getting hold of me, got onto a public road. And as I got on the footpath, this guy caught up with me and flung me in front of an oncoming car. Simon, targeted on the basis of his sexual orientation One of them kicked the door, with both feet and it slammed against my hand … they had spat all over my back. And I didn’t even know they’d done it. How disgusting is that? … They spit on the windows or they throw eggs. Damien, targeted on the basis of his race I was walking through an estate on a sunny day and I simply looked at them. They didn’t like the look of me, so they ran up to me and they chased me and I fell and I couldn’t get up because I’d broken my knee, and they kicked me in the face, and they kept doing that repeatedly. Will, targeted on the basis of his mental ill-health If you pull up on the side of the road you have all sorts chucked at trailers, sometimes bricks … [One time] a car had driven by, it had slowed down, somebody had thrown something the size of an orange, lobbed it at her, it had missed her, whizzed by her head, hit the ground and it was an explosive device. Rosie, targeted on the basis of being an English Gypsy

94  S.-J. HARDY AND N. CHAKRABORTI These people with bikes and two golf clubs came and we just ran … They ended up getting my friend in the face with a golf club … they were just tearing at his clothes, beating him up … That was a very frightening experience. We were shit scared of going there again. Callum, targeted on the basis of his alternative appearance One day I was walking with my friend and this person came in and kicked my friend and started swearing. My friend just asked him “Why are you swearing at me?” And he threw his bike and came for a fight. Adil, targeted on the basis of his asylum seeker status

These quotations demonstrate the complexity surrounding the process of hate crime victimisation: often a victim’s interaction with a perpetrator begins with a verbal exchange which escalates into property damage and/or physical assault. As evidenced so far within this chapter and in official data (Home Office 2018), those affected by hate crime are more likely than other victim groups to be repeat victims, with the underpinning hostility manifesting through different forms of victimisation. The following extract attempts to present a more comprehensive and realistic account of the ongoing abuse, harassment and violence suffered by many hate crime victims, and yet still falls short of capturing the victim’s experience in its entirety. We’d been on holiday and the house had been burgled… there was nothing left. Later the same year, my house was burgled again … A group of youths stopped my son because they had had a fight with a member of the Kurdish community and they stopped anybody from that community. They attack my son and threaten him saying ‘we will kill you at any point’ and ‘we know where you live’. They bullied him at school. My son was scared to tell me … I came outside and my windscreen had been smashed completely … I called the insurance, the insurance came and sorted out the windscreen … the next evening they came back and they broke it again … We had an appointment at the dentist and my son went to buy a fizzy drink and saw one of the young people from the group and his mum, both of them attacked him lots and his face was completely blood and bruises … I went outside the dentist and there were so many people with the mum and son and I had an attack on my face near my eye but I don’t know what tool they used. They attacked my wife and cut her lip and were using all nasty words … They drove into my wife, 20 days she’s been with the support of the cast … The car was completely damaged and it had our baby in who was only weeks old. Beyani, targeted on the basis of his race

6  EVERYDAY HATE 

95

The regularity in which victims experience hate crime was especially apparent within the sample who reported being targeted on the basis of their learning difficulty and/or physical disability and/or mental illhealth (Vedeler et al. 2019; Beadle-Brown et al. 2014). Overall, disabled respondents had been subjected to some of the highest levels of verbal abuse, harassment, physical assault and sexual violence when compared to other groups of hate crime victims. For example, 92 per cent of disabled respondents had experienced intimidating and threatening behaviour—and half of these respondents had encountered it frequently—while 50 per cent had encountered violent crime and 22 per cent sexual violence. The persistent and ongoing nature of the hostility, abuse and harassment experienced by this victim group was especially vivid during interviews, as illustrated by this extract from the research field diary: I met a group today which was made up of people with multiple physical and learning disabilities. I have conducted many group based interviews with disabled people and it never fails to shock me how routinely they are targeted. Today alone I heard about the countless names that disabled people are called on a daily basis, including ‘spag’, ‘retard’ and ‘paedo’; I heard about multiple incidents in which individuals within the group were followed and intimidated as they walked down the street or through the city centre; and I heard about numerous cases in which disabled people were exploited, such as one male who had allowed local youths to frequent his flat only to be robbed and physically assaulted. The scale of the abuse suffered by disabled people is truly hideous.

Conclusion This chapter was designed to capture the lived reality of hate crime victimisation which is often characterised by repeated episodes of abuse, harassment and threats, and which frequently culminates in a violent episode. Often research findings depict experiences of victimisation in a retrospective manner, but it is important to recognise that within the context of hate crime, future victimisation is a real and significant threat. We found that amongst the total sample concern about being a victim of hate crime again in the future was high, though there was more apprehension about certain types of crime than others. While this may come as a surprise to many readers, respondents were most likely to be fearful

96  S.-J. HARDY AND N. CHAKRABORTI

about being targeted through the more ‘everyday’ forms of victimisation in the future. Two-thirds reported that they were very or fairly concerned about being harassed and suffering verbal abuse (67 per cent and 64 per cent respectively), and when asked about how much the fear of hate crime had affected their quality of life 91 per cent of respondents observed that they had been affected in some way. This chapter has also drawn attention to the diverse profile of hate crime victims, highlighting the experiences of those groups and communities who occupy a marginalised and to some extent, ‘invisible’ presence within policy and academic frameworks, and within society more broadly. The adoption of a more inclusive construction of ‘difference’ within this research permitted a richer insight into the specificities of experiences, the intersections of identity and lifestyle, and the everyday realities associated with hate crime victimisation. The following chapter seeks to further illustrate the complex processes involved in hate crime victimisation by showcasing the broader structural, economic and situational factors which can exacerbate a victim’s vulnerability to hate crime victimisation.

Note 1. The names of research participants have been changed for the purposes of this book in order to preserve their anonymity.

References Beadle-Brown, J., Richardson, L., Guest, C., Malovic, A., Bradshaw, J., & Himmerich, J. (2014). Living in fear: Better outcomes for people with learning disabilities and autism. Canterbury: Tizard Centre. Home Office. (2018). Hate crime, England and Wales, 2017/18. London: Home Office. ODIHR (Office for Democratic Institutions and Human Rights). (2018). Hate crime reporting. http://hatecrime.osce.org. Vedeler, J. S., Olsen, T., & Eriksen, J. (2019). Hate speech harms: A social justice discussion of disabled Norwegians’ experiences. Disability & Society, 34(3), 368–383.

CHAPTER 7

Everyday Contexts

Abstract  This chapter begins by uncovering the profile of hate crime perpetrators in relation to the characteristics of age, gender and ethnicity, and by teasing out the varying patterns which emerge within the context of different forms and strands of hate crime. It also provides new insights into the relationship between the victim and the perpetrator(s) and by doing so, it challenges the prevailing assumption that hate crimes are committed by people who are qualitatively different from the ‘rest of us’: hate-fuelled strangers. This chapter also offers a contextual analysis of the environmental, situational and socio-economic factors which are proven to be especially relevant to the process of hate crime perpetration. Keywords  Hate crime Identity

· Victims · Perpetrators · Vulnerability ·

Within this chapter, the analytical lens shifts from offering a more holistic depiction of the nature and scale of hate crime victimisation which for too long has been hidden from scholars, professionals and policy-makers, to revealing who commits hate crime and in what contexts. While some degree of caution should be exercised as the data is based upon the victim’s perception of the perpetrator in terms of their demographics and their motives, these findings are significant because of the dearth of administrative data and empirical evidence on those who perpetrate these types of offences. This chapter begins by uncovering the profile of hate © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_7

97

98  S.-J. HARDY AND N. CHAKRABORTI

crime perpetrators in relation to the characteristics of age, gender and ethnicity, and by teasing out the varying patterns which emerge within the context of different forms and strands of hate crime. It also provides new insights into the relationship between the victim and the perpetrator(s) and by doing so, it challenges the prevailing assumption that hate crimes are committed by people who are qualitatively different from the ‘rest of us’: hate-fuelled strangers. This chapter also offers a contextual analysis of the environmental, situational and socio-economic factors which are proven to be especially relevant to the process of hate crime perpetration. It illustrates that there are a wide range of contextual and lifestyle variables which contribute to certain groups of victims being especially vulnerable to hate crime victimisation. These findings have implications for how hate crime is conceptualised, further reinforcing the importance of developing a more inclusive and nuanced framework which better captures the complex processes involved in hate crime victimisation

Perpetrators of Hate Crime Hate crime perpetration is a relatively elusive area of scholarship within the broader field of hate studies. Equally, administrative data on the demographic profile and motives of hate crime perpetrators is sparse. Through this research, we had the opportunity not only to transform understanding of hate crime victimisation but also to gain new knowledge about hate crime perpetrators from the perspective of a broad and diverse sample of victims. As part of the survey, a series of questions were posed pertaining to the profile of the perpetrator who was deemed responsible for committing the respondent’s most recent experience of hate crime. The findings illustrate that group dynamics play an important role within the commission of hate crime, with just one in five respondents being targeted by a single offender (22 per cent). In nearly half (45 per cent) of the incidents reported within the survey, between two and five offenders were involved and in 6 per cent of cases in excess of five perpetrators took part in targeting the respondent. Some key differences in relation to the type of victimisation and the number of offenders involved emerged within the data, with respondents whose most recent experience had involved sexual violence being more likely than others to say it had been perpetrated by one offender; those whose most recent experience involved violence were more likely to say there had been five

7  EVERYDAY CONTEXTS 

99

or more offenders; and those whose most recent experience involved property crime were more likely than others to say they did not know how many offenders had been involved. There were also a number of nuances which surfaced in relation to where the offence had taken place and how many perpetrators had been involved. For instance, respondents who had experienced their most recent hate crime in or near their place of work were more likely to say that it had been carried out by one offender, whereas those whose most recent hate crime involved online harassment were more likely to report that five or more offenders had taken part. In terms of the demographic profile of hate crime perpetrators, the survey findings illustrate that the majority of incidents are committed by males (68 per cent), which aligns with administrative data published by criminal justice agencies around the world. However, deeper insights into the gendered dimensions of hate crime perpetration are revealed when an examination of different types and strands of hate crime is undertaken. For instance, males are significantly more likely to be involved in incidents motivated by the victim’s sexual orientation (86 per cent), transgender identity (77 per cent) or gender (76 per cent). Further contextual data provided during the course of the interviews indicated that participants within these groups of victims felt that they had been targeted not only on the basis of their gender or sexual orientation, but also because of the way in which they had performed that element of their identity. For some victims who have a particular sexual orientation or gender identity, their perceived ‘difference’ is more readily identifiable and is interpreted by some—particularly males—as being an affront to societal ideals on gender performance, evoking a hypermasculine response which is designed to signal an inacceptance of behaviour which transgresses heteronormative practices. The role which women play within the commission of hate crime is often overlooked as on the surface their involvement is relatively minor. However, the survey findings demonstrate that female perpetrators play a more prominent role in specific strands of hate crime, namely those involving Gypsies and Travellers (45 per cent compared with 26 per cent), asylum seekers and refugees (45 per cent), and disabled people (38 per cent). Again, contextual factors appear to influence this pattern with the survey data indicating that females are more likely to offend in social spaces which are familiar to both the victim and the perpetrator. For instance, respondents whose most recent hate crime had occurred at school, college or university were more likely to state that a female

100  S.-J. HARDY AND N. CHAKRABORTI

offender had been involved (53 per cent compared with 26 per cent). However, those respondents who encountered hostility in more open and public environments were more likely to say that a male offender had been responsible (85 per cent compared with 68 per cent overall), as were those who had been targeted within the night-time economy (88 per cent compared with 68 per cent overall). These findings point to a much more complex picture then is commonly assumed when it comes to hate crime perpetration, with a number of situational factors influencing the profile of the perpetrator, including the victim group, the type of victimisation, the location and the relationship between the victim and the perpetrator. There were also some interesting findings to emerge in relation to the age of hate crime perpetrators. Just over a third of the most recent hate crimes reported through the survey had involved an offender aged between 20 and 30 years old (37 per cent), and a further third had involved a teenage offender (33 per cent), which again corresponds with administrative data published by criminal justice agencies. There were, however, subtle differences when it came to strands of hate crime and the age of perpetrators, with younger offenders—those between the ages of 13 to 30 years old—appearing to play a more prominent role in targeting people based upon their sexual orientation, their subcultural status and their transgender status. Additionally, respondents were also more likely to have experienced a hate crime involving an offender who was of a similar age to themselves. For instance, respondents aged between 25 and 34 were more likely than respondents overall to report that their most recent experience of hate crime had involved offender(s) aged between 20 and 30. It’s normally like four or five white, young, early twenties, maybe late teens, men. And they will be the ones who will usually start shouting stuff. Paula, targeted on the basis his transgender status All of them were underage, they were seventeen or eighteen and all of them were totally drunk. You can smell it when they talk to you. Yehuda, targeted on the basis of his religion I think people under a certain age – teenagers or early 20s – they do it because they’re with two or three people and I’m on my own. It doesn’t happen if I’ve got somebody with me. Paul, targeted on the basis of his visual impairment

7  EVERYDAY CONTEXTS 

101

The people that pester me the most are the people in the category of 20 to 30 years old, there’s usually about two or three of them, and they’re usually male. If there’s just a male it is very rare that they will actually do anything violent. And if there’s a female in the company, then I’m usually quite safe. Cath, targeted on the basis of her transgender status

In relation to the ethnic profile of perpetrators, there has been a prevailing assumption that hate crimes are typically committed by people who belong to majority or well-established groups, which in domestic terms within a UK context would be people who identify as White British. However, the findings from this research reveal that both majority and minority groups, and both established and new and emerging communities can express hostile views and commit acts of hate crime. What I hear now is about Eastern Europeans, what I hear now is about Romanians, Polish and it’s the same rhetoric what was being said in the 70s: “They’re coming for our jobs, our houses, our schools, our NHS system” and even sometimes “Watch out for our young girls”. Tony, targeted on the basis of his race Even in supermarkets Somalians can be particularly aggressive, I think it all stems from their background and where they’ve come from. That’s the problem, there’s a large community in this area that causes issues. We have a similar problem in our area. There’s a very large Polish community, and they’re very aggressive. And that’s hard, because we’ve lived here since 1966 and we never had a problem. Now the communities that have come in seem to put barriers up. Rajani, targeted on the basis of her religion She said “I know, I’m Sikh, and we know about Islam, we know about your religion, pathetic. I can’t believe in this day and age you’re allowed in this 21st century”. She just went on and on. Oh another bit I missed out, she said “You’re terrorists, you’re terrorists!” Nasar, targeted on the basis of his religion I’d find I’d only be walking down the street and I’d have Black people verbally abuse me, because I’m half-caste. It’s really weird … White people, Black people they’d try and pick on you cos, you know, you’re mixed race. Ezra, targeted on the basis of his learning disabilities and race

102  S.-J. HARDY AND N. CHAKRABORTI

The survey data illustrate that hate crime perpetrators come from a range of different ethnic backgrounds, with three fifths of respondents reporting that their last experience of hate crime had involved a White offender (61 per cent), one in six an Asian offender (16 per cent), and one in eight a Black offender (12 per cent). The diversity among perpetrators becomes more vivid when examining specific strands of hate crime victimisation. Taking gender-based hate crime as an example, just over half of respondents described the offender as White British (55 per cent), over a quarter stated that an Asian British, Indian, Pakistani or Bangladeshi (27 per cent) individual had been involved in their victimisation, and 17 per cent reported that a Black British, Caribbean or African perpetrator was responsible. It is also a common misconception that hate crimes are committed exclusively by strangers, but it is likely that this supposition is rooted in an unwillingness to accept that someone who has a pre-existing relationship with the victim—whether this be in the capacity of a carer, a friend, a neighbour or a work colleague—could engage in such hurtful behaviour. Crucially, only half (54 per cent) of survey respondents had experienced their most recent hate crime at the hands of a stranger. Examining the responses from those who knew their offender(s), respondents cited acquaintances (9 per cent), neighbours (6 per cent), friends (5 per cent), work colleagues (5 per cent), family members (3 per cent) and carers (1 per cent) as being responsible for their victimisation. Of note is that those targeted on the basis of their mental ill-health (48 per cent compared to 25 per cent) and physical disability (36 per cent) were more likely to know the offender, and that those targeted on the basis of their subcultural status (79 per cent) and sexual orientation (64 per cent) were least likely to know the offender. This finding is likely to be connected to the location in which an incident takes place as there are subtle variations in trigger environments for different strands. For instance, those who were targeted on this basis of their mental ill-health and physical disability were more likely to cite the location of their victimisation as being outside or near their home, and therefore it is more likely that the perpetrator(s) will be known to the victim in some capacity, such as a neighbour or a local acquaintance. Additionally, disabled people were more likely than other groups of victims to report ongoing harassment, and those respondents who suffer this type of victimisation are more likely to say that they had known the offender.

7  EVERYDAY CONTEXTS 

103

I think they targeted me because they knew I was the soft one, the nervous one. Deborah, targeted on the basis of her age They target me because I’m Asian and by myself. Amandeep, targeted on the basis of her race and religion I think they targeted me because they could see that I’m not well and vulnerable or that I might be gay and an easy target. Joel, targeted on the basis of his sexual orientation and mental ill-health It’s when people aren’t confident about being Black or about being a Muslim woman or about being trans. People spot the vulnerable ones and that’s who they go for. Caroline, targeted on the basis of her transgender status

Again, this data reveals the gendered patterns of perpetration which are often underappreciated or to some extent overlooked altogether. Illustrative of the intersections between gender, location and the strand of hate crime, is the finding that males are more likely to commit acts which are violent and sexual in nature, and to perpetrate homophobic hate crime which is more likely to take place in openly public settings such as in a street, a city centre or a bar or nightclub, and these are social spaces in which the offender is more likely to be a stranger. In comparison, female offenders have a greater tendency to engage in ongoing patterns of abuse and harassment, to commit hate crimes in familiar locations, to target those who are in especially vulnerable situations such as those with mental ill-health, learning difficulties and physical disabilities, and to be known to the victim in some capacity. These contextual variables offer an important insight into the causal factors and trigger environments for different profiles of perpetrators and for different strands of hate crime. In short, this research illustrates that there is ‘no one size fits all’ type of hate crime perpetrator. In comparison to the diversity that exists with respect to the profile of perpetrators, when victims were asked for their perceptions of what motivated perpetrators to commit hate crimes, the responses were strikingly similar. Participants consistently referred to a sense of unfamiliarity and intolerance towards ‘difference’ as being key motivating factors underpinning their experiences of abuse, harassment

104  S.-J. HARDY AND N. CHAKRABORTI

and violence. Perceived vulnerability was also identified by participants as being central to why they thought the perpetrator had targeted them specifically. While it is difficult to unpick the specific variables which motivate this form of crime as this research is based upon the perceptions and experiences of victims, the following section considers the environmental, situational and socio-economic factors which were proven to be especially relevant to the process of victimisation which in turn offers insight into the contextual variables that facilitate hate crime perpetration.

Drivers of Hate Crime This section examines the interplay between identity and lifestyle characteristics, and a broad range of situational variables which can exacerbate the vulnerability of certain groups, and which can motivate ordinary people to commit hate crimes. One of the major variables which is considered to increase a person’s susceptibility to victimisation and which was discussed within the previous chapter is the presence of obvious markers of ‘difference’ such as walking aids, facial hair, religious dress and displaying affection towards a same-sex partner. It was also noted within the previous chapter that a significant proportion of participants felt that in the eyes of the perpetrator they were an ‘easy target’, ‘soft’, ‘too nice’ and ‘weak’ and that this perceived vulnerability put them at greater risk of repeated victimisation. Of note, however, is that just four in ten victims had been targeted while they were on their own—which is often considered to heighten a person’s vulnerability—with those targeted on the basis of their transgender status (64 per cent), their mental ill-health (51 per cent) and their physical disability (51 per cent) at greatest risk of being targeted while alone. In addition to these variables, socio-economic status also emerged as a recurring theme in relation to increasing a victim’s exposure to hostility. Of the total survey sample, a third were in employment, including self-employment (33 per cent), and another third were in full or parttime education (34 per cent). Smaller proportions were unemployed and actively seeking work (7 per cent), unable to work due to sickness or a disability (7 per cent), retired (6 per cent), or were looking after children or their home (4 per cent). Of note is that a substantial number of respondents reported receiving very low or no yearly income. A quarter said that they did not receive any income (23 per cent) and one-fifth had

7  EVERYDAY CONTEXTS 

105

an income of £15,000 or less annually (19 per cent). One in five stated that they received benefits (19 per cent). There are numerous ways in which a victim’s socio-economic status can increase their risk of hate crime victimisation, chiefly because those who are most economically marginalised are more likely to have insecure accommodation or to live in social housing, to work antisocial hours, and to rely on public transport which emerged as a particularly risk-laden environment for certain groups of hate crime victims (see below for further discussion). Furthermore, there were particular groups of hate crime victims who felt that the recent government focus on welfare reform had contributed to a rise in their experiences of abuse, harassment and violence. Specifically, those with physical disabilities, people with mental illhealth and the homeless commented on the mounting hostility directed towards them. One car drove by me shouting ‘What are you fucking like, you can see really!’ That same evening, a couple of guys shouted, “If you’re blind, we all are!” … It’s definitely got worse. Another time I walked out of here, two guys, two people walked past me, one turned around and said, “You fucking skank, you can see really”, saying that I was doing it for benefit reasons. Lee, targeted on the basis of his visual impairment Because they think we’re scum. They say “You scumbags, you’re begging!” We weren’t begging. We were just sitting there in the doorway waiting. And they walked up and said “Do you want some money?” and we said “No”, and he went “Get a fucking job!” and one of them just walked up to me and just kicked me. My brother pushed me out of the way and got kicked in the face. Carl, targeted on the basis of being homeless Before I had my accident I had been working for 25-30 years, so I contributed to the benefit system. Now I’m on benefits and I’m being targeted because of it. We feel as though there’s a bit focus on it at the minute that is causing everybody to target people on benefits. David, targeted on the basis of his physical disabilities People say I’m a waste of space because I’m bone idle, lazy. They say I’m screwing the system for everything I can get … What people say plays on my mind a lot. If you’ve got depression you sort of go over and over what people say. Ken, targeted on the basis of his mental ill-health

106  S.-J. HARDY AND N. CHAKRABORTI

The broader political climate was also revealed to exacerbate intolerance of and tensions towards particular minority ethnic communities and new and emerging communities specifically. In order to substantiate this assertion, it is necessary to focus exclusively on the survey findings of the most recent study (Study Four) which was commissioned following the EU Referendum in 2016. Survey respondents (n = 360) were asked to consider whether they had become more concerned about hate crime following the EU referendum and 21 per cent of victims reported feeling more concerned, with 29 per cent feeling much more concerned. This perception appeared to be based upon their observations of prejudiced views becoming more overt, which they felt politicians and the media had legitimised. Brexit changed a lot in a bad way. As a polish national I feel scared about the future of myself and my daughters as it’s become clear that we are no longer welcome in UK … After Brexit people feel allowed to show their aggression and that people like me are not welcome. Lena, targeted on the basis of her race On a day-to-day basis we have comments passed, people giving us weird looks, saying something. We’ve always had that, but since high profile incidents such as Woolwich it does spike up. Taliha, targeted on the basis of her religion

A significant number of participants suggested that recent global events, including ‘Brexit’ and ‘Trump’, had not only ‘validated’, ‘reinforced’ and ‘fuelled’ hostile views but had also ‘emboldened’ people by permitting them think that it is acceptable to express such views in both online and offline settings. Participants, who identified as belonging to a minority ethnic and/or faith community spoke of feeling ‘concerned’, ‘fearful’ and ‘scared’ about their future within the UK. Experiences of hate crime within a work setting further point to the contributory role that environmental factors plan in increasing the risk of victimisation. The interviews in particular highlighted that those who work within the night-time economy frequently experience fraught and difficult situations when dealing with drunk and hostile customers, with many taxi drivers, restaurant workers, takeaway owners and sex workers feeling especially prone to being harassed while carrying out their jobs.

7  EVERYDAY CONTEXTS 

107

When I was growing up, I did experience a lot of racism. But a lot of it when I think back was down to ignorance of culture, because we were different. As an adult I went into the restaurant trade and this is when I had my worst racial experiences. Nilesh, targeted on the basis of his race I’ve been here 18 years so I’m used to it. It’s one of those where… it’s not nice. I don’t think anybody, for any reason, getting shouted at down the phone, or sworn at, or threatened is nice but unfortunately it’s part of the trade. Randhir, targeted on the basis of his race Like I said before I worked at a Chinese takeaway and had eggs thrown. You always have half-drunk people come in, and then they start, you know, throwing a tantrum. It could be broken glasses, broken windows or insults. All sorts. Li Wei and Chen targeted on the basis of their race

The environments which can be especially treacherous for certain groups of hate crime victims to traverse are those which are often hidden within an everyday social setting. For instance, the survey data illustrates that hate crime victims most commonly encounter hate crime in a public street (32 per cent) but for some this risk-laden street might be the walk to work, to school or to the local supermarket. Furthermore, while the night-time economy in general is a particularly dangerous environment for those with visible markers of difference, the toilets within bars, clubs and restaurants emerged as being an especially risky social space. At night-time when everybody’s drinking, the easier it is to say comments they wouldn’t normally say. I’d say definitely out at night in clubs and stuff when people are drinking is the worst time for me. Ollie, targeted on the basis of his sexual orientation There are certain times that we go out now. Like Saturday evening, we don’t go out, no, it’s dangerous for my kids, I tell them not to go outside at night. Jaali, targeted on the basis of his race The most dangerous part of my daily life is using washrooms in public places like restaurants and bars and pubs and hotels, clubs and all these things. That is really, really dangerous for me. Liz, targeted on the basis of her transgender status

108  S.-J. HARDY AND N. CHAKRABORTI

Finally, a sense of increased vulnerability on public transport was referred to by a substantial proportion of the victims who took part within this research. Public transport brings together groups of people in a confined space, and for certain groups of victims, it was found to be an environment which heightened their risk of targeted hostility. Respondents targeted because of their learning disabilities were more likely to have suffered their most recent hate crime on public transport compared to any other victim group. Throwing papers and swearing at us on the bus … they’re calling names and hitting people. Anne, targeted on the basis of her learning difficulties If a bus is full of kids, the kids start shouting and then, on transport, when they start shouting at you, when you’ve done nothing wrong and then they’re shouting at you… Jason, targeted on the basis of his learning difficulties I use regular buses in the morning and some bus drivers are horrible to me. Wendy, targeted on the basis of her physical and learning disabilities

Being targeted on public transport is not exclusive to people with physical or learning disabilities and we heard many accounts from people who had been singled out because of their appearance or mode of dress, including veiled women, trans people, asylum seekers and people who belong to an alternative subcultural. However, we found that those with physical and/or learning disabilities and/or mental ill-health who were able to travel independently experienced regular verbal abuse and harassment. It is worth reiterating that social and economic factors are highly relevant within this context, with evidence suggesting that people with less material wealth are at greater risk of being targeted on public transport. This is attributable to the fact that people who can afford to buy a car or who are able to pay for taxis can avoid the types of situation that place them at risk of targeted hostility, whereas those on a low income often have no choice but to take buses and then ‘run the gauntlet’ of verbal abuse and harassment that comes their way.

7  EVERYDAY CONTEXTS 

109

Conclusion This research has offered new insights into the profile of hate crime perpetrators, illustrating the importance of conducting deeper analysis which takes different types and strands of victimisation into account. This chapter has highlighted that both majority and minority groups, and both established and new and emerging communities can perpetrate acts of hate crime. It also illustrated that gendered patterns of victimisation are evident, with female perpetrators playing a more prominent role within the context of ongoing patterns of abuse and harassment, and within familiar environments such as workplaces and educational spaces. In comparison, male perpetrators were almost exclusively responsible for committing violent expressions of hate crimes and were more likely to perpetrate incidents which take place within a street-based situation or the night-time economy. This chapter has demonstrated that the victim-offender relationship is often more complicated than is commonly thought, with victims recounting experiences of being targeted by their neighbour, by a work colleague, by a support worker, by a fellow pupil or by a ‘friend’. Worryingly, this evidence reveals that for the most part, perpetrators are ordinary people who commit hate crime within the context of their everyday life. In addition to generating new knowledge on who commits hate crime, the chapter also exposed the reader to the ‘hidden’ environments where incidents are most likely to take place. By doing so, it highlighted the complex interplay between identity and lifestyle characteristics, and a broad range of economic and situational variables which can increase the likelihood of exposure to hostility, abuse and harassment. The chapter highlighted that victims are often socially and economically disadvantaged, and that environments in which victims navigate as part of their everyday life, including public transport, bars, takeaways and restaurants, toilets and supermarkets are ones where they face particularly high levels of risk in relation to hate crime. Broader political discourse and trigger events not only exacerbate the fear and vulnerability felt by particular minority groups, but also lead to an increase in the hostility directed towards them.

CHAPTER 8

Invisible Harms

Abstract  This chapter presents an unflinchingly vivid account of the ways in which victims’ lives are blighted by bruises, broken bones and scars, by the emergence or exacerbation of debilitating health conditions, and by anxiety, fear and depression. It also illustrates that experiences of hate crime can often be commonplace that the impact has, to some extent, been ‘neutralised’ or normalised as part of an unwanted but inevitable feature of being ‘different’. This chapter will also show that the damage caused by hate crime extends well beyond the primary victim, often evoking fear, alarm and distress among family members, and in some contexts, within wider communities. Keywords  Hate crime Coping strategies

· Victims · Physical harms · Emotional harms ·

Within the context of hate crime, victims are being targeted typically because of who they are, because of the community to which they are perceived to belong or because of the way in which they live their life. For these reasons, this form of victimisation can feel inherently personal. As a result of an ever-growing body of research, the additional and long-lasting harms associated with hate crime are more widely recognised by scholars, policy-makers and professionals (Paterson et al. 2019; Iganski and Lagou 2015; Williams and Tregidga 2014). © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_8

111

112  S.-J. HARDY AND N. CHAKRABORTI

This chapter seeks to enhance this knowledge-base by focusing more explicitly on the physical, emotional and community-level impacts generated by hate crime. Drawing upon evidence from a broad and diverse sample of hate crime victims, this chapter presents an unflinchingly vivid account of the ways in which victims’ lives are blighted by bruises, broken bones and scars, by the emergence or exacerbation of debilitating health conditions, and by anxiety, fear and depression. It also illustrates that experiences of hate crime can often be commonplace that the impact has, to some extent, been ‘neutralised’ or normalised as part of an unwanted but inevitable feature of being ‘different’. This chapter will also show that the damage caused by hate crime extends well beyond the primary victim, often evoking fear, alarm and distress among family members, and in some contexts, within wider communities. Nearly every participant who took part in these studies reported that being a victim of hate crime had had some impact upon their quality of life. This chapter reveals, however, that hate crime affects victims differently depending upon the type and frequency of victimisation, and upon the identity strand that they have been targeted. Within this chapter, we explore the layers of damage caused by repeat victimisation, and the variations similarities and specificities evident the ways in which victims respond to hate crime in order to reduce the risk of future victimisation. These are nuances which until now have been invisible to and thus underappreciated by scholars, professionals and policy-makers.

Physical Harms As noted in Chapter 6, a sizeable proportion of respondents had experienced property crime. Through the course of the interviews, we heard from participants who had been forced to endure offensive words such as ‘cunt’ scratched into their car and their car windows smashed, sometimes repeatedly; graffiti spray-painted over their house; rubbish, fireworks and faeces pushed through their letter box; eggs, stones and bricks thrown at their house; their plant pots smashed, fences pulled down and gardens ruined; and their houses set alight. While the financial cost of repairing or replacing property is often appreciated, the harm caused by property crime is much more far-reaching. As evidenced by this research, the location in which a hate crime takes place can heavily influence how the incident affects the victim and their family. This is particularly relevant when incidents occur in or near the victim’s home as is often the case within

8  INVISIBLE HARMS 

113

the context of property crime. It is difficult to comprehend what it must feel like when the hostility, abuse and threatening behaviour appears to be inescapable, and when these incidents are taking place in close proximity to children and young people. As part of these studies, we met victims who had been pushed, punched, kicked, hit with bottles, sticks, bats and poles, stabbed, groped, raped and driven into. These incidents left victims with cuts, bruises, broken bones and scars which were still visible years later. For some participants, their experiences had been so extreme that they had been admitted to hospital to receive the medical treatment they needed to repair, heal and recover. While the physical injuries caused by hate crime are widely recognised, the detrimental impact of experiencing this form of victimisation upon a victim’s physical health is underappreciated. Many participants within this research reported having good physical health before being subjected to persistent abuse and intimidation. For these participants the underlying anxiety induced by their experiences manifested through the emergence of new ailments, impairments and conditions, including heightened blood pressure, panic attacks, psoriasis, sleep disruption and deprivation, irritable bowel syndrome, and inflammation of arthritis. Similarly, we heard from participants who had ongoing physical and mental health conditions which in their own right were debilitating, and which were exacerbated by being subjected to hate crime victimisation. With MS one of the things we find is that the higher your stress levels are, the more problems you can have with MS. It can progress quicker … I ended up having to go back into hospital. Karen, targeted on the basis of her physical disabilities I’m on antidepressants. My blood pressure was up. Kat, targeted on the basis of her sexual orientation Every time they throw a punch at me, it’s another scar. It takes a while to heal from that. Paula, targeted on the basis of her transgender status He’s on the sick at the moment through depression. He’s not dealing with it. He has panic attacks when he goes into crowds. Sharon, describing her son’s experiences of being targeted on the basis of his sexual orientation and physical disabilities

114  S.-J. HARDY AND N. CHAKRABORTI I dreaded every day of my life, I didn’t get any sleep. I had a massive arthritis attack, for nine months I was crippled. Jo, targeted on the basis of her gender

The interviews revealed the upheaval caused by being admitted to hospital to treat the flare-up of an ongoing health condition, but also the respite it provided from the persistent abuse, and the growing sense of trepidation as the discharge date grew closer. This research found that opportunities were missed to identify hate crime, to support victims and to involve criminal justice agencies. Although many of those who experienced new symptoms or a deterioration of an existing condition visited a general practitioner, health centre or hospital, very few were asked about the underlying causes of these health issues. It is important to consider the wider repercussions of a victim’s illhealth. Through interviews, participants spoke of needing to take time off from work or having to leave work completely which often resulted in financial hardship; of feeling unable to adequately care for their children or to play an active role in taking them to school, to parties and to other recreational activities; and of avoiding social gatherings or public settings generally. This research highlights that the experience of living with physical injuries and health conditions also has a detrimental effect upon a victim’s emotional well-being. As highlighted by the survey, the level of impact varied according to the type of hate crime that victims were subjected to, with those victims whose most recent experience of hate crime had involved property crime or violent crime being more likely to say it had affected their quality of life significantly.

Emotional Harms Within the survey, the impact of victimisation was analysed according to different types of offence, thereby enabling us to unpick the specificities in the harms associated with different forms of victimisation. Additionally, due to the rich demographic data collected through the survey, we were able to identify significant differences between victim groups in relation to the impacts of hate crime. Only 5 per cent of the total sample stated that their experience of verbal abuse and harassment in both online and offline settings had had no impact upon them, with the most common response being that victimisation had caused upset. It is important not to underestimate or to downplay this impact; through

8  INVISIBLE HARMS 

115

the interviews we met participants who still felt consumed by sadness and hurt—sometimes years after an incident had taken place—and who fought back tears when recounting their experiences. I’ve been called all sorts of names because I’ve got disabilities. It makes you feel like a lesser form of human being and makes you feel empty, angry, upset, confused, all of the above. And it just leaves you feeling hollow for so long. Keith, targeted on the basis of his learning difficulties It makes you feel quite upset and angry. You know there was a time when I couldn’t leave the house without crying. Frances, describing her daughter’s experiences of being targeted on the basis of being autistic It makes you feel demoralised. It makes you feel hated. It makes you feel isolated, unwanted. Ahmed, targeted on the basis of his religion I was very, very angry. I felt a fool … I just feel sick now … after you’ve been attacked, you feel so small and so … I can’t even tell you how I felt. Lorna, targeted on the basis of her age I have never experienced a single, serious trauma, but the pervasive effect of being frequently harassed and intimidated has affected me in terms of my mental and physical health. Trish, targeted on the basis of her sexual orientation It made me not feel safe at home. It made me dread coming home. If I heard her car pulling up, I would immediately tense up. It was horrible … I dreaded every day of my life. Margaret, targeted on the basis of her gender

It is important to highlight that the proportion of victims who reported feeling upset as a result of their victimisation was similar for both those who had experienced the more ‘everyday’ forms of hostility, as well as for those who encountered the physical manifestations of hate crime—including property crime, violence and sexual assault. However, those who experienced violent expressions of hate crime were more likely to say that they felt angry (41 per cent) and now distrusted others as a result (37 per cent).

116  S.-J. HARDY AND N. CHAKRABORTI

Survey findings suggest that the impact of experiencing ongoing abuse and harassment is similar to the harms caused by the more extreme acts of hate crime. For instance, 41 per cent of those who had experienced the more ‘everyday’ forms of hate crime reported feeling anxious as a result of their experiences, compared to 44 per cent of those who were subjected to a physical act. Similar patterns were visible in relation to vulnerability (36 per cent compared to 42 per cent) and feeling fearful (33 per cent compared to 45 per cent) as a result of victimisation. Notable differences in relation to the effects of hate crime upon a victim’s emotional well-being were apparent, however, between different groups of victims. For instance, of the overall sample of survey respondents who had experienced the more violent expressions of hate crime, those targeted on the basis of their transgender status (77 per cent), physical disabilities (73 per cent), mental ill-health (72 per cent) and learning disabilities (66 per cent) were more likely to describe feeling vulnerable as a consequence (compared to 42 per cent of the overall sample). Another standout finding to emerge from the survey was that a substantial proportion of victims reported feeling depressed and suicidal as a result of their victimisation. Of particular interest is that those who had experienced verbal abuse and harassment in both online and offline contexts were more likely to report feeling depressed compared to those who had been on the receiving end of property crime or violent assault (29 per cent compared to 24 per cent). Furthermore, the same proportion of victims in both categories reported feeling suicidal (7 per cent). It was terrible, it was very dark point in my life. I felt like I just wanted to end it. It was just hatred every single day. To make it even worse my family weren’t even that supportive. They were probably the worst. Drew, targeted on the basis of his sexual orientation Years ago I was at breaking point, it really made me ill. If I’d been on my own I think I’d have … well I don’t know what I would have done. Lorna, targeted on the basis of her age In my case the bullying, harassment and verbal abuse has led to serious depression, and it became almost impossible to leave my home where I feel safe. Claire, targeted on the basis of her physical disabilities

8  INVISIBLE HARMS 

117

Again, some significant variations were evident within the data in relation to the impact upon different groups of hate crime victims. For example, of the total sample who experienced a more ‘extreme’ manifestation of hate crime, those targeted because of their mental ill-health (70 per cent), their transgender status (54 per cent), their physical disability (51 per cent) and their sexual orientation (43 per cent) felt depressed to a much greater extent. Specifically, those who were targeted on the basis of their mental ill-health, transgender status and learning disabilities were more likely to report feeling suicidal as a result of their victimisation (46 per cent, 38 per cent and 26 per cent, respectively). These figures are significantly higher than the corresponding proportion of the total sample who felt the same (7 per cent). When analysing findings across a broad range of impacts, it is evident that the emotional harm caused to transgender respondents who have experienced both the more ‘everyday’ and ‘extreme’ acts of hate referred is acute, with high proportions citing feeling upset, vulnerable, distrustful, angry, depressed and suicidal. These variations in the impacts reported by different groups of victims highlight the importance of recognising the influence of the broader context and the situational factors which can exacerbate the harms associated with hate crime. For instance, many of the groups of hate crime victims listed previously are socially, economically and politically disempowered, encountering virulent discrimination in relation to education, employment and housing. In this respect, hate crime is one aspect of a broader continuum of prejudice faced by these groups in everyday life, and thus the impact of this hostility in all its guises is cumulative. Through the course of interviews, it was revealed that many victims grew up feeling different, feeling inferior and feeling like they do not belong. Therefore, when a perpetrator singles out a victim on the basis of a particular feature of their identity or lifestyle, they further reinforce and magnify this notion of difference which is already an embodied status for hate crime victims because of prior, entrenched experiences in school, employment and everyday life. The stuff they’re saying about you is the same stuff you have fought for years. To be comfortable in your own skin. To come out and say “This is who I am and I’m happy with it”. So when it does happen it’s something you can’t change but it’s something you struggle to accept. Ollie, targeted on the basis of his sexual orientation

118  S.-J. HARDY AND N. CHAKRABORTI It’s a particularly painful form of crime. It’s an attack on your very being, and that’s both extraordinarily painful and humiliating. Nadia, targeted on the basis of her gender I think growing up gay, it’s true to most people, my generation and even now growing up gay, your experiences at school are horrible, they can scar you and have an effect on you for the rest of your life. Michael, targeted on the basis of his sexual orientation It’s all of those little micro stresses that, in themselves, don’t feel like they mean much because you kind of let them go, but actually do have an impact over a period of time on you, about how you feel about yourself. And it does have an impact on your mental health. Tony, targeted on the basis of race You try to forget these things but somehow it becomes almost a part of you. You can’t get rid of it. It’s like a bad smell: you try and try and try and you just can’t get rid of it. Gillian, targeted on the basis of her mental ill-health

Within the total sample of participants there were certain groups of victims whose experiences of hate crime victimisation formed part of their day-to-day life. It was within this smaller sample of participants that the impact of hate crime had, to some extent, been ‘neutralised’. In part, this can be seen as a coping strategy, whereby participants had come to accept that hostility was a ‘normalised’, ‘routine’ feature of their everyday life. When you’re young, at first it upsets you, because you don’t understand so much when you’re young. You think, “What have I done? Why are you doing that to me, I haven’t done anything wrong?” When you get older you understand more and you live with it, you toughen up against it. And then obviously you fight back. But when you’re young it is horrible, I’m not gonna tell you no lies. Kayla, targeted on the basis of being an English Gypsy The first couple of times that you lose somebody from the community you get upset and whatever else, but then eventually you come to the point where I can’t remember half their names now … Since I’ve been out on the scene anywhere between seven and twelve people have killed themselves. Cath, targeted on the basis of her transgender status

8  INVISIBLE HARMS 

119

I have lived a lifetime with hate crime. I’m lucky in the sense that I can move on from incidents. It’s not easy to be able to do this but it’s a survival technique learnt over time. Yvonne, targeted on the basis of her physical disabilities When I told my friends about what happened to me, they found it hilarious that I reported it to the police. Because they think it’s something weird to do, they’re just so used to it they no longer view it as a hate crime at all. Dariya, targeted on the basis of her religion People shouting or spitting or throwing bottles from cars or, you know, stuff like that. But that stuff happened all the time, it was nothing out of the ordinary at all. Laura, targeted on the basis of her alternative appearance These experiences have made me a stronger person. Maybe I’ve just become accustomed and developed a thick skin. Faizah, targeted on the basis of her religion I suppose because I grew up over here, in the old days it was quite bad but in a way I’ve got used to it. Rish, targeted on the basis of his race

Within interviews, participants spoke about having the ‘strength’ to deal with being targeted because they had developed a ‘thick skin’ and had ‘accepted’ that they will encounter hostility as a result of being ‘different’. Within this context, informal support networks and individual-level resilience were identified as being important factors that can help victims to cope with hate crime as illustrated within the next chapter.

Wider Harms Hate crimes are often described as ‘message crimes’ designed to intimidate not only the victim but also their family members and the broader community to whom they are perceived to belong (Paterson et al. 2019; Perry and Alvi 2012; Perry 2001). Through interviews, we were able to explore the wider impacts of hate crime, and discovered that participants’ family members could often be directly and indirectly targeted (as illustrated in Chapter 6 through the extract of an interview with Beyani). We also heard from respondents who felt guilty because their family and

120  S.-J. HARDY AND N. CHAKRABORTI

friends had been affected by their experiences, or fearful because their family was at risk of further victimisation. Well the harassment and hate was really levelled at my daughter and her family, but then it overflowed onto us. So if I’m out in my car and I’m seen by the perpetrators, I’m given abuse. Cut throat signs. The ‘V’ signs. Laughed at but not in a joking way, in a mocking way … It’s had such a profound effect on Mary, her husband and the kids, that it automatically then rolls on down the family. Mandy, targeted on the basis of husband’s mental ill-health They did stuff to my brother as well. They used to whack him on the way home, with sticks and stuff, just because he was my brother. Anne, targeted on the basis of her learning difficulties When it comes to throwing eggs at the door, it’s not very nice. I have my kids here. I’m scared for my kids because what if they just come and hurt my children when I’m not here? Jaali, targeted on the basis of his race I don’t feel myself or my children are safe, because I know that the group are going to attack me again. In my house they attacked me twice, and then they attacked my wife and car and the children and everything has been damaged. Our health is now going down. I don’t feel my children are safe if I leave home and when I’m outside all I think about is hoping that my home has not been attacked again. Beyani, targeted on the basis of his race I think twice about going some places and I spend time educating and preparing my children about what a potentially dangerous world this is. Anthony, targeted on the basis of being in a mixed race relationship

This led some parents to adopt a protectionist approach, monitoring their child’s whereabouts, and curtailing their freedom to play in the garden, the street or the park. While this theme was not explored in sufficient depth to offer conclusive assertions, evidence suggests that growing up in this environment in which a parent suffers with mental ill-health or in which controlling or overbearing tactics are employed can affect a young person’s emotional development and well-being (Fitzsimons et al. 2017; Stafford et al. 2015; Waite et al. 2014).

8  INVISIBLE HARMS 

121

The survey findings further demonstrated the impact of hate crime upon families, with similar proportions of victims who had experienced ‘everyday’ and more ‘extreme’ forms of victimisation reporting that it had caused arguments within their family (11 per cent and 12 per cent, respectively). Of note is that within the total sample of victims who had experienced a violent expression of hate crime, those targeted on the basis of their learning difficulty (34 per cent), their mental ill-health (33 per cent), their physical disability (27 per cent), their transgender status (23 per cent) and their asylum seeker status (20 per cent) were more likely to refer to this impact. Moreover, the survey revealed that family arguments were more likely to occur when the respondent knew the perpetrator who carried out their most recent hate crime, regardless of whether this was in the capacity of an acquaintance, neighbour, work colleague or family member. The impact of the most recent hate crime upon respondents’ quality of life was found to correlate with whether or not the incident had caused arguments within their family. Past studies have referred to the wider community harms associated with hate crime; namely the process by which the impact extends beyond the individual victim to create a sense of fear and apprehension among people with similar identity characteristics to the victim. These wider harms were not consistently evident within this study. For example, during the early stages of the project many of the thousands of people with whom we engaged were unaware of the broader problems encountered within their own community and did not feel as though other people’s experiences of hate crime had any bearing on their own quality of life. This sense of detachment from what are commonly, and simplistically, assumed to be homogenous and fully connected communities came through especially strongly in some of the interviews with gay men. I think in a city like Leicester where we’ve got such a broad, diverse, ethnic community, and a strong faith background, if you come from one of those faith backgrounds, and you’re gay, or lesbian, bisexual, trans you’re facing a double whammy of isolation and prejudice. The gay commercial scene probably won’t cater for you; it’s very White and if you’re not into drinking, then it definitely won’t cater for you. And your own cultural group, this ethnic and faith group will not embrace your sexuality, you’re completely isolated. Dennis, targeted on the basis of his sexual orientation

122  S.-J. HARDY AND N. CHAKRABORTI I hate the gay community, they’re the most utterly malicious people, not all of them, but the majority of them are malicious. Malicious and evil. Will, targeted on the basis of his sexual orientation I think the LGBT communities, if they exist, have to take a certain responsibility themselves. The scene is very small, it’s tiny and it tends to be young, White and tends to be very working class. The middle class professional gay people, they’re elsewhere. And also the scene is shrinking, not just in Leicester but in other communities because technology has changed, people don’t need to go to bars and clubs to meet people. Martin, targeted on the basis of his sexual orientation

However, the wider community impact of hate crime became more apparent in the context of religiously motivated victimisation and within some of the smaller, more marginalised minority groups, including the trans, homeless and English Roma communities. We’re all tarred with the same brush. We’re used to it. We’ve grown up with it all. Even a newborn child in our community, it’s hated because it’s been born. Now that child, it has to come into the world, the parents let that child in and they love that child, so why should everybody else hate it? It’s not done anything wrong. Jackie and Pauline on the basis of being English Gypsies We have to stick together. We are like one big family. If there’s somebody new we look out for them. Mark, targeted on the basis of being homeless We all know somebody who wears a veil. And in my case it’s my sister and her daughters. So it affects us all when someone is targeted … it just demoralises you, because you feel even more vulnerable. Nadima, targeted on the basis of her religion

It is widely acknowledged that experiences, particularly if they are negative in nature, can travel quickly and deeply within certain communities. As discussed in greater detail within the next chapter, the role of familiar and community networks has particular relevance in relation to decision-making when it comes to reporting, and in accessing support services.

8  INVISIBLE HARMS 

123

Coping Strategies This section reveals the steps which victims take in attempt to cope and to recover from their experiences of victimisation, before moving on to highlight the tactics employed by victims to reduce their risk of future victimisation. When asked about the impact of hate crime, similar proportions of respondents who had encountered the more ‘everyday’ and ‘extreme’ forms of hate crime reported that they had turned to alcohol or to prescription or non-prescription drugs to deal with their experiences (8 per cent and 5 per cent compared to 8 per cent and 6 per cent). There were variations in groups of victims reporting these coping mechanisms, with those targeted on the basis of their learning difficulty, their mental ill-health, their physical disability or their transgender status being most likely to have turned to substances regardless of the form their victimisation took. In terms of variations in the impact caused by different forms of hate crime, respondents who had experienced verbal abuse and harassment were more likely to say that it had led them to avoid certain areas (35 per cent compared to 28 per cent of those who had experienced violent crime), to conceal their identity (11 per cent compared to 8 per cent), and to change their appearance (11 per cent compared to 8 per cent). Again, those targeted on the basis of their physical and/or learning disability, their mental ill-health or their sexual orientation were most likely to report these impacts as a result of their experience. Of particular significance, however, is that hate crime victimisation appears to result in transgender people making more considerable changes to their daily routines, which includes avoiding certain areas (69 per cent compared to 28 per cent of the total sample), concealing their identity (46 per cent compared to 8 per cent) and changing their appearance (46 per cent compared to 8 per cent). The survey findings also revealed a slight variation in terms of those who had experienced a violent crime being more likely to move house (22 per cent compared to 15 per cent of those who encountered verbal abuse or harassment) or to leave the city altogether (15 per cent compared to 11 per cent) as a result of their victimisation. Furthermore, in instances where the participants knew the perpetrator who had committed their most recent hate crime, the victim was more likely to have changed their mobile phone number or to have moved home.

124  S.-J. HARDY AND N. CHAKRABORTI

Within the interviews we were able garner further insight into the strategies that victims employ to feel safer and to reduce their risk of victimisation. These ranged from personal changes in dress and/or appearance, including Sikh men cutting off their hair, Muslim women removing the veil and Muslim men shaving off their beards. We also captured the more practical strategies employed by our sample such as crossing the road to avoid large groups of people, carrying safety devices, improving their home security and installing CCTV, and bypassing certain areas altogether including public parks, transport, places of leisure and other ostensibly ‘safe’ spaces for those not at risk of hate crime. In some cases, respondents had undertaken more extreme measures to minimise the risk of repeat victimisation, such as attempting to hide their nationality, their religion, their sexual orientation, their transgender status or their asylum seeker status. I dress down whenever I go out, rather than dress like I would actually like to, you know there’s things that I would do differently if it were acceptable. To be honest even if I lived in London I’d probably do things differently. Liam, targeted on the basis of his sexual orientation I went to B&Q and bought CCTV cameras from my own pocket and put a CCTV camera on the front of the house, on the back of the house, and the side. Hasan, targeted on the basis of his asylum seeker status I started drinking heavily, ended up with a drink problem, because that was the only way I could get to sleep. So it was either get absolutely hammered so I can at least fall asleep but have a horrific hangover, or get no sleep at all. Sally, targeted on the basis of her mental ill-health I have one of the numbers of my mobile phone as a video and another as a voice- recorder. So if you do get into trouble, or you think you’re getting into trouble, you can just put your hand in your pocket, hold the key down and start recording. Cath, targeted on the basis of her transgender status You’re always living in fear. I mean a woman in a burka or a man with a beard, will try not to go out at night, avoid pubs, avoid dark alleyways, avoid stopping near a car and speaking with someone looking for directions. It’s all just fear. Amandeep, targeted on the basis of her race and religion

8  INVISIBLE HARMS 

125

I’m really, really anxious, and you know, in the last three months I haven’t been out at all, apart from coming out to see you here today because I’m too damn scared to go out of my house. Iffy, targeted on this basis of his physical disabilities and religion Oh yeah, if we see trouble we change our way. Simple as that. And not talking with other people because maybe they make trouble for us. And we know we haven’t got any right in the country. Adil, targeted on the basis of his asylum seeker status And it was then I thought, I’ve just got to sell the house, I’m never going to beat this. And so that’s what I did, I stuck a For Sale sign up, took a 50 grand loss and bought a detached house. That was agony because I had to live somewhere else to start afresh … the police just said, ‘you’re going to have to move out’. Jo, targeted on the basis of her gender It’s that feeling that gets left with you. You do feel scared. Now whenever I go to town I’m extra careful what I do and where I go. Even when you go down an escalator I’ll wait until no one’s there before I get on the escalator. If there are lots of people in the lift I won’t go in there. It affects you in different ways. Taliha, targeted on the basis of her religion

Among the sample there were differences in the measures adopted by participants to make themselves feel safer, with women more likely than men to have taken such steps. For instance, women were more likely to avoid walking in certain areas or going to certain places (67 per cent of women used these avoidance strategies compared with 54 per cent of men) and to avoid going out at night (43 per cent compared with 27 per cent). They were also more likely to have improved their home security (31 per cent of women compared with 20 per cent of men), started carrying personal security devices (19 per cent compared with 10 per cent) and changed their mobile phone number (13 per cent compared with 9 per cent). There were also differences between the forms of victimisation experienced and the strategies implemented to reduce the risk of re-victimisation. Respondents who had been victims of verbal abuse, harassment or violent crime were more likely than respondents who had not been victims of such crimes to avoid walking in certain areas or going to certain places. Similarly, those who had been victims of violent crime were more likely to avoid going out at night than respondents who had not been targeted in these ways.

126  S.-J. HARDY AND N. CHAKRABORTI

Conclusion This chapter has sought to convey the devastating reality of being subjected to hate crime victimisation. It revealed that hate crime affects people in different ways, from the injuries endured as a result of victimisation, to the deterioration of someone’s physical health. It also illustrated that the emotional harms suffered by hate crime victims can be profound, with almost every single research participant feeling that hate crime had adversely affected their quality of life. Of particular importance is the revelation that there are certain victims groups who suffer more acute impacts as a result of their victimisation, including higher levels of anxiety, isolation and depression. These happen to be the same groups and communities— including disabled people, people with mental ill-health and transgender people—who reported experiencing abuse, harassment and physical assault to a greater extent when compared to other groups of hate crime victims. These groups often lack a sense of established or recognised civic presence and find themselves marginalised from the awareness-raising, policymaking and advocacy evident within ‘mainstream’ society. As illustrated through a series of quotations, for those groups of victims whose experiences form ‘part and parcel’ of their everyday life, the only way in which the impacts of hate crime can be mediated is through the adoption of coping strategies and defence mechanisms. The steps taken by victims in an attempt to reduce the risk of repeat victimisation ranged from avoiding certain areas, changing dress and appearance, installing or carrying safety equipment, moving house or changing jobs or school. Unfortunately, for many victims the feeling of vulnerability and the fear of future victimisation remained a constant feature of their day-to-day reality even after taking such steps, and this was often because they had been denied access to support services or had received an inadequate and ineffective response from criminal justice agencies and other relevant organisations.

References Fitzsimons, E., Goodman, A., Kelly, E., & Smith, J. P. (2017). Poverty dynamics and parental mental health: Determinants of childhood mental health in the UK. Social Science & Medicine, 175, 43–51. Iganski, P., & Lagou, S. (2015). The personal injuries of ‘hate crime’. In N. Hall, A. Corb, P. Giannasi, & J. G. D. Grieve (Eds.), The Routledge international handbook on hate crime (pp. 34–46). London: Routledge.

8  INVISIBLE HARMS 

127

Paterson, J. L., Brown, R., & Walters, M. A. (2019). Feeling for and as a group member: Understanding LGBT victimization via group-based empathy and intergroup emotions. British Journal of Social Psychology, 58(1), 211–224. Perry, B. (2001). In the name of hate: Understanding hate crimes. London: Routledge. Perry, B., & Alvi, S. (2012). ‘We are all vulnerable’: The in terrorem effects of hate crimes. International Review of Victimology, 18(1), 57–71. Stafford, M., Kuh, D. L., Gale, C. R., Mishra, G., & Richards, M. (2015). Parent–child relationships and offspring’s positive mental wellbeing from adolescence to early older age. The Journal of Positive Psychology, 11, 326–337. Waite, P., Whittington, L., & Creswell, C. (2014). Parent–child interactions and adolescent anxiety: A systematic review. Psychopathology Review, 1(1), 51–76. Williams, M. L., & Tregidga, J. (2014). Hate crime victimization in Wales: Psychological and physical impacts across seven hate crime victim types. British Journal of Criminology, 54(5), 946–967.

CHAPTER 9

Invisible Victims

Abstract  This chapter is framed around three key barriers which victims face in relation to accessing justice and support. This chapter highlights problems associated with levels of awareness and confidence amongst hate crime victims; with the appropriateness and effectiveness of existing criminal justice interventions; and with knowledge and uptake of support services. It demonstrates that there is no ‘one-size fits all’ approach to delivering support to hate crime victims. Keywords  Hate crime Justice

· Victim support · Reporting · Recording ·

During the past two decades, and as outlined within the early part of this book, a series of steps have been taken by lawmakers, non-governmental organisations, activists and professionals within and beyond the criminal justice sector to develop improved responses to hate crime through the introduction of relevant legislation and evidence-based policy. However, within this chapter we show that these developments have been only partially effective in addressing victims’ needs and lived experiences and that the damage caused by hate crime can often be reinforced—and not alleviated—by the continued failures of policy-makers, practitioners and scholars to respond effectively. This chapter is framed around the three key barriers which victims face in relation to accessing justice and support. There has been a tendency to consider the perceptions, experiences © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_9

129

130  S.-J. HARDY AND N. CHAKRABORTI

and support needs of hate crime victims as a collective rather than teasing out the specificities in the factors which affect the capacity of victims from different identity groups and backgrounds to access appropriate support. It is important to understand these discreet experiences because it is only through greater acknowledgement of these shortcomings that we can begin to address them.

Barriers to Reporting One of the common goals across our research studies has been to assess hate crime victims’ experiences and expectations of agency responses to their victimisation and to identify ways of improving the quality of support offered to them. Criminal justice and other partner agencies are somewhat hampered in their ability to investigate and prosecute hate crimes without being notified of an incident, which is why victims are considered to be the ‘gatekeepers’ of the criminal justice system (Gottfredson and Gottfredson 1988: 16). As outlined within Chapter 3, the importance of reporting in the context of generating effective prioritisation and service delivery is well documented, both within previous academic research on hate crime and within key sources of policy guidance; indeed, the Hate Crime Action Plan issued by the UK Government in 2016 pledged to ‘ensure that everyone has the opportunity and information needed to report hate crime’ (point 89) while in its Foreward the Secretary of State for Communities and Local Government stated that ‘if we report every incident of hate crime, we can drive it from our streets’ (2016: 6). Despite such commitments, our research highlights that there are still a wide-range of structural, social, situational and individual barriers which generate a deep-rooted resistance to reporting amongst hate crime victims. Within the large-scale survey administered in Study One, respondents were asked if they had reported their most recent experience of hate crime to a criminal justice organisation or partner agency or to an individual who is in a position to offer support. We found that only 24 per cent of respondents had reported their most recent experience of hate crime to the police. There were some subtle variations in patterns of reporting to the police between respondents who had been targeted on the basis of different aspects of their identity or lifestyle. For example, only 14 per cent of those targeted because of their sexual orientation and 17 per cent of those targeted because of their membership of

9  INVISIBLE VICTIMS 

131

an alternative subculture had reported their most recent experience, both of which are notably lower than the 24 per cent overall reporting level. In comparison, significantly higher reporting rates were evident amongst respondents who had been targeted because of their physical disability (56 per cent) and/or learning difficulty (44 per cent). There were also some nuances in reporting levels amongst survey respondents who had experienced different types of victimisation. Victims of verbal abuse were the least likely to have reported the crime to the police (16 per cent), followed by offline harassment (33 per cent), online harassment (36 per cent), sexual violence (41 per cent), violent crime (60 per cent) and property crime (62 per cent). When this pattern of reporting was probed further within Studies Three and Four, 84 per cent of survey respondents stated that they were ‘unlikely’ or ‘highly unlikely’ to report being verbally abused to the police; 58 per cent stated that they would be ‘unlikely’ or ‘highly unlikely’ to report being harassed online; and 48 per cent stated that they would be ‘unlikely’ or ‘highly unlikely’ to report being harassed either in person. The survey findings also revealed, however, that people would be much more inclined to report their victimisation if it were to involve deliberate damage to property or physical attack (80 per cent and 84 per cent, respectively). It appears that the ‘tipping point’ for reporting relates to situational-level factors such as the presence of violence, repeat victimisation involving the same perpetrator, and the reality of needing a crime number for insurance claims. Additionally, the greater the impact that hate crime had upon the respondent at the time, the more likely they were to have reported it to the police. Within Study One, we found that those who said that their experience had affected them very significantly were more likely to have reported their experience of hate crime to the police (55 per cent compared with 24 per cent overall). Further factors of influence revealed by the survey data included having knowledge of who the offender was (31 per cent of those who knew the offender(s) reported the incident to the police compared with 20 per cent of those who did not) and the incident having occurred in or near their home (45 per cent compared to 24 per cent overall). When asked why they had reported their most recent experience of hate crime to the police, almost two-thirds of survey respondents stated that they had done so because it was a serious crime (63 per cent). Half said that they had reported it because it is important to report any experience of crime (50 per cent); 48 per cent had done so in the hope

132  S.-J. HARDY AND N. CHAKRABORTI

that the offender(s) would be brought to justice; and 29 per cent had reported because it had happened before. Three-quarters of respondents felt that the police had recorded their most recent experience of hate crime after they reported it (72 per cent). This finding is concerning as it suggests that nearly a third of respondents who had reported a hate crime to the police did not believe that the police had recorded it as such (28 per cent). Also of concern is the fact that fewer than half of respondents thought that the police had investigated their most recent incident (43 per cent). In just one-sixth of cases did respondents think that the police had arrested someone (16 per cent). There were some slight differences between the circumstances surrounding the incident of hate crime and the response provided by police. For instance, survey respondents who had reported a hate crime involving violence were more likely than others to have said that the police arrested someone in relation to the case (31 per cent said this compared with 16 per cent overall). Respondents who had known the offender(s) involved in the reported hate crime were more likely to have said that the incident culminated in a warning being issued (24 per cent compared with 12 per cent overall). In terms of victims’ satisfaction with the response received from the police, just over half of our sample who had reported their last hate crime to the police said that they were satisfied with the police’s response (55 per cent). Interestingly, this finding closely mirrors the most recent Crime Survey for England and Wales which highlighted that half (51 per cent) of victims of hate crime were very or fairly satisfied with how the police handled the incident, which is a lower proportion than for all CSEW crime (69 per cent) (Home Office 2018). Around a third of respondents within Study One were dissatisfied with the response that they had received from the police (35 per cent). It is noteworthy that respondents who felt that they had been very significantly affected by their experience of hate crime were more likely to be dissatisfied with the response of the police than others who were less affected (54 per cent compared with 35 per cent overall). This is also true for those whose most recent hate crime had involved violence (53 per cent of such respondents were dissatisfied with the police response compared with 35 per cent overall). The contrast between our participants’ expectations of the police response and the actual level of service they had received emerged as a significant theme within the interviews. Unfortunately, positive encounters with the police were heavily outweighed by negative ones.

9  INVISIBLE VICTIMS 

133

We are all told to report incidents, [and so] we went and reported this to the police. Our experience following that was actually extremely negative … we were told that because they couldn’t find anybody that they were not going to record it as a hate crime, but as a drunk and disorderly incident, public affray or whatever they call it. That has made me feel extremely sad … I started to think I would have been better off not reporting it because then I would not have had the added stress of going through all of that and nothing happening. Nira, targeted on the basis of her religion Their services were appalling. The police promised, I don’t know how many times, to follow up and come round to my flat, and all sorts of things. I don’t even have a crime reference number … they just completely ignored me, which is really, really frustrating. Cath, targeted on the basis of her transgender status Our police service is totally shit. They don’t help at all. Kumar, targeted on the basis of his race and religion I called the police and they said there’s nothing we can do because you’re not under threat. I reported again, to say I wasn’t satisfied. They sent an officer out who said, “Can’t you find a different way to walk to the club you go to?” I said “No - I haven’t done anything wrong”. Gavin, targeted on the basis of his physical disabilities I’m quite surprised that I’ve even heard the police do anything. They never helped me no matter how many times I go in and complain. … So like I say, my experience with the police is that they are just useless and worthless. They’ll investigate every other crime but that [type of] crime. John, targeted on the basis of his mental ill-health and sexual orientation I know now I cannot trust the police to deal effectively and sensitively with any hate crime. Ameera, targeted on the basis of her race and religion

Over three-quarters of survey respondents had not reported their most recent experience of hate crime to the police, a worryingly high figure (76 per cent). When asked why, the most frequently cited reasons were that they did not feel the police would take it seriously (30 per cent), that they dealt with it themselves or with the help of others (27 per cent) or

134  S.-J. HARDY AND N. CHAKRABORTI

that the police could not have done anything (20 per cent). The perception that an incident would not be taken seriously was particularly prominent amongst those groups who are often not associated with hate crime policy, including those targeted on the basis of their membership of an alternative subculture (61 per cent), their asylum seeker status (47 per cent) and their gender (41 per cent). These structural and situational barriers also emerged prominently during the course of the interviews. There’s not much they can do anyway because you never come across those people again. I would never recognise them. There’s no way they could get them. It would just be a waste of time on my side to report it to the police. Eniola, targeted on the basis of her asylum seeker status and race I mean, what’s the point in telling them? They won’t listen. Someone like you can [directed at the researcher]. If we go to anyone and report it, they make assumptions about who you are. And they won’t understand. Keith, targeted on the basis of his learning difficulties If I went to the police to report every single incident I wouldn’t be doing anything else. I would be spending half my day being insulted and humiliated. Then the other half a day I would spend in the police station reporting things. Liz, targeted on the basis of her transgender status If you’ve not got an independent witness then the police aren’t interested. If it’s just slagging you off, smirking, rude signs you’re told to ignore it. Which if it happens to you occasionally, you can. When it’s happening to you all the time and over a long period of time, it really, it really gets to you. Mandy, targeted on the basis of her husband’s mental ill-health When I’ve spoken to gay people and they’ve mentioned things have happened to them I’m like “Did you report it?” And in nearly all of the cases they haven’t, they’re either “What’s the point?” or they don’t realise it is something they can report. Like lots of victims of hate crime they think you’ve just got to put up with that. No-one is going to take it seriously. Andrew, targeted on the basis of his sexual orientation Ideally I should report it, and I know that I should report it. Will I go to the police and say somebody was being verbally abusive, being racist towards me, chanting these words? Do I expect them to do anything? It’s a waste of time. If they cannot deal with Stephen Lawrence in 20 years what chance do I have with verbal abuse? It hurts but it doesn’t kill. Baakir, targeted on the basis of his race

9  INVISIBLE VICTIMS 

135

One of the most prominent themes to emerge from the interviews was a sense that victims felt as though they were receiving an especially poor level of support from the police because of who they are. We heard from many participants from different backgrounds who felt that the police response had been influenced by negative attitudes towards their identity or perceived ‘difference’, be it their disability, ethnicity, mental ill-health, religion, sexual orientation or transgender status. We are scared of the police, even if they said they would change their ways. If we see somebody fighting, we would just change the way we are going, we wouldn’t want to be involved. They might start blaming us. Once the police know we are asylum seekers then the way they are talking to us is changed. You can’t call the police, if you call the police maybe they arrest you instead. Kidane, targeted on the basis of his asylum seeker status and race A lot of BME people that I speak to find the level of confidence in the police is not that high. Nobody believes that these guys can do anything anyway. Unless it’s the other way round, and you do silly stuff and see how quickly they descend on you! Aiyden, targeted on the basis of his race I’m beginning to think there is now some sort of, what do they call it, institutionalised racism. So is this institutionalised Islamophobia? We have two ladies who have reported incidents, conveniently the CCTV is the wrong way. It’s not recorded as a hate crime but as a public affray. Let’s report another one and conveniently the CCTV footage is lost. Veiled Nadima, targeted on the basis of her religion If the police are standing and someone calls me a nigger, they are straight over there. If someone calls me a filthy dyke, like, they just wait for whatever witty retort I come back with and stand and have a giggle. Lou, targeted on the basis of her race and sexual orientation The majority of them [the police] are very ignorant with anything to do with mental health. George, targeted on the basis of his mental ill-health and physical disabilities I think a lot of it is the fact that they [transgender people] are frightened that the police will not take them seriously and perhaps even laugh behind their backs. Lucy, targeted on the basis of transgender status

136  S.-J. HARDY AND N. CHAKRABORTI

A person’s perception of the value of reporting a specific form of crime to the police is also shaped by the frequency with which they are experiencing victimisation. We found that there was a sizeable number of participants for whom micro-aggressions, harassment and abuse were “part and parcel” of their daily life. From this perspective, everyday forms of hate such as verbal abuse and threatening behaviour are viewed as being ‘ordinary’ and ‘routine’. When a victim conceptualises experiences in this way it means that such incidents are often not thought of as deserving a criminal justice intervention. Another prominent situational barrier that was identified relates directly to the environment of reporting. Many participants described reporting structures as being “confusing”, “complicated”, “time-consuming”, “daunting”, “stressful” and “emotionally draining”. For victims with work and childcare commitments or caring responsibilities, taking time off to report hate crimes was simply not a viable option, particularly when these experiences are a regular occurrence. Concerns were also raised by members of new and emerging communities and those for whom English was not a first language, that cultural and linguistic barriers prevented them from reporting. Equally, the level of courage, patience and resilience needed in order to share harrowing experiences with an unfamiliar and potentially sceptical third party was something that participants believed was commonly overlooked by practitioners, and thus created further resistance to reporting. As noted, only one in four of respondents had reported their most recent experience of hate crime to the police. The lack of confidence in the police’s ability to take hate crime seriously and the disconnect between the response that victims expect and the service that they receive from the police is concerning. However, responses to hate crime require understanding and intervention from other agencies too, and for this reason, we also explored which other organisations victims may be reporting to or accessing support from. Over the last decade, there has been an increase in the use of third party reporting centres that are located in non-criminal justice organisations. The aim of such centres is to offer an alternative reporting outlet to that of the police while also recognising that hate crime is not just the responsibility of the criminal justice system. These centres offer victims the opportunity to share experiences of hate crime with a wide-range of non-statutory organisations. We found that just one-fifth of those who took part in the survey had reported their most recent experience of hate crime to an individual or organisation(s) other than the police (18 per cent).

9  INVISIBLE VICTIMS 

137

Today I visited a group for elderly Sikh women, a mental ill-health support group and conducted three individual interviews. Not one person out of the 100 individuals had heard of third party reporting centres. I reeled off local libraries, community centres and online reporting mechanisms but not one individual had heard of them, let alone used them. I am shocked because I know that both the City and the County Councils have awareness raising campaigns, as well as the police. There appears to be a clear disconnect between those who decide where to place reporting centres (and how to raise awareness of them) and those who are directly affected by hate, prejudice and targeted hostility. Field diary entry

Aside from police officers, teachers were the most likely individuals with whom respondents had shared their experiences of hate crime (4 per cent), compared to social care workers (2 per cent), a doctor or nurse (2 per cent) or a community leader (1 per cent). In terms of the types of organisations that respondents had reported to, the locally commissioned victim support and the local council were most frequently cited (both 3 per cent). Only 1 per cent of respondents had reported their most recent experience of hate crime to a community support organisation, such as those for lesbian, gay and bisexual groups, or those relating to disability, race equality and/or religion/belief. Of those participants who conveyed at least some awareness of these schemes, most were highly critical of the perceived inaccessibility of third party reporting options, and specifically the inappropriate locations of reporting centres and the potential for digital formats such as websites and mobile phone apps to exclude those who are unfamiliar with or unable to use such technologies.

Barriers to Justice One of the areas that we were keen to explore was how victims felt about conventional criminal justice outcomes, and whether this response met their expectations with respect to offering an appropriate level of punishment and rehabilitation. As highlighted in Chapter 3, amongst the reasons for having hate crime laws in place is their capacity to act as a deterrent to potential offenders; to convey a message of solidarity to minority victims; to signal society’s condemnation of such acts; and to acknowledge the additional harms caused by hate crimes. However, of the thousands of people that we engaged with over the lifecycle of

138  S.-J. HARDY AND N. CHAKRABORTI

these studies only a small proportion knew what a hate crime was and barely anyone had any knowledge of existing hate crime laws. This poses a challenge with respect to the effectiveness of hate crime legislation, as it appears that very few people at a grassroots level—and in environments where people are most likely to require protection from the law—know that this legislation even exists. This in turn made exploring victims’ opinions on hate crime legislation much more difficult. In addition to having little prior knowledge of hate crime laws, relatively few survey respondents had any prior experience of seeing offenders brought to justice through the courts system. Only 4 per cent of respondents’ most recent experiences of hate crime had gone to court. Of those cases that had got that far, just over half (54 per cent) had seen the offender found guilty, while 5 per cent had seen the case dismissed/ discontinued and in a similar proportion the offender had been found not guilty. 18 per cent of respondents stated that their case had not gone to court yet, and a further 15 per cent did not know the outcome of their case. These figures indicate that a surprisingly high proportion of those participants whose cases went to court did not know the outcome. When this issue was probed, further it was evident that this was the result of victims not being kept up to date with developments within their case and/or victims finding the whole procedure of going to court too confusing. Study Four provided an opportunity to assess hate crime victims’ preferences with regard to which approaches should be prioritised in order to tackle hate crime perpetration and to rehabilitate offenders. It transpired that the vast majority of participants would like to see greater priority given to designing and developing initiatives that educate young people about diversity and hate crime (85 per cent). Within interviews, an increased use of tailored programmes of education within schools and local communities was commonly referred to as a mechanism to inform young people about positive aspects of diversity, to challenge fears and stereotypes, to connect divided and segregated communities and to raise awareness of the harms of hate. Equally, a significant proportion of participants stated that they would like to see more hate crime cases going to court (71 per cent). Concerns over how daunting and distressing the criminal justice process can be did not appear to affect participants’ willingness to pursue a conviction within the right context. Approximately half of the respondents cited that community ‘payback’ orders and/or

9  INVISIBLE VICTIMS 

139

diversity awareness courses should form part of a suite of responses used with hate crime perpetrators (56 per cent and 48 per cent, respectively). The ignorance in some of the young boys and girls nowadays is there because they’ve not been educated. They’ve not had support from their parents in terms of knowing what’s right, what’s wrong, what you can say, what you can’t say or what you shouldn’t say. Isaac, targeted on the basis of his race I think that education and awareness needs to start younger … So much needs to be done in schools that isn’t being done. But the whole integration and acceptance of others surely needs to come from young, so that it’s natural. Sophie, targeted on the basis of her gender I don’t think anybody ever is born racist or sexist, or disablist or hostile, or anything. Everything is taught. How to be bad is taught. Therefore, surely to God we can teach people how to be good. I just think with a national curriculum that forces them to learn about English grammar and Henry VIII, it’s missing the real stuff that will keep us safe and secure and optimistic and efficient in life, is kind of being left behind. Chris, targeted on the basis of his sexual orientation I think it’s more about having interaction with different faiths and undertaking different awareness and educational programmes, especially in schools. Of course, there is RE subject, but there has to be a special focus on respecting different faiths, so that a child going to school, from the first day, they have a positive message in their mind. Nazir, targeted on the basis of race and religion The only punishment for those people is to educate them. We are all human beings, what they did is wrong. So this is my only request: if they be punished be it through education. Hafsa, targeted on the basis of his race and religion I went to the LGB tent and I filled out a form to see if there’s anything they could change and I said, yeah, get more people into schools. For the younger generation. Ollie, targeted on the basis of his sexual orientation

Nearly a third (32 per cent) of actual and potential hate crime victims would like to see more opportunities for face-to-face supervised

140  S.-J. HARDY AND N. CHAKRABORTI

mediation between the victim and the offender. Restorative practices aim to facilitate an inclusive and meaningful dialogue between those who have been involved in or affected by a crime in order to heal the harm caused or to prevent it from happening again (Walters 2014). Although restorative practices have become relatively familiar within the criminal justice system, its use for hate crime cases continues to face resistance. The limited application of restorative justice in the context of hate crime has resulted in a paucity of data on its effectiveness in rehabilitating offenders, but research does indicate that it can empower victims by providing a platform for their voices to be heard (Walters and Brown 2016; Walters 2014). One of the key concerns expressed by participants with regard to restorative justice was in relation to coming face-to-face with the perpetrator. However, face-to-face interaction between the victim and the perpetrator is just one approach that falls within the range of available restorative practices. When participants were asked for their opinions on a restorative approach that involved the perpetrator and a ‘champion’—someone who would belong to the same community or possess the same identity trait as the victim—the response was overwhelmingly positive. It was perceived that this form of restorative practice would readdress the power imbalance which can exist between the victim and the perpetrator while still providing an opportunity to convey the harm caused by the incident. Within the last decade, academics and policy-makers have increasingly focused on the purpose and application of hate crime legislation, and in particular whether it should be extended to include other identity characteristics or lifestyles beyond the monitored five strands. Although these debates are important, the findings from our research illustrate how few cases of hate crime proceed to court. Rather than wishing to see perpetrators punished more severely through the higher sentencing tariffs available under hate crime legislation, hate crime victims appear to be overwhelmingly in favour of educational approaches which offered perpetrators—and specifically young people who had or who were at risk of committing hate crimes—an opportunity to learn about the positive elements of diversity and the harms associated with hate crime through a programme of tailored intervention. As mentioned above, many also spoke positively about the ideas behind alternative forms of justice, and in particular the capacity of restorative interventions to encourage offenders to acknowledge the consequences of their actions for victims and their families.

9  INVISIBLE VICTIMS 

141

Barriers to Support As part of Studies Three and Four, we were able to undertake a more in-depth inspection of victims’ experiences and expectations of support services. One of the most troubling findings to emerge from these studies—particularly in light of the considerable range of harms caused by hate crime—is that very few hate crime victims have accessed any form of support service to assist them in their recovery. Just 4 per cent had accessed support provided by the police, 3 per cent from a locally commissioned victim support service and 1 per cent from their local authority. Interviewees were especially expressive about locally commissioned victim support services. Victim Support was fantastic. They arranged times for us when he [the perpetrator] was out. I was in quite a state by the time I got to them, but she [Victim Support worker] would phone me every other day to make sure I was coping. Karen, targeted on the basis of her physical disabilities I’ve been under Victim Support for quite a time now … I have got somebody there who listens, and if I don’t think the police are taking it seriously, I’ve got somebody who is happy to push it along as well. The police seem to take them more seriously then what they would take me. Paul, targeted on the basis of his visual impairment I was offered counselling. I kinda went, no, no, no, I’m a closed person. But part of me wishes I had of done that, because I would have got it all out then, rather than let it fester. Because it was a long time before I could talk about it. Margaret, targeted on the basis of her gender I think Victim Support probably would have been helpful. I didn’t feel that I really needed to talk to anybody about it at the time. But looking back I maybe should have done,’cause it did change the way I acted and if I went out. I didn’t take it at the time because of the way it was offered. It’s a good thing that it’s offered, but the way it’s offered it’s like the police, and certainly in my case, I was at the police station, they asked “Do you want to talk to anybody about this attack?” Well no you don’t, you just want to go home. So maybe if there was some follow up, more follow up than there already is, then that would be a lot better. Simon, targeted on the basis of his sexual orientation

142  S.-J. HARDY AND N. CHAKRABORTI

Locally commissioned victim support services were often organisations with which victims were generally satisfied once they had accessed their services. One of the positive aspects of these services commonly referred to by participants was the speed with which support is provided. However, many participants spoke of not wanting to use victim support services immediately after they had experienced a hate crime, but instead to have the opportunity to access their services later down the line. Unfortunately, when they were ready to talk to somebody about their experiences they found the process of accessing victim support to be difficult, primarily because they had not received a follow-up notification to remind them of how to access victim support which made participants feel as if they had missed their opportunity to use the service. This was particularly true for some of the more marginalised and isolated groups who often had very little, if any, knowledge of victim support services. Several interviewees had also turned to other organisations, such as housing associations, the local authority or a tailored community support service, to report their experiences of hate crime and/or to receive support. When my wife was sitting with a handful of tablets, I phoned the doctor and the doctors were absolutely brilliant. They said “Get her down here in ten minutes”. It wasn’t a case of “You need to book an appointment” or anything like that, no, “Get her down here in ten minutes”. And got her down to see a doctor, got her put on antidepressants. That was when it was really at its worse. Basically the doctors kept an eye on us. Jeremy, targeted on the basis of his physical disabilities A lot of doctors know nothing about us, a lot of health professionals know nothing, a lot of psychiatrists know nothing. How many psychiatrists are there trained in transgender people? Cath, targeted on the basis of her transgender status The City Council were awful. They didn’t want to get us away from the person who was abusing us, every day of our lives. And it was every day. The worker actually said “I don’t see a problem, he’s a nice guy”. David, targeted on the basis of his physical disabilities I remember last time when I was assaulted, I was really desperate to sit down and have a talk with somebody, but I didn’t really get much attention from the LGBT centre I went to just afterwards. I didn’t get much attention from Victim Support either. Lyndsey, targeted on the basis of her transgender status

9  INVISIBLE VICTIMS 

143

I’m satisfied with the people at [name of locally-run charity]. They are such good people. They are helping a lot and they are doing plenty to help us, to help everyone actually. Because of their behaviour as well and they have a passion, they listen completely to you and they are trying to find a solution for your problems. Adil, targeted on the basis of his asylum seeker status

The views of victims with respect to the response received from their local authority and housing associations were overwhelmingly negative. The most commonly cited criticisms in this context were that staff working within those organisations lacked knowledge of what a hate crime is, had not taken victims’ experiences of hate crime seriously and had failed to intervene early on within the victimisation process. Those participants who had engaged with such organisations to access support often found that the response did not help them but instead simply reinforced their sense of victimisation, despair and isolation. To date, no study has sought to explore the reasons as to why such a small number of hate crime victims utilise existing support provisions. Within Studies Three and Four, respondents were asked why that had not sought support after experiencing a hate crime, and the most commonly cited reason was that the victim felt that they were able to deal with it themselves or with the help of others (46 per cent). It is important, however, to consider the implications for those who do not have informal support networks at their disposal. Respondents to this question referred to the invaluable support that they had received, and continue to receive, from their family members, friends, carers and colleagues. The second most commonly cited explanation for not accessing a support service was a lack of awareness when it came to knowing what services existed (23 per cent). As the quotations that follow illustrate, hate crime victims are commonly unaware of which organisations they can turn to in order to access emotional and/or practical support. I didn’t even know any support mechanisms existed. Syed, targeted on the basis of his religion Where support can be accessed needs to be well communicated.Aat present people don’t know where to go to get support. Minah, targeted on the basis of her religion and appearance

144  S.-J. HARDY AND N. CHAKRABORTI Currently there is a confusion about who exactly one should go to for support. Hafsa, targeted on the basis of his religion If there are services available, these need to be publicised more widely as I was not aware there were support systems. Preet, targeted on the basis of her race and religion We know [where we can access support] because we’re lucky enough and privileged enough to be in the know, that the police is part of a wider network of organisations both statutory and voluntary. But I’m not sure that comes across clearly to service users. Rahul, targeted on the basis of his physical disabilities and race

Awareness of hate crime support services tends to be limited to those in ‘privileged’ positions, including those who work in an environment related to hate crime and/or those who are socially and economically empowered. However, the findings from this study suggest that the level of knowledge even amongst those who occupy these ‘privileged’ positions is often inadequate. It is vital that front-line professionals who engage with those communities and groups who are most vulnerable to victimisation—and who are least likely to be aware of support services including asylum seekers, people with learning difficulties and older people—are equipped with the knowledge to be able to signpost victims to appropriate provision. These studies also provided invaluable opportunities to ask victims to consider what features of a support service would be most important to them in terms of repairing the harms caused by hate crime. The most frequently cited expectations were being able to access support quickly (77 per cent); being treated with kindness and compassion (75 per cent); being able to access practical support (e.g. safety advice, personal safety equipment) (60 per cent); being made aware of local support groups (50 per cent); being able to access emotional support face-to-face (45 per cent); and being able to access emotional support on the phone (33 per cent). When participants were asked about who they would want to provide them with emotional support the response was mixed, with participants referring to mainstream organisations, specialist services, community and voluntary-run groups and charities. For those participants who felt that they had more complex support needs as a result of repeat victimisation, the suggestion was that emotional support should

9  INVISIBLE VICTIMS 

145

be delivered by a specially trained counsellor. However, it became evident that regardless of which organisation was providing the support, the most important characteristic of an effective support service was that the member of staff responsible for providing support was trained to ensure that they fully comprehend how impactful hate crime can be.

Conclusion This chapter has highlighted problems with levels of awareness and understanding amongst hate crime victims; with feelings of confidence in criminal justice agencies and partner organisations; with the appropriateness and effectiveness of existing criminal justice interventions; and with knowledge and uptake of support services. Reporting enables victims to become ‘visible’ to criminal justice agencies and to other relevant organisations, which should result in victims being able to access practical and/ or emotional support. Those victims who do not report to the police and who are in need of support are often left to suffer in silence by virtue of not possessing the knowledge or resource to access support services. This chapter demonstrated that there is no ‘one-size fits all’ approach to delivering support to hate crime victims. Participants expressed a desire for services that are tailored to meet their needs, which will vary greatly on the basis of a range of different individual and situational factors, including the type of hate crime experienced by the victim or how often they have been targeted; the availability, or otherwise, of existing support networks for the victim; their social and economic position within society; and the presence of physical or mental health issues. This chapter has also revealed that many hate crime victims are receptive to the extended use of alternative interventions beyond conventional punitive measures. This includes calls for more information about, and access to, restorative approaches to justice so that perpetrators can begin to understand the impacts of their actions through the use of mediation processes. To a lesser extent, participants suggested that rehabilitation programmes should be used more readily within prisons and within communities in order to address the underlying prejudices that give rise to individual acts of hate and to reduce the likelihood of perpetrators committing similar acts in the future.

146  S.-J. HARDY AND N. CHAKRABORTI

References Gottfredson, M. R., & Gottfredson, D. M. (1988). Decision making in criminal justice: Toward the rational exercise of discretion (2nd ed.). New York: Plenum. Home Office. (2016). Action against hate: The UK Government’s plan for tackling hate crime. London: Home Office. Home Office. (2018). Hate crime, England and Wales, 2017/18. London: Home Office. Walters, M. (2014). Hate crime and restorative justice: Exploring causes, repairing harms. Oxford: Oxford University Press. Walters, M. A., & Brown, R. (2016). Preventing hate crime: Emerging practices and recommendations for the improved management of criminal justice interventions. Brighton: University of Sussex.

PART IV

Transforming Responses

CHAPTER 10

Implications for Scholarship

Abstract  Within this chapter, we reflect upon how our findings have contributed to a significant step change within the field of hate studies. We put forward a new, more inclusive definition of hate crime which accounts for the breadth of experiences that have emerged during the course of conducting each of the four studies and which captures the intersectional nature of hate crime victimisation. We also highlight the ways in which our research has advanced scholarly understanding of hate crime, particularly in relation to the diverse profile of victims, the complex process of victim selection and the ‘everyday’ nature of victimisation. Keywords  Hate crime Intersectional

· Victims · Targeted hostility · Prejudice ·

Within this chapter, we reflect upon how our findings have contributed to a significant step change within the field of hate studies. We put forward a new, more inclusive definition of hate crime which accounts for the breadth of experiences that have emerged during the course of conducting each of the four studies and which captures the intersectional nature of hate crime victimisation. We also highlight the ways in which our research has advanced scholarly understanding of hate crime, particularly in relation to the diverse profile of victims, the complex process of victim selection and the ‘everyday’ nature of victimisation. Finally, this © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_10

149

150  S.-J. HARDY AND N. CHAKRABORTI

chapter considers the implications of our research for future studies of hate crime and illustrates why the process of thinking more creatively about methods of access and engagement is central to capturing the voices of hidden victims and to facilitating a more comprehensive understanding of diversity and ‘difference’.

Defining Hate Crime If you put four academics in a room and asked them to define hate crime they would emerge days later with five different definitions. (Giannasi 2014: 35)

The concept of hate crime has been subjected to a myriad of definitions (see, inter alia, Gerstenfeld 2013; Hall 2013; Petrosino 2003; Perry 2001; Sheffield 1995). While there is no universal consensus, these academic contributions have helped to illustrate the complexities associated with hate crime by demonstrating the relationship between structural hierarchies, institutionalised prejudices and acts of hate. Perry’s (2001) definition in particular has enhanced our understanding of the relevance of power dynamics within modern societies by suggesting that expressions of violence are qualitatively different when motivated by bigotry and directed towards already disempowered and marginalised populations (Perry 2001). As Giannasi goes on to acknowledge, few practitioners with any knowledge of this field and its associated complexities would ever deny the value of academic debate which is grounded in different theoretical perspectives, nor would they expect a universal standpoint to be reached. Nonetheless, if such debate is to shape policy and ‘real-life’ responses to hate crime victims, definitions need to be constructed and presented in ways that have meaning to people who are affected by, and who are responsible for responding to, hate crime. Based on the considerable body of evidence presented within preceding chapters, we would argue that hate crime is a much more expansive notion than existing academic frameworks suggest. Definitions rooted within group identity politics have the propensity to create and reinforce divisions between communities of identity rather than illustrating the shared nature of their victimisation. We propose a new empirically grounded definition which sees hate crime as any act of violence, intimidation or hostility directed towards people on the basis of their identity, perceived ‘difference’ or vulnerability. The rationale for framing

10  IMPLICATIONS FOR SCHOLARSHIP 

151

hate crime in these broad but simpler terms is to ensure that targeted micro-aggressions which might not be criminal acts in themselves—but which can have damaging impacts upon the victim, their family and wider communities—are covered by this definition. Equally, while this definition is deliberately concise, it acknowledges that some victims can be subjected to hate not exclusively because of their membership of a particular identity group but also because they are seen as vulnerable or somehow ‘different’ in the eyes of the perpetrator (this nascent theme is discussed in greater depth within the next section). Furthermore, our definition has the potential to advance conventional conceptualisations of hate crime by offering insight into the motivations which underpin such acts without seeking to determine or categorise those who are responsible for committing hate crimes. Existing definitions constrain our understanding of how this phenomenon materialises in the ‘real-world’ by suggesting that hate crime is exclusively a form of ‘stranger-danger’ crime. As highlighted in Chapter 7, our research evidence challenges this assumption as we found that a substantial proportion of hate crime victims knew the perpetrator in some capacity, whether as a neighbour, a colleague, a classmate or a carer, to name just some examples. As a result of our research and the new knowledge that we have generated, we no longer need to draw generic assumptions about the relationship between victims and perpetrators but instead we can begin to explore the nuances in different groups’ and communities’ experiences. For instance, we found that people with mental illhealth and physical disabilities are most vulnerable to being targeted by an individual with whom they are familiar, whereas members of alternative subcultures and lesbian, gay and bisexual people are much more likely to be attacked by a stranger. Additionally, conventional thinking is that hate crimes are typically committed by people who belong to majority or well-established groups but this research has revealed that both majority and minority groups, and both established and new and emerging communities can express hostile views and commit acts of hate crime. In short, we have established a definition which provides a more realistic account of the wide-ranging nature of hate crime victimisation, the complex interplay between identity and lifestyle characteristics, and a broad range of economic and situational variables which can increase exposure to hate crime, and the dynamic relationship between the victim and the perpetrator.

152  S.-J. HARDY AND N. CHAKRABORTI

Theorising Hate Crime Over the past decade, hate crime scholarship has occupied a more prominent position within the discipline of criminology and within social policy more broadly. However, despite the growth in hate crime research there remained a dearth of knowledge in relation to the process and experience of victimisation, particularly in relation to those groups who have traditionally occupied a marginal position within research and policy frameworks. By undertaking the most comprehensive studies of hate crime victimisation ever undertaken within the UK, we now know so much more about who is affected by hate crime; more about the prevalence and forms that hate crime can take; more about the physical and emotional impacts of hate crime upon victims, their families and communities; and more about who commits hate crime. One of the most salient findings from these studies is that for many victims micro-aggressions are a routine ‘everyday’ feature of their life. Within Study One, we found that 48 per cent of the total sample are called abusive names related to their identity, lifestyle or perceived ‘difference’ on a frequent basis. It is worth noting that this proportion increased to 54 per cent within our most recent study which may indicate that experiences of targeted hostility have become more prevalent since the EU referendum in 2016 and explain why more than half of participants within Study Four felt more concerned about becoming a victim following the referendum. Additionally, our research has revealed that certain groups appear to be more at risk of encountering hostility on a routine basis when compared to others, including those with mental ill-health, people with physical disabilities and trans individuals. These groups also emerged as being more likely to be on the receiving end of threatening, intimidatory and violent behaviour, further evidencing variations in the types and frequency of victimisation experienced by different groups of hate crime victims. Before we conducted Study One, little was known about how hate crime plays out in the ‘real-world’. We have generated new and important knowledge on the everyday social spaces which can be risk-laden environments for hate crime victims. Our research shows that hate crime victims commonly encounter hate crime in public streets, at home, in the night-time economy, at school, college or university and in the workplace. In particular, public transport was found to be a particularly problematic environment for those who possess visible markers of

10  IMPLICATIONS FOR SCHOLARSHIP 

153

‘difference’, including disabled people, veiled women, trans people and members of alternative subcultures. It is difficult to imagine a lived reality which entails living in constant fear of being harassed, abused or attacked whilst travelling to work, doing the weekly shop, going to school or simply returning home. In this sense, our research has revealed the ‘ordinariness’ of hate crime, not in relation to the harms it causes—and Chapter 8 highlighted that these are considerable and wideranging—but in terms of its infusion into the daily routines of many hate crime victims. Our research also challenges conventional wisdom in relation to the types of people affected by hate crime. Existing academic frameworks are predicated upon the notion that hate crime victims come from and are targeted because of their membership of a minority group (see Perry 2001 for instance). This is likely to be the result of the sheer weight of academic research devoted to race and racist victimisation in comparison with the relatively small body of work on other strands of hate crime. However, one of the shortcomings of existing literature on racist hate crime and within the field more broadly has been a common failure to see beyond simplistic constructions which depict minority groups as one seemingly homogeneous victim group. This approach dismisses, or at best underplays the differences in experiences and need between people grouped together within such a framework, as well as those who are typically excluded from such a framework. Our research has enhanced scholarly thinking by revealing the hostility suffered by those groups who have been overlooked or who are often not associated with the five identity strands monitored by the criminal justice system within England and Wales (CPS 2018; College of Policing 2014). For instance, with regard to racist hate crime we highlighted the diversity of experiences which are often subsumed within this category and encountered by asylum seekers, Gypsies and Travellers, international students, those in mixed-race relationship, new and emerging communities, and refugees. Similarly, our research has advanced understanding of the nuances in the patterns of victimisation for the other monitored strands, including the differences in experiences for those targeted on the basis of their learning difficulties and/or their physical disabilities, for those who practice different religions and for those who identify as lesbian, gay and bisexual. The findings from our research add weight to the criticism posited towards hate crime scholarship—and hate crime policy—in that it has been exclusionary in its approach, resulting in many groups of ‘Others’

154  S.-J. HARDY AND N. CHAKRABORTI

being unable to claim recognition for their victimisation under the umbrella of ‘hate crime’. Within each of the four studies, we came across significant numbers of victims who had been targeted on the basis of aspects of their identity, lifestyle or appearance which are not commonly associated with hate crime victimisation, including a distinctive or strong accent, age, body shape, dress, education status, gender, homelessness, immigration status, living alone, mental ill-health, membership to an alternative subculture and social status. Importantly, the experiences of the less familiar groups within the five monitored identity strands, and of those who were targeted on the basis of identity and lifestyle features which sit outside of those strands altogether, bore all the hallmarks of the more recognised and established hate crime victim groups. As mentioned above, these findings challenge conventional theories of hate crime which frame such incidents as linked exclusively to the victim’s membership of a particular group and instead require scholars to recognise the importance of visibly different identity markers and of vulnerability. For instance, in Chapter 6 we highlighted the central role that dress and appearance plays within the context of victim selection. Dress and appearance includes visual indicators of identity traits which can contribute to people being more easily identifiable by virtue of a specific ethnicity, physical disability or sexual orientation and which mark someone as being visibly ‘different’ such as those who wear unconventional modes of dress or those who have physical imperfections such as birthmarks and scars. This reinforces the relevance of intersectionality which has traditionally been underplayed by scholars within the field of hate studies. Our research enhances understanding of the interplay of identities with one another and with other personal, social and situational characteristics. We found, for example, that half of hate crime victims have been targeted on the basis of more than one identity or lifestyle characteristic, with many targeted because of their race and their religion; their mental ill-health and their physical or learning disability; or their sexual orientation and their dress and appearance. Indeed, we found that the vast majority of hate crime victims are concerned that they will be targeted again in the future on the basis of more than one identity or lifestyle characteristic. Significantly, our research has revealed that being targeted on the basis of more than one variable is more likely to have a significant impact upon a victim’s quality of life. The relevance of class and economic marginalisation to the commission of hate crime has also been underexplored to date, with our research highlighting that hate crimes

10  IMPLICATIONS FOR SCHOLARSHIP 

155

can often be triggered and exacerbated by socio-economic conditions. Ultimately some victims will invariably be better placed than others to avoid persecution by virtue of living at a greater distance from prejudiced neighbours or in less overtly hostile environments. This contribution to scholarship has the capacity to transform understanding of the lived realities of hate crime victimisation, particularly for those groups whose experiences have been rendered invisible and will shape new thinking in relation to the diverse profile and needs of hate crime victims and the intersectional nature of victimisation. We have been able to produce a more complete and accurate account of hate crime victimisation and to generate fresh knowledge on the patterns, specificities and nuances facing different communities, primarily by adopting a less formulaic, grassroots-based approach to conducting research with diverse communities. As such, our research also has methodological implications for the social sciences more broadly, as discussed within the following section.

Researching Hate Crime Our experiences over the life cycle of these studies have shown that the use of more flexible and creative approaches to locating and engaging with ‘hidden’ communities is key to developing more inclusive and comprehensive analytical frameworks. Such approaches enable researchers to access diverse communities at a grassroots level, as opposed to being reliant upon the limited access points provided by official community leaders or gatekeepers. For instance, when we spent prolonged periods of time immersed within everyday spaces such as international supermarkets, taxi ranks, food outlets and community centres (to name but a few) within Study One, we were able to informally engage with new and emerging communities who often occupy a marginalised and to some extent, ‘invisible’ presence. We began to connect with Somali, Congolese, Kurdish, Turkish, Polish, Lithuanian and other growing populations in different areas of the city, becoming more attuned to the social spaces which they use to interact with each other. Furthermore, and taking the Somali community as an example, it was only through using a range of different routes to grassroots engagement with a variety of Somali networks and groups that we learnt of the diversity that exists within this one category, such as the differences in background of those who have migrated from the Netherlands, Sweden and those seeking

156  S.-J. HARDY AND N. CHAKRABORTI

asylum directly from Somalia. By using more imaginative and meaningful approaches to engagement, a more inclusive construction of diversity can be developed which begins to capture the specificities and intersections in identity and experience. Moreover, using more nuanced approaches to engaging with ‘hidden’ and diverse communities is crucial to reducing both the real and perceived disconnect between researchers, their research and ‘everyday’ people. As mentioned in Chapter 4, during the initial stages of our grassroots engagement we realised that the term ‘hate crime’, which we had (naively) assumed to be widely known, was one which the vast majority of community members had never heard of. There are certain groups and communities who lack an established presence and who find themselves marginalised from the awareness-raising, policy-making and advocacy evident within ‘mainstream’ society. While hate crime is part and parcel of everyday life for many of these victims, our research has highlighted that they do not associate their experiences of micro-aggressions, intimidation and violence with the term hate crime which is why the vast majority do not report their experiences to the police or other partner organisations. Although many victims, particularly those from ‘hard to reach’ communities, may not initially recognise their experiences as hate crime or even as a form of victimisation, researchers can seek to address this challenge by familiarising themselves with the discourse used by different groups and communities. These are the kinds of lived realities which grassroots-based methods of engagement can enable researchers to become attuned to. It was only through the adoption of a multi-method pragmatic approach that we were able to develop a greater understanding of the intersectional nature, lived experience and impacts of victimisation. Grassroots engagement offered invaluable insights into the methods of data collection which were more appropriate for specific groups and communities and reinforced the importance of being agile in our approach. Within Study One—and as outlined in Chapter 4—we designed the survey strand of our methodology to be accessible and inclusive, with respondents being asked to self-select which aspect of their identity they felt they had been targeted because of. Rather than relying solely on the five protected identity characteristics of race, religion, sexuality, disability and transgender status, we expanded the question to cover 14 identity and lifestyle markers, which included broadening out catch-all categories and acknowledging identity

10  IMPLICATIONS FOR SCHOLARSHIP 

157

characteristics which often slip under the radar such as age, gender and asylum seeker or refugee status. This approach provided respondents with a broader range of identity choices, the option to select multiple variables and an opportunity to cite additional factors which were not listed, such as a distinctive or strong accent or body shape. Subsequently, we used a second strand of in-depth qualitative interviews to further explore intersections between identity characteristics and situational factors and the relevance of those intersections to people’s experiences of victimisation. It is within these interviews that we were able to develop our understanding of how certain socio-economic and cultural factors contribute to victims being in vulnerable situations, and therefore at higher risk of being exposed to hate and prejudice. The use of auto-ethnographic field notes completed our triad of data collection methods and enhanced our understanding of the reality and context of hate crime victimisation. In particular, documenting informal discussions and participant observations facilitated a greater recognition of the impact that hate crime can have upon the victim when understood in relation to the nature of the incidents, the context in which they took place and the interplay between identity, circumstance and vulnerability. To an outsider, ‘low level’ incidents such as being stared at, being threatened and being called abusive names may seem relatively minor. However, if researchers are able to achieve meaningful engagement with a specific group, the use of detailed field notes and reflections can provide greater insight into the lived reality and pervasive nature of targeted micro-aggressions. As already mentioned, often it is the more extreme acts of hate crime which tend to attract media, political and academic attention while the experiences and cumulative harms of the more ‘ordinary’, everyday forms of abuse, hostility and harassment are often underappreciated. We would urge researchers to utilise data collection and dissemination methods that capture the range of human emotions and the lived realities of victimisation that are all too often underappreciated and overlooked. This enables us to ensure that the needs and sensitivities of our research participants, and particularly those who occupy marginal positions within society, are at the forefront of decisions about access, engagement and data collection methods. These steps have been essential in facilitating a much more comprehensive understanding of victimisation which, in turn, has generated evidence-based recommendations which have the potential to shape new and improved policy responses to hate crime.

158  S.-J. HARDY AND N. CHAKRABORTI

Conclusion If research is to represent the voices of vulnerable, marginalised and ‘hard to reach’ communities, then the use of more innovative and unorthodox methods of grassroots engagement and data collection is pivotal. In addition to uncovering ‘hidden’ truths about their experiences, needs and perceptions, these methods enable researchers to circumnavigate the bureaucratic challenges associated with access through self-appointed gatekeepers and community leaders whose presence can sometimes be a hindrance to the process of undertaking meaningful research in this field. Although there will be contexts where this mode of access is appropriate, an over-reliance on such a strategy can reinforce the marginalisation of those less vocal voices within a given community. Adopting this engagement approach which factored in more imaginative modes of access, more inclusive constructions of ‘difference’ and more complete accounts of victimisation has generated new knowledge on the specific issues and problems facing different groups and communities who for too long have been peripheral to hate crime scholarship as a result of the deployment of narrow and inadvertently exclusive frameworks. This book has sought to make the ‘hidden’ worlds of hate crime victims ‘less hidden’ and by doing so has transformed scholarly understanding of the lived realities of hate crime victimisation and the failings of policy responses.

References College of Policing. (2014). Hate crime operational guidance. Coventry: College of Policing. CPS (Crown Prosecution Service). (2018). Hate crime annual report 2017–18. London: CPS. Gerstenfeld, P. B. (2013). Hate crimes: Causes, controls and controversies (3rd ed.). London: Sage. Giannasi, P. (2014). Academia from a practitioner’s perspective: A reflection on the changes in the relationship between academia, policing and government. In N. Chakraborti & J. Garland (Eds.), Responding to hate crime: The case for connecting policy and research (pp. 27–38). Bristol: Policy Press. Hall, N. (2013). Hate crime (2nd ed.). London: Routledge. Perry, B. (2001). In the name of hate: Understanding hate crimes. London: Routledge.

10  IMPLICATIONS FOR SCHOLARSHIP 

159

Petrosino, C. (2003). Connecting the past to the future: Hate crime in America. In B. Perry (Ed.), Hate and bias crime: A reader (pp. 9–26). London: Routledge. Sheffield, C. (1995). Hate violence. In P. Rothenberg (Ed.), Race, class and gender in the United States (pp. 432–441). New York: St. Martin’s Press.

CHAPTER 11

Implications for Policy

Abstract  The need for fresh responses to hate crime has become all the more apparent at a time when numbers of incidents have risen to record levels, both within the UK and beyond. Despite progress within the domains of scholarship and policy, these escalating levels of hate crime cast doubt over the effectiveness of existing measures and their ability to address the needs of hate crime victims. This chapter explores the practical implications of our research and is designed to generate improved policy within and beyond the domain of criminal justice by providing evidence-based recommendations which acknowledge the diverse needs of diverse groups as well as the very real practical constraints facing all professionals tasked with responding to hate crime. Keywords  Hate crime

· Policy · Policing · Law · Victim support

This book has identified a series of failings within existing policy-level responses to hate crime which undermine their credibility and effectiveness in the eyes of victims. It has drawn from rigorous, in-depth fieldwork to identify a set of specific problems—namely a failure to dismantle barriers to reporting, a failure to prioritise engagement with diverse communities and a failure to provide meaningful criminal justice interventions—which affect the capacity of hate crime victims from different identity groups and backgrounds to access appropriate support. © The Author(s) 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7_11

161

162  S.-J. HARDY AND N. CHAKRABORTI

The need for fresh responses to hate crime has become all the more apparent at a time when numbers of incidents have risen to record levels, both within the UK and beyond. Despite progress within the domains of scholarship and policy, these escalating levels of hate crime—and the associated increase in tensions, scapegoating and micro-aggressions that accompanies such spikes—cast doubt over the effectiveness of existing measures and their ability to address the needs of hate crime victims. This chapter explores the practical implications of the preceding discussion and is designed to generate improved policy within and beyond the domain of criminal justice by providing evidence-based recommendations which acknowledge the diverse needs of diverse groups as well as the very real practical constraints facing all professionals tasked with responding to hate crime.

Improving Front-Line Responses As charted in Chapter 3, progress within the context of domestic hate crime policy has been considerable over the course of the last two decades. The principles underpinning these developments within the domains of policy have been supported by successive UK Government Action Plans (HM Government 2012, 2014, 2018; Home Office 2016) which are designed to reinforce its commitment to tackling hate crime. The most recent of these, Action against hate, was published in the wake of the post-Brexit rise in recorded hate incidents across many parts of England and Wales and refers to the importance of protecting ‘the shared values that underpin the British way of life’ (Home Office 2016: 5) through a series of actions and priority areas which seek to identify best practice across all strands of hate crime. However, while the Action Plan sets commendable goals in relation to prevention, reduction and support for victims, it fails to specify how these goals will be achieved and evaluated or how the empirical evidence produced by academic research will be used to promote best practice. For instance, while the Action Plan pledges to ‘ensure that everyone has the opportunity and information needed to report hate crime’ (point 89), the findings from our research suggest that we have a very long way to go before these words can become a reality. Within Study One, for example, fewer than one in four survey respondents had reported their experiences of hate crime to the police and fewer still had shared their

11  IMPLICATIONS FOR POLICY 

163

experiences with other organisations who could potentially have offered support such as a teacher, Victim Support, their local authority or a social care worker. Only 1 per cent of respondents had shared their experiences of victimisation with a community support organisation, such as an LGBT, disability or race equality network, while none had ever utilised any of the third party reporting options available through local libraries or online via True Vision, despite third party reporting schemes (and the use of True Vision as an online information resource and reporting option) being heralded within successive Government Hate Crime Action Plans. Collectively these findings present an alarming picture of how often victims of hate crime victims suffer in silence and how few incidents come to the attention of the police and other agencies. A wide range of factors underpinned victims’ reluctance to disclose their experiences, with one of the most common across each study being a perception that reporting would simply be a waste of time, either for the victim (because the police and other partner organisations would fail to grasp the seriousness and prevalence of hate incidents) or for the police (because of a perceived tendency on the part of the police to prioritise other types of offence with easily identifiable perpetrators and more obvious resolutions). As a result, most victims tended to normalise their experiences of repeat harassment and hostility as a routine feature of being ‘different’, which in turn reinforced their sense of alienation. Equally, many victims expressed a sense of trepidation about the prospect of reporting which stemmed from fears about being outed, drawing attention to themselves or retaliatory attacks. This highlights how the process of reporting can in fact exacerbate existing concerns on the part of the victim, rather than provide the necessary levels of reassurance. Although some victims could take a degree of comfort from sharing experiences with friends or family members, the overwhelming majority across all strands of victims and types of incident would have welcomed the opportunity to tell the police or another appropriate ‘authority’ figure if the barriers to reporting did not feel so insurmountable. One of the most concerning findings to emerge from our research is that victims’ previous bad experiences of liaising with the police and other agencies—and the experiences shared by family members, friends and members of their own community—can reinforce a lack of trust and the sense that front-line professionals fail to show sufficient empathy or kindness to people who find themselves in positions of extreme

164  S.-J. HARDY AND N. CHAKRABORTI

vulnerability and difficulty. An inability to offer relevant insight, or the perception that practitioners were ‘winging it’ in terms of their background knowledge of or training in hate crime, was also a common criticism. The most effective way to improve understanding of hate crime and its impacts is through tailored and up-to-date training. In the context of escalating levels of hate crime across the UK, changing demographics within local areas and increased demands upon public services, it is crucial that the UK Government and national policing bodies commit to providing evidence-based diversity and hate crime training as a mandatory requirement for professionals at all levels. Despite successive UK Government Action Plans committing to improving police training, and the most recent update in 2018 citing that ‘The National Police Chiefs’ Council, with support from agencies including the Home Office and College of Policing is delivering hate crime training for police call handlers’, no package has yet been produced some 18 months on. Moreover, the weight of evidence from our series of studies suggests that organisations should be encouraged to publicise the training that staff receive to the general public in order to increase levels of confidence amongst those who have experienced—or who are at risk of experiencing— hate crime. Our research has highlighted that many hate crime victims lack awareness of what hate crimes are and where they can access support. This lack of knowledge also extends to the procedures and practices that are involved when the police or another relevant organisation are investigating a hate crime. It is widely recognised that hate crime cases can be particularly challenging and complex to deal with, which can result in investigations being drawn out and many cases concluding in no further action. Although the outcome cannot be changed, the way in which a victim feels about it and the service that they have received is often based on their interactions with front-line officers or professionals. Within each study criticisms of the initial response tended to stem from a feeling of not being listened to, not being taken seriously and not being treated with an appropriate level of decency and empathy. Interviews with victims in each area revealed that problems encountered at the pre-reporting and initial response stages were compounded by what they perceived as a slow, intimidating and at times incomprehensible criminal justice system. Many spoke at length about the difficulties that they had encountered when trying to make sense of what police officers were telling them to do; or the complex, unfamiliar

11  IMPLICATIONS FOR POLICY 

165

terminology used in reference to hate crime and accompanying legislation; or the seemingly endless delays in waiting to receive any form of communication or follow-up; or the hostile manner and tone adopted by front-line practitioners. These flawed responses are likely to explain why the Crime Survey for England and Wales found that hate crime victims are more likely to be very dissatisfied with the police handling of their case than victims of other forms of crime (25 per cent compared to 15 per cent) (Home Office 2018). These problems are deeply concerning given that victims are likely to have already experienced considerable emotional and physical distress as a result of their hate incident and to have negotiated various barriers in order to report the incident. As the first set of studies to explicitly ask victims about their support needs, this research has highlighted the intrinsic value of treating victims with kindness and compassion above all else. It is, therefore, incumbent on professionals working within and beyond the criminal justice sector to consider what steps they can take to ensure that the investigation process is explained in an accessible way and that case updates are communicated in a manner that is both transparent and empathetic. Clearly under-reporting remains a major challenge in relation to hate crime, and national and local policy has done little in reality to improve front-line responses for victims. The importance of simplifying reporting structures and making the reporting and investigation process more victim- and witness-friendly cannot be overstated. The vast majority of the victims within these studies who referred to various challenges around reporting had nothing against the principle of reporting per se and ideally would gladly report incidents in order to access support for the emotional and physical harms experienced and to feel a sense of justice. However, the barriers referred to above—in addition to the cultural and linguistic barriers which particularly affected those from new and emerging communities and those for whom English was not a first language—were significant enough to deter most victims from seeking much-needed help to alleviate their distress. As such, and despite continued references to improvements in reporting structures and front-line responses within state-level narratives, operational practice continues to be framed in a way that fails to acknowledge the needs and concerns of victims.

166  S.-J. HARDY AND N. CHAKRABORTI

Improving Justice Outcomes Collectively the four studies have revealed that an added frustration for those victims who had reported hate crimes to the police was a perceived failure on the part of criminal justice agencies to bring perpetrators to justice or to even keep them informed on the progress of their case. Within Study One, for example, only 4 per cent of survey respondents had seen their most recent experience of hate crime go to court. The difficulties surrounding interpretations of hostility and motive, and the stringent evidential proof required for prosecutions, have been explored elsewhere (see Walters et al. 2017). At the time of writing the Law Commission is undertaking a review of the scope and effectiveness of hate crime laws in England and Wales which is a welcome opportunity to shape a more effective, victim-centred legislative framework. Although we cannot pre-empt the outcome of this review, our research makes a compelling case for legal reform. Certainly a move to extend the range of existing aggravated offences—which currently only cover incidents motivated by hostility on the grounds of race and religion—to incorporate those offences which demonstrate hostility towards a person’s disability, sexual orientation or transgender identity would be welcome. Equally valuable would be an extension to existing stirring up offences in order to protect disabled and trans people from threatening words or behaviour, as would a legal framework which enables the courts to impose enhanced sentences in cases of targeted hostility which sit outside of the monitored strands. These reforms would ensure parity in legal protection across the monitored and unmonitored strands of hate crime and would facilitate a much more accessible, logical and consistent legal framework. Within the prevailing climate of rising levels of hate crime and victim dissatisfaction, addressing the complexity and inequality associated with hate crime would reinforce the government’s commitment to generating meaningful and sustainable improvements to the policy domain. It is also important to recognise that while legislation has considerable value, there is mounting research evidence which raises questions over the appropriateness and effectiveness of such responses (see Walters and Brown 2016; Hall 2013). Prison does not prevent or deter people from committing hate crime and may actually reinforce prejudicial attitudes, and it offers little opportunity for meaningful rehabilitation work to take place (ibid., 2013). A salient finding to emerge from our research is that

11  IMPLICATIONS FOR POLICY 

167

many victims were receptive to the extended use of alternative interventions beyond conventional punitive measures. Within Study Four, for example, fewer than half of survey respondents referred to longer prison sentences as their preferred response to hate crime. Instead, the vast majority of survey respondents called for greater use of tailored programmes of education within schools and local communities as a mechanism to inform young people about positive aspects of diversity and the harms of hate crime. Smaller but still significant proportions of respondents wanted to see an increased use of community ‘payback’ orders for hate crime perpetrators and were in favour of more face-to-face supervised mediation between the victim and the offender. These feelings were shared amongst victims of different types of violent and non-violent hate crime and from different communities, ages and backgrounds, and illustrate strong levels of support for utilising interventions outside of the formal criminal justice routes within responses to hate crime. However, despite the extended use of such interventions in the context of other criminal offences—and despite a growing body of empirical evidence to support the growth of restorative practices alongside educational and rehabilitative programmes as part of a package of preventative measures to combat hate crimes—their use in the context of hate crime remains relatively limited. Although the most recent UK Government Action Plan names a selection of education programmes to evidence its support for early interventions, it is difficult to see how these programmes can deliver sustained success in the context of the myriad other complex challenges within schools and in the absence of wider investment to take the pressure of delivery away from teaching staff who may be ill-equipped to engage pupils on issues relating to hate crime. Such indifference not only undermines the aims enshrined within national and local strategies, guidance documents and Action Plans to deliver effective hate crime responses; it undermines the importance attached to addressing the needs and expectations of the many thousands of hate crime victims who are failing to receive adequate support from existing criminal justice interventions. Within the UK and beyond, effective responses to hate crime perpetration require government bodies and policy-makers to invest in smarter forms of punishment, and not harsher punishment. Currently, there is little understanding of why people commit hate crimes, whether there are different factors that motivate different forms of hate crime, and what works in the context of rehabilitating hate offenders. It is only

168  S.-J. HARDY AND N. CHAKRABORTI

through the generation of a solid evidence base that police forces, local governments and other relevant partner agencies will be able to develop and deliver educational and rehabilitative programmes which challenge underlying prejudices and prevent future offending. In this context, the need for large-scale academic research which explores the profile and motivations of hate crime perpetrators remains pressing but this will only garner the level of evidence required if researchers are granted access to perpetrators involved in the criminal justice system which typically has been denied to them. This should be followed by a large-scale, indepen­ dent evaluation of existing educational and rehabilitative programmes in order to identify ‘what works’ in addressing offending behaviour and to share good practice which will improve responses within the context of education, youth work, probation and prison.

Improving Support Provision The final issue identified as a priority area for improvement relates to victim support. Within Studies Three and Four, we found that fewer than 10 per cent of hate crime victims had received support following their experience of victimisation. Despite the limited number of victims who accessed support, their responses highlighted the broad range of organisations who are in a position to provide support to victims, including the police, locally commissioned victim support services, local authorities, housing associations, health and social care organisations, and voluntary community groups. As discussed in Chapter 9, our evidence raises concerns not just in relation to the low numbers of victims who tend to access support services, but also with respect to the low levels of satisfaction with these services. Within our analysis of the failings of existing support provision, it was especially worrying to discover that a significant proportion of hate crime victims were simply unaware of the existence of various support services and vice versa. Victims often only become visible to criminal justice agencies, and by extension support services, through the formal process of reporting which is likely to account for the low uptake of support services for hate crimes. In recent years, public and third sector agencies have devoted a significant amount of time and effort to developing awareness-raising campaigns at a local and national level. However, the findings from each of our studies suggest that these initiatives are failing to reach and to connect with people at a grassroots level, particularly

11  IMPLICATIONS FOR POLICY 

169

those who belong to socially and economically disadvantaged communities. Awareness-raising campaigns need to involve representatives from a diverse range of communities in order to ensure that key messages resonate with specific groups, including those who are wary about the existence of hate crime policy. Moreover, those key messages should be tailored to address the main barriers identified by our research, including improving awareness of what forms hate crimes take; how hate crime can affect victims and their families; what responses victims and witnesses can expect from the police and other relevant organisations; and why it is important to report hate incidents. As illustrated by the methodological approach employed within our research, community representatives can play a key role in this process by identifying appropriate community-based locations to disseminate and publicise awareness-raising material and by working as partners to reinforce and translate messages in a style that carries weight within their community. Our research has demonstrated that there is no ‘one-size fits all’ approach to delivering support to hate crime victims. There are a number of factors which would contribute towards victims feeling more ‘satisfied’ and supported by services. These include sharing information in a timely manner, which ensures the victim is well informed of the criminal justice process, and advice on how to contact specialist organisations to support their recovery. We have found that a combination of contact methods is necessary to meet the diverse needs of hate crime victims: not all victims want to receive support in person, and offering telephone support and information sharing via a letter or email work well to ensure that advice is shared in a way which is accessible. Advocacy also emerged as being viewed in particularly positive terms, especially in cases where victims require emotional support over a longer period. Having a single point of contact or advocate enables the victim’s needs to be identified more readily and supports recovery from the harms caused by hate crime. As with police responses to hate crime, our evidence indicates that those in positions to support victims often possess a poor understanding of hate crime and of hate crime victims. This is likely to impact upon the provider’s ability to appreciate the diverse profile and heightened vulnerability of hate crime victims and to deliver a tailored support package which meets the victim’s needs. It is not only essential that these professionals have awareness of what support services are available within their local area, but also that they play a more prominent role in encouraging

170  S.-J. HARDY AND N. CHAKRABORTI

victims and witnesses to utilise those services through a more meaningful referral process. This could be achieved in a number of ways, including providing the potential service user with a specific name and the contact details of a professional within that organisation; asking the organisation offering the support to contact the victim directly; or arranging and attending the first meeting with the victim. Each of these steps would contribute to the victim feeling that their incident is being taken seriously, and would increase the likelihood of the victim making use of a support service. Additionally, we would recommend that police forces adopt an ‘opt-out’ approach to boost the visibility and uptake of support services, thereby ensuring that every hate crime victim who reports to the police is automatically contacted by a locally commissioned support service unless the victim specifies otherwise. With such low numbers of hate crime victims accessing support services, limited opportunities exist for organisations and commissioning bodies to pilot new interventions or to test ‘what works’. With Ministry of Justice day-to-day spending falling by more than a third in the last ten years, the pressure is building on the UK Government and Police and Crime Commissioners to ensure that resources are used in costeffective ways. This has increased the demand for more reliable and robust forms of evidence and has reinforced the need for criminal justice agencies to work ‘smarter’ by ‘testing’ adapted or new p ­ olicies and practices. We found that hate crime victims are more likely to access emotional and practical support from small, localised community-based groups rather than the more familiar, mainstream organisations. Frustratingly for victims and professionals working within this space, these services have been acutely affected by ongoing government spending cuts with locally run, voluntary support groups for asylum seekers and refugees, women, people with mental ill-health, and lesbian, gay, bisexual and trans communities, to name just some examples, reducing their services or having to stop running altogether. Although the impact will be felt in different ways across different areas, many towns and cities have lost vital services such as mental health organisations and services for asylum seekers. Given the key role that voluntary and community services play in providing support to hate crime victims and to some of the most vulnerable and marginalised members of society, it is imperative that governing bodies provide resource to ensure the continued existence of these organisations and services.

11  IMPLICATIONS FOR POLICY 

171

Conclusion This book has illustrated the need to move beyond the prevailing shallow, repetitive cycle which sees leading political figures simply condemn hate acts as ‘utterly unacceptable’ and as ‘having no place in our society’ (to coin recent phrases used by the then Home Secretary (Home Office 2016) only to issue well-intentioned Action Plans and associated criminal justice strategies which outline actions without detailing the process through which they will be achieved, monitored and evaluated (see also Walters and Brown 2016). This chapter in particular has highlighted a series of failings in relation to dismantling barriers to reporting, facilitating justice through conventional or alternative routes and delivering empathetic, timely and meaningful support for hate crime victims. These problems compound the extensive physical and emotional harms that victims will already have to contend with as part of the process of experiencing hate crime and the sense of alienation that is typically felt within groups who encounter targeted hostility as a routine feature of being ‘different’. In the context of economic, political and social conditions which act as enabling factors for the demonisation of ‘marginal’ communities and with levels of hate crime escalating to increasingly alarming levels both within the UK and beyond, academic evidence has a pivotal role to play in generating improvements to existing policy responses. This has been the overarching goal of our empirical work. Within this chapter, we have set out a set of evidence-based, feasible and potentially transformative recommendations which have the potential to shape new and improved thinking amongst senior figures in political and criminal justice spheres, the public and voluntary sector, and within the domains of activism, campaigning and grassroots community work. If implemented, these recommendations have the capacity to make a sustained difference to the ways in which hate crime victims can receive support, respect and recognition. But equally, a refusal to adapt existing mindsets and policy provision is likely to extend and exacerbate the enforced ‘invisibility’ of countless hate crime victims. The heartbreaking testimonies shared within preceding chapters underline why this must not be allowed to happen.

172  S.-J. HARDY AND N. CHAKRABORTI

References Hall, N. (2013). Hate crime (2nd ed.). London: Routledge. HM Government. (2012). Challenge it, report it, stop it: The government’s plan to tackle hate crime. London: HM Government. HM Government. (2014). Challenge it, report it, stop it: Delivering the government’s hate crime action plan. London: HM Government. HM Government. (2018). Action against hate: The UK Government’s plan for tackling hate crime—Two years on. London: Home Office. Home Office. (2016). Action against hate: The UK Government’s plan for tackling hate crime. London: Home Office. Home Office. (2018). Hate crime, England and Wales, 2017/18. London: Home Office. Walters, M. A., & Brown, R. (2016). Preventing hate crime: Emerging practices and recommendations for the improved management of criminal justice interventions. Brighton: University of Sussex. Walters, M. A., Wiedlitzka, S., & Owusu-Bempah, A. (2017). Hate crime and the legal process: Options for law reform. Brighton: University of Sussex.

Index

A Age, 16, 31, 55, 69, 83, 87, 98, 100, 154, 157 Alternative subcultures, 27, 33, 131, 134, 151, 153, 154 Anti-Muslim, 5, 32, 82 Anti-Semitism, 5, 32, 82 Asylum, 18, 54, 55, 70, 83, 84, 87, 92, 99, 108, 121, 124, 134, 135, 144, 153, 156, 157, 170 Australia, 29 Auto-ethnography, 61 B Bigotry, 13, 150 C Canada, 29, 34 College of Policing, 15, 27, 32, 35, 153, 164

Community, 5, 13, 16, 32, 46–49, 51, 53, 55, 59, 60, 68, 70, 71, 73, 94, 101, 106, 111, 118, 119, 121, 122, 137, 138, 140, 142, 144, 155, 156, 158, 163, 167–171 Community engagement, 20, 47, 48, 71 Community leaders, 46, 62, 69, 137, 155, 158 Copeland, David, 26 Coping strategies, 118, 126 Crime Survey for England and Wales (CSEW), 5, 34, 35, 132, 165 Criminal justice, 3, 7, 12, 15, 26, 28, 29, 31, 34–37, 47, 59, 72, 82, 83, 99, 100, 114, 126, 129, 130, 136–138, 140, 145, 153, 161, 162, 164–171 Criminal Justice Joint Inspection (CJJI), 36 Crown Prosecution Service (CPS), 26, 36, 153

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2020 S.-J. Hardy and N. Chakraborti, Blood, Threats and Fears, Palgrave Hate Studies, https://doi.org/10.1007/978-3-030-31997-7

173

174  Index D Difference, 3, 11, 12, 14, 16, 17, 19, 20, 34, 44, 45, 52, 58, 70, 81, 84, 85, 87, 88, 92, 96, 99, 103, 104, 135, 150, 152, 153, 158 Disability, 15, 16, 27, 29, 31, 32, 35, 36, 47, 55, 83, 86, 104, 135, 137, 156, 163, 166 Disablist hate, 82 Discrimination, 12, 15–19, 33, 84, 117 E Economic and Social Research Council, 44, 54, 56 Education, 7, 12, 18, 20, 84, 104, 117, 138, 139, 154, 167, 168 Empathy, 68, 163, 164 Engagement, 12, 19, 20, 46–49, 51–53, 59, 62, 67, 68, 70, 71, 150, 156–158, 161 Equality and Human Rights Commission (EHRC), 18, 19, 44, 54, 56 Ethnicity, 14, 18, 26, 29–31, 55–57, 69, 84, 98, 135, 154 EU referendum, 4, 5, 18, 28, 59, 106, 152 European Union (EU), 30, 32 Extremism, 3, 5, 6 F Facebook, 31, 51 Faith, 18, 55, 62, 70, 106, 121 Field diary, 46, 49, 52, 60, 61, 68, 69, 71, 73, 75, 95

G Gatekeepers, 46, 48, 53, 62, 71, 72, 76, 130, 155, 158 Gender/gendered, 19, 30–32, 55, 69, 70, 82–84, 87, 92, 93, 98, 99, 103, 109, 134, 154, 157 Government, 18–20, 26, 28, 30, 35, 37, 47, 105, 130, 166–168, 170 Grassroots engagement, 62, 155, 156, 158 Gypsy/Traveller, 19, 87, 92 H Harassment, 15, 17, 26, 27, 58, 74, 89–91, 94, 95, 99, 102, 103, 105, 108, 109, 114, 116, 120, 123, 125, 126, 131, 136, 157, 163 Hard to reach, 20, 43–45, 51, 52, 54, 55, 76, 156, 158 Harm(s), 6, 7, 20, 28, 37, 44, 89, 111, 112, 114, 116, 117, 121, 126, 137, 138, 140, 141, 144, 153, 157, 165, 167, 169, 171 Health, 5, 7, 12, 18, 19, 31, 33, 51, 55, 90, 112–114, 126, 168 Her Majesty’s Inspectorate of Constabularies and Fire and Rescue Service (HMICFRS), 35, 36 Homeless, 18, 27, 33, 53, 70, 87, 105, 122 Homophobic hate, 35, 103 Hostility, 13, 15, 20, 21, 27, 28, 34, 52, 73, 82–84, 86–90, 92–95, 100, 104, 105, 108, 109, 113, 115, 117–119, 150, 152, 153, 157, 163, 166, 171 Housing, 12, 17, 19, 47, 84, 92, 105, 117, 142, 143, 168 Hypervisible, 18

Index

I Identity, 3, 5, 13–17, 19, 27–33, 37, 44, 45, 54, 55, 57, 69, 70, 73, 81–90, 92, 93, 96, 99, 104, 109, 112, 117, 121, 123, 130, 135, 140, 150–154, 156, 157, 161, 166 Immigration, 88, 154 Insider, 68, 70, 76 Intersectional/intersectionality, 17, 83, 89, 149, 154–156 Intolerance, 4, 12, 18, 103, 106 L Lawrence, Stephen, 26, 134 Learning difficulties/disabilities, 49, 51, 55, 83–85, 91, 92, 95, 103, 105, 108, 116, 117, 121, 123, 131, 144, 153, 154 Legislation, 25, 28–31, 129, 138, 140, 165, 166 Lesbian, Gay, Bisexual and Transgender (LGBT), 54, 57, 122, 142, 163 Lifestyle, 5, 83, 85–88, 90, 92, 93, 96, 98, 104, 109, 117, 130, 140, 151, 152, 154, 156 M Macpherson, 26 Marginalisation, 6, 20, 47, 154, 158 Mediation, 140, 145, 167 Mental health/ill-health, 47, 55, 75, 85, 87, 90–93, 102, 104, 105, 108, 113, 114, 116, 117, 120, 121, 123, 126, 135, 137, 145, 152, 154, 170 Message crime, 13, 16, 119 Methodology, 7, 43, 156

  175

Methods, 46, 48, 54, 55, 57, 59, 61, 62, 67, 150, 156–158, 169 Micro-aggressions, 3–5, 7, 13, 15, 33, 74, 151, 152, 156, 157, 162 Misogyny, 27 Muslim, 18, 69, 73, 83, 124 N New Zealand, 29 O Offender(s), 28, 31, 98–100, 102, 103, 131, 132, 137, 138, 140, 167 Office for Democratic Institutions and Human Rights (ODIHR), 14, 28, 30–32, 34, 82 Office for Security and Cooperation in Europe (OSCE), 14, 30, 32, 34, 82 Office for the Police and Crime Commissioner (OPCC), 44, 54–56 Otherness, 12 Other(s), 4, 13–16, 18, 20, 26–34, 37, 44, 52, 53, 55, 70, 82, 83, 85, 88, 90, 92–95, 99, 102, 108, 114, 115, 124, 126, 130, 132, 133, 136, 142, 143, 145, 152–156, 163, 165, 167, 169 Outsider(s), 17, 69, 70, 76, 90, 157 P Physical assault, 17, 94, 95, 126 Physical disability, 83, 84, 90, 91, 95, 102, 104, 117, 121, 123, 131, 154

176  Index Police, 4, 15, 26, 27, 29, 34–36, 73, 87, 119, 130–137, 141, 144, 145, 156, 162–166, 168–170 Policy, 3, 7, 11–14, 16, 25–30, 32–37, 48, 71, 73, 87, 89, 96, 129, 130, 134, 140, 150, 152, 153, 157, 158, 162, 165–167, 169, 171 Power, 13, 20, 46, 68, 70, 76, 140, 150 Prejudice, 12, 13, 15, 16, 20, 21, 27, 28, 30, 37, 52, 73, 82, 87, 117, 121, 137, 145, 150, 157, 168 Prison, 145, 166–168 Property crime, 92, 99, 112–116, 131 Prosecution, 29, 36, 166 R Race, 14, 15, 26, 27, 29–31, 83–86, 137, 153, 154, 156, 163, 166 Racially motivated hate, 49 Racist hate, 5, 35, 83, 153 Recording, 4, 15, 31, 32, 37, 61, 73 Refugees, 5, 54, 55, 70, 83, 84, 87, 92, 99, 153, 170 Rehabilitation, 137, 145, 166 Religion, 14, 15, 27, 29–31, 55, 56, 83–86, 88, 101, 124, 135, 137, 153, 154, 156, 166 Religiously motivated hate, 35 Reporting, 4, 31, 34, 35, 55, 59, 102, 121–123, 130, 131, 133, 136, 137, 145, 161, 163, 165, 168, 171 Resilience, 45, 62, 68, 73, 75, 76, 119, 136 Restorative, 7, 140, 145, 167

S Sample, 44, 47, 55, 58–60, 62, 67, 82, 83, 85, 88, 90–93, 95, 98, 104, 112, 114, 116–118, 121, 123–125, 132, 152 Sentencing, 28, 31, 140 Sexual assault, 93, 115 Sexual orientation, 15, 19, 27, 29, 31, 32, 55, 82–84, 86, 93, 99, 100, 102, 117, 123, 124, 130, 135, 154, 166 Sex workers, 27, 33, 106 Social media, 4, 31, 51, 52 Socio-economic, 17, 18, 104, 105, 155, 157 Subcultural, 84, 85, 87, 90, 92, 93, 100, 102, 108 Support, 7, 20, 27–29, 31–34, 36, 44, 48, 51, 55, 59, 73–75, 94, 109, 114, 119, 122, 126, 129, 130, 135–137, 139, 141–145, 161–165, 167–171 Sweden, 5, 32, 155 T Training, 31, 72, 164 Trans, 17, 19, 26, 32, 55, 73, 86, 108, 122, 152, 153, 166, 170 Transgender, 15, 27, 83, 90, 92, 99, 100, 104, 116, 117, 121, 123, 124, 126, 135, 142, 156, 166 Transphobic hate, 32 Trigger event(s), 109 Twitter, 31, 51, 52 U United States, 29

Index

V Verbal abuse, 17, 49, 74, 85, 90, 91, 95, 96, 108, 114, 116, 123, 125, 131, 136 Victimisation, 3, 12, 14–18, 34, 37, 44, 46, 54, 57–59, 61, 62, 67, 71, 73, 81, 82, 84, 86, 88–98, 100, 102, 104–106, 109, 111– 118, 120–126, 130, 131, 136, 143, 144, 149–158, 163, 168 Victims Code, 28, 36 Victim selection, 16, 149, 154 Victim support, 28, 137, 141, 142, 163, 168

  177

Violence, 3–5, 7, 13, 20, 26, 27, 33, 34, 45, 94, 95, 98, 104, 105, 115, 131, 132, 150, 156 Visibility, 12, 16, 84, 170 Visual impairment, 49, 55 Vulnerability, 3, 5, 17, 18, 21, 45, 70, 83, 88, 89, 92, 96, 104, 108, 109, 116, 126, 150, 154, 157, 164, 169 X Xenophobia, 4, 30, 32