Social Class and Routledge Advances in

1. Family Life and Youth Offending Home Is Where the Hurt Is Raymond Arthur

2. China’s Death Penalty History, Law, and Contemporary Practices Hong Lu and Terance D. Miethe

3. The Politics of Antisocial Behaviour Amoral Panics Stuart Waiton

4. Hooked Drug War Films in Britain, Canada, and the United States Susan C. Boyd

5. The Violence of Incarceration Edited by Phil Scraton and Jude McCulloch

6. Violence, Prejudice and Sexuality Stephen Tomsen

7. Biology and Criminology The Biosocial Synthesis Anthony Walsh

8. Global Gambling Cultural Perspectives on Gambling Organizations Edited by Sytze F. Kingma

9. Social Class and Crime A Biosocial Approach Anthony Walsh Social Class and Crime A Biosocial Approach

Anthony Walsh

New York London First published 2011 by Routledge 270 Madison Avenue, New York, NY 10016

Simultaneously published in the UK by Routledge 2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN

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

This edition published in the Taylor & Francis e-Library, 2010. To purchase your own copy of this or any of Taylor & Francis or Routledge’s collection of thousands of eBooks please go to www.eBookstore.tandf.co.uk.

© 2011 Taylor & Francis

The right of Anthony Walsh to be identified as author of this work has been asserted by him in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.

All rights reserved. No part of this book may be reprinted or reproduced or utilised in any form or by any electronic, mechanical, or other means, now known or hereaf- ter invented, including photocopying and recording, or in any information storage or retrieval system, without permission in writing from the publishers.

Trademark Notice: Product or corporate names may be trademarks or registered trade- marks, and are used only for identification and explanation without intent to infringe.

Library of Congress Cataloging-in-Publication Data Walsh, Anthony, 1941– Social class and crime : a biosocial approach / by Anthony Walsh. — 1st ed. p. cm.—(Routledge advances in criminology ; 9) Includes bibliographical references and index. 1. Criminology. 2. Criminal behavior—Social aspects. 3. Criminal behavior— Genetic aspects. 4. . 5. Social classes. I. Title. HV6025.W3657 2010 364.2'5—dc22 2010008524

ISBN 0-203-84424-6 Master e-book ISBN

ISBN13: 978-0-415-88347-4 (hbk) ISBN13: 978-0-203-84424-3 (ebk) This book is dedicated to my wife, Grace, my parents Lawrence and Winifred, my sons Robert and Michael, stepdaughters Heidi and Kasey, and my grandchildren, Robbie, Ryan, Mikey, Randy, Christopher, Stevie, Ashlyn, and Morgan, and my great-granddaughter Kaelyn.

Contents

Preface ix Acknowledgments xiii

1 The Biosocial Approach 1

2 Genes, Environments and Behavior 13

3 , Crime and Status 26

4 The Neurosciences, Conscience and the Soft-Wired Brain 39

5 Social Class and Criminal Behavior: Myth or Reality? 52

6 The Class-Crime Relationship in Criminological Theories 65

7 Social Class and Socialization 77

8 Poverty, Crime and Developmental Neurobiology 90

9 Social Stratifi cation, the Genome, and Social Structure 103

10 The Nature and Nurture of Intelligence 116

11 Class Mobility: Ascription or Achievement? 128

Epilogue 140

References 143 Index 163

Preface

Social class has been at the forefront of sociological theories of crime and criminal behavior from their inception. Indeed, one would be hard pressed to think of a theory in which class is not explicitly central to the theory (/strain theory) or nipping aggressively at its periphery (social con- trol theory). However, none of these theories engage in a systematic explo- ration of what social class is and how individuals come to be on one rung of the class ladder rather than another. If these questions were to be explored in depth perhaps the class-crime relationship would become clear. That is, rather than simply adding class as a covariate to our causal models as if class itself required no explanation, perhaps the same factors that help to determine a person’s class level also help to determine that person’s risk for committing criminal acts. This book is about the “causal” factors shared by class attainment and criminal activities. Of course, this requires reducing ’s “master variable” to a lower level analysis, which is stoutly resisted by many traditional sociolo- gists, and by extension, criminologists, since the vast majority of them are sociologically trained (Cooper, Walsh, and Ellis 2010). While most crimi- nologists would probably agree in the abstract that criminology is inher- ently interdisciplinary, in practice they tend to ignore any perspective other than the one they were trained in, and with which they have become com- fortable. They are particularly likely to ignore biology because they see it as deep, mysterious, and even racist, sexist, and other infuriating names one person calls another person’s position to silence him or her. There is a growing cadre of criminologists today, however, who argue that the causes of criminal behavior should be sought at many levels as long as each level is part of a coherent and mutually reinforcing whole. Chapter 1 introduces the reader to the biosocial approach, an approach that integrates insights from multiple biological and socio-cultural disci- plines. It is not a biological approach to human behavior because such an approach is not possible. It is called a biosocial approach because of the realization that all human behavior is the result of the complex interac- tion of nature and nurture over the life course. The biosocial approach is x Preface an extremely robust approach. As eminent Canadian philosopher of sci- ence Ian Hacking has characterized it: “‘Biosocial’ is a new word, but its pedigree, although brief, is the best” (2006: 81). The chapter looks at the major assumptions of biosocial theory (there is a human nature, it is both reductionist and holistic, it is developmental, and it eschews ideology), and briefl y introduces the three major divisions of biosocial criminology (genet- ics, evolutionary psychology, and the neurohormonal sciences). Chapter 2 takes a brief look at concepts and methods of behavior genet- ics, molecular , and epigenetics. Few scientists today ask whether genes affect behavior; they ask how they do. Twin and adoption study designs that allow geneticists to separate genetic from environmental affects are explored, and then the genetic infl uences on criminality are examined using the concepts of gene x environment interaction and gene/environment correlation. The changing affects of genes in different environments and on personality and behavior over the lifespan are also explored. I then explore how molecular genetics is having an impact in criminology through studies of various genetic polymorphisms. I end the chapter with a look at epige- netics, a branch of genetics that studies how environmental experiences may be captured by the genome without altering the DNA sequence. The more we know about genetic effects on behavior, the more we know about environmental effects. Chapter 3 examines what evolutionary psychology has to say about sta- tus seeking and criminal behavior. Evolutionary approaches are fundamen- tally environmental in that they describe how environments have shaped the behavior of organisms by as they strategically adapt to their environments. Altruism and empathy are addressed as polar opposites of criminality, as is parenting effort versus mating effort. Traits (aggression, deception, low empathy, etc.) that lead to behaviors we now call criminal evolved to aid mating effort, not to aid criminality, but once in place they are useful in that regard. Similarly, traits that evolved to foster parent- ing effort such as empathy and altruism, also foster prosocial behavior. The relationship between violence, status, and reproductive success is also discussed. Finally, the importance of the family as the nursery of human nature is examined as a species-expected . Chapter 4 looks at the contributions to the neurosciences to understand- ing criminal behavior. It is emphasized that although the brain has many assumptions built into it over eons of evolutionary time (it is no tabula rasa), it physically “captures” the environment as it wires itself in response to it. This experience-dependent wiring is explored in terms of early expo- sure to violence. Conscience is a concept rarely addressed by sociological criminologists, but it has obvious importance in terms of the probability of committing or not committing . Thus I examine the neurobiology of conscience as emotions interact with lessons learned to form it. I conclude the chapter with an examination of the concept of the “craving brain” and a brief look at prefrontal dysfunction theory. Preface xi Chapter 5 looks at whether the class-crime relationship is real or a myth, as some criminologists have maintained (Tittle, Villemez, and Smith 1978). I consider this issue by looking at the strengths and weaknesses of self- report data, at ecological data, and at offi cial police and victimization data and conclude that an inverse class-crime relationship is very real. I then address the issue of white-collar crime and how it coheres with the class- crime relationship. How different white-collar criminals are from street criminals and how biosocial criminologists explain occupational and cor- porate white-collar crimes conclude the chapter. Chapter 6 examines Charles Tittle’s (1983) contention that criminologi- cal theories that posit a negative class-crime relationship do not provide adequate theoretical rationales for doing so, and that they only do so by assuming that lower-class individuals have negative characteristics that lead them to crime. I look at this criticism with respect to the theories Tittle himself examined—anomie strain theory and its extensions, Walter Mill- er’s focal concerns, the Chicago school of ecology, , Marxist/confl ict theories, utilitarian/ theories, and social con- trol. Upon examining them I fi nd that all provide a great deal of theoretical substance based on empirical evidence to posit a negative class-crime rela- tionship. I conclude that Tittle’s objections are an example of the moralistic fallacy, the fallacy of confusing a moral should with an empirical is. Chapter 7 examines the empirical evidence supporting the theoretical position that a negative class-crime relationship should exist. The chapter concentrates on class-based socialization practices and strategies. A brief discussion of the concept of socialization is followed by a look at the role of the family in the socialization process which emphasizes its recipro- cal nature as opposed to the old idea of a top-down parent-to-child pro- cess. I then look at the role of the schools in the socialization process, and fi nally the role of peers during adolescence. I maintain that adolescent antisocial behavior as infl uenced by peers cannot be understood without understanding what is going on in the adolescent body (the pubertal surge of testosterone) and brain (the changing ratio of excitatory to inhibitory neurotransmitters, synaptic pruning, and myelination), both of which are being sculpted into their adult form. Chapter 8 looks at poverty, crime, and developmental neurobiology. I fi rst examine the causes of poverty and then the issue of crime as a cause of poverty rather than vice versa. Although I begin the chapter with an emphasis on agency and choice, I reverse myself when it comes to individu- als who experienced truly horrendous childhood environments. I examine these experiences in terms of their effects on the central and peripheral nervous systems, which typically means a down-regulated neurophysiology that places individuals at high risk for criminal behavior. The experiences examined include maternal substance abuse and their attendant terato- genic effects. Also examined are the effects of lead exposure and of being deprived of the multiple benefi ts of breastfeeding. xii Preface Chapter 9 examines the changing effects of the genome on social strati- fi cation with changing social structures. I fi rst examine the “founders” of stratifi cation theory—Marx and Weber—and how their thinking cohered more with biosocial thinking than the kinds of so prevalent in modern sociology. Then the primordial origins of stratifi cation and the neurohormonal underpinnings of status hierarchies are examined. The effects of social structure on the extent to which the genome is allowed free expression in terms of the genomeÆability and abilityÆstatus rela- tionships as cultures moved from hunter-gatherer to agrarian to modern industrial societies should be of particular interest to sociologists. Finally, I briefl y discuss the concepts of regression to the mean, the Hardy-Weinberg equilibrium, and reduced assortative mating to show that social class can never become genetically fi xed. Chapter 10 explores the nature and nurture of intelligence. I fi rst look at the issues of what intelligence is, how it is measured, and if IQ tests are biased. I then look at what neurobiology and genetics have to tell us about intelligence. After examining the “nature” of intelligence I examine the “nurture” of intelligence via various environmental effects, particularly the Flynn Effect. The Flynn Effect refers to the secular increase in IQ observed in numerous countries over the past 50 or so years that cannot be attributed to genomic changes. I explore various explanations for this phenomenon, which at bottom shows us how genes and environments work in concert to develop intellectual abilities. The chapter concludes with a look at the IQ- crime relationship and how it is explained. Chapter 11 pits the ascription and achievement models of socioeconomic status attainment against one another. We fi nd that intelligence (as mea- sured by IQ) is the major determinate of status achievement (occupational mobility) in modern societies, and that it explains more variance than all background variables (including parental socioeconomic status) combined. Other determinants of social class attainment are temperament and per- sonality. Individuals with irritable, miserable, and angry temperaments are not very likely to be successful in any area of life. The heritable person- ality traits of conscientiousness and agreeableness are particularly useful in achieving occupational success. Across numerous studies in the United States and Great Britain, IQ and conscientiousness appear to be by far the most important variables determining the socioeconomic status individuals will achieve. Acknowledgments

I would fi rst of all like to thank editor Max Novick for his faith in this proj- ect from the beginning. Thanks also for the commitment of his very able assistant Jennifer Morrow. These tireless two kept up a most useful three- way dialog between authors, publisher, and a number of excellent review- ers. The copy editor, Sarah Black, spotted every errant comma, dangling participle, and missing reference and misspelled word in the manuscript, for which we are truly thankful, and our production editor, Ryan Kenney, made sure everything went quickly and smoothly thereafter. Thank you one and all. Thanks to my colleagues, geneticist Greg Hampikian and sociologist Arthur Scarritt, for reading and commenting on relevant chapters. I am also most grateful for the reviewers who spent considerable time providing me with the benefi t of their expertise during the writing/rewriting phase of the book’s production. Their input and encouragement have undoubtedly made the book better than it would otherwise have been. These expert criminologists were Matt DeLisi and one reviewer who wishes to remain anonymous. Most of all I would like to acknowledge the love and support of my most wonderful and drop-dead gorgeous wife, Grace Jean; aka “Grace the face.” Grace’s love and support has sustained me for so long that I cannot imagine life without her; she is a real treasure and the center of my universe.

1 The Biosocial Approach

A new paradigm for behavioral science called the biosocial has landed and is disgorging research, books, and articles in many disciplines at a frenetic pace (Wright and Boisvert 2010). It is strange that it took so long to arrive because it is surely obvious that human are biosocial beings possessing genes, brains, hormones, and an evolutionary history who live in societ- ies that cultivate and reinforce their biology. The eminent philosopher of science Ian Hacking has written that: “’Biosocial is a new word, but its pedigree, although brief, is the best” (2006: 81), but what is it? The bioso- cial approach is one that recognizes that any trait or behavior of any living thing is always the result of biological factors interacting with environmen- tal factors. As such, the approach integrates relevant fi ndings, concepts, principles, methods, and theories from the biological sciences with the fi ndings, concepts, principles, methods, and theories from the social and behavioral sciences. The history of science shows that immature sciences fl ounder around among mounds of undigested data and disconnected theo- ries and are openly hostile to the more fundamental adjacent science. The same history also shows that their greatest leaps forward were made when they jettisoned their hostility to the more advanced science and become integrated with it (Walsh 1997). The information received by students in a typical criminology class today differs little from what students received many decades ago. This is to be expected in the humanities since the lessons they teach us are timeless, but it is stagnation and scandalous in a discipline claiming to be a science. Criminology is in desperate need of help from the biologi- cal sciences because decades of research have not produced an accumu- lation of taken-for-granted knowledge that is the hallmark of a scientifi c discipline. Part of this problem has been that our parent discipline, sociol- ogy, possesses “a conceptual scheme that explicitly den[ies] the claims of other disciplines potentially interested in crime” (Gottfredson and Hirschi 1990: 70). Because of this legacy, most criminologists are poorly trained in biology (Walsh and Ellis 2004). This is a pity, for as Matthew Robinson (2004: ix–x) has opined: “the biological sciences have made more progress in advancing our understanding about behavior in the last 10 years than sociology has made in the past 50 years.” 2 Social Class and Crime Things are changing as more infl uential sociologists bemoan their dis- cipline’s lack of progress. In his 2001 Presidential address to the Ameri- can Sociological Association, Douglas Massey (2002: 1) lamented that: “Sociologists have allowed the fact that we are social beings to obscure the biological foundations upon which our behavior ultimately rests.” Since that time there have been numerous biosocial articles printed in sociology’s major journals such as American Sociological Review, American Journal of Sociology, and Social Forces, even special editions entirely devoted to the approach. Noting the explosive growth of the genomic and neurosci- ences over the last two or three decades, a growing number of criminolo- gists have come to realize that if they are to capture the dynamic nature of criminal behavior they must span multiple levels of analysis and multiple disciplines. I am not averse to the sociological notion that society “prepares the stage for crime,” but I would point out in response that crime is committed by fl esh and blood human beings with brains, genes, hormones, and an evolu- tionary history. We therefore have to get beyond trying to understand the behavior of these human beings as if they were disembodied spirits blown hither and thither by the environment like dead leaves in November. Bio- social criminology recognizes the tremendous role the environment plays in all aspects of human life, but it also recognizes that environments act on diverse human material. There is an old pithy adage that declares: “The heat that melts the butter hardens the egg.” In a nutshell, that is what bio- social criminology is all about: how similar environments have different affects on different people, and vice versa.

THE BIOSOCIAL PERSPECTIVE POSITS A UNIVERSAL HUMAN NATURE

One of the most fundamental philosophical questions of interest to social science is: “What is human nature?” All criminological theories contain an underlying vision of human nature, although these visions differ radi- cally. Different views of human nature logically lead to different theories and different interpretations of the data. Having contradicting views of the nature of a discipline’s fundamental subject matter is rightly seen as an impediment in the more fundamental sciences, which then set about the task of reconciling them. The path of least resistance for criminology in the past was to invoke the notion that humans have no nature—a blank slate view that has largely lost its intellectual cachet outside of postmod- ern and strong social constructionist versions of sociology and cultural anthropology. If we take this “no nature” view, we are arrogantly plac- ing humanity above nature—a pleasant enough thought but a false one. Humans are certainly unique, but we are not uniquely unique. Every ani- mal species is unique in many ways vis-à-vis other species. Indeed, every The Biosocial Approach 3 member of Homo sapiens is in some way unique from every other member given the almost infi nite number of possible permutations of genes and environments that shape each person. However, the uniqueness of each person lies in variation in the component parts of human nature, not in the central tendency of the whole. To describe the nature of anything we list its special features and nomi- nate those that are unique, or quantitatively enhanced, that differentiates it from everything else that is not it. The diffi culty arises because we are try- ing to defi ne our own natures, and because our mental machinery operates so automatically we are often blind to what our nature is. Yet most people lacking the trained incapacity to see the obvious can enumerate many of the features of human nature once its contemplation is required of them. The more radical wing of the political left, because of training and tempera- ment, remains the most insistent that there is no human nature, claiming that it is nothing more than the ensemble of social relations within a given “mode of production.” If this is so, then all we have to do to change human nature and achieve secular salvation is to change the mode of production. The left denies a universal and timeless human nature because the concept militates against the left’s dream of the perfectibility of humanity and soci- ety, but it is this universal human nature that chaffs to break free when megalomaniacs assuming they can mold people to their ideological visions emerge (Pinker 2002). Modern Marxists tend to forget that on several occasions Marx strongly implied the existence of a human nature: “Man is directly a natural being . . . furnished with natural powers of life—he is an active natural being. These forces exist in him as tendencies and abilities—as impulses” (Marx 1978: 115). For Marx the essential being of any species is the activity that distinguishes it from every other species, and humans distinguish themselves from other animals by consciously creating their environment instead of merely submitting to it. This free and creative activity, wrote Marx, is the distinguishing human species being, “man’s spiritual essence, his human essence” (in Sayers 2005: 611). Marx was right, but if we allow this, we must also allow that component design features such as language, large brain size, intelligence, rationality, self-consciousness, foresight, sociality, continuous sexual receptiveness, status striving, moral sensibility, and a myriad other that allow humans to create their environments. The sum of these features is human nature. In Capital Marx wrote: “To know what is useful for a dog, one must study dog nature . . . Applying this to man, [we] must fi rst deal with human nature in general, and then with human nature as modifi ed in each historic epoch” (1967: 609). We may deduce from this passage that Marx is saying that a social theory must necessarily start with a theory of human nature, that there is a universal human nature as well as one modifi ed by a particu- lar mode of production, and that the explication of human behavior must be predicated on an understanding of that nature. 4 Social Class and Crime There is no doubt that human nature is developed or constrained con- tingently, and therefore we should not take human nature as we observe it instantiated in a particular culture as exhausting all that human nature is. It is doubtless true that hunter-gatherers, agrarians, and industrial workers living in vastly different times and in greatly different cultures express their natures differently, but these expressions are only variations on a common theme running through time and place. While this is true, no humans have ever lived in a culture such as the one presented by (2001: 973):

They live in peace and harmony and don’t get in into social confl icts; sex roles are reversed, with women being masculine and aggressive, men being feminine; husbands don’t care if their wives have sex with other men, and women don’t care if their husbands give the bulk of the meat from the hunt to their lovers; they lack envy, jealousy, and avarice; men fi nd older women who are grandmothers to be more sexu- ally attractive than young women; they lack status hierarchies and are perfectly egalitarian; and they channel acts of altruism as much toward other people’s children as their own.

If some anthropologist came out of the jungle with such a description it would really make news, but few people would believe it other than some older cultural determinists and strong social constructionists whose world stopped spinning with the unmasking of much of the foolishness and skull- duggery that came out of the Boas/Benedict/Mead tradition of cultural arbitrariness (see Pinker 2002). Grosvenor (2002: 434) maintains that there is a universal human nature by appealing to the timelessness of the lessons of literature: “It is the exis- tence of perennial traits that enables us to understand, for example, the motivations of characters in the plays of Shakespeare or Sophocles, even though they were written in times radically different from our own.” Gros- venor is saying that if there is no universal human nature underlying cultural variation, then the stories from ancient and distant cultures would mystify us, but they do not. How is it that we can understand Antigone’s struggles against King Creon to secure a decent burial for her brother; how is it that male students smirk lasciviously upon hearing ’s allegory of the Ring of Gyges and asked what would they do if they had such a ring, and how is it that across the ages military leaders and businessmen the world over have found much to learn in Sun-Tzu’s Art of War, written in 6th century BCE China? If we all did not have the same hopes, aspirations, character traits, emotions, feelings, goals, needs, moral strengths, and weakness of humans in all cultures at all times—if culture was a more or less arbitrary selec- tion from a grab-bag of possibilities—we would have nothing in common save the necessity to satisfy our physical needs, nor could we differentiate human from non-human save through morphological details. The Biosocial Approach 5 THE BIOSOCIAL PERSPECTIVE IS BOTH REDUCTIONIST AND HOLISTIC

Biosocial approaches are often met with one of sociology’s favorite boo words—reductionism. Reductionism is nothing more sinister than the pro- cess of examining a complex phenomenon at a more fundamental level. Critics of reductionism tend to fear it because they believe that it privileges biological mechanisms at the expense of more holistic factors. Biosocial sci- ence is not reductionist in this sense because it emphasizes that reduction- ist accounts augment and strengthen, not compete with, holistic accounts. Paradoxically such accounts more forcefully underline the importance of the environment than do strictly environmental accounts. Although bio- social criminology is not reductionist, it does take note of and integrate relevant reductionist information, as all good science must. In one sense, biosocial criminology is the antithesis of reductionism because it recognizes the interplay of genes, brains, hormones, nurture, and the social, cultural, and physical environment context, and how all of this is underlain by the evolutionary history of Homo sapiens. To explain something we have to discover its causes, and to do this we have to look at its constituent parts. Causes are typically not discovered by looking up the hierarchy at things that are larger than that which is to be explained. Criminologists often claim that demographic variables such as class, age, race, or gender explain X percentage of variance in antisocial behavior. While these things are predictors of antisocial behavior, identify- ing which of the categories is more prone to antisocial behavior is a descrip- tive statement, not an explanation. What we would like to know is why, for instance, young males everywhere and always are more prone to antisocial behavior than females and older males. Such explanations require proximal genetic and neurohormonal explanations of age and sex differences, and evolutionary explanations of why they exist in the fi rst place. There are holistic accounts that are more coherent and satisfying than reductionist accounts, and it is true that the interaction of the individual elements of a system (atoms, molecules, cells, people) produce effects not readily predictable a priori from their constituent parts. Propositions about biological entities such as genes, hormones, and neurons do not contain terms that defi ne the most meaningful aspects of the human condition such as love, , and morality. Nevertheless, those biological entities iden- tify and elucidate mechanisms underlying these things and provide evolu- tionary explanations of why they exist in the fi rst place in terms of their adaptive functions. Only when mechanisms are discovered and understood can we begin to reasonably understand them and the emergent phenom- ena they underlie. Even so, we must be careful not to lose meaning as an essential component to understanding behavior by exclusive emphasis on mechanistic accounts. Biosocial scientists would condemn Dennett’s (1995: 82) “greedy reductionists” (persons who skip over several layers of higher 6 Social Class and Crime complexity in a rush to fasten everything securely to a supposedly solid foundation) just as surely as they would naive antireductionists who yearn for what Dennett calls “sky hooks” (metaphysical constructs suspended on nothing that miraculously lift them out of diffi culty). Nonetheless, science has made its greatest strides when it has picked apart wholes to examine the parts to gain a better understanding of the wholes they constitute. As already noted, many social scientists are now following the natu- ral and physical sciences and accepting reductionist and holistic accounts as complementary accounts rather than as zero-sum competitors. Holistic accounts of many phenomena existed in the natural sciences long before their underlying mechanisms were discovered, and they are not aban- doned when reductionist theories arrive as long as they remain consistent with them. Newtonian physics is still useful for building automobiles and bridges, relativity and quantum mechanics notwithstanding. Cell biolo- gists know that at some level they are dealing with atoms, but they also know that there are properties of the cell that cannot be easily deduced from atoms a priori. The cell requires an atomic structure, but cell biolo- gists are less interested in that than they are about discovering functional explanations of the whole cell and how cells fi t together into networks to form the organism. The whole is certainly more than the sum of its parts, but there is no other way to understand the whole or to verify statement about it other than by coming to know the parts. Useful observations and hypotheses go in both reductionist and holistic directions in the hard sci- ences, such as from quarks to the cosmos in physics and from nucleotides to ecological systems in biology. Likewise, criminologists should be just as comfortable with neurons as they are with neighborhoods. As Matt Ridley (2003: 163), the heavyweight champion of nature via nurture, has opined: “Reductionism takes nothing from the whole; it adds new layers of wonder to the experience.”

THE BIOSOCIAL PERSPECTIVE IS DEVELOPMENTAL

Biosocial theory is developmental and dynamic in that it emphasizes that individuals develop along different pathways. As individuals develop, fac- tors that were previously meaningful to them such as acceptance by antiso- cial peers no longer are, and factors that previously meant nothing to them such as marriage and a career suddenly become important. But biosocial theories go beyond merely noting social turning points over the life course; they attempt to explain why they are turning points, why they are mean- ingful at certain life junctures and not at others, and why individuals are differentially affected by them. Sampson and Laub (2005) do not like the term developmental because to them it implies an overly deterministic process similar to a photographic image by which only the process of developing the negative is required to make manifest what is already there. For them developmentalism implies a The Biosocial Approach 7 predetermined latent form slowly unfolding as an internal programming is activated by internal or external environmental factors. The concept that Sampson and Laub appear to have in mind is preformationism, which does indeed imply a latent predetermined form waiting only for developmental process to make it apparent. A preformationist (if such persons still exist) is one who believes that organisms are fully formed miniature versions of what they will eventually become at the moment of their conception. To call someone a “preformationist” today is to substitute an old and gentle term implying only erroneous thought for the more modern and hostile “genetic determinist,” implying malevolence as well as error. The alternate explanation for the development of the organism from embryo to mature form is epigenesis, which posits meandering rather than linear causality and emphasizes the effi cacy of the environment to mold indi- viduals in non-predestined ways. Sampson and Laub conceive of development as lifelong interactions between individuals and their environments peppered with human agency and random noise (pure chance). This is exactly the way biosocial theorists think of development. Biosocial theorists are developmen- tal in the sense that they believe genes set us on particular developmental trajectories, but they are aware that the vagaries of the environment can send those trajectories askew as things frequently zig when they should have zagged and vice versa. Without speaking for other biosocial criminologists, I also strongly support Sampson and Laub’s faith in purposeful human agency within the constraints and opportunities that confront individuals; and who can deny the role of pure chance in anyone’s life? Where Sampson, Laub, and I part company is in their belief that child- hood experiences deserve no special place in the pantheon of causes of later behavior. Development is indeed a lifelong interaction between individuals and their environments, but this does not mean that each environment is virginal. The environment a person is currently in and how he or she will respond to it is conditioned by experiences in previous person-environment interactions. Events experienced in our earliest years when the brain is most pliable, chisel neural networks to respond to subsequent experiences in ways that are not easily altered by later life experiences. Later experiences (Sampson and Laub’s “turning points”) are interpreted differently by dif- ferent neural networks because their infl uence is channeled along pathways laid down during childhood. Within the normal range of environments of rearing, these neuronal pathways will not differ much, but outside of that range truly horrendous environments will result in brains calibrated to expect violence and hostility and to act accordingly.

THE BIOSOCIAL PERSPECTIVE ESCHEWS IDEOLOGY

Arguments about the substance of criminological theory are seldom settled with data. This is because while all sides of the argument are looking at the same data, like the blind men feeling the elephant in the 8 Social Class and Crime ancient Indian parable, they tend to see different things. They see differ- ent things because they each hold what economist/philosopher Thomas Sowell (1987) calls visions. Sowell avers that two contrasting visions in constant confl ict with each other have shaped thoughts about the subject matter of the human sciences throughout history: the constrained and unconstrained visions. The constrained vision views human activities as constrained by an innate human nature that is largely unalterable; the unconstrained vision denies an innate human nature and believes in the perfectibility of humanity. The many differences between the two visions are summed up by the constrained vision’s assertion that “this is how the world is,” and the unconstrained vision’s assertion that “this is how the world should be.” Visions are akin to ideology but are more general and pervasive; they are a way of looking at the world—a general picture of “how things should be” that apply across numerous domains. Each vision has its own peculiar logic to complement the emotional attachment persons have to them. Like ideol- ogy, a commitment to a vision almost guarantees a selective interpretation of evidence rather than an objective and rational evaluation. One’s way of looking at the world intrudes more on behavioral scientists than on natural and physical scientists because the subject matter of the behavioral sciences are suffused with political and moral issues, and there are few places in which this is more true than in criminology, dealing as it does with highly disvalued behavior. One way to overcome visionary parochialism is to force one’s self to feel all parts of the criminological elephant and then, and only then, attempt to describe it. For Sowell, unconstrained visionaries are moralists and idealists who search for solutions and good intentions, and constrained visionaries are pragmatists who look to prudent trade-offs rather than solutions and judge things good or bad by their consequences, regardless of their inten- tions. Adherents of both visions agree about the desirability of such things as equality and justice, but they disagree about what these things mean. They ask different questions about the same issues that both sides agree are problematic and require attention. Because they see human nature as basically good, “believers in the unconstrained vision seek the special causes of war, poverty, and crime”; that is, such things are aberrations to be explained. On the other hand, because constrained visionaries believe that humans have to learn to be good but frequently are not, and believe that these things are historically normal and inevitable, albeit regrettable: “believers in the constrained vision seek the special causes of peace, wealth, or a law-abiding society” (Sowell 1987: 31). Because criminologist differ in their visions, it is not surprising that their theories differ on how they approach social problems such as crime and inequality. A theory of criminal behavior or the origins of social class is at least partly shaped silently by the vision of the person who formu- lated it. Sowell avers that a vision “is what we sense or feel before we have The Biosocial Approach 9 constructed any systematic reasoning that could be called a theory, much less deduced any specifi c consequences as hypotheses to be tested against evidence” (1987:14). Those who are drawn to a particular theory owe most of their attraction to it to the fact that they share the same vision as its originator than to its level of empirical support. Thus “visions” more than hard evidence often lead criminologists to favor one theory over another more strongly than most of them care to acknowledge. For example, Walsh and Ellis (2004) found socio-political ideology to be the major factor in explaining which of 23 theories a sample of pro- fessional criminologists favored. Self-identifi ed conservatives, moderates, liberals and radicals were cross-tabulated with the theory they believed had the most empirical support. The results were highly signifi cant (χ2 = 177.23, P < .00001, Cramer’s V = .65). Conservatives and moderates tended to favor either social- or self-control theories, which exemplify the constrained vision in that they do not ask what causes crime, but rather what prevents most of us from committing it. Liberals and radicals favored theories that ask for the causes of crime, and that attribute those causes to factors such as an unfair economic system. Based on their fi ndings, Walsh and Ellis (2004) characterized ideology as criminology’s “Achilles’ heel” which could be fatal to the discipline’s scientifi c cachet, for surely only one theory can have the “most empirical support.” This study supports sociologist Orlando Patterson’s (1998: ix) complaint that liberals and conservatives “play tiresome and obfuscating games” when they focus on a single area of the causal net. According to Patterson, con- servatives believe that only “the proximate internal cultural and behavioral factors are important (‘So stop whining and pull up your socks, man!’),” and “liberals and mechanistic radicals” believe that “only the proximate and external factors are worth considering (‘Stop blaming the victim, rac- ist’!).” I like to play my cards face-up and therefore confess to being more of a “pull up your socks” conservative than a “Stop blaming the victim” liberal. Admitting one’s biases and striving to confront them will hope- fully produce a more balanced account than pretending that one can be entirely neutral. Thus when I claim that the biosocial perspective eschews ideology I certainly do not mean that those who work within this approach have escaped the clutch of visions. I mean that the approach itself, being interdisciplinary, eschews ideology because it is not committed to a naive either/or zero-sum dichotomy of nature or nurture. Rather it is committed to a strong emphasis on nature via nurture because nature and nurture are inextricably linked cofactors in producing our behavior.

THE THREE MAJOR DIVISIONS OF BIOSOCIAL CRIMINOLOGY

Because the biosocial paradigm is relatively new, it is necessary to devote the next three chapters to introducing its foundational theories, concepts, 10 Social Class and Crime and methodologies before applying them to the issue of social class and criminal behavior. There are three general biosocial approaches to the study of human behavior: the genetic, evolutionary, and neurobiological. Although these approaches employ different theories and methods, work with different units of analysis, and invoke different levels of causation, they are vertically integrated; that is, their principles are conceptually consistent across all three levels of analysis. All three approaches are so environment-friendly that they may be considered biologically informed environmental approaches. That is, they recognize that the environment turns genes on and off, soft-wires the brain, and regulates the secretion patterns of hormones and neurotransmitters. Each of these approaches is examined in separate chapters and only briefl y introduced here. Genetic approaches have been considered a danger to sociology and criminology because for those disciplines genes connote hard determin- ism. The solution to such an erroneous belief is to learn something about genetics, which might be something of a necessity for future generations of social scientists because we are increasingly seeing statements such as: “If sociologists ignore genes, will other academics—and the wider world— ignore sociology? Some in the discipline are telling their peers just that. With study after study fi nding that all sorts of personal characteristics are heritable—along with behaviors shaped by those characteristics—a see-no- gene perspective is obsolete” (Shea 2009: B6). Social scientists have nothing to fear; there are far too few genes for them to exercise control over our behavior. Complex organisms in com- plex environments cannot possibly be hard-wired for every possible situa- tion they could encounter. The almost limitless kinds of situations humans could possibly face would require many millions of genes carrying a series of “if this/then do that” algorithms rather than the approximately 30,000 that we have. We could never possibly be preprogrammed to respond to all the possibilities in fi xed, undeviating ways. Genes simply provide us with some very general rules, such as “avoid snarling creatures,” “favor close kin,” “enjoy sex,” “cooperate with others,” “fear heights,” or even more generally, “seek pleasure and avoid pain,” and then leave the specifi cs to the judgment of the whole organism. Evolutionary is the second important biosocial approach. We are all descendant from ancestors who successfully overcame the adaptive prob- lems of all living thing—how to survive and reproduce. The behavioral propensities that solved these problems are still with us today, and some of them can and are used for purposes other than those for which they were designed. These broad behavioral categories have to be identifi ed and explored in detail, and that is why we need a common framework from which to proceed. For this reason, a growing number of leading social/ behavioral scientists are “gaining enthusiasm for a Darwinian framework, which has the potential to tie together the forest of hypotheses about human behavior now out there” (de Waal 2002: 187). The Biosocial Approach 11 Writing about criminal behavior and socioeconomic status necessitates an emphasis on the negative side of human nature because a criminologist’s stock in trade is vice, not virtue. At the same time I am well aware that in social species is more about cooperation than confl ict; in The Decent of Man, mentioned cooperation three times more than he did competition (Levine 2006). But social cooperation is contingent on reciprocity, a tit-for-tat strategy favored by natural selection because of the benefi ts it confers. While individual are not adapted to behave “for the good of the group,” acting as though they were enhances their survival and reproductive success as well as that of those who share the coopera- tive enterprise. Cooperation confers benefi ts on all cooperators; we feel good when we cooperate, and cooperating identifi es us as reliable, altruis- tic, and trustworthy, which by defi nition are traits conspicuously absent in criminals. The third biosocial approach examines the neurological and hormonal basis of behavior. Regardless of the evolutionary origins or of the genetic and environmental instigation to behave one way or another, everything has to fi rst be channeled through the brain, the magnum opus of natural selection. As the executor of all that we do and think, the brain and its processes must be a vital part of the criminologist’s repertoire of knowledge. We can no longer afford to view the brain as a mysterious and foreboding “black box” that can be safely ignored as irrelevant to our work. Although we need not concern ourselves with the minutia of brain anatomy and physiology any- more than with the minutia of genetics or , we should learn neuroscience’s basic language so that we understand the relevance of its fi ndings for criminology. Examining the structure and function of the brain will fi ll in so many blanks and prevent criminologist from inventing improbable scenarios to account for traits that are risk factors for crime such as self-control, sensa- tion seeking, fearlessness, intelligence, aggression, status seeking, and many others. All of these variables are heritable and have identifi ed neuroana- tomical and chemical correlates. Again, this does not lead to a dismissal of environmental experiences related to these variables. We will see that these environmental experiences are physically captured in the brain’s circuitry, and that to a great extent the experiences one has literally wires the brain accordingly. We are designed to extract information from the environment into our genomes and brains: “The extended period of neuroplasticity is an aspect of human nature that allows and requires environmental input for normal human development” (Wexler 2006: 16).

CONCLUSION

Karl Marx wrote long ago that “Natural science will in time subsume under itself the science of man, just as the science of man will subsume under itself 12 Social Class and Crime natural science: there will be one science” (1978: 91). Biosocial criminologists agree with Marx that rather than viewing biology as a threat to our discipline we should welcome it as an opportunity to collaborate with a very robust ally. The biological sciences have a bounty of treasures to offer criminology, and criminology should seek theoretical integration with them. The social and biological sciences need each other and belong together. As the history of the physical and natural sciences demonstrates, cross-fertilization of concepts, methods, and theories breeds hybrid vigor into disciplinary domains. I believe that a paradigm shift is occurring in criminology, moving it away from its strict environmentalism imposed on it by its sociological progenitor. Francis T. Cullen, one of the leading luminaries of sociology and criminology/ is also “persuaded that sociological crimi- nology has exhausted itself as a guide for the future study on the origins of crime. It is a paradigm for the previous century, not the current one” (2009: xvi). He believes the paradigm that will take the place of sociology for 21st- century criminology will be biosocial, which he calls “a broader and more powerful paradigm” (2009: xvii). This shift, while intellectually exciting, is extremely scary for some because it necessitates the learning of some genetics, neuroscience, and evolutionary biology. A recent survey of 770 American criminologists (Cooper et al. 2010) found that the modal category of biology classes taken by these criminolo- gists was a fat zero. On the bright side, however, the same survey found that 46.4% agreed or strongly agreed that “biology has a lot to offer criminol- ogy,” as opposed to 30.9% who either disagreed or strongly disagreed. We cannot, of course, expect to be experts in these biological sciences, or indeed in any one of them. We do have to be well acquainted with them, however. This is really no different than having to learn statistics well enough to use it intelligently when doing our own research or evaluating the research of others. Although some would rather continue to rail against biology than learn some, Thomas Kuhn (he of “paradigm shift” fame) inti- mates, those who fail to do so will fi nd themselves irrelevant: “retooling is an extravagance reserved for the occasion that demands it,” and the wise scientist knows when “the occasion for retooling has arrived” (1970: 76). Kuhn goes on to describe what we will fi nd in the new paradigm:

Led by a new paradigm, scientists adopt new instruments and look in new places. Even more important, during revolutions scientists see new and different things when looking with familiar instruments in places they have looked before. It is rather as if the professional community has been suddenly transported to another planet where familiar objects are seen in a different light and are joined by unfamiliar ones as well. (1970: 111)

I hope that examining the question of socioeconomic status and criminal behavior from a biosocial perspective will at least provide a taste of what Kuhn promises. 2 Genes, Environments and Behavior

WHAT ARE GENES?

A simple defi nition of a gene is that it is a segment of DNA that codes for the amino acid sequence of a protein (enzymes, neurotransmitters, hor- mones, or cell-structure proteins). These proteins do not cause us to behave or feel one way or another; they facilitate our behavior and our feelings and modulate how we respond to the environment. Genes are not automatons pulling strings in our heads and determining the directions of our lives; rather they help us to get there once we decide on the direction, although they do bias us to prefer one direction over the other. They are part of an organism’s machinery for responding and adapting to its environment composed of the sensory, evaluative, production, and motor systems. The afferent nerves sense and transmit information about the state of a person’s internal or external environment to the central nervous system (CNS) to be evaluated. Genes then respond to the information evaluated by the brain by manufacturing the appropriate proteins. Efferent nerves then carry impulses away from the CNS to the motor systems (muscles or glands) that initiate the organism’s overt reactions. There are two elements controlling gene functioningy—transcription factors and promoters (Orians 2008). Transcription factors control the transfer (transcription) of genetic information from DNA to messenger RNA (mRNA) which then takes the information into the cell’s protein manufacturing factory outside of its nucleus. Less than 2% of human DNA codes for proteins (the exon segments). Promoter factors are stretches of DNA that regulate gene expression. The interesting thing about promot- ers for behavioral scientists is that differences among species are not about what proteins are manufactured in their cells (transcription factors) but rather about genetic regulation (promoter factors). If we were to lump all genes from all primate species together such that the pool contained 100% of the genes contained in the entire primate order, we would fi nd that all species shared an average of about 98% of their genes with all others (Yoav et al. 2003). Yet the gulf between humans and our closest genetic relatives, the chimpanzees, is enormously greater than the 2% (or less) genetic dif- ference implies. The different morphological and behavioral characteris- tics of humans and chimps are based on quantitative (the amount of gene 14 Social Class and Crime products) rather than qualitative (different gene products) genetic differ- ences. All humans and chimps have hair and brains, but chimps have more hair and humans have more brains. All these differences have to do with where and when in the genome the regulatory genes are turned on and off (Orians 2008).

THE CONCEPT OF

To disentangle genetic from environmental effects on phenotypic traits we compute a heritability coeffi cient (h2), which is a quantifi cation ranging between 0.0 and 1.0 of the proportion of variance in a trait in a population attributable to genes. It is an estimate that varies among populations and within the same population as it experiences different environments. All cognitive, behavioral, and personality traits have been shown to be herita- ble (Rutter 2007), which led Turkheimer and Gottesman (1991) to suggest that h2 ≠ 0.0 be enshrined as the fi rst law of behavior genetics. Heritability coeffi cients are typically obtained by comparing inter-class correlations for a trait between monozygotic (MZ)twin pairs and same-sex dizygotic (DZ) twin pairs. MZ twins are genetically identical; they share 100% of their genes, and DZ twins share 50% of their genes on average. If genes are a source of trait variation, the more individuals are geneti- cally related the more similar they should be on any given trait. If this were not so, it would be logically impossible to calculate meaningful heri- tability coeffi cients. Heritability does not tell us how much of the trait itself is genetic; it only tells us how much of the variance in the trait is accounted for by genes at this time and in this population. Genetic potential can only be fully real- ized in environments that are conducive to its realization, so h2 should not be read as placing constraints on environmental affects or as setting limits on creating new environments that could infl uence the trait. Heritability tells us only about actualized genetic potential, and nothing about what the un-actualized potential may be (Bronfenbrenner & Ceci 1994). Even if h2 is 1.0, it does not mean that the trait in question is beyond the power of the environment to change. Genes cannot be expressed at all without an environment, so it would be absurd to read “h2 = 1.0” to mean that the environment has no effect. High heritability tells us that the pres- ent environment at the present time accounts for very little variance in the trait, it does not tell what other environments may affect variance in the trait. Likewise, h2 = 0.0 does not mean that genes have nothing to do with the phenotype because there would be no phenotype without the genes that gave it life and form; h2 = 0.0 only means that genes did not contribute to phenotypic variation. The environmental variance in human traits is decomposed into shared and non-shared effects. Shared environment refers to the environment Genes, Environments and Behavior 15 experienced by children reared in the same family and include such vari- ables as parental socioeconomic status (SES), religion, values and attitudes, parenting style, family size, intactness of home, and neighborhood. Non- shared environment can be familial or extra-familial. Familial non-shared variables include gender, birth order, perinatal trauma, illness, and parental favoritism. Extra-familial non-shared factors include experiences unique to the individual such as having different peer groups and teachers. Behavior genetic studies do not typically endeavor to determine what those factors are; they leave that to the social scientists. Because variation can only be attributed to sources that vary, it follows that the more environments are equalized (the less variability there is within them), the more trait variability will be a function of genes and h2 will be larger. Conversely, the more variable the environment, the lower h2 will be. A high heritability coeffi cient for a given trait such as IQ indicates that a society is doing a good job with respect to equalizing the environment relevant to the expression of that trait. There is no such thing as a single heritability estimate for any trait. Much depends on the population being studied and when it is being studied, the age and gender of the subjects, and the way in which the traits of interest and their covariates are measured. Because SES constitutes an environment, it is important to explore shifting across different SES popula- tions. For instance, the heritability of IQ is always higher in advantaged environments (≈ .70) than in disadvantaged environments (≈ .25) (Rowe, Jacobson, and Van den Oord 1999; Turkheimer et al. 2003). Similar results are found with respect to the heritability of antisocial behavior (Bartels and Hudziak, 2007; Rowe, Almeida, and Jacobson 1999). Because higher heritability for antisocial behavior are found in advan- taged environments, it should not be assumed that individual living there possess stronger genetic dispositions toward antisocial behavior on aver- age than those in disadvantaged environments. Neighborhood crime rates refl ect mean effects of a variety of risk factors in those neighborhoods but say nothing about their variance. Although the mean level for genetically infl uenced risk factors will be low in advantaged environments, individuals in them who commit crimes have to overcome strong crime-constraining forces to do so, which means they probably possess a greater genetic liabil- ity for such behavior than most others in their environment. The relation- ship between genetics and antisocial behavior will be less in disadvantaged environments because the strong environmental push toward antisocial behavior will mask genetic contributions, even though there are doubtless a greater percentage of individuals with a stronger genetic load for anti- social behavior in those environments than in advantaged environments. Antisocial behavior is like a weed, and intelligence is like a fl ower. Flowers are delicate and need cultivating to bloom, but weeds are robust and will proliferate wildly if left alone. Parents in advantaged environments typi- cally endeavor to cultivate the fl ower of intelligence and uproot weeds of 16 Social Class and Crime antisocial behavior. Parents in disadvantaged families may be unwilling or unable to cultivate the intellectual potential of their children, and thus it does not fl ower; they do little to monitor their children, so antisocial behavior grows like weeds between the cracks. This observation leads us to a discussion of how genes and environments are related.

GENE-ENVIRONMENT INTERACTION AND CORRELATION

Individuals create micro-environments attuned to their genetic proclivities via constant interactions of genes and environment through the processes of gene-environment interaction (GxE) and gene-environment correlation (rGE). The concepts of GxE and rGE not only tell us something about the action of genes, they also benefi t our understanding of the role of the envi- ronment in shaping behavior. As Baker, Bezdjian, and Raine (2006: 44) put it: “the more we know about genetics of behavior, the more important the environment appears to be.” GxE interaction involves different phenotypes responding to environ- ments in different ways. The concept is aptly captured by the saying “the heat that melts the butter hardens the egg.” That is, the same “heat” (envi- ronment) that provides a fearless and impulsive child (the “egg” genotype) opportunities for antisocial behavior does not for a more fearful and less impulsive child (the “butter” genotype). The infl uence of an environment on the person depends on the person’s phenotype at the time, and the per- son’s current phenotype is the cumulative result of numerous previous interactions of his or her phenotype with the environment (Turkheimer and Waldron 2000). Although the genotype underlies and guides phenotypic development, once on a particular trajectory, it is genes + experience (i.e., the phenotype) that interact with the environment. The concept of rGE avers that genotypes and the environments they encounter are not random with respect to one another and that genetic factors infl uence complex psychosocial traits by infl uencing the range of individuals’ effective experiences (Moffi tt 2005). The concept thus enables us to conceptualize the indirect way that genes help to determine what aspects of the environment will be salient and rewarding to us. There are three types of rGE: passive, evocative, and active. Passive rGE refers to the association between the individual’s genotype and the environment provided for him or her. Infants are provided by their biological parents with genes that facilitate certain traits and an environ- ment favorable for their expression—genotype and its environment are positively correlated. A child born to intellectually gifted parents is likely to receive genes conducive to above average intelligence and an environment in which intellectual behavior is modeled and reinforced, setting the child on a trajectory independent (passively) of his or her actions. Genes, Environments and Behavior 17 Evocative rGE refers to the way others in the social environment react to the individual on the basis of his or her behavior. Children bring traits with them that increase or decrease the probability of evoking certain kinds of responses from others. A pleasant, well-behaved child evokes different reactions from others than a moody and bad-tempered child. Some children may be so resistant to parental control that parents either resort to extreme forms of or give up on the child altogether. Either punitive or permissive responses will serve to exacerbate the child’s already antisocial personality and drive him or her to seek social environments in which such behavior is accepted, which are typically populated by similarly disposed individuals. Evocative rGE thus serves to magnify differences among phe- notypes (Dickens and Flynn 2001). Active rGE refers to people actively seeking of environments compatible with their genetic dispositions. Within the range of cultural possibilities and constraints, our genes help to determine what features of the environ- ment will be rewarding to us. The reactions we evoke from others will help us to determine the environments we will seek. Thus evocative and active rGE feed back on one another. Active rGE gains momentum as individuals acquire the ability to take greater control of their lives. The power of genes on forming these environments can be gauged by studies showing that the intelligence, personalities, and attitudes of MZ twins are essentially unaf- fected by whether or not they were reared together. That is, MZ twins reared apart construct their environments about as similarly as they would have had they been reared together and considerably more similarly than DZ twins reared together (Plomin 2005).

BEHAVIOR GENETICS AND CRIMINAL BEHAVIOR

Meta-analyses and reviews of the heritability of antisocial behavior pro- vide a broad estimate of 50% heritability, 20% shared environment, and 30% non-shared environment (Guo, Roettger, and Shih 2007; Miles and Carey 1997; Moffi tt 2005; Rhee and Waldman 2002). Heritability is higher among adult offenders than among juvenile delinquents because of the high base rate among juveniles. Criminality is a quantitative variable that is itself an amalgam of other quantitative variables such as low self-control, low empathy, sensation seeking, and many others traits that make a person less than desirable as a friend, mate, or employee. Quantitative traits always require the shared operation of more genes than do qualitative traits, thus h2s computed for any quantitative trait such as criminality are actually capturing a wide variety of correlated sub-traits. Geneticists call these sub-traits endophenotypes which means “within the phenotype.” Endophenotypes are essential parts of the chain leading from genotype to phenotype, with each having only minor effects on the 18 Social Class and Crime phenotype, and then only if combined with other relevant endophenotypes and the right environments. The point is, when it is time to appeal to molec- ular genetics, reducing the criminality phenotype to the constituent and less complex endophenotypes such as impulsiveness or low empathy, it will be easier to identify and study their genetic basis than if we tried to map the more elusive and amorphous phenotype itself (Glahn, Thompson, and Blangero 2002). This is because intermediate traits (the endophenotypes) “sit closer to the genotype in the development scheme” (Gottesman and Hanson 2005: 268). Studies of rGE and antisocial behavior have focused on parenting prac- tices evoked by the behavior of adopted children (evocative rGE). It is invariably found that adopted children at genetic risk for antisocial behav- ior (as assessed by antisocial behavior of their biological parents) consis- tently receive more negative parenting from their adoptive parents than do children not at genetic risk. In each case negative parenting was seen as parental reaction to the behavior of their adopted children (Moffi tt 2005; O’Connor et al. 1998; Riggins-Caspers et al. 2003). GxE studies typically examine the effects of aversive home environ- ments (marital discord, divorce/separation, abuse/neglect, and so forth) on adopted children who are or are not at genetic risk for antisocial behavior as indexed by antisocial behavior of birth parent or parents (Cadoret et al. 1995; Riggins-Caspers et al. 2003). Adverse home environments lead to signifi cant increases in antisocial behavior for adoptees at genetic risk, but not for adoptees who were not at risk. In these studies, neither genetic risk nor aversive home environment was powerful enough to produce antisocial behavior independently, but genes and environment operating in tandem (interacting) produced signifi cant antisocial behavior. That is, children genetically at risk for antisocial behavior reared in positive family environ- ments did not display antisocial behavior, and children not at genetic risk did not become antisocial in adverse family environments. Because genes affect differential exposure to environmental risks via active rGE and differential susceptibility to environmental risks via GxE, both processes are always operating and diffi cult to untangle. In other words, people self-select themselves into different environments on the basis of their genetic preferences when they are able (rGE), and because they self-select a particular environment, they will be more susceptible to its infl uence (GxE) than will those there by happenstance. Once in contact with an environment inhabited by antisocial others, the environment may have unique causal effects on those in them.

ASSORTATIVE MATING FOR ANTISOCIAL BEHAVIOR

Assortative mating is non-random mating based on couple resemblance for one or more characteristics. Assortative mating for antisocial behavior is Genes, Environments and Behavior 19 positive, which is troubling in terms of increasing the probability of inter- generational transmission of criminality. A British sample found a correla- tion of .50 between husbands’ and wives’ criminal convictions (Rowe and Farrington 1997); a New Zealand study found an average correlation of .54 between mates for a variety of antisocial attitudes and traits (Krueger et al. 1998); and an American study found positive correlations between part- ners/mates for general and serious offending, delinquent peers and beliefs, and problem drinking (Knight 2008). All three of these studies were based on longitudinal cohort data. Krueger and colleagues interpret positive assortative mating for antiso- cial behavior as creating “criminal families” via active rGE in the parental generation and passive rGE in the offspring generation. Thus offspring with two antisocial parents receive both paternal and maternal genes for traits that increase the probability of antisocial behavior and a home environ- ment modeling such behavior, so nature and nurture “becomes a tightly tied bundle” (Krueger et al. 1998: 183). Similarly, Jaffee and her colleagues (2003: 120) write that when an antisocial father resides with mother and offspring “children experience a double whammy of risk for antisocial behavior. They are at genetic risk because antisocial behavior is highly her- itable. In addition, the same parents who transmit genes also provide the child’s environment.”

BEHAVIOR GENETIC CAVEATS

Behavior genetics studies show that shared environmental effects on cog- nitive and personality traits in childhood disappear almost completely in adulthood. This should not be taken to mean that parents have no effect on children apart from the genes they provide them. Behavior genetic stud- ies only show that parental effects on personality and cognitive traits that made siblings somewhat similar while they shared a home fail to survive the period of common rearing, and this is decidedly not the same as saying parents don’t matter. Parental effects on their children’s attitudes, values, behavior, and choice of leisure activities and professions do not disappear, and the expression of many heritable traits often depends to various degrees on parental factors. The weak effects of shared environments on older sub- jects means only that shared environments have few similar effects on sib- lings for a number of personality and cognitive traits; it does not mean they have had no effect. Carey (2003) offers an example of how family affects sibling’s traits without making them similar on those traits by looking at the effects of an aggressively dominant father on two of his sons. One of the sons shared his father’s personality, which resulted in frequent confrontations between father and son. Because of this the son became more confron- tational than he might otherwise have been. The other son had a more 20 Social Class and Crime submissive personality and always acceded to his father’s demands, and as a result becomes more submissive than he would otherwise have been. This example of GxE interaction magnifi ed the difference between the two sons with respect to dominance and submissiveness more than their genetic dif- ferences predict they should have been. This effect is called “non-shared” because only effects that make siblings similar are called shared, but it is nevertheless a family effect. The non-shared features of the environment are more salient with respect to the development of an individual’s personality and cognitive traits than shared features, but parents have a huge say in what those non-shared envi- ronments are. Parental SES strongly determines the neighborhoods their children grow up in, and thus the kinds of peers they will be exposed to, which is the major source of non-shared effects. Parental monitoring of their children’s friendship patterns, and patient and caring supervision of their activities, can go a long way in assuring they take the straight and nar- row. The lower h2s found in lower SES environments refl ect family effects, and those effects are large. Families matter, and they matter more than anything else in a child’s life. Note also that heritability can remain stable even while the mean level of a variable is rising or falling (Maccoby 2000). In other words, if some environmental factor raises or lowers all (or almost all) scores on a trait to a similar degree thus altering the trait mean, it does not necessarily effect the trait variance, the basis for computing heritability. For example, a recent study found that individual differences in executive functions (self-control being one such function) are almost entirely genetic (Friedman et al. 2008). This does not mean that executive functioning per se is almost entirely genetic, that it is impervious to targeted training, or that parenting does not matter. Executive functioning, or most anything else, can be strongly related to parental training efforts and thus affect the population mean, but individual differences (i.e., population variance) can still be almost entirely genetic. In other words, if parental socialization factors raise or lower all scores on a trait to a similar degree thus altering the trait mean, it does not affect the trait variance, the basis for computing heritability.

MOLECULAR GENETICS

Heritability studies show only that “something genetic” is involved; they do not tell us what genes are involved. Fortunately, we are now able to go beyond heritability and into the world of molecular genetics. At the molec- ular genetic level, quantitative trait differences are the result of quantitative differences in gene products due to genetic polymorphisms. refers to the differences in allelic combinations (an allele is an alternate form of a gene; one inherited maternally and one paternally) that make us different from one another even though we share the same genes. That is, Genes, Environments and Behavior 21 all humans have all the same genes for species-defi ning physical and mental structures that are coded for in the species-wide pool of DNA, but they are not all heritable. Intelligence is 100% transmitted by genes (inherited), but its expression depends on both genes and environment, so heritability can never be 100% in the same sense that inherited is. The traits with zero heritability are those which are most genetically determined. There are no heritability estimates for the presence or absence of human intelligence, aggression, altruism, noses, sex organs, and so on. These are universal human traits that function the same way for us all. It is the variance in the quantitative properties of phenotypic traits that makes us all different. Polymorphisms are thus variations on a common theme. There are three different types of polymorphisms: single nucleotide polymorphisms (SNPs), minisatellites and microsatellites (Altukhov and Salmenkova 2001). SNPs are the most frequently occurring (about 90%), and while most are nonfunc- tional, it has been estimated that about 85% of the genetic causes of most behavioral and psychological disorders are attributable to SNPs (Plomin et al. 2001). As the term implies, a difference in just one nucleotide is all that differentiates one allele from another. A nucleotide is one of the four nitrog- enous bases (adenine, cytosine, guanine, thymine or uracil) plus a sugar and phosphate backbone that make up DNA and RNA. When we consider that the human genome contains approximately 3 billion base pairs, we can appreciate the potential number of polymorphisms it contains. Microsatellites and minisatellites are sections of DNA that are repeated a different number of times rather than having a nucleotide substitution. An example of interest to criminologists is the dopamine receptor gene (DRD4) that has a 48 base pair sequence repeated from 2 to 11 times (Ding et al. 2002). The shorter the repeat, the more responsive the brain is to dopamine; the longer the repeat, the less responsive it is. Individuals with two or three repeats tend to be over-stimulated by events that most people (people with the four-repeat version) fi nd optimally stimulating and seek to withdraw from them or tone them down. Individuals with the 7-repeat allele are sub-optimally aroused by the same events. Sub-optimal arousal is experienced as boredom, and bored people seek to raise the level of stimuli to alleviate it, which sometimes results in antisocial behavior. The 7-repeat allele has been associated with many phenotypic traits linked to antisocial behavior such as ADHD, sensation seeking, and impulsivity (Canli 2006). Geneticists search for candidate genes that are may be associated with a quantitative trait via quantitative trait loci (QTL) mapping. QTLs are regions containing stretches of DNA linked to targeted quantitative traits on one or more chromosomes (Plomin, 2005). QTLs are detected using either linkage or association analysis. Linkage analysis is a “top-down” approach that compares groups with and without a given disease or prob- lem and tries to establish a link between it and specifi c polymorphisms. Association studies use a “bottom-up” approach that begins with a candi- date polymorphism thought to be linked to a trait and tries to establish a 22 Social Class and Crime signifi cant association between the trait and the candidate gene. When we have discovered QTLs related to a phenotypic trait, we are obviously still a long way from understanding the route from genes to complex behaviors— the QLT being just another piece in the puzzle that we hope one day to assemble into a picture we can recognize.

GENE X ENVIRONMENT INTERACTION REDUX

In this postgenomic era there have been a number of studies in which both specifi c genes and environments are identifi ed in GxE. One of the best of these studies of interest to criminologists is the longitudinal cohort study of Caspi and his colleagues (2002). The environmental risk in this study was verifi ed child maltreatment, and the identifi ed genetic risk was the mono- amine oxidase A (MAOA) polymorphism. The dependent variables were a variety of antisocial outcomes; the most serious being a conviction for a violent crime by the age of 26. All self-reported measures were verifi ed via offi cial police and social service records. MAOA is an enzyme that metabo- lizes a variety of neurotransmitters and comes in long (high transcriptional activity) and short (low transcriptional activity) alleles. No main effects for MAOA were found, but there were strong GxE effects. The low MAOA/ maltreatment males were only 12% of the cohort but were responsible for 44% of its violent convictions; the odds of such males having a violent crime conviction being 9.8 times greater than the odds of the high MAOA/ non-maltreated males having one. The maltreatment/non-maltreatment groups did not differ signifi cantly on MAOA activity, thus ruling out low MAOA activity contributing to child maltreatment via children’s evocative (rGE) behavior. A meta-analysis of this body of research concluded that the interaction between MAOA and maltreatment is a statistically signifi cant predictor of antisocial behavior across studies (Kim-Cohen et al. 2006). However, Widom and Brzustowicz (2006) found that while the high activity MAOA allele buffered whites from the effects of childhood abuse and neglect on antisocial behavior later in life, it did not protect non-whites. The authors suggest that other environmental stressors may have negated the protective power of the genotype among non-whites, which is another indication of the inherent diffi culties of nailing down human behavior. Genes interact with one another also, and some of these interactions (GxG) are related to some forms of antisocial behavior. Carrasco et al. (2006) examined the effects of the DRD4 and DAT1 (dopamine trans- porter gene) polymorphisms on ADHD and found no independent effects. However, individuals who possessed both the 7-repeat allele of the DRD4 and the 10-repeat allele of the DAT1 were signifi cantly more likely to be diagnosed with ADHD (odds ratio = 12.7) than subjects possessing nei- ther or only one of these alleles. Similarly, Beaver et al. (2007) found no Genes, Environments and Behavior 23 signifi cant main effects for either the DRD2 or DRD4 polymorphisms on conduct disorder or antisocial behavior, but the GxG interaction produced signifi cant effects on both outcome variables.

EPIGENETICS: HOW THE ENVIRONMENT GETS INTO YOUR GENOME

Epigenetics means “any process that alters gene activity without chang- ing the DNA sequence” (Weinhold 2006: 163). Epigenetic modifi cations of DNA make the code more readily accessible or inaccessible, and thus they infl uence the reaction range of a gene (Gottleib 2007). DNA itself only spec- ifi es for transcription into mRNA which leaves the cell nucleus and moves into the cell’s protein factory where it is translated and assembled. Epige- netic modifi cations mean that genes are permanently or semi-permanently switched on and off by signals from the internal chemical environment and/ or by its external physical and social environment. The epigenetics of gene expression emphasize the interplay of nature, nurture, and pure chance in making us what we are, and may have as much or more infl uence on the development of individual differences than differences in DNA polymor- phisms (Kramer 2005). Epigenetic regulation is accomplished a number of processes, the most common being DNA methylation. By the attachment of a methyl group of atoms to a cytosine base, methylation prevents the translation of DNA into mRNA; hence the protein the gene codes for is not manufactured which, depending on what the protein does, is good or bad for the organism. Another process is histone acetylation, which loosens the DNA wrapped around the histones (the protein core around which the DNA is wrapped), which increases the likelihood of genetic expression (Lopez-Rangel and Lewis 2006). Both processes occur in response to internal and external signals (or spontaneously), and the resulting regulatory alterations are often heritable; thus changes occurring in one generation are passed onto the next without altering DNA sequences (Lopez-Rangel and Lewis 2006). Epigenetics heralds the idea of genomic plasticity calibrated to environ- mental events by providing the software by which it can assimilate those events without changing the DNA hardware (Pigliucci, Murren, and Schlichting 2006). Epigenetics provides insight to the elusive processes of non-shared envi- ronment that make even MZ twins reared together sometimes dramatically different. For instance, an MZ twin with a co-twin with schizophrenia will have about a 50% probability of developing schizophrenia compared with a 1% probability in the general population. While this indicates a large genetic effect, the concordance rate is lower than expected given that MZ twins age genetically identical. Looking for specifi c genes that differ- entiate people with psychosis has not been productive, and the search for 24 Social Class and Crime environmental effects has been even less so (Crow 2007). A study of healthy MZ twins reared together found that while they are epigenetically indis- tinguishable in early life, they diverge considerably as they got older. Fifty- year-old twins averaged four times the epigenetic differences as 3-year-old twins (Fraga et al. 2005). Because these twins were reared together and thus shared 100% of their genes and 100% of their rearing environment, epigenetic modifi cation have to be attributed to non-shared environmental (internal or external) events or to stochastic events. Some of the most spectacular epigenetic fi ndings have involved nurturing studies. Nurturing is critical for the healthy development of children, and thus the epigenetics of nurturing is probably the research that will prove most salient to criminologists (Rutter 2007). Infant nurturing produces more than warm cuddly feelings; it changes the brain wiring and may also permanently alter the genome. The work in this area has so far been lim- ited to animal studies, but nurturing behavior doubtless has similar effects on human infants; once a biological mechanism has been demonstrated in one mammalian species, it is almost always found to be applicable to others (Ridley 2003). While it is true that we are probably decades away from discovering the range of human cognitive, personality, and behav- ioral traits subject to epigenetic alteration, some epigeneticists are making statements suggesting that the fi eld may have profound meaning for human development. Epigeneticist Michael Meaney, for instance, is quoted by Watters (2006: 75) as saying: “We’re beginning to draw cause-and-effect arrows between social and economic macrovariables down to the level of the child’s brain.” Watters also quoted Lawrence Harper as saying: “If you have a generation of poor people who suffer from bad nutrition, it may take two or three generations for that population to recover from that hardship and reach its full potential.” Epigenetic effects are heritable, meaning that they operate across gen- erations, although they are reversible because they regulate, not change, the DNA sequence. In other words, healthier and more nurturing envi- ronments may reverse negative epigenetic modifi cations and generate positive ones. But if criminologists and other social scientists are to share in the exciting world of epigenetics, they will have to shed their fear of genetics. Fear of genetics is unwarranted, as Matt Ridley (2003: 6) points out:

Genes are not puppet masters, nor blueprints. They may direct the construction of the body and brain in the womb, but they set about dismantling and rebuilding what they have made almost at once in re- sponse to experience. They are both the cause and consequence of our actions. Somehow the adherents of the “nurture” side of the argument have scared themselves silly at the power and inevitability of genes, and missed then greatest lesson of all: the genes are on our side. Genes, Environments and Behavior 25 CONCLUSION

A basic understanding of genetics must be an integral part of any social scientist’s intellectual armamentarium. We cannot discuss the problems of crime and SES without it being so. As we have seen, the power of genes takes nothing away from the power of the environment; indeed, it under- lines it. Certainly, genes modulate environmental effects, but environments also modulate genetic effects. Genes and environments are the heads and tails of our existence. This alone is reason enough for criminologists to learn the rudiments of genetics at least as well as they learn the rudiments of statistics. Then there is the depressing accusation of genetic determinism, implying that we are slaves to DNA. Far from being molecular puppet masters, Colin Badcock (2000: 71) asserts that our genes “positively guarantee” human freedom and agency. Likewise, because so many things that we do in life affect the expression of our genes, epigenetic researcher Randy Jirtle asserts that “Epigenetics introduces the concept of free will into our idea of genet- ics” (in Watters 2006: 34). These scientists are able to say this because, unlike our environments, our genes belong to us; they are ours. They are what make us uniquely ourselves and thus resistant to environmental infl u- ences which grate against our natures. This view of humanity is far more respectful of human dignity than the view that we are putty in the hands of fi ckle environments arranged by others. By constantly extracting infor- mation from the environment and manufacturing the proteins we need to navigate it, our genes are enablers, not constrainers. 3 Evolutionary Psychology, Crime and Status

EVOLUTIONARY PSYCHOLOGY AND GENE-CULTURE CO-EVOLUTION

The human genome is the DNA library of the accumulated wisdom of mil- lions of years of natural selection. Genes that sit on the shelves of that library are there because they provided advantages to our ancestors in the pursuit of the shared goals of all life forms: survival and reproduction. Evo- lutionary psychology is an approach that utilizes the synthesis of natural selection and to test hypotheses regarding the func- tional advantages conferred by those traits. It is interested in distal “why” questions rather than proximate “how” questions. In terms of our interests, evolutionary psychologists are interested in why social status is so impor- tant, why seeking it may lead to criminal activity in some circumstances, and how status seeking and criminal behavior are functionally related to evolutionary goals. Evolutionary psychology is even more environmentally friendly than behavior genetics because it looks for environmental causes of behavioral variation and focuses on human similarities rather than dif- ferences. Evolutionary psychologists do not deny human genetic differences distributed around adaptive means; they simply focus on central tendency rather than variation. Evolution is succinctly defi ned as changes in a population’s gene pool over time by selective retention and elimination of genes. The nature of any living thing is the sum of its evolutionary adaptations, and so it follows: “Human nature may be defi ned as our collection of adaptations” (Ken- nair 2002: 27). An adaptation is a design feature that arose and promoted its increased frequency through an extended period of natural selection because it functioned to increase survival and/or reproductive success. Behavioral analysis is at the very heart of evolutionary processes because behavior is evoked in response to environmental challenges, and natural selection passes judgment on behavior that has fi tness consequences. Selec- tion for behavioral traits is almost certainly more rapid than for physi- cal traits because organisms play an active part in the selection of their behavior. This is why Plomin and his colleagues have asserted that “the behavioral genomic level of analysis may be the most appropriate level of understanding for evolution because the functioning of the whole organ- ism drives evolution. That is, behavior is often the cutting edge of natural Evolutionary Psychology, Crime and Status 27 selection” (2003: 533; emphasis added). After all, it is behavior that creates new variants, and then, and only then, can natural selection work its pro- cess of selective retention and elimination. When we speak of the evolution of behavior, we are talking about behavioral traits and general propensities to behave in one way rather than in another, not about specifi c behaviors. Evolutionary psychology does not ignore culture; it simply reminds us that ultimate level explanations complement, not compete, with proximate level explanations, and that “psychology underlies culture and society, and biological evolution underlies psychology” (Barkow 1992: 635). Paraphras- ing Theodosius Dobzhansky’s oft-cited observation about biology and evo- lution, Richerson and Boyd, the champions of gene/culture co-evolution, remark that “nothing about culture makes sense except in the light of evo- lution” (2004: 237). Gene-culture co-evolution means that nature (genes) and nurture (cultural learning) constitute a fully integrated reciprocal feed- back system. Both genes and culture are information transmission devices: the former laying the foundation (the capacity) for the latter, and the latter then infl uencing the former (what polymorphisms are useful in this culture at this time?). In other words, if a novel trait or behavior that happens to be culturally useful and desirable emerges, those displaying the trait or behavior will be advantaged in terms of securing resources and mates. Indi- viduals thus blessed will advance their fi tness as the alleles underlying the trait or behavior will be preserved and will proliferate in the population gene pool. One important example of gene/culture co-evolution is studies of a large range of animal species that show group size to be related to brain size, leading a number of evolutionary biologists and neuroscientists to propose that the intellectual demands of living in large groups drove the selection for what they term the “social brain.” The social brain enables us to negoti- ate relationships and understand what others are thinking and feeling, and leads us to cooperation to secure resources and for group defense (Dunbar and Shultz 2007). It has also been shown how genes underlying altruism can become a fi xed part of the genome because the trait is culturally valued and thus fi tness enhancing for those who demonstrate it (Gintis 2003). It used to be thought that the process of gene-culture evolution was most salient during the Pleistocene epoch when cultural change was slower than genetic change, and that culture evolution is too rapid for natural selection to track in more modern times. However, using new genetic technology John Hawks and his colleagues (2007) have shown that the rate of genomic change has been an incredibly 100 times greater over the last 40,000 years than it was during the 1.8 million-year-long Pleistocene. They attributed this to the greater challenges posed by living in larger and larger social groups: “[T]he rapid during the Late Pleistocene created vastly more opportunities for further genetic changes, not fewer, as new avenues emerged for communication, social interaction, and creativity” (Hawks et al. 2007: 20757). Although our most human characteristics evolved during 28 Social Class and Crime the Pleistocene, we do not operate with brains forged exclusively during that epoch as some evolutionary psychologists aver. New genetic varia- tions affecting the brain’s structure and function have been discovered as it continues to evolve in response to new ecological and cultural conditions (Evans et al. 2005; Mekel-Bobrove et al. 2005).

ALTRUISM AND EMPATHY

To understand criminality we have to understand altruism (Tibbetts 2003). Altruism is an active regard for the well-being of another and is the epitome of prosocial behavior, and thus the polar opposite of criminality. Reproduc- tive success is the spoils of an unconscious competition to contribute more of one’s genes to subsequent generations than other members of the breed- ing population. Because altruism involves extending a benefi t to others at a cost to the altruist, it used to be thought that it did not make sense that an organism would risk its own fi tness for the benefi t of others. William Ham- ilton’s (1964) theory of inclusive fi tness (the direct fi tness of the individual and the indirect fi tness of its genetic kin) was one answer to this paradox, but this kind of altruism is defi ned exclusively by its fi tness consequences and is exercised without any sort of conscious intentionality. Although inclusive fi tness and its close relative, theory (the tendency to favor close genetic relatives over others), does help to explain how altruism could have been selected for in any species, they do not constitute a satis- factory explanation for human altruism. Unconscious gene-based theories cannot help us to explain altruistic acts in which the benefi ciaries have no genetic link to their benefactors. The theory of is one attempt to explain intentional altruism (Trivers 1971). Reciprocal altruism is an act in which an organ- ism provides some benefi t to another without the expectation of immedi- ate payment, but with an unconscious expectation of future payment. In a hunter-gatherer band, for instance, a hunter who has been successful that day will share his meat with one who was not, thereby extending a great benefi t to another at little cost to himself (his family cannot eat all the meat, and it would soon spoil). The unspoken expectation is that if the tables are turned tomorrow, he and his family will be repaid in kind by the recipient of today’s altruism. Reciprocal altruism is ultimately designed to benefi t the altruist because of the expectation of future reciprocity, and is thus not selfl ess (Lehmann and Keller 2006). But reciprocal altruism is also unable to explain situations in which an individual confers some benefi t on strangers, perhaps even risking his or her life to save another. Altruism extended to non-kin others in situations where reciprocity is not likely is called “psychological altruism” (Kruger 2003) and has a dif- ferent evolutionary history. We feel good when we extend some benefi t to others without expectations of reciprocity; dropping money in a beggar’s Evolutionary Psychology, Crime and Status 29 hat may make us feel superior, and giving presents to deprived children may assuage any guilt we have about our privileged position. Psychological altruism is thus apparently motivated by internal rewards (Brunero 2002). We act altruistically because we feel good when we do, and because it con- fers valued social status on us by identifying us as persons who are kind, reliable, and trustworthy. In an evolutionary sense, we are altruists because our distant ancestors who were cooperative enjoyed greater fi tness ben- efi ts than those who were not, and the neural mechanisms that produces rewarding feelings when we do good things for others is the physiological adaptation. All evolutionary biologists are constrained visionaries when it comes to natural selection. They point out that individuals are adapted to act in ways that tend to maximize their own fi tness, not to behave for the good of the group, but also that individuals’ unconscious fi tness goals are best realized by adhering to the rules of cooperation, and that is “for the good of the group.” Actions are judged good or bad by natural selection according to their consequences, not by their intentions, just as actions are judged by constrained visionaries. That is, although intentional psychological altru- ism is ultimately (but unconsciously) self-serving, this does not diminish its value to its benefi ciaries one bit. Selfi shness as understood by evolutionists is both individually and socially adaptive. It is by cooperating with others and being actively concerned with their well-being (and others behaving likewise) that we simultaneously serve our own best interests and the best interests of our communities. Frans de Waal (2008) posits that the more ancient trait of empathy channels altruism in social species without reliance being placed on cog- nitive ruminations about such things as reciprocity concerns. Empathy is the cognitive and emotional ability to understand the feelings and distress of others as if they were our own. The cognitive component allows us to understand the distress of others and why they are feeling it, and the emo- tional component allows us to “feel” their distress. The feeling of empathy is an evolved visceral motivator moving us to take action to alleviate the distress of others if we are able. The basis of empathy is the distress we ourselves feel when witnessing the distress of others, and if we can assuage the distress of others we thereby ease our own. Thus empathy also has a selfi sh component, but it is a very good thing that it is because if we lacked in emotional connectedness to others, we would be like psychopaths, cal- lously indifferent to their needs and suffering. Empathy evolved rapidly in the context of parental care because care- givers must quickly and automatically relate to the distress signals of their offspring (de Waal 2008). This is perhaps the reason that females are invariably found to more empathetic than males regardless of the tools and methods used to assess empathy (Campbell 2006). Parents who were not alerted to or who were unaffected by their offspring’s distress signals or by their smiles and cooing are surely not among our ancestors. Like the 30 Social Class and Crime diffusion of adaptive love and nurturance of offspring to the non-adaptive love and nurturance of the children of others and to pets, the capacity for empathetic responses, once locked into the human repertoire, diffused to a wider network of social relationships and is made manifest in acts of con- scious intentional altruism.

CRIMINALITY

If altruism is defi ned as an active concern for the well-being of others by conferring some benefi t on them at an expense to one’s self, criminality can be defi ned as its polar opposite: as an active concern exclusively for one’s own needs and wants and a willingness to meet those needs and wants at the expense of others. Criminality is selfi shness as understood in the ver- nacular (i.e., a crabbed egoism shorn of any concern for others). This kind of selfi shness is ultimately self-destructive because non-cooperators in a population of cooperators are treated as pariahs. Yet evolutionary psychol- ogists agree with Durkheim that although it is morally regrettable, crime is normal behavior (Kanazawa 2003). If behavior we now call criminal is normal in the sense that it is in the behavioral repertoire of all humans, it must have conferred some evolutionary advantage on our distant ances- tors. Judith Harris (1998: 299–300) speculates about the probable traits of hunter-gatherer leaders and how they would have been useful:

Almost all the characteristics of the “born criminal” would be, in wa- tered-down form, useful to a male in a hunter-gatherer society and useful in his group. His lack of fear, desire for excitement, and impul- siveness made him a formidable weapon against rival groups. His ag- gressiveness, strength, and lack of compassion enable him to dominate his groupmates and give him fi rst shot at hunter-gatherer perks.

The perks at which he had fi rst shot were those that were most perti- nent to survival and reproductive success—resources and women. Women would have been attracted to such men, not because they were sensitive “nice guys,” but because they had status and resources within the group and were good protectors (Buss 2005). Such traits can certainly overshoot their optimum and become liabilities rather than assets, which is often the case when exercised too freely in modern evolutionarily novel societies. Modern environments are so radically different from the hunter-gatherer environments in which we evolved our most human traits that many traits selected for their adaptive value at the time may not be adaptive today. Con- versely, behaviors that may be adaptive in some modern environments may not be adaptations in the sense that they have an evolutionary history of selection. To call some feature of human nature an adaptation is to make a claim about the past, not the present, and defi nitely not about the future. Evolutionary Psychology, Crime and Status 31 Criminal behavior is a way to acquire valued resources by force or fraud. Evolutionary psychologists refer to such behavior (whether it is defi ned as criminal or not) as cheating. Cheats gain resources from oth- ers by signaling their cooperation but then failing to follow through. In the absence of internal (guilt, shame) or external (punishment, ostracism) deterrents it is in an individual’s interests to obtain resources under the assumption of reciprocity and then to default. Cheating behavior has been observed among numerous animal species, which implies that it has had positive fi tness consequences (Alcock 2005). Although all humans may have the potential to exploit and deceive others, at the same time we have evolved to be a highly social and cooperative species with minds forged by evolution to form cooperative relationships built on trust. Among humans, criminal behavior may be viewed as an extreme form of default- ing on the rules of cooperation. But cheating comes at a cost, so before deciding to do so individuals must weigh the costs and benefi ts of coop- erating versus defaulting. Cheating is only rational in circumstances of limited interaction and communication, but frequent interaction breeds trust and bonding, and cheating becomes an irrational strategy because cooperators remember those who cheated them. Cheating ruins reputations and invites retaliation, which is why cooperation is the predominant strategy in any social species even if cheating is rational under some circumstances. The vast majority of social animals, including human beings, are condi- tional cooperators, who can be cheated because they believe in mutual trust and cooperation and expect the same from others. If they are victimized by non-cooperators, however, they retaliate by not cooperating with their victimizer in the future, and may repay him or her in kind. Under these cir- cumstances cooperation is the rational strategy because each player reaps in the future what he or she has sown in the past. Exposure and retaliation are threats only if cheats are constrained to operate where their reputations are known. Cheats can move from location to location meeting and cheating a series of grudgers who are unaware of their reputation. This is the pattern of many career criminals who move from place to place, job to job, and relationship to relationship, leaving a trail of misery behind them before their reputation catches up. Cheats are thus more likely to prosper in large cities in modern societies than in small traditional communities where the threat of exposure and retaliation is great. However, factors such as the stability of the group and cultural dynamics must be considered when discussing reputation. In socially dis- organized neighborhoods there is very little of the trust upon which coop- erative relationships are forged, and so many of their inhabitants resort to cheating behavior because that is what they have learned to expect from others. But even in communities in which males value a “bad ass” reputa- tion more than anything else, there must be a certain level of group loyalty and cooperation. 32 Social Class and Crime PARENTING EFFORT VERSUS MATING EFFORT

Evolutionary theories of crime focus on sexuality as the prime mover of criminal behavior because ultimately the imperative of all living things is reproductive success. These theories posit two strategies that members of any animal species can follow to maximize reproductive success: parenting effort and mating effort (Campbell 2009). Parenting effort is that propor- tion of an organism’s total reproductive effort invested in rearing offspring, and mating effort is that proportion allotted to acquiring sexual partners. Although humans invest more energy and other resources in parenting than any other species, there is considerable variation within the species, with gender accounting for the vast proportion of it due to different levels of obligatory . Female investment necessarily requires an enormous expenditure of time and energy, but the only obligatory invest- ment of males is the time and energy spent copulating. Male reproductive success increases in proportion to the number of females they have sexual access to, and thus males have an evolved propensity to seek multiple part- ners. Mating effort emphasizes maximizing the number of offspring rather than nurturing a few, although maximizing offspring is obviously not a conscious motive of any male seeking sex. The proximate motivation is sexual pleasure, with more offspring being a natural consequence in pre- contraceptive days when the strategy proved successful. Ancestral females could not increase their reproductive success by mating with multiple partners given the male reluctance to invest in females with a reputation for promiscuity. Females could increase their success by mating with multiple partners in some circumstances, not by increasing the number of offspring, but by obtaining resources from each partner, thus increasing the probability that offspring they already have will survive to reproductive age (Hrdy 1999). Reproductive success among our ancestral females rested primarily on their ability to secure mates to assist them in raising offspring in exchange for exclusive sexual access, and thus human females evolved a much more discriminating attitude about sexual behavior (Geary 2005). Evolutionary theory maintains that the inherent confl ict between the reckless and indiscriminate male mating strategy and the careful and dis- criminating female mating strategy drove the evolution of traits such as aggressiveness and deception that are useful in helping males to overcome male competitors and/or female reticence. Although these male traits were of course designed by natural selection to facilitate mating effort, they are also useful in gaining non-sexual resources via illegitimate means. In other words, the neurohormonal processes underlying these traits were selected to foster reproductive success not criminality, but once the mechanisms are in place, they can serve purposes other than those for which they were designed (Quinsey 2002; Walsh and Wu 2008). Likewise, traits that facilitate parenting effort are conducive to proso- cial activity. As David Rowe (1996: 270) states: “crime can be identifi ed Evolutionary Psychology, Crime and Status 33 with the behaviors that tend to promote mating effort and noncrime with those that tend to promote parenting effort.” Because female reproductive success hinges on parenting effort rather than mating effort, females have evolved higher levels of the traits that facilitate it (e.g., empathy, altruism, and constraint), and lower levels of traits inimical to it (e.g., aggressiveness, risk-taking) than males (Campbell, Muncer, and Bibel 2001). Empirical research strongly supports the notion that an excessive con- centration on mating effort is a robust predictor of criminal behavior. A review of 51 studies examining the relationship between number of sex partners and criminal behavior found 50 of them to be positive; they also found age of onset of sexual behavior to be negatively related to criminal behavior (the earlier the age of onset, the greater the criminal activity) in all 31 reviewed studies (Ellis and Walsh 2000). A British cohort study found that the most antisocial 10% of males in the cohort fathered 27% of the children (Jaffee et al. 2003), and a number of gang studies (e.g., Palmer and Tilley 1995) show that gang members have more sex partners than non-gang members in the same community, and that gang leaders have the most of all. Finally, there are striking differences in behavior between males in cultures that emphasize parenting versus mating strategies; males in cultures that emphasize mating effort exhibit low-level parental care, hypermasculinity, and transient bonding (Ember and Ember 1998). These are of course the kinds of behaviors we also fi nd in socially disorganized lower-class communities. Molecular genetic studies are also fi nding statistically signifi cant rela- tionships between sexual and criminal behavior. Guo, Tong, and Cai’s (2008) study of 674 white males found that males homozygous for the 10-R allele of the dopamine transporter gene (DAT1) combined with heterozy- gous males (10-R/9-R) had signifi cantly more sex partners (M = 5.66) than 9-R homozygotes (M = 2.94), as well as signifi cantly higher delinquency scores and scores on other kinds of risky behaviors. However, 9-R/9-R homozygotes were not “protected” from promiscuity if they had low cogni- tive ability and attended schools with a culture of early sexual activity, indi- cating once again how important it is to uncover GxE interactions. Another genetic study found that the same DAT1 polymorphism (10-R/10-R) that was signifi cantly related to number of sex partners was also signifi cantly related to antisocial behavior among a sample of 2,574 males between the ages of 18 and 26 (Beaver, Wright, and Walsh 2008).

VIOLENCE AND STATUS

Violence is one of criminology’s deepest concerns, and thus we should understand how it was adaptive in evolutionary environments and what environmental circumstances are likely to evoke it. Violence would not now be part of the evolved human behavioral repertoire if it were not helpful 34 Social Class and Crime in solving some set of adaptive problems. Evolutionary psychologists view aggression and violence as a context-contingent strategy triggered by evo- lutionarily relevant contexts (i.e., circumstances akin to the adaptive prob- lems our distant ancestors experienced in which its use resulted in a desired benefi t (Buss and Duntley 2006)). The human brain evolved in the context of overwhelming concerns about resources and mate acquisition. When food and sex partners are plentiful, using violence to obtain them is risky and unnecessary, but when resources are scarce, the use of violence may be worth the risk. The credible threat of violence is related to reproductive success in almost all species of social animals. When others seek to deprive us of our resources or mates it would have been very useful in evolutionary environments to react with violence or threats of violence. Acquiring a reputation for violence in environments where you have to take care of your own problems is useful because others who are aware of your reputation will avoid your resources. A violent repu- tation serves to let any potential challenger know that it would be in his best interests to avoid you and your resources and to look elsewhere. This is why a “bad ass” reputation is so highly valued in many lower-class neighbor- hoods, why those with such a reputation are always looking for opportuni- ties to validate it, and why it is craved to such an extent that “Many inner city young men . . . will risk their lives to attain it” (Anderson 1994: 89). Risky violent confrontations among males over status issues are most often observed in environments lacking fi rmly established status hierar- chies and in which social restraints have largely dissolved. These environ- ments have been termed “subcultures of violence” or “honor subcultures” in which taking matters into one’s own hands is seen as the only way to obtain the all-important “juice” (status) on the street (Anderson 1999). Serious assaults and homicides among members of honor subcultures are often the result of trivial altercations over matters of honor, respect, and reputation in the context of a culture where the violent defense of such intangibles is a major route to status (Mazur and Booth 1998). Assaults and homicides often take place in front of an audience composed of friends of both the killer and the victim, thus squeezing the maximum amount of “juice” from the incident (Buss 2005). Reacting violently to even minor threats to one’s status lets others know that “You can’t push me around!” Having a violent reputation not only protects what one has, it is also use- ful in gaining more in contexts in which there is little social control. Because high status typically brings more copulation opportunities, genes inclining males to aggressively pursue their interests enjoyed greater representation in subsequent generations. Cultures where mating effort is strongly favored have homicide rates greatly exceeding those of any modern society, such as the Yanomamo rate of 166 per 100,000 (Ellis and Walsh 2000: 71). Among the Yanomamo, males who have killed the most in inter-village warfare have three times as many wives and children on average as those who have done little or no killing (Buss, 2005). Evolutionary Psychology, Crime and Status 35 The use of violence imposes costs such as the possibility of death or severe injury, so perhaps sharing resources with others might be a better strategy. Mutual sharing is what has shaped us into a cooperative species, but we are concerned with cheats who never signed on to that particular social contract. In socially disorganized neighborhoods where a tradition of set- tling one’s own quarrels without involving the authorities is entrenched, a person who is unprepared to match violence with violence may loose every- thing. Natural selection has provided us with brains that quickly switch to a violence mode when we have reason to believe that things we value may be taken from us (Kelly 2005). Lower SES individuals are often said to possess traits such as fatalism, and impulsivity which are considered maladaptive. Wilson and Daly (1997: 1271), however, suggest that discounting the future: “may be a ‘rational’ response to information that indicates an uncertain or low probability of surviving to reap delayed benefi ts, for example, and ‘reckless’ risk taking can be optimal when the expected profi ts from safer courses of action are negligible.” What Wilson and Daly are saying is that males compete for sta- tus and resources by whatever means are available to them in the environ- ments in which they live, and when many people they know die at an early age, and that when they perceive little opportunity for legitimate success, living for the present and engaging in risky violence to obtain resources makes evolutionary (if not moral) sense. Using data from 77 neighborhoods in Chicago for the years 1988 through 1995, Wilson and Daly hypothesized that neighborhoods with the low- est income levels and the shortest life expectancies (excluding homicides) would have the highest homicide rates. Statistically removing the effects of homicide mortality, life expectancy ranged from 54.3 years in the poorest neighborhood to 77.4 years in the wealthiest. Homicide rates ranged from 1.3 per 100,000 in the wealthiest neighborhood to 156 in the poorest, a huge 120-fold difference. Wilson and Daley interpreted the data as sup- porting the view that natural selection has equipped humans to respond to extreme levels of inequality and short life expectancy by creating risky, high-stakes male-male competitions.

THE IMPORTANCE OF THE FAMILY

The most important feature of the environment for the healthy develop- ment of human beings is love and nurturance. These features are found within the family, the nursery of human nature. Lancaster and Lancaster (1987: 188) view the family as a “specialized and very basic adaptation that greatly extended the investment parents could make in their offspring.” The idea of extended parental investment in the form of bi-parental care remains pivotal in evolutionary theorizing about the origins of the human family (Gross 2005). As Pillsworth and Haselton (2005:101) explain: 36 Social Class and Crime “Comparative, physiological, and cross-cultural evidence supports the hypothesis that humans have a suite of adaptations for forming conjugal pairs . . . the motivations to form couples emerges from adaptations with deep evolutionary roots, and thus assumes a central motivational status for most humans.” Family forms vary widely across time and culture, but with regard to healthy human development the “best” family rearing environment is one in which children are surrounded by many kinfolk, which was the case for most of the evolutionary history of Homo sapiens (Daly and Wilson 2005). The term nuclear is aptly chosen because the triad of man, women, and children is the core (nucleus) around which the web of kinship in its various family forms revolves. The triad of man/woman/child is thus the minimal core of the environment of rearing for the human species. The extreme level of dependency of human infants required selection for strong bonds between mother and infant, and because of this male involvement in pro- visioning and parenting would have been an important source of parental investment in our ancestral environments (Gross 2005). Males and females who bonded to jointly provide parental investment unconsciously increased their reproductive success by increasing the probability of their offspring surviving to reproductive age. To help in the bonding process natural selec- tion seized on some of the same substrates of mother-infant bonding such as oxytocin and other neurological reward systems that provided joy and pleasure (see Walsh 2009 for a more detailed discussion). These mecha- nisms were designed to keep parents together for at least long enough for offspring to become minimally self-suffi cient in the context of rearing envi- ronment that contained numerous kin. Infants, children, and adults need and expect to be attached to others because natural selection has built in those deep needs because of the vital role attachment has played over our evolutionary history. The universality of marriage refl ects the deep recognition of the impor- tance of the family in rearing healthy offspring: “Marital alliance and bipa- rental care are part of the human adaptation” (Daly and Wilson 1988: 187). Because the family is a product of gene-culture co-evolution, we should expect signifi cant social problems to arise when it is disrupted. Meta-analy- ses have concluded that children from broken homes had signifi cantly poorer psychological adjustment, lower self-esteem, poorer academic achievement and social relations, and more psychological maladjustment and marital instability than adults reared in intact homes (Amato and Keith 1991). Given that the heritability of divorce is about 0.50 (Booth, Carver, and Granger 2000), we should not too hastily attribute these negative outcomes to divorce per se. There is no “divorce gene” of course; the coeffi cient is the result of people with traits that make them diffi cult to live with and that increase the likelihood of unfavorable outcomes in all areas of their lives, including divorce. Divorce is a phenotype, and the traits underlying it are endophenotypes. The genes underlying these endophenotypes are passed on to Evolutionary Psychology, Crime and Status 37 their children who then become susceptible to antisocial behavior because of their inherited traits, not their parent’s divorce. But to the extent that divorce independently increases children’s risk of antisocial behavior, a father-absent upbringing is said to breed a “supermasculine” male identity (Hall 2002) which constantly needs to be validated. The effects of father absence are not due to its non-normativeness in Western societies because similar effects are found in prestate polygynous cultures where children typically grow up in mother-child households separate from their fathers. “Societies in which children are reared in mother-child households or the father spends little time in child care tend to have more physical violence by males than do societies in which fathers are mostly around” (Ember and Ember 1998:14). Single-parent families increase children’s risk of future offending in a variety of ways, such as decreased supervision, low levels of social sup- port, and an increased probability of abusing and neglecting their children. A nationwide study of children ages 2 through 9 found that children of single parents were 6.7 times more likely to witness family violence, 3.9 times more likely to be maltreated, and 2.7 times more likely to be sexually assaulted than children with both biological parents present. The fi gures for stepparent families were even worse at 9.2, 4.6, and 4.3, respectively (Turner, Finkelhor, and Ormrod 2006). The vast majority of stepparents do not abuse their stepchildren, but the risk for abuse is greatly elevated in stepfamilies. Another study found that stepchildren were between 9 and 25 times more likely (depending on age, with the risk being greater the younger the child) to be abused than children residing with both biological parents (Daly and Wilson 1985). A child living with a stepfather or live-in boyfriend is approximately 65 times more likely to be fatally abused than a child living with both biological parents (Daly and Wilson 1996). The lessons of history tell us the same things about the importance of families. The efforts of the Soviet Union in 1928 (Hosking 1985) and in Communist China in 1949 (Fletcher 1991) to abolish the “bourgeois fam- ily” led to such disastrous effects (abandoned women and children, ille- gitimacy, rampant crime) that both countries did a 180 degree turn and began extolling the virtues of the family. These two “natural experiments” tell us more about the importance of the family for the development of healthy prosocial human beings than all social science studies of the fam- ily combined. As Alfred Blumstein (1995: 12) has opined: “teenage moth- ers, single-parent households, divorced households, unwed mothers,” all constitute high risk for criminality, “and certainly don’t bode well for the future of the nation.”

CONCLUSION

To understand status (including socioeconomic status) and why it is so important, it is imperative to understand what evolutionary purposes it 38 Social Class and Crime serves. The one single evolutionary goal of all sexually reproducing crea- tures is to pass on their genes to future generations. Therefore, any behav- iors that arose that would have given an organism a reproductive edge would have spread in the mating population. Mother Nature does not mor- alize; any behavioral trait that proved useful in this regard, however mor- ally repulsive or attractive we may now consider it to be, would be selected for. Due to different levels of obligatory parental investment, males and females have evolved quite different mating strategies, and quite different levels of the behavioral traits underlying mating and parenting strategies. The traits underlying male-typical mating effort (aggression, deception, low fear and empathy, etc.), while designed for gaining sexual access, are useful in seeking other resources illegitimately. Conversely, the traits underlying female-typical parenting effort (high empathy and nurturance, low level of risk taking) are conducive to prosocial activities. This is the ultimate reason why females in every culture throughout history have committed far less crime than males, and the more serious and violent the crime, the bigger the male-female gap. The ultimate reason that high status in one’s social group is so important for males is that status brought with it greater access to females in ancestral environments, and still does today. Of course, this is not necessarily a con- scious motivation for status seeking, which is a valuable asset quite inde- pendent of its evolutionary origins. Status striving in combination with the evolutionary importance of the family helps us to make sense of the kinds of behavior often observed in lower-class neighborhoods where altercations are frequently sparked over trivial matters of street reputation. 4 The Neurosciences, Conscience and the Soft-Wired Brain

THE BASIC BRAIN

If the genome is the production center of the organism’s response machin- ery, the brain is where all the decisions about what to produce are made. The brain is where genetic dispositions and environmental experiences are integrated and become one. Much more so than the genome, the brain physically captures the environment in its circuitry, so if we want to under- stand the impact of different environments on behavior, criminology can ill afford to continue to ignore the brain. As Robinson (2004: 72) asserts, any theory of behavior “is logically incomplete if it does not discuss the role of the brain.” The insights we can derive from the neurosciences will not only strengthen criminology, they will also strengthen our claims for preventa- tive intervention. Bruce Perry (2002: 81) points out that there are three key brain systems relevant to survival and reproductive success: mechanisms that (1) facili- tate responses to threats to our well-being, (2) facilitate mate selection and reproduction, and (3) facilitate the protection and nurturing of the young. Each one of these mechanisms has relevance for criminology. The mecha- nisms are composed of a number of sub-mechanisms of hundreds of bil- lions of nerve cells called neurons. All our thoughts, feelings, emotions, and behavior are the result of networks composed of these neurons, which are nourished and supported by other brain cells called glial cells. Each neuron consists of the cell body containing the nucleus that carries out the meta- bolic functions of the neuron, an axon that serves as a transmitter sending signals to other neurons, and a number of dendrites that serves as a receiver picking up information from neighboring neurons. The brain’s more than 100 billion neurons “form over 100 trillion connections with each other— more than all of the Internet connections in the world!” (Weinberger, Elvevag, and Giedd 2005:5). Lower level brain structures come with neural connections “hard-wired” at birth, but development of the higher brain areas depends on making connections between neural networks after birth in response to experience (“soft-wired”). There are debates are about the relative contributions to brain develop- ment of processes intrinsic to the brain (in the genome) and extrinsic to 40 Social Class and Crime it (in the environment). Neuroscientists who favor innateness are called selectionists or nativists, and those who favor of the power of extrinsic factors are called constructivists or connectionists. Both positions agree that the environment is hugely important in the development of the brain; the argument is not “whether the environment thoroughly infl uences brain development, but how it does” (Quartz and Sejnowski 1997: 579; emphasis original). Both sides are also equally adamant that the typical social sci- ence blank slate view of the human mind is a neurological impossibility. Both see neonates as born with a large number of built-in assumptions that frame the world for them (Mitchel 2007). Neuroscientists distinguish between two brain developmental processes that physically capture environmental events in the organism’s lifetime: experience-expected and experience-dependent (Schon and Silven 2007). Everyone inherits species-typical hard-wired brain structures and func- tions produced by a common pool of genetic material; these are the experi- ence-expected mechanisms that refl ect the phylogenic history of the brain. However, individuals will vary in brain functioning as their genes inter- act with the environments they encounter to construct those brains; these are experience-dependent mechanisms that refl ect the brain’s ontogenic plasticity (Gunnar and Quevedo 2007). A concrete example of the differ- ence between the two processes is language. While the capacity for lan- guage is entirely genetic (a hard-wired experience-expected capacity), what language(s) a person speaks is entirely cultural (soft-wired in experience- dependent fashion). Experience-expected processes have evolved as neural readiness during “critical” developmental periods to incorporate environmental informa- tion that is vital to an organism. Natural selection has recognized that cer- tain processes such as sight, speech, depth perception, affectionate bonds, mobility, various aversions, and sexual maturation are vital and has pro- vided for mechanisms (adaptations) designed to take advantage of experi- ences occurring naturally within the normal range of human environments. Pre-experiential brain organization (built-in assumptions) frames or orients our experiences so that we will respond consistently and stereotypically to vital stimuli (Geary 2005). A large part of the variability in the wiring patterns of the brains of different individuals depends on experience-dependent processes (i.e., the kinds of physical, social, and cultural environments they will encounter). As Depue and Collins (1999: 507) wrote: “experience-dependent processes are central to understanding personality as a dynamic developmental con- struct that involves the collaboration of genetic and environmental infl u- ences across the lifespan.” Although brain plasticity is greatest in infancy and early childhood, it is maintained to a lesser degree across the lifespan as we shape and reshape the brain in ways that could never have been genetically pre-programmed. The Neurosciences, Conscience and the Soft-Wired Brain 41 SYNAPTOGENESIS

The process of wiring the brain is known as synaptogenesis, a process that occurs both according to a genetic program and the infl uence of the envi- ronment. During the fi rst few months of an infant’s life, dendrites prolifer- ate, and specialized glial cells wrap around axons to begin the process of myelination (myelin is a fatty substance that insulates the axons and makes for speedier transmission of electrical impulses). Dendrite growth and axon myelination continues throughout life, but proceeds at an explosive rate during infancy and toddlerhood. The experience-expected “lower” brain regions (spinal column and limbic regions) are the fi rst to be myelinated, and some “higher” brain regions are not fully myelinated until well into adulthood (Sowell, Thompson, and Toga 2004). The birth of a set of synapses is less important than whether they will survive the competition for synaptic space. The most active period of synaptogenesis is infancy and early childhood, although about half of these connections will eventually be eliminated. The brain creates and eliminates synapses throughout life, but creation exceeds elimina- tion in the fi rst two years after which production and elimination are roughly balanced until adolescence when elimination exceeds production (Giedd 2004). This process of selective production and elimination has been termed neural Darwinism by Gerald Edleman (1992), who posits a selection process among competing brain modules (populations of dendrites and synapses). Neuronal connections are selected for retention or elimina- tion according to how functionally viable (adaptive) they prove to be in the person’s environment in the same way that environmental challenges select from a population’s reservoir of in evolutionary time. The brain’s neuronal populations thus evolve in somatic time like species evolve in geological time by the selective elimination or retention of genes. Neural networks retention is a use-dependent process in which con- nections that are sustained are those that exchange information fre- quently and strongly (Penn 2001). Experiences with strong emotional content are accompanied by strong electro-chemical impulses so that the neurons become more sensitive and responsive to similar stimuli in the future (Shi et al. 2004). Frequently activated neurons are thus primed to fi re at lower stimulus thresholds once neurological tracks have been laid down. This process is summed up in the neuroscientists’ saying: “The neurons that fi re together, wire together; those that don’t, won’t” (Penn 2001: 339). Figure 4.1 illustrates the explosive growth of neurons and dendritic branching from birth to age one. The cell bodies increase in size, dendrite branching becomes more elaborate, and axons become myelinated. 42 Social Class and Crime

Figure 4.1 The growth of neurons and neuronal networks from birth to age one.

BRAIN DEVELOPMENT AND EARLY EXPOSURE TO VIOLENCE

Thus, we can conclude that the brain is always a “work in progress,” and that its development is “use-dependent.” The message is that we and the experiences we encounter will largely determine our neuronal connection patterns. To paraphrase , “Men make their own history, but they do not make it as they please.” All cortical modules are not equal candidates for selection because rGE assures us that some modules have a greater probability of selection than others via channeling effects (Mitchell 2007). Additionally, and perhaps more importantly, when we are assessing the effects of a lower SES upbringing, we are much less the masters of our own fates during our formative years. Some rearing environments are truly horrendous and leave permanent neuropsychological scars. As Perry and Pollard (1998: 36; emphasis added) point out: “Experience in adults alters the organized brain, but in infants and children it organizes the develop- ing brain.” Brains organized by stressful and traumatic events tend to relay future events along the same neural pathways etched out by those early events and be interpreted in negative or even hostile ways. For good or bad, well-grooved synaptic pathways established in early life are more resistant to pruning than pathways laid down later in life. These pathways have The Neurosciences, Conscience and the Soft-Wired Brain 43 been stabilized, and thus intrude into our transactions with others across the lifespan. This is why I took issue with Sampson and Laub’s (2005) claim that childhood experiences deserve no special place in the pantheon of causes of later behavior in Chapter 1. We have seen that from an evolutionary perspective violence is some- thing we should expect to see in environments in which social control has largely broken down. Because the brain physically captures our experi- ences by molding and shaping neuronal circuitry in ways that make our behavior adaptive in the environments in which we fi nd ourselves, we have further insight as to how violence begets violence. Impulsiveness is the proximate behavioral expression of a brain wired by consistent expo- sure to violence (Niehoff 2003). If our brains develop in violent environ- ments, we expect hostility from others and behave accordingly (“Do onto others before they do it to you”). By doing so we invite the very hostility we are on guard for, thus confi rming our belief that the world is a hostile place and setting in motion a vicious circle of negative expectations and confi rmations. Chronic vicarious exposure to violence through violent video games has been shown to desensitize viewers to it and makes them callous and indif- ferent to the suffering of others (Cooley-Quille et al. 2001). Subjects who view these types of games frequently show lower physiological responses when exposed to depictions of violence than do non-viewers (Carnagey, Anderson, and Bushman 2007). Functional MRI (fMRI) studies also show decreased activity in brain structures that regulate aggression and increased activity in structures associated with increased aggression in fre- quent viewers (Murray et al. 2006). Experimental studies such as these are conducted with subjects exposed to chronic vicarious violence, not real violence aimed at themselves or at loved ones. Witnessing and experienc- ing violence “upfront and personal” is presumably captured on the neural circuitry of children more strongly and plants a warning sign there that the world is a dangerous place in which they must be prepared to protect their interests by violent means if necessary. How much real violence do the children of our poorest citizens witness? In one study in Chicago, 33% of school children had witness a homicide, and 66% had witnessed a serious assault (Osofsky 1995). Another study found that 32% of Washington, DC, children and 51% New Orleans chil- dren had been victims of violence, and 72% of Washington, DC, and 91% of New Orleans children had witnessed violence (Osofsky 1995). Although intimate partner violence occurs in all classes, over 30 years of research has consistently shown that the prevalence is greater in lower SES families (Cunradi, Caetano, and Schafer, 2002). Witnessing and experiencing so much violence cannot help but stamp on the neural circuitry of these chil- dren that the world is a dangerous place in which one must be prepared to protect one’s interests by violent means if necessary. It is in this sense that violence begets violence. 44 Social Class and Crime THE NEUROBIOLOGY OF CONSCIENCE

Criminal behavior refl ects a failure of socialization. A large part of success- ful prosocial socialization revolves around the person’s relative sensitivity to reward and punishment and is rooted in emotion. The person must of course learn cognitively what is expected of him or her, but how well the lesson is learned is more a function of how it engages limbic system emotion than how it engages rational refl ection. Reinforcement sensitivity theory (RST) is the major neurobiological personality theory focusing on the con- ditioned emotional suppression or expression of behavior (Cooper, Perkins, and Corr, 2007). RST posits three interacting systems of emotional regula- tion located within separate brain circuits that underlie motivated behavior: the behavioral activating (or approach) system (BAS), the behavioral inhi- bition system (BIS), and the fi ght-fl ight-freezing system (FFFS). The BAS and BIS are part of the limbic system, and the FFFS is part of the peripheral nervous system, specifi cally, the autonomic nervous system (ANS). RST has migrated to criminology where it is known as reward dominance theory. The BAS is sensitive to signals of reward from both conditioned (e.g., alcohol, gambling) and unconditioned (e.g., food, sex) appetitive stimuli. The BAS is reminiscent of the Freudian id which obeys the pleasure prin- ciple, craves instant gratifi cation of its desires, and cares not whether the means used to satisfy them are appropriate or injurious to self or to oth- ers. The BIS is sensitive to conditioned (e.g., violations of social rules) and unconditioned (e.g., heights, snarling creatures) threats of punishment (Corr 2004). The BIS can be likened to the Freudian superego striving for the ideal and representing all the moral “brakes” (the dos and don’ts) inter- nalized by the person during the process of socialization. Unlike the Freudian id, ego, and superego, the biological substrates of the BAS and BIS have been neurologically mapped out. The BAS is primar- ily associated with dopamine and with mesolimbic system structures such as the nucleus accumbens, a structure which is rich in neurons that produce and respond to dopamine (Day and Carelli 2007). The BIS is associated with serotonin and with limbic system structures such as the hippocampus and the amygdala that feed their memory circuits into the prefrontal cor- tex (Goldsmith and Davidson 2004). Certain polymorphisms of the dop- amine gene and of the dopamine-deaminating enzyme COMT have also been shown to be related to dimensions of RST (Reuter et al. 2006). Figure 4.2 shows the prefrontal cortex, the nucleus accumbens, and the ventral tegmental area. If the accelerator gets stuck and/or the brakes are faulty, we may pro- duce a “craving brain” that leads one into all sorts of physical, social, moral and legal diffi culties, such as addiction to gambling, food, sex alcohol, and drugs (Ruden 1997). A faulty BAS may be seen as a brain less effi cient in binding dopamine, thus leading its owner on an incessant search for more pleasure inducing feelings. The genetic polymorphism The Neurosciences, Conscience and the Soft-Wired Brain 45

Figure 4.2 Brain areas relevant to reward and executive functions. Source: National Institute of Drug Abuse, The Brain’s Drug Reward System, 1996. underlying such ineffi ciency is most likely the 7-repeat allele of the DRD4 gene. The more typical repeat alleles make shorter receptor proteins that are more effi cient in binding dopamine (Propper and Moore 2006). Defec- tive brakes, on the other hand, are associated with the short allele of the serotonin transporter gene (5-HTTLPR), which leads to low 5-HTTLPR expression and thus lower serotonergic expression (Propper and Moore 2006). The Freudian ego’s function is to sort out the confl ict between the antisocial demands of the id and the overly conformist demands of the superego. The ego can be viewed as a BAS/BIS-balanced system that obeys the reality principle: It does not deny the pleasure principle: it sim- ply adjusts to the demands of reality. The fi nal dimension of RST is the FFFS, which is part of the ANS. The FFFS also includes the amygdala and hippocampus, which suggests that in many ways the BIS and the FFFS may be thought of as a single integrated inhibition system. Although most researchers interested in the behavioral biology of arousal to refer to the whole ANS, the FFFS was coined to dis- tinguish the ANS’s fi ght-fl ight function from its “housekeeping” functions such as automatically maintaining homeostasis in the visceral parts of the body (heart, lungs, etc). The ANS has two complementary branches: the sympathetic and parasympathetic systems. When an organism perceives a threat, its amygdala-associated neural networks send signals to the sym- pathetic branch to mobilize the body for vigorous action (the “fi ght or fl ight” reaction). Pupils dilate for better vision, the heart and lungs accel- erate their activity, and digestion stops, among other things, all of which is aided by pumping out the hormone epinephrine. The parasympathetic 46 Social Class and Crime system restores the body to homeostasis after the organism perceives the threat to be over. Because the BIS and the FFFS are engaged simultaneously in approach- avoidance situations, the reactivity or response level of the ANS is a major determinant of a person’s innate temperament. The avoidance of fear and anxiety mediated by the ANS has great importance for the development of our consciences. The conscience is the basis for forming moral judgments and for moving one’s behavior into conformity with those judgments, but it is more than a reasoning mechanism since it typically moves us in one direc- tion or another before we have had much of an opportunity to contemplate options. The conscience has an emotional component that precedes cogni- tion that Barkow has (1989: 121) described as an involuntary and invasive “limbic system override” that serves to adjust our behavior in social situ- ations. These so-called “social emotions” (guilt, embarrassment, shame, pride, etc.) animate, focus, and modify neural activity in ways that lead us to choose moral responses over other possible responses in approach- avoidance situations. The social emotions move us to behave in ways that enhanced our distant ancestor’s reproductive success by overriding neocor- tical decisions suggesting alternatives to cooperation (i.e., cheating) which may be rational in the short term (since they supply free resources) but which are ultimately fi tness reducing. Conscience is thus a mix of limbic emotional and rational cognitive mech- anisms that enables individuals to internalize the moral rules of their social groups. Individuals with a strong conscience will feel guilt, shame, and anxiety when they violate these rules; those with an impaired conscience will experience none of these emotions. As Kochanska and Aksan (2004: 304) put it: “An impaired conscience is the core aspect of conduct disor- ders, antisocial development, and . Conversely, the capacity for remorse and empathy, an appreciation of right and wrong, and engaging in behavior compatible with rules all mark successful adaptation.” Indeed, the major defi ning characteristic of psychopaths at the neurobiological level is their greatly reduced ability to “tie” the brain’s cognitive and emotional networks together (Pitchford 2001; Wiebe 2004). Differences in the emotional component of conscience are observed in infancy, long before children are able to cognitively refl ect on their behav- ior as morally right or wrong. Variation in ANS arousal patterns lead to variation in how well the prescriptions and proscriptions of moral behav- ior are learned via classical conditioning (Kochanska and Aksan 2004). Unlike operant conditioning which is active and cognitive in that it forms a mental association between a person’s behavior and its rewarding or punishing consequences, classical conditioning is mostly passive and is visceral in nature. Classical conditioning simply forms a subconscious association between two paired stimuli, the strength of which depends on the level of ANS arousal more than on conscious deliberation and purposeful action. The Neurosciences, Conscience and the Soft-Wired Brain 47 ANS arousal thresholds (as measured by heart rate, blood pressure, and electrodermal activity) are normally distributed, with most individuals by defi nition clustered around the mean and with a small minority on each tail. Individuals with a hyperarousable ANS are easily aroused by situa- tions they perceive as threatening, and thus they condition easily because arousal produces punishing visceral feelings. When hyperarousable indi- viduals contemplate behaving in a way they might want to but which is contrary to the expectations of others, their sympathetic system is quickly engaged. When they make the decision not to act on the antisocial impulse (the “limbic system override”) the parasympathetic system engages to restore the ANS to homeostasis. Acquiescing to the conformity demands of one’s socializers (parents, teachers, peers, and so on) thus prevents sympa- thetic ANS arousal, which powerfully reinforces conforming behavior via the reduction of incipient feelings of fear and anxiety. The visceral associations between behavior and the response of others is the mechanism by which we develop the “gut level” social emotions that make up the “feeling” superstructure of our consciences. These social emo- tions are retrofi tted to the same machinery that drives the hard-wired pri- mary emotions of fear, anger, sadness, and joy (Tibetts 2003). Retrofi tting the primary emotions involves elaboration and refi nement by cognition dur- ing the socialization process. This is the “knowledge” part of conscience; the lessons we learn relating to what things we should feel guilty, ashamed, and embarrassed about, and what things we should feel good about. Social- ized individuals will fear the pain of punishment and the shame of rejec- tion when they transgress and welcome the joy of acceptance and affection when they behave well. A hyperarousable ANS is a powerful protective factor against criminal behavior. A review of 40 studies of ANS activity and criminal and other antisocial behaviors found 38 studies that supported the link and 2 that had non-signifi cant results (Ellis and Walsh 2000). A number of subsequent studies have shown that young males living in criminogenic environments who remain free of their infl uences (i.e., they do not become delinquent) show hyperactive ANS arousal under conditions of threat (Lacourse et al. 2006). Youths whose ANSs are highly reactive are actually less likely to commit antisocial acts than youths with hyporeactive ANS arousal reared in non-criminogenic environments (Brennan et al. 1997). Individuals with a hyperactive ANS are easily socialized. They learn their moral lessons well because sympathetic ANS arousal is subjectively experienced as fear and anxiety. Children learn that when they behave well they do not incur the wrath of others, and fear and anxiety do not occur. Individuals with relatively unresponsive ANSs are diffi cult to condi- tion because they experience little fear or guilt when they transgress, even when discovered and punished, and thus they have no visceral restraints against further transgressions. They are not viscerally reinforced for con- forming behavior because they are not aroused to fear in the fi rst place and 48 Social Class and Crime thus cannot receive positive reinforcement for conforming behavior by way of a return to ANS homeostasis (Bell and Deater-Deckard 2007). Having knowledge of what is right or wrong without that knowledge being paired with emotional arousal is like knowing the words to a song but not the music—not much good for singing in the social choir.

BIOSOCIAL INTERACTIONS

The same sort of biosocial interactions found in behavior genetic studies (high heritability in advantaged SES environments and low heritability in disadvantaged SES environments) are found in the arousal literature. It is reasonable to assume that the rational cognitive components of con- science are strongly stressed in the socialization process of children from advantaged environments. If such children become delinquent and crimi- nal, then it follows that because they lack the environmental instigation to such behavior they will be less autonomically (emotionally) conditionable than their peers. Variance in physiological measures will therefore be more salient in accounting for antisocial behavior in higher SES individuals than in lower SES individuals whose physiological risks are masked by psycho- social risk factors. This is consistently found and has been called the “social push hypoth- esis” (Raine 2002). Scarpa and Raine (2003: 213) defi ne the social push hypothesis thusly: “[I]f an individual lacks psychosocial risk factors that predispose toward antisocial behavior yet still exhibits antisocial behavior, then the causes of this behavior are more likely to be biologically than socially based.” Children from lower SES environments who are diffi cult to condition because of a hyporeactive ANS are at greatest risk for antisocial behavior because they are also more likely, on average, to also lack the cognitive conditioning necessary for acquiring a conscience. Both the ANS and another component of the stress response called the hypothalamic- pituitary-adrenal axis (to be discussed in Chapter 8) are liable to down- regulation (to become less responsive) by experiences in severely stressful environments of rearing.

PREFRONTAL DYSFUNCTION THEORY

The prefrontal cortex (PFC) is responsible for a number of executive func- tions such as making moral judgments, planning, analyzing, synthesizing, and modulating emotions. The PFC provides us with knowledge about how other people see and think about us, thus motivating us to adjust our behav- ior to consider their needs, concerns, and expectations of us. The PFC can be considered the physical substrate of George Herbert Mead’ s concept of the “generalized other,” which is essentially a theory of mind that attempts The Neurosciences, Conscience and the Soft-Wired Brain 49 to account for the social origin of self-consciousness through role-taking and role-adjustment. PFC functions are collectively referred to as executive functions (Fishbein 2001) and are clearly involved in prosocial behavior if functioning normally. The association between the frontal lobes and behavior is often illus- trated by the dramatic 19th-century case of Phineas Gage who had a tamp- ing iron blasted through his the PFC, and this terrible accident drastically changed his personality. This case gave early neuroscientists their fi rst clues about the functions of the PFC, although it is obviously an extreme exam- ple. Damage to the PFC does not have to be massive or even anatomically discernible to negatively impact behavior. PFC damage could be at the cel- lular level and be the result of genetic factors that affect neuron migra- tion during the earliest stages of frontal lobe development, or the result of maternal substance abuse during pregnancy that likewise adversely affect neuron migration. (See Chapter 8.) Brain imaging studies consistently fi nd links between PFC activity and impulsive criminal behavior. A PET (positron emission tomography) study comparing impulsive murderers with murderers whose crimes were planned found that the former showed signifi cantly lower PFC and higher limbic sys- tem activity (indicative of emotional arousal) than the latter and other con- trol subjects (Raine et al. 1998). An MRI study (Raine et al. 2000) found that males with antisocial personality disorder (APD) had 11% less pre- frontal gray matter volume than subjects in two control groups (a “healthy” group and a substance-dependent group) matched for SES, ethnicity, IQ, and head circumference. The antisocial group also evidenced reduced ANS activity during exposure to stress. The APD group self-reported a signifi - cantly greater number of violent crimes than either control group. The PFC and ANS measures predicted APD versus non-APD status independent of 10 demographic and psychosocial risk factors and were by far the most powerful predictors. Another MRI study found that psychopaths (n = 16) labeled “unsuccessful” (caught) had a 22.3% reduction in prefrontal gray matter compared with 23 control subjects. “Successful” (uncaught) psycho- paths (n = 13), however, did not differ in gray matter from controls. Higher psychopathy scores were associated (r =–.39) with low gray matter volume (Yang et al. 2005). It should be noted that these fi ndings relate to acts of impulsive reactive violence, not planned proactive violence. Cauffman, Steinberg, and Piquero (2005) combined the reward domi- nance and PFC dysfunctions theories in a large scale study of incarcerated youths in California and found that seriously delinquent offenders have slower resting heart rates (indicative of low fear and low ANS arousal) and performed poorly relative to non-delinquents on various cognitive functions performed by the PFC. In another study combining both theories, Yacubian (2007) and her colleagues monitored activity in the PFC and the ventral striatum areas via fMRI scans among 105 healthy male volunteers. Some subjects showed high levels of activity and some showed blunted activity in 50 Social Class and Crime anticipation of similar (monetary) rewards. Examining subjects’ genotypes, the researchers found that variation in neural activity was mediated by vari- ation in genetic polymorphisms related to dopamine activity, namely the dopamine transporter gene and COMT, the primary enzyme responsible for the degradation of dopamine. Studies such as this provide solid neurological and genetic evidence of the bases for the reward dominance theory of addic- tion to all kinds of things, including addiction to crime. Neuroscientists have studied how the MAOA variants discussed in Chap- ter 2 in the context of child maltreatment and violence. Meyer-Lindenberg and his colleagues (2006) examined neural structure and function via fMRI of healthy (non-criminal, non-substance abusing) volunteers identi- fi ed as possessing the low (n = 57) or high (n = 85) variant of the MAOA polymorphism. Subjects, particularly males, with the low variant showed hyperactivity in the amygdala (fear) and hippocampus (memory) cir- cuits and diminished activity in the prefrontal and cingulate (regulatory) regions during induced emotional arousal. Not only was the function of these areas different across the different MAOA polymorphisms, but so was their structure (the level of enzymatic activity is thought to infl uence the wiring of the brain during development). Although these subjects were non-criminal and non-violent, the structural and functional profi les of the low MAOA individuals refl ect a vulnerability to impulsive violence (quick arousal/poor regulatory inhibition) under the right circumstances such as perceived threats.

CONCLUSION

The chapter on genetics showed how the genome responds to environmen- tal challenges by manufacturing the appropriate proteins to help us to meet them. The evolutionary chapter showed how the human genome itself was shaped by the challenges to their survival and reproductive success our dis- tant ancestors faced. In this chapter we complete the circle by showing how the physical architecture of the higher brain centers is laid down according to the environments we encounter. Most people’s brains develop normally because they have been exposed to the normal range of experience-expected rearing environments, but others have been exposed to environments that have thrown the developmental process off. But it is certainly no more pru- dent to conclude that antisocial behavior is solely a function of neurological abnormalities than it is to conclude the same thing about different genetic polymorphisms or about criminogenic environments. These things are risk factors that require the presence of other risk factors, both biological and environmental, to manifest themselves as aberrant behavior. A dopamine- serotonin imbalance certainly puts a person at great risk for a “craving brain” and all the antisocial activities such a brain implies, but thankfully it does not guarantee it. The Neurosciences, Conscience and the Soft-Wired Brain 51 One very positive aspect of a neuroscience perspective on aberrant behavior is that the evidence produced by the “hard” laboratory/high-tech- nology sciences is taken more seriously than that produced by the “soft” questionnaire/low-technology social sciences. This is good because both hard and soft methods tend to tell us the same things—neglect, abuse, pov- erty, violence—must be prevented if we are to prevent antisocial behavior. Kagan (2007: 367) notes how diffi cult it is to get legislators to provide funds to battle poverty, abuse, and neglect with appeals to the heart based on soft social science data, but “these same legislators become enthusiastic advocates of prevention if they are told that these same environmental con- ditions alter children’s brains.” Social science opponents of “biologizing” should at least see this as a very positive side of taking biology seriously even if they believe (and I think sometimes quite correctly) that we tend to get overly enthusiastic and to over interpret data from the hard sciences relative to human behavior. 5 Social Class and Criminal Behavior Myth or Reality?

TITTLE’S MYTH

The relationship between social class, or socioeconomic status, and crimi- nal behavior has been a central focus of criminology since its inception. Of course SES does not directly cause crime and delinquency, or anything else for that matter. SES is only a convenient label conceptualized and measured in different ways to categorize people so they can be compared across various domains of interest. It is shorthand for a composite of variables used to place individuals, families, and neighborhoods into categories according to their current ability to legitimately gain access to valued resources. The variables used to defi ne SES are themselves the result of complex interactions of many factors that we will examine in future chapters. The issue here is whether or not SES is related to crime and delinquency, not why it is or is not. Despite its long-term focus on SES, there is a school of thought in crimi- nology that considers the negative relationship between social class and criminal behavior to be a myth (Tittle 1983). To call something a myth is to claim that it is without determinable basis in fact or that it defi es natural explanation. What it means in the present context is that empirical tests have allegedly consistently failed to fi nd signifi cant relationships between SES and criminality. The major basis of this claim was an article by Tittle, Villemez, and Smith entitled “The Myth of the Social Class and Criminal- ity” (1978). This relationship/no relationship issue is one that should be addressed at some length before we proceed further since I have no desire to perpetuate a “myth.” Tittle et al. based their conclusion about the mythical status of the class- crime relationship on 35 self-report studies from which they calculated an average gamma of–0.09. No overall measure of statistical signifi cance was reported for this average gamma, but the message is clear—the relationship, signifi cant or not, is indeed very weak (but hardly mythical since it is in the hypothesized direction). Comments on the Tittle et al. paper ranged from praising it as a defi nitive work (Stark 1979) to writing that the no-class/ crime relationship “has the same status (in my view) as fl at earth theories” (Rossi 1990: 623). Harris and Shaw (2001: 137) were even more unkind, calling the conclusion that there is no class-crime relationship “extraordi- narily senseless.” Social Class and Criminal Behavior 53 John Braithwaite’s (1981) response was to compile a signifi cantly larger number of studies (90) than Tittle et al.’s, only 21 of which failed to support a negative class-crime relationship. All 34 of Braithwaite’s studies of adult crime and SES based on offi cial statistics found signifi cant inverse rela- tionships. The most recent contribution to the debate is Ellis and McDon- ald’s (2001) collection of 273 studies from around the world. This study also produced mixed results. Looking at studies assessing the relationship between parental social class and crime/delinquency, they found that 57.7% reported a signifi cant inverse relationship, 28.2% found no relationship, and 14.1% reported a positive relationship. Using offi cial statistics, how- ever, only 15.1% failed to fi nd a signifi cant inverse relationship between parental social class and subjects’ delinquency. When looking at studies using offi cial statistics and respondents’ own social class (rather than par- ents’) only 2 out of 72 failed to fi nd a signifi cant inverse relationship.

SELF-REPORTS: THEIR STRENGTHS AND WEAKNESSES

The controversy over the class-crime relationship proved useful by sparking a great deal of interest in examining the strengths and weaknesses of self- report studies. Self-report surveys typically provide a list of offenses and request respondents to check each type of antisocial act they recall having committed and how often, or if they have ever been arrested and, if so, how many times. These studies have relied in the past primarily on students in colleges and high schools for subjects, although inmates and proba- tioners/parolees have also been surveyed and interviewed. Several studies have addressed the issue of the accuracy and honesty of self-reported offending in various ways with generally encouraging results, at least for uncovering the prevalence of minor offenses. Studies have shown that offi cially identifi ed delinquents and criminals self-report substantially more antisocial conduct on anonymous questionnaires than do non-delinquents and non-criminals (Cernkovich, Giordano, and Pugh, 1985). Had this not been the case serious doubt would have been cast on the self-report methodology. The evidence thus appears to indicate that self-report measures provide fairly accurate information about some forms of antisocial behavior and that almost everyone has committed some sort of illegal act sometime in their lives. The method also captures the rough extent of illegal drug usage among high school and college students, some- thing that neither of the two major sources of offi cial crime statistics—the Uniform Crime Reports (UCR) and the National Crime Victimization Sur- vey (NCVS)—attempts to do. However, there are a number of reasons why self-report crime surveys also provide a highly distorted picture of criminal activity. As mentioned previously, most self-report studies in the past surveyed largely class- homogeneous “captive” students. High school and college students are 54 Social Class and Crime not populations in which one expects to fi nd many seriously criminally involved individuals; those that are so involved have dropped out, are in juvenile facilities or jails, or are the most likely to be truant. Such studies thus do not cover the very people we are most interested in gathering infor- mation about. Further muddying the waters, self-report studies typically only ask about trivial antisocial acts such as fi ghting, stealing items worth less than $5, disobeying parents, smoking, and playing truant, most of which are offi - cially classifi ed as status offenses (if classifi ed as offenses at all) rather than delinquency. Almost everyone has committed one or more of these acts, and they are hardly acts that help us to understand the nature of serious crime. Some researchers compound this problem by lumping respondents who report one delinquent act together with adjudicated delinquents who break the law in many different ways many different times. Braithwaite (1981) points out, for instance, that one study he examined only required students to admit to underage drinking plus having heterosexual sex to be placed in the most delinquent category! Venial sins were translated into mortal sins by a zealously Jesuit vision of iniquity in this study. Even though most people are fairly forthright in revealing their peccadil- loes, most people do not have a serious criminal history, and those who do have a distinct tendency to underreport their crimes (Hindelang, Hirschi, and Weis 1981). Furthermore, as the number of crimes people commit increases, so does the proportion of offenses they withhold reporting (Cernkovich et al. 1985). For instance, researchers have asked respondents with known arrest histories whether they have ever been arrested and found that 20% to 40% replied negatively, with those arrested for the most serious offenses being most likely to deny having been arrested (Hindelang et al. 1981). Another problem is that reporting honesty appears also to vary across race and gender; males tend to report their antisocial activities less hon- estly than females, and African Americans less honestly than other racial/ ethnic groups (Kim, Fendrich, and Wislar, 2000; Maxfi eld, Weiler, and Widom, 2000). This evidence renders any statements about gender or racial differences regarding antisocial behavior based on self-report data highly suspect. In a similar vein, almost all studies assessing the issue fi nd that lower-class youths tend to score higher on “lie scales” (faking either being good or bad) built into survey questionnaires to detect false reporting (Braithwaite 1981). If everyone has committed trivial offenses of the kind included in almost all self-report studies, there is no point in searching for group or individual differences with respect to those offenses. If one does, one should not be surprised to fi nd null or very weak relationships. Subjects from relatively class-homogeneous schools being asked about trivial offenses results in severely truncated variances in both measures that will by defi nition reduce the value of a correlation that may be extremely strong when assessed across the entire population range of both criminal offenses and SES. Social Class and Criminal Behavior 55 Perhaps at the heart of all this confusion about the class-crime relation- ship is the tendency to equate parental SES with offspring SES. Self-report studies utilizing students assign them a class position on the basis of their parents’ class position, but as various researchers note (Ellis and McDonald 2001), adolescents are already on their way to establishing their own SES, which may be quite different from that of their parents. If we have a sample with a modicum of SES heterogeneity, we will have individuals from the lower classes moving up and individuals from the middle classes moving down, with the upwardly mobile lower-class youths probably having lower “delinquency” scores than downwardly mobile middle-class youths. There is some support for the idea that middle-class students that fail in school are more rebellious than lower-class boys that fail (Chen and Kaplan 2003). “Delinquency” is thus better predicted from the class the individual will enter as an adult than from the individual’s class of origin. This is why the Braithwaite (1981) and Ellis and McDonald (2001) studies of the class- crime relationship among adults reported consistent and robust negative correlations. Self-report studies have their place, but like the drunk who knew he hadn’t lost his keys under the lamppost but looked for them there anyway “because that’s where the light is,” criminologists who use convenience samples and mistake peccadilloes for crimes are fooling themselves without the excuse of inebriation. Surely we should be looking for the things being sought where they are most likely to be found. Lest I be accused of being too harsh, recall that Tittle et al. entitled their article the myth of social class and criminality, not the myth of parental social class and minor misbehav- iors. Modern theorists recognize that criminality and minor misbehaviors are two very different things. There are a number of longitudinal studies being carried on today such as the Dunedin Multidisciplinary Health and Development Study (Caspi et al. 2002), the National Longitudinal Study of Adolescent Health Study (Udry 2003), and the National Youth Survey (Menard and Mihalic 2001). Studies such as these use a variety of methods to verify self-report data, and many of them are collecting DNA from their subjects, thus adding a very exciting and robust dimension to criminologi- cal studies.

ECOLOGICAL STUDIES

Fearing the ecological fallacy trap, one source of data assessing the class- crime relationship dismissed by Tittle et al. is ecological studies. The eco- logical fallacy is the fallacy involved in making inferences about individuals based on aggregated statistics collected from some group (e.g., race) or area (e.g., city) to which those individuals belong. A misreading of what the fal- lacy means is that it is always wrong to draw inferences about sub-units based on the aggregates to which they belong. What the fallacy actually 56 Social Class and Crime means is that the process of aggregating data conceals a great deal of varia- tion. Ecological data often provide valuable clues about individual behav- ior, particularly in medical epidemiology and economics. Furthermore, the ecological fallacy gets less problematic as the size of aggregate is reduced; it is much more problematic to make inferences about individual Protestants and suicide based on Durkheim’s (1951) fi nding that Protestant countries have higher suicide rates than Catholic countries than it is to make infer- ences about individuals and homicide in each of the 77 Chicago neighbor- hoods in the Wilson and Daly (1997) study discussed in Chapter 3. Ecological studies invariably show far more crime in lower-class areas than in middle-class areas. Every year the NCVS reports victimization rates to be highest in the lowest income areas and that they drop monotonically as the average income of an area increases. The ecological fallacy notwith- standing, unless we are to believe that these differences are the result of middle- and upper-class folks trolling for victims in lower-class areas, we must conclude that the crimes recorded in these areas are committed by people who inhabit them. The Wilson and Daly (1997) study showing the 120-fold difference in murder rates between the highest and lowest income neighborhoods in Chicago would surely not be dismissed by anyone on the basis of the ecological fallacy or the possibility that murders are being committed by middle-class outsiders. As Braithwaite (1981: 37) comments: “Perhaps Tittle et al. take their own fi ndings seriously and adopt no extra precautions when moving about the slums of the world’s great cities than they do when walking the middle-class areas of such cities.” Of course we all know that they would, as would we all.

DO OFFICIAL STATISTICS REVEAL POLICE ARREST BIAS: THE CASE OF RACE?

Because of the alleged discrepancy between self-report data (no or weak class-crime relationship) and offi cial data (consistent and robust class-crime relationship), a number of supporters of the superior validity of self-report measures have claimed that the latter relationship is the result of police bias (Uggen 2000). The UCR does not break down arrest statistics by SES because, unlike its other categorizations of race, age, and gender, SES is not a natural category. We are able to circumvent this problem using ecologi- cal studies of different class areas, but if this still smells of the ecological fallacy we may substitute race for class to look at arrest rates. Harris and Shaw (2000: 136) claim that the correlation between race and SES “is so strong that any relation observed between race and crime might be nothing more than a cloaked relation between class and crime.” Because of this oft- heard claim, we are probably on safe ground examining the bias argument using black-white differences in the probability of arrest as a proxy for SES differences in the probability of arrest. Social Class and Criminal Behavior 57 When all we had in terms of nationwide offi cial statistics was UCR data on which to base our pronunciation, it was almost impossible to deter- mine if the highly disproportionate number of blacks arrested for the index offenses represented disproportionate black involvement with crime or police bias in making arrests. Those who make such claims never ask if the overrepresentation of whites relative to Asians in arrests is the result of anti-white racism, or if the overrepresentation of young males relative to older males and to females is the result of sexism and ageism. With the advent the NCVS, researchers were able to compare offi cial arrest data from the UCR (real crimes) with NCVS data because victims of personal attacks are asked to identify the race of their assailants to NCVS interviewers. Victimization studies yield essentially the same racial differen- tials as UCR statistics. For example, about 60% of robbery victims describe their assailants as black, and about 60% of the suspects arrested for robbery are black (O’Brien 2001). Nevertheless, Tittle at al. ignored this valuable source, even though the NCVS uses probability sampling of the entire United States (thus encompassing all SES levels) and interviews far more people each year than in all of the 35 studies looked at by Tittle et al. combined. D’Alessio and Stolzenberg (2003) brought National Incident-Based Reporting System (NIBRS) data, which combines the best features of the UCR and NCVS, to bear on the issue of differential black-white arrest rates. D’Alessio and Stolzenberg used NIBRS data from 17 states and 335,619 arrests for rape, robbery, and aggravated and simple assault and found results directly opposite of the police bias hypothesis. Blacks were indeed arrested in far greater numbers than their proportion in the population would lead us to expect, but less often than their proportion of offenses they commit would lead us to expect. For those crimes in which the race of the perpetrator was known, 30.8% of the white robbers were arrested versus 21.4% of black robbers, and the same was found for aggravated assault (53.1% versus 42.5%) and simple assault (46.8% versus 36.8%), but no signifi cant race difference was found for rape (27.2% versus 28.1%). D’Alessio and Stolzenberg conclude that the disproportionately high black arrest rate is attributable to their disproportionately higher involvement in crime. Pope and Snyder (2003), also using NIBRS data, analyzed 102,905 incidents of violent crime committed by juveniles and found essentially the same thing (i.e., white youths were more likely to be arrested than black youths despite the greater overall seriousness of the crimes committed by black youths). If the assumption that race is a good proxy for social class is true, the same should hold for class differentials in arrest rates also.

WHAT ABOUT WHITE-COLLAR CRIME?

Crimes occur on America’s Main Streets and Wall Streets as well as on her Mean Streets, although Americans seem to be less concerned about 58 Social Class and Crime the former until an ENRON-type scam comes along. Unless a white-collar crime does something like deplete the retirement funds of someone near and dear, we tend to think of it as less scary than “in-your-face” street crime. Victims may not even be aware that they are being victimized for months or even years after a white-collar crime is initiated, but they are acutely aware of the gun in their face and the alien hand on their wallet. This does not mean that many white-collar crimes are not just as (or more so) injurious than common street crimes. As far as I know, former NASDAQ chairman Bernie Madoff never pulled a gun or knife on anyone or threatened anyone face to face, yet his $65 billion Ponzi scheme caused more misery than any 10 Mafi a hit men combined. Victimized charities and foundations had to shut their doors to the needy, thousands of inves- tors lost their shirts, families were wiped out fi nancially and spiritually, and a number of the victims committed suicide. Then there was the savings and loan (S & L) scandal of the 1980s that cost the U.S. taxpayer $473 bil- lion. This staggering amount is many times greater than losses from all the “regular” bank robberies in American history put together (Calavita and Pontell, 1994). These crimes were crimes of the morally depraved, not the economically deprived, and must also be accounted for when discussing class and crime. The term white-collar crime was coined in the 1930s by Edwin Suther- land who defi ned it as crime “committed by a person of respectability and high social status in the course of his occupation” (1940: 9). Although this defi nition became enshrined for a long time as the defi nition of white-collar crime, it is problematic in three ways: (1) Many (perhaps most) white-col- lar criminals are not of “high social status,” (2) many are not otherwise respectable people, and (3) it fails to distinguish between crimes committed by individuals acting for personal gain and crimes committed on behalf of the employer with the employer’s blessing and support. An alternate defi nition of white-collar crime is that of the U.S. Congress in the Administration Improvement Act (AIA) of 1979. The AIA defi ned it as “an illegal act or series of illegal acts committed by non-physical means and by concealment or guile, to obtain money or property, or to obtain business or personal advantage” (in Weisburd et al. 1991). This defi nition focuses on the characteristics of the offense rather than Sutherland’s focus on the characteristics of the offender. Some criminologists delight in drawing U-shaped curves to show that while the lower classes are overrepresented in street crime, the upper classes are overrepresented in white-collar crime. These tend to be the same crimi- nologists who accept Sutherland’s defi nition of the white-collar criminal rather than the AIA’s defi nition of the white-collar crime. In our enthusi- asm to spread malevolence evenly across the classes, we tend to forget that while we can intelligently speak of lower-class overrepresentation in street crime, we cannot likewise speak of the upper classes being overrepresented in white-collar crime if we have already defi ned the lower classes out of the Social Class and Criminal Behavior 59 running by accepting Sutherland’s defi nition. After all, anyone is eligible to commit a common street crime, but to commit many white-collar crimes one must be in a position to do so; only a medical professional can commit Medicare fraud, and only a corporate accountant is in a position to make false statements to the Securities and Exchange Commission. Thus, if one accepts Sutherland’s defi nition, to say that the middle and upper classes commit more white-collar crime than the lower classes is to state something that could not be otherwise because the lower classes have been defi ned out of the running. Class is a constant in the restrictive world of white-collar crime but a variable in the democratic world of street crime. This being so, it is impossible for the middle and upper classes to be overrepresented vis- à-vis the lower classes since the latter are not represented at all.

HOW DIFFERENT ARE STREET AND WHITE-COLLAR CRIMINALS?

The answer to the question of how different white-collar criminals are from street criminals depends on whether we focus on all acts of concealment and guile as defi ned by the AIA or confi ne ourselves to Sutherland’s more restrictive defi nition. Except for forgery/counterfeiting, fraud, and embez- zlement, white-collar crimes are conspicuously absent from the yearly crime tally contained in the UCR. However, NIBRS contains a laundry list of 58 different offenses it classifi es as white-collar crimes—such as the adultera- tion of food, drugs, or cosmetics; bribery; health care provider fraud; Ponzi schemes; and wire fraud—but again does not breakdown offenses by SES (Barnett 2002). Are the UCR “white-collar” crimes really white-collar? True to Suther- land, Harris and Shaw reserve the term white-collar for “the offenses of middle- and upper-middle-level managers and executives who work in offi ces and suites,” and claim that embezzlement, fraud, and forgery are “not occupationally related,” and not “committed by individuals working as legitimate employees in legitimate fi rms” (2001: 155). However, these crimes can be occupationally related (embezzlement is, almost by defi ni- tion), but Harris and Shaw only have that sub-category of the genre called in mind when they talk about white-collar crime. Why we should confi ne ourselves only to corporate crime, and what we should call the other forms of crime outlined in the AIA and in NIBRS if not white- collar, is not addressed. The motives of white-collar criminals are the same those of street crimi- nals—to obtain benefi ts quickly with minimal effort. Furthermore the age, sex, and race profi les of occupational criminals are not that much different from those of street criminals. Using UCR data, Hirschi and Gottfredson determined that rates for fraud, forgery, and embezzlement peak in the early 20s and decline with increasing age to approximately one-half by age 60 Social Class and Crime 41. They also found that male rates were higher than female rates, and that African American rates were higher than white rates. From these data they concluded: “When opportunity is taken into account, demographic dif- ferences in white collar crime are the same as demographic differences in ordinary crime” (Hirschi and Gottfredson 1987: 967). The claim that the demographic correlates of street and white-collar criminals are or are not the same is important for criminologists such as Hirschi and Gottfredson who see their theory as applicable to all offenders. No study of the issue fi nds that the demographics of the street criminals and UCR white-collar criminals are different in kind (i.e., old, female, and white, rather than young, male, and black), but only in degree. Those arrested for embezzlement, fraud, and forgery do tend to be disproportionately young, male, and black, but older persons, females, and whites are proportionately more represented in them than they are in street crime. When assessing this argument, however, we have to real- ize that the great majority of the UCR white-collar crimes are relatively minor ones such as passing bad checks, welfare fraud, credit card fraud, and minor scams that are committed primarily by people who defi nitely do not enjoy white-collar status. A U.S. Department of Justice study of 22,580 federal white-collar defendants found that gender, race, and age representation varied with the type of crime (Barak 1998). For all eight federal white-collar crimes com- bined, 84% were male, 26% black, and 75% under 40 years of age. Those committing lending and credit fraud were the most likely to have attended college (49%, and therefore presumably not members of the lower class), and those committing forgery were the least likely (17%). Another study of federal non-corporate white-collar criminals showed that most of them were much less involved in crime and than “common” offend- ers, but that chronic white-collar offenders (about 16% of the sample) were similar to common street criminals in their patterns of prior deviance (Benson and Moore 1992). Yet another study of federal prison inmates found that white-collar criminals with prior arrests for non-white-collar crimes were largely indistinguishable from street criminals in their demo- graphics, lifestyle, and endorsement of criminal thinking patterns but were quite different from white-collar criminals with no history of arrest for non-white-collar crimes (Walters and Geyer 2004). It would seem that the great majority of white-collar criminals share more social space with street criminals than with members of the privileged classes, that most of them are “one shot” offenders, and that most commit relatively banal offenses (Benson and Kerley 2001). When we look at white-collar crimes that do require high-status occu- pations for their commission, we see a much different picture. In Weis- burd et al.’s (1991) of 1,094 white-collar criminals processed through the federal courts found that anti-trust offenders were 99.1% white, 99.1% male, and averaged 53 years of age. Securities fraud offenders were 99.6% Social Class and Criminal Behavior 61 white, 97.8% male, and had an average age of 44. Males were even more heavily overrepresented than they are in street crime, and nonwhites were also overrepresented (ranging from 17% of the bribery offenders to 38% of the false claim offenders), although not to the extent that they are present among street criminals,

EXPLAINING WHITE-COLLAR CRIME

The data seem to indicate that most white-collar criminals occupy a middle ground between the general public and street criminals. In terms of trying to explain the crimes of their sample of offenders, Weisburd and his col- leagues (1991: 192) explain white-collar crime in the anomie/strain tradi- tion, stating that it is “normal” in a business environment that encourages competition and innovation: “we have about the rate of white-collar crime that we ‘need’ in order to encourage the amount of freedom, aspiration, and upward mobility that we seem, as a society, to want.” The authors also note that many white-collar criminals were trying to hold on to the positions and wealth they had worked so hard to gain and were in danger of losing. The great Greek philosopher Plato understood human nature better than most. His allegory of the ring of Gyges is particularly apropos to under- stand white-collar crime. Gyges was a young shepherd in the service of the king of Lydia. While he certainly was not a high-status person, he was an apparently a quite respectable chap. The gist of the story is that Gyges came upon a gold ring in a cave that allowed him to become invisible whenever he turned it on his fi nger. When Gyges realized the possibilities this opened for him, he went to the court, seduced the queen, murdered the king, and took over the kingdom. Plato’s allegory makes the point that with the ring’s gift of invisibility no one is so virtuous that they could resist all temptation to act unjustly in the name of self-interest. Putting the words into the mouth of Glaucon, Plato asserts that:

No man can be imagined to be of such an iron nature that he would stand fast in justice. No man would keep his hands off what is not his own when he could safely take what he liked out of the market, or go into houses and lie with anyone at his pleasure, to kill or release from prison whom he would, and in all respects be like a God among men. (1960: 44)

Plato knew what the classical theorists knew: humans are designed to attempt to manipulate the world to the extent that they are able to maximize their pleasure and minimize their pain. We obey the dictates of these two sovereign masters differently according to the tensions that exist between our opportunities and constraints. Middle- and upper-class white-collar 62 Social Class and Crime criminals are different from the street criminal only in degree. Before the middle and upper classes commit crimes, the rewards have to be much greater and the risks of discovery, apprehension, and punishment much lower because they have so much more to lose if they are discovered and arrested (Yu and Zhang 2006). Figure 5.1 illustrates the biosocial position that criminal activity is always a function of variation in individual propensities toward criminality inter- acting with variation in the level of criminogenic infl uences in the environ- ment (GxE). The horizontal line represents individual propensity ranging from low to high, and the vertical axis represents the level of inducement present in the environment ranging from low to high. Person A has high criminal propensity and will cross the threshold from non-crime to crime at even the lowest level of environmental inducement. Person B is in a crime- constraining environment, but because he crossed the crime threshold at the same place as person A, we can assume he has more or less the same level of criminal propensity as person A. Unfortunately, people like A and

Figure 5.1 Illustrating the interaction of individual propensity and criminogenic environments in the risk of crossing the criminal threshold. Social Class and Criminal Behavior 63 B are usually found in criminogenic environments, which enhances their propensity (rGE). Person D has an average propensity for criminal behav- ior, so the environmental inducement (the ratio of reward to risk) has to be much stronger before he crosses the threshold. Person C is the “respectable” white-collar criminal with a good quality education, is married with children, and has a job bringing him a hand- some salary as an upper-level executive. Plato would have had no diffi culty appreciating why C crossed the crime threshold given the high rewards and low risks involved. Opportunities abound in corporate America to gain wealth beyond what individuals could earn legitimately. We may expect business executives to have assimilated the egoism, acquisitiveness, and competitiveness cultivated by capitalism more completely than most. Yu and Zhang (2006) maintain that the capitalist ideology of business edu- cation (maximizing profi ts) desensitizes students to ethical issues. Com- bine this with the absence of meaningful controls and the seductive lure of big money, and corporate crime is more easily understood than is street crime with its more meager rewards and more severe penalties (at least until recent times which have seen corporate crooks actually receiving sentences commensurate with the enormity of their crimes). One of the major environmental factors contributing to fraudulent busi- ness practices is the executive compensation package, which frequently loaded with stock-based compensation. Stock-based compensation is designed to increase executives’ focus on stockholder’s profi ts, but it also means that a substantial portion of his or her yearly compensation is tied to stock prices. While this method of executive compensation is a major incentive for executives to engage in creative enterprises, as in the ENRON case it creates a situation ripe for fi nagling stocks through insider trading, falsifying accounts, participating in fraudulent trading, and emphasizing short-term earnings rather than the long-term success of the corporation (Tang, Chen, and Sutarso 2008). A criminogenic environment is any environment conducive to criminal activity no matter how affl uent the ambiance; it is not only sleazy slum neighborhoods. Newcomers entering corporate environments are social- ized into the prevailing “way of doing things.” Just as slum children adopt the criminal defi nitions of their peer groups, the professional socialization of a newcomer into a business organization where unethical and illegal practices are fairly common will infl uence the newcomer’s behavior in the same direction. If the newcomer does not “fi t in” and conform, he or she is not likely to remain employed there very long. This process can produce a sort of moral apathy in well-socialized executives striving to do their jobs in their own and their company’s best interests (Yu and Zhang 2006). If their consciences are pricked by what is happening, they have ready access to techniques of neutralization to quiet them and to make further wrongdoing that much easier. If each tiny bending of the rules that brings profi t to the company also brings the rule bender appreciation and bonuses, his or her 64 Social Class and Crime behavior will be slowly, almost imperceptibly, molded in a process of oper- ant conditioning to engage in ever-greater patterns of wrongdoing.

CONCLUSION

The issue of the connection between social class and delinquent and crim- inal activity has been bedeviled by semantic and methodological deceit. Semantically, researchers have examined trivial misbehaviors and called it criminality, delinquency, and crime. Methodologically they have searched for a type of subjects in places where they are not likely to fi nd them and substitute another type simply because they are readily available. Those who deny the class-crime relationship ignore offi cial statistics and ecologi- cal studies even though both data sources reveal a robust and ubiquitous negative class-crime relationship. Their explanation for ignoring the former is police bias and the latter is the ecological fallacy. NCVS and NIBRS data show that (with respect to blacks, not social class per se) the police bias argument has no real basis. The ecological fallacy argument is based on an overly strict reading of what it actually means, and unless we are to believe that the wealthier classes troll lower-class neighborhoods for victims, the ecological data must be interpreted as meaning that the inhabitants of those neighborhoods are responsible for the crime committed there. The middle and upper classes engage in crime also, but the benefi t-cost ratio must generally be much greater before people in those classes take the risk. Such people usually commit “white-collar” crimes rather than street crimes. Depending on whether we accept the Sutherland defi nition of white-collar criminals in terms of the status of the criminal or the AIA and NIBRS defi nition in terms of the crime, we may or may not speak of the upper classes being overrepresented in white-collar crime. 6 The Class-Crime Relationship in Criminological Theories

After attempting to demolish the empirical foundation of the class-crime relationship with his colleagues Wayne Villemez and Douglas Smith (1978), Charles Tittle (1983) went on to critique the theoretical foundations for expecting such a relationship. His examination of a variety of theories claiming that a negative relationship should exist between social class and crime led him to conclude that none of them provide an adequate theoretical rationale for such a prediction. Tittle claims that the theories he discusses only lead researchers to hypothesize a negative class–crime relationship because those who formulated the theories held unwarranted assumptions about the lower classes. These assumptions contained in their theories are seen by Tittle as little more than as pejorative stereotypes about lower-class culture, and the mentality and character of lower-class individuals. This chapter examines how the class–crime relationship is conceptual- ized by the criminological theories that Tittle (1983) discussed in his cri- tique. I fi rst briefl y introduce the bare essentials of each theory and then examine the concepts embedded in them that serve as their foundations for expecting a negative class-crime relationship. No attempt is made to either defend or criticize these theories. I only want to show that they all make entirely logical cases for expecting such a relationship by addressing criti- cisms made by Tittle implying that they do not.

ANOMIE/STRAIN THEORY

According to Tittle (1983) the tradition that most clearly posits a negative class–crime relationship is the anomie/strain tradition. The tradition began with Emile Durkheim’s theory of anomie, a societal situation that occurs in times of rapid social change and is characterized by normative deregulation. Durkheim set the stage for later extensions of anomie theory by address- ing “strains” when he wrote that although human beings are as similar in their “essential qualities, one sort of hereditary will always exist, that of natural talent” (1951: 251). This presents a problem for the individual and for society because for Durkheim human appetites are insatiable, and “No 66 Social Class and Crime living thing can be happy or even exist unless his needs are suffi ciently pro- portioned to his means” (1951: 246). Durkheim’s happiness can be seen as a ratio of expectations to accomplishments, which must at least equal unity because anything less breeds unhappiness, or in more modern criminologi- cal terminology, “strain.” When people get less than they expect they are ripe for criminal behavior. Robert Merton (1938) expanded anomie theory to develop an explana- tion of crime in an American context. The central feature of Merton’s the- ory is that American culture defi nes monetary success as the predominant cultural goal and defi nes self-worth in terms of whether or not one achieves that goal. Merton also maintained that legitimate avenues to that goal are denied to some segments of society: “It is when a system of cultural values extols, virtually above all else, certain common success-goals for the popu- lation at large while the social structure rigorously restricts or completely closes access to approved modes of reaching these goals for a considerable part of the of the same population, that deviant behavior ensues on a large scale” (1968: 200). Merton viewed anomie as a permanent condition caused by this the dis- junction between the cultural goal and structural impediments to achieving it. It is in this anomic gap that crime is bred. Being unable to attain cultur- ally defi ned wants legitimately invites frustration (strain) and may result in efforts to obtain them illegitimately. Merton claims that American culture and social structure actually exert pressure on some people to engage in nonconforming behavior rather than conforming behavior. Thus, society is the cause of anomie, not the victim of it as in Durkheim’s view. Merton identifi ed fi ve modes of adaptation that people adopt in response to this anomic condition. Conformity is the most common mode because most people have the means of legally attaining cultural goals at their dis- posal. Conformists accept the success goals of American society, and the prescribed means of attaining them (hard work, education, persistence). Ritualism is the adaptation of the nine-to-fi ve slugger who has given up on ever achieving material success, but who nevertheless continues to work within legitimate boundaries because he or she accepts the legitimacy of the social opportunity structure. Retreatism is adopted by those who reject both the cultural goals and the institutionalized means of attaining them. Retreatists drop out of society and often take refuge in drugs, alcohol, and transience and are frequently in trouble with the law. Rebellion is the adap- tation of those who reject both the goals and the means of American society but wish to substitute alternative legitimate goals and alternative legitimate means. Innovation is the adaptation of the criminal who accepts the cul- tural goals of monetary success but rejects legitimate means of attaining them. Crime is thus an innovative avenue to success—a method by which the deprived get what they have been taught by their culture to want. Merton makes it clear in several places that those from the lower classes were most likely to be the innovators and retreatists because it is on them The Class-Crime Relationship in Criminological Theories 67 that the social structure and the culture make “incompatible demands” (1968: 200). However, he also believed that the lower-middle class also had the same incompatible demands made on them: “If we should expect lower-class Americans to exhibit [innovation]—to the frustrations enjoined by the prevailing emphasis on large cultural goals and the fact of small social opportunities, we should expect lower-middle class Americans to be heavily represented among those making Adaptation III, ‘ritualism’” (1968: 205). The difference between these two sets of individuals faced with the same “incompatible demands,” according to Merton, is the strong empha- sis on conformity stressed in the socialization experiences of lower-middle- class children that produces a different “character structure” (1968: 205) from that which is produced in the process of lower-class socialization. Tittle claims that Merton assumes that “[i]nnovative [criminal] behavior only occurs when structural strain is experienced by those who are mor- ally uncommitted to society’s norms . . .” and that this “moral inhibition is most prevalent among the lower classes, hence they should commit more crime” (1983: 337). Tittle then goes on to list the assumptions of anomie theory relevant to the class–crime relationship: (1) poor people are less moral because of ineffective parenting, (2) the lower classes envy the higher classes and want to be like them, and (3) that upper class individuals rarely aspire beyond their means (1983: 339). These assumptions are the result of class-linked prejudice according to Tittle, although he does not say, much less show, that they lack empirical support.

SUBCULTURAL AND OTHER EXTENSIONS OF THE ANOMIE/STRAIN TRADITION

Albert Cohen (1955) proposed a explaining how lower-class youths adapt to the supposed limited avenues of success open to them. Cohen maintained that most lower-class crime is not a rational method of acquiring assets, but is rather an expression of short-run hedo- nism in which actors seek immediate gratifi cation of their desires without regard for any long-term consequences. He also asserts that much delin- quent behavior is non-utilitarian, malicious, and negative in the sense that it turns middle-class norms of behavior upside down by destroying rather than creating. Because many lower-class boys cannot adjust to what Cohen calls mid- dle-class measuring rods, they experience a status frustration and spawn an oppositional subculture with behavioral norms that are consciously contrary middle-class norms. Cohen saw criminal subcultures as a kind of mass reaction formation to the problem of status frustration, not to blocked opportunities to fi nancial success. Cohen saw that lower-class youths desire approval and status like everyone else, but seek it via alter- nate means “new norms, new criteria of status which defi nes as meritorious 68 Social Class and Crime the characteristics they do possess, the kinds of conduct of which they are capable” (Cohen, 1955: 66; emphasis original). These status criteria are most often physical in nature, such as being ready to respond violently to challenges to one’s manhood, or gaining a reputation as a “stud.” Tittle (1983: 342) criticizes Cohen’s work as simply taking “for granted that lower class people have undesirable traits . . . low aspirations, no abil- ity to defer gratifi cation, no sense of individual responsibility. . . .” He goes on to say without the “class mythology” serving as a starting point Cohen’s theory would not predict a class-crime relationship. Again, Tittle does not show that Cohen is wrong in claiming that these traits are more common among the lower classes, but rather he takes it for granted that they do not, thus falling into the same assumptive trap that he accuses Cohen and others of falling into. Robert Agnew’s (1992) extensions of the anomie/strain tradition (gen- eral strain theory [GST] and his “super traits” theory) did not exist when Tittle wrote his critique, but both continue the tradition of predicting a negative class-crime relationship. For Agnew strain is primarily the result of negative emotions that arise from negative relationships with others, not from blocked opportunities to success, but these negative relationships impact individuals in the lower classes most. Multiple strains are expe- rienced throughout life, but their impact differs according to their mag- nitude, recency, duration, and clustering (miseries that cluster together produce a whole greater than the sum of its parts and may overwhelm coping resources). Agnew tells us that strain can result in crime and delin- quency through the development of a generally negative attitude about other people, a lowering of social control, and a tendency to respond to strain in an aggressive manner. GST asserts that the presence of strain is less important than how one copes with it. The traits that differentiate people who cope poorly from those who cope well include “ . . . temperament, intelligence, creativity, problem-solving skills, interpersonal skills, self-effi cacy, and self-esteem. . . . These traits affect the selection of coping strategies by infl uencing the indi- vidual’s sensitivity to objective strains and the ability to engage in cognitive, emotional, and behavioral coping” (Agnew 1992: 71). Not only do these traits help to determine how one copes with strain, but, for obvious rea- sons, they also help to determine the level of success—the class level—that one achieves. Agnew’s (2005) “super traits” theory identifi es fi ve life domains— personality, family, school, peers, and work—that interact and feedback on one another across the lifespan that contain possible crime generating factors. He suggests that personality traits set individuals on a particular developmental trajectory that infl uence how other people in the various social domains react to them in evocative rGE fashion. In other words, personality variables “condition” the effect of social variables on criminal behavior. The Class-Crime Relationship in Criminological Theories 69 Agnew identifi es the traits of low self-control and irritability as his super traits and asserts that people who possess them are likely to evoke negative responses from family members, schoolteachers, peers, and workmates that feedback and exacerbate those tendencies (the multiplier effect in evocative rGE). Agnew states that “biological factors have a direct affect on irritabil- ity/low self-control and an indirect affect on the other life domains through irritability/low self-control” (2005: 213). He goes on to state that “Indi- viduals from low-SES families may be more likely to inherit these traits, as these traits may be more common among low-SES individuals” (2005: 143). Agnew further asserts that low-SES individuals may not only be more likely to inherit irritability/low self-control but are also be more likely than higher SES individuals to suffer biological insults such as prenatal exposure to toxic substances, birth complications, poor maternal health, and so on, that contribute to these traits.

WALTER MILLER’S FOCAL CONCERNS

Anthropologist Walter Miller (1958) took issue with Cohen’s assertion that gangs are formed as a reaction to status deprivation. Criminals and delinquents may resent the middle class, but It is not a matter of “If you can’t join ‘em, lick ‘em,” because their resentment is born out of envy for what middle-class people have, not for what they are. Middle-class traits such as hard work, stable habits, and responsibility are not appealing to them. Miller asserted that lower-class values must be viewed on their own terms and not as simple negations of those of the middle class. He identi- fi ed six focal concerns that are part of a value system and lifestyle that has emerged from the realities of life on the bottom rung of society. These interrelated focal concerns are trouble, toughness, smartness, excitement, fate, and autonomy. The hard-core lower-class lifestyle typifi ed by these focal concerns catch those engaged in it in a web of situations that virtually guarantee delinquent and criminal activities. The search for excitement can lead to confl ict with other gangs and with the law (trouble), and to sexual adventures in which little preventative care is taken (fate). The desire to be free of restraints and responsibilities (autonomy) is likely to preclude marriage if pregnancy results. Many lower-class males thus grow up in homes lacking a father fi g- ure, leaving them with little supervision and leading them to seek their male identity in what Miller (1958: 14) calls “one-sex peer units” (male gangs) as the culture recycles over and over. Tittle criticizes Miller’s lack of a random sample, even though it was a huge study sponsored by the National Institute of Health and conducted by Miller and seven trained social workers who maintained daily contact with black and white groups “for a total time period of about thirteen worker years” (Miller 1958: 6). In one of his oddest criticisms, Tittle (1983: 70 Social Class and Crime 341) asks if the middle and upper classes embrace values opposite those of the lower class: do they “value weakness, stupidity, boredom, and depen- dency?” Miller (1958: 6) explicitly states that these focal concerns are not confi ned to the lower classes; it is their patterning (their meaning and the ways they are expressed) “that differs both in rank order and weighting.” Presumably we all understand that toughness, excitement, smartness, and autonomy mean vastly different things to middle- and lower-class individu- als. Miller’s team spent years interacting with their subjects (not simply surveying them in classrooms), yet their 8,000 pages of observations and interviews are dismissed as class mythology that pathologizes lower-class culture. Surely asking people to tell you how they behave by marking a spot on a questionnaire is a less effective strategy than observing how they actually do behave. Miller’s work is given strong support by Elijah Anderson’s ethnographic study of African American neighborhoods in Philadelphia (1999). Ander- son asserts that the concentration of disadvantages found in those areas has spawned a hostile oppositional culture spurning most things valued by middle-class America, as in “rap music that encourages its young lis- teners to kill cops, to rape, and the like” (1999: 107). He points out that there are many “decent” families in these neighborhoods, but that the cul- tural ambiance is set by “street” families, which often makes it necessary for decent people to “code-switch” (adopt street values) to survive. Striv- ing for education and upward mobility is viewed as “dissing” the neigh- borhood, and street people do what they can to prevent their “decent counterparts from . . . ‘acting white’” (Anderson 1999: 65). Anderson supports Cohen and Miller in his contention that the street code is pri- marily a campaign for respect (“juice”) achieved by exaggerated displays of manhood, defi ned in terms of toughness and sexuality. These displays contribute greatly to the high rates of violence and out-of-wedlock births in the kinds of neighborhoods Anderson (and Cohen and Miller before him) describes.

THE CHICAGO SCHOOL OF ECOLOGY

Chicago school of human ecology was developed in the 1920s and 1930s primarily through the works of Clifford Shaw and Henry McKay (1972). Ecologists view the city as a kind of super organism with natural areas that are adaptive for different ethnic groups (little Italy’s, Chinatowns, etc.). When natural areas are eroded by “alien” ethnic groups, increased deviant behavior ensues. Viewing the Cook County Juvenile Court records spanning the years from 1900 to 1933, Shaw and McKay noted that the majority of delinquents always came from the same neighborhoods and that the ambience of some neighborhoods facilitated crime independent of other factors. It was not claimed that residential areas “caused” crime, The Class-Crime Relationship in Criminological Theories 71 but rather that it was heavily concentrated in lower-class neighborhoods regardless of the ethnic identities of their residents. Shaw and MacKay identifi ed social disorganization (the breakdown of the power of informal rules to regulate conduct) as the mechanism through which neighborhoods exerted their criminogenic infl uence. Social disor- ganization is created by the continuous redistribution of neighborhood populations as outsiders bring with them a variety of cultural traditions sometimes at odds with traditional American middle-class norms of behav- ior. A neighborhood invaded by members of an alien ethnic group became rife with confl icting values and conduct norms and loses its sense of com- munity. Social disorganization impacts antisocial behavior both by the lack of informal social controls which inhibit it, and by the generation of a set of antisocial values that develop to fi ll the vacuum left by the loss of prosocial values. Slum youths thus have both negative and positive inducements to crime and delinquency, represented by the absence of social controls and the presence of delinquent values, respectively. These conditions and values are transmitted across generations until they become intrinsic properties of the neighborhood. The basis for predicting a negative class-crime relationship is crystal clear in this theory, yet Tittle argues against it by saying that “lower class people become middle class people simply by change in economic resources and geographic movement” (1983: 344). I absolutely agree, and so did Shaw and McKay. People who possess the personal qualities needed to move up the class ladder shed one class mantel and don another. Indeed, the major premise of this book is that theories that posit the negative class-crime relation must come to grips with the fact that some people possess traits that move them in prosocial directions and up the class ladder, and oth- ers possess traits that dispose them to antisocial behavior which tend to move them down the class ladder, or to remain there if already present. But this fact cannot be used to deny that ecological theory logically predicts a negative class-crime relationship. Recall that the theory itself was driven by Shaw and McKay’s multiyear observations that the highest rates of offi - cial delinquency were always found in the same neighborhoods, and that these neighborhoods were inhabited by the lowest classes. Thus, predict- ing a negative class-crime relationship is “predicting” something Shaw and McKay already knew existed. Shaw and McKay’s theory was devised to explain the observed class-crime relationship, and they made no pejorative “assumptions” about anything. Tittle also criticizes the theory on the grounds that it assumes that all lower-class people live in cities and neglects the “enormous amount of rural and small town poverty” (1983: 345). A study of 264 rural counties in four states by Osgood and Chambers (2003) rectifi es this for us. Osgood and Chambers also traced social disorganization to high rates of population turnover and high levels of ethnic diversity, but the most important fac- tor was high rates of female-headed households. Single-parent households 72 Social Class and Crime were highly related to the kinds of problems Shaw and McKay previously identifi ed as related to social disorganization (i.e., low income, dilapidated housing, and poor supervision of children). The most notable fi nding of the study was that “a 10 percent increase in female-headed households was associated with a 73- to 100-percent higher rates of arrest for all offenses except homicide [a 10% increase in female-headed households was associ- ated with a 33% increase in homicide]” (Osgood and Chambers 2003: 6). Again, this clearly shows a negative class-crime relationship mediated by the presence of a powerful intervening variable (female-headed households).

DIFFERENTIAL ASSOCIATION THEORY

Differential association theory, the brainchild of Edwin Sutherland, focuses on the process of becoming delinquent via attitude formation. Sutherland’s theory takes the form of nine propositions outlining the process by which individuals come to acquire attitudes favorable to criminal or delinquent behavior, as well as the specifi c techniques involved in it. The following four points adequately summarize his nine propositions:

(1) Criminal behavior is learned just like any other behavior. (2) Learning takes place in intimate personal groups. (3) The content of that learning includes both behavioral techniques and a set of supporting motives, attitudes, and values. (4) The social setting in which this learning takes place is one of norma- tive confl ict and differential social organization.

The key proposition in the theory is: “A person becomes delinquent because of an excess of defi nitions favorable to violations of law over defi ni- tions unfavorable to violations of law” (Sutherland and Cressey 1974: 77). Associations with others holding defi nitions favorable to violation of the law vary in frequency, duration, priority, and intensity. That is, the earlier individuals are exposed to criminal norms of conduct, the more often they are exposed to them, the longer those exposures last, and the more strongly they are attached to people who supply us with the defi nitions favorable to law violation, the more likely they are to commit criminal acts when oppor- tunities to do so present themselves. Children associate with and became friendly with other children in the neighborhoods they fi nd themselves in. If children fi nd themselves in a neighborhood characterized by what Suther- land called differential social organization (he found the phrase social dis- organization to be insulting to lower-class people), they will be surrounded by peers with “defi nitions favorable to law violation” and cannot but help but be infl uenced by them regardless of their individual traits. Sutherland’s theory predicts a negative class-crime relationship for much the same reasons as ecological theory. Poorer people are thrown together in The Class-Crime Relationship in Criminological Theories 73 run-down neighborhoods lacking strong informal controls and lacking what modern ecologists call “collective effi cacy”—the ability of the community to bring its members together to organize strategies to combat community problems such as crime. Sutherland simply added the social-psychological process of learning “defi nitions favorable to law violation” to the ecologi- cal tradition. For Sutherland and Cressey (1974: 227) social class “affects crime and criminality [by determining] associations with criminal behavior patterns or isolation from anticriminal behavior patterns.”

THE MARXIST AND CONFLICT TRADITIONS

The core of Marxist philosophy is the concept of class struggles between the haves and have-nots throughout recoded history (Marx and Engels 1948). The haves in Marx’s time were the owners of the means of production (the bourgeoisie) and the have-nots were the workers (the proletariat). The bourgeoisie oppressed the workers by paying them essentially starvation wages while working them long hours in dangerous unhealthy conditions. Wage-labor literally dehumanizes human beings by taking from them their creative advantage over other animals and reducing them to their level. This is the essence of Marx’s concept of alienation. When individuals become alienated from their work and from themselves they also become alienated from others and from their society in general. Alienated individuals may then treat others as mere objects to be exploited and victimized as they themselves are exploited and victimized. Marx and his collaborator, Freidrich Engels, made plain their disdain for criminals, calling them “The dangerous class, the social scum, that rotting mass thrown off by the lowest layers of the old society” (1948: 22). This “social scum” came from the very lowest class in society—the lumpenproletariat. Because Marx and Engels produced no coherent theory of crime, Tittle focused his criticisms of Marxist criminology on the work of Willem Bonger, the fi rst Marxist criminologist. Although the roots of crime for Bonger lay in the exploitive and alienating conditions of capi- talism, he believed that some individuals are at greater risk than others because of variation in their “innate social sentiments” of altruism and egoism. Bonger believed that capitalism accentuates egoism and blunts altruism because it relies on competition for wealth, status, and jobs, set- ting person against person and group against group, leaving the losers to their fates. Bonger believed that poverty was the major cause of crime, but traced poverty’s effects on family structure (broken homes, illegitimacy), poor parental supervision of their children, and “the lack of civilization and education among the poorer classes” (1969: 195). Bonger certainly makes a number of comments in the less politically cor- rect times in which he wrote (“lack of civilization,” etc.) that would be con- sidered pejorative today. He clearly believed that what he was writing was 74 Social Class and Crime the truth bluntly stated, and his theory clearly and unambiguously predicts a negative class-crime relationship based on high levels of female-headed households, low levels of supervision of children, low levels of education, and “lack of civilization” (poor socialization practices) among the lower classes. All of the factors are realities, as a cascade of research has docu- mented for decades, and not merely baseless pejorative assumptions. The confl ict tradition traced to the works of also predicts a negative class-crime relationship, but for different reasons. Even though individuals and groups enjoying wealth, prestige, and power have the resources necessary to impose their values and vision for society on oth- ers with fewer resources, Weber also viewed the class divisions in society as normal, inevitable, and acceptable, as do many contemporary confl ict theorists. The confl ict tradition posits a negative class-crime relationship by virtue of the fact that the upper classes have the power to criminalize the acts of the lower classes that are contrary to their interests and to ignore acts of the upper classes that are contrary to the interests of the lower classes. Confl ict theorists see social life as a continual struggle to maintain or improve one’s own group’s interests—workers against management, race against race, ecologists against land developers, and the young against adult authority—with new interest groups continually forming and disbanding as confl icts arise and are resolved. Confl icts between youth gangs and adult authorities are of particular concern because such gangs are in confl ict with the values and interests of just about every other interest group, including those of other gangs. Gangs are examples of minority power groups, or groups whose interests are suffi ciently on the margins of mainstream soci- ety that just about all their activities are criminalized.

THE UTILITARIAN/DETERRENCE TRADITION

The fi nal theoretical tradition that Tittle takes to task is the utilitarian/ deterrence tradition derived primarily from the classical work of (1789/1948). Bentham’s model of human motivation and behavior was straightforward: “Nature has placed mankind under the governance of two sovereign masters, pain and pleasure. It is for them alone to point out what we ought to do, as well as to determine what we shall do” (1789/1948: 125). The traditional view of human nature is that it is hedonistic, rational, and endowed with free will. Before committing themselves to any course of action, including criminal action, people weigh the ratio of costs (what they stand to lose) to benefi ts (what they stand to gain) and will proceed if the ratio favors the benefi t side or desist if it favors the cost side. Rational choice theory is a modern exemplar of the utilitarian/deterrence tradition. Rational choice theorists view criminal acts as specifi c examples of the general principle that all human behavior refl ects the rational pursuit of pleasure and avoiding pain. People are conscious social actors free to choose crime, and they will do so if they perceive that its utility exceeds The Class-Crime Relationship in Criminological Theories 75 the pains they might conceivably expect if discovered. The theory does not assume that we are all equally at risk to commit criminal acts or that we do or do not commit crimes simply because we do or do not “want to.” It rec- ognizes that personal factors such as temperament, intelligence, and cog- nitive style, as well as background factors such as family structure, class, and neighborhood, affect our choices, but it largely ignores these factors in favor of concentrating on the conscious thought processes involved in mak- ing decisions to offend (Cornish and Clark 1986). Rationality is thinking and behaving in accordance with logic and rea- son such that one’s reality is an ordered and intelligible system (i.e., there is a logical correspondence between the goals sought and the means used to obtain them). It is both subjective and bounded; we do not all make the same calculations or arrive at the same game plan when pursuing the same goals, for we contemplate our anticipated actions with incomplete knowl- edge, with different mind-sets, and with different reasoning abilities. Our emotions (guilt, shame, anxiety, etc.) also function to keep our temptations in check by overriding purely rational calculations of immediate gain. All people have mental models of the world and behave rationally with respect to them, even if others consider the behavior irrational. Criminals behave rationally from their private models of reality, but their rationality is con- strained, as is everyone’s, by ability, knowledge, emotional input, and time (Cornish and Clark 1986). According to Tittle (1983: 349), “The utilitarian notion leads to a nega- tive class/crime prediction when theorists provide auxiliary arguments which explain (1) That the objective cost-benefi t ratio varies directly with class posi- tion, or (2) that the classes misperceive cost and benefi t so that the perceived risk/reward ratio varies directly with socioeconomic position.” In the fi rst case, the idea drawn from the utilitarian position is we expect more crime among people from the lower classes because they have more to gain from it and less to lose. Given the similar risk/reward ratio for a given crime, we should expect few middle- and higher-class individuals to commit it because, given their earning capacity, the rewards of the crime are relatively much less than they are for the lower classes and the potential losses are relatively much greater (loss of a valued career, the shame of exposure, and so forth). In the second case, the idea is that people differ in their ability to calcu- late the cost-benefi t ratio, and that this difference varies with social class. The assumption is that even if the cost-benefi t ratio of criminal activity was the same for all social classes, the lower classes would still commit more crimes because of their inability to calculate a “true” ratio. This is because lower-class individuals are considered to be impulsive, tend to grat- ify immediate needs and neglect long-term goals, and have lower IQs. Tittle (1983: 350) asserts that these are pejorative assumptions and that the utili- tarian/deterrence model can only claim to predict a negative class-crime relationship by the “addition of external considerations, many of which are logically weak and some of which make the assumption of criminogenic lower class characteristics.” 76 Social Class and Crime CONCLUSION

Tittle (1983: 353) concludes his critique by claiming that the negative class- crime relationship is “both empirically and theoretically problematic,” and the fact that most social scientists remain “wedded to the idea that social class is a key variable in explaining criminal behavior . . . [is] a riddle.” One explanation he gives for this “riddle” is that generations of criminologists simply cite the same literature over and over again without checking origi- nal sources until statements contained in that literature become self-evident. There are certainly cases of this in criminology (how many criminologists who rail against Lombroso have really read him, or how many have delved into the concept of IQ as it is understood by psychologists, geneticists, and neuroscientists before dismissing the concept as repugnant?), but the litera- ture regarding Tittle’s “auxiliary assumptions” about social class is enor- mous, well read by social scientists, and supports theoretical predictions of a negative class-crime relationship. These factors will be discussed in the next chapter, as well as touched upon in other chapters. In the face of overwhelming evidence that it does, the argument that a negative class-crime relationship does not exist must be seen as an example of the moralistic fallacy. The moralistic fallacy is the reverse of the better- known naturalistic fallacy, the fallacy of confusing is with ought. No dis- tinction is being made between an empirical fact and the moral evaluation of it when one leaps into the naturalistic fallacy. Nature simply is; what ought to be is a moral judgment. An example of the naturalistic fallacy would be to say that because violence is part of human nature (i.e., natural), it should be approved of as a good thing. The moralistic fallacy is the opposite of the naturalistic fallacy in that it confuses ought with is (i.e., deducing a fact from a moral judgment). Those who commit this fallacy are arguing in effect that if something is morally offensive to them it is a priori false: violence is a bad thing; therefore it cannot be part of human nature. We often see other examples of this fal- lacy with such issues as race or gender differences: There ought not be such differences; therefore no such differences do “in fact” exist. Even though human equipotentiality is known to be counterfeit beyond our domain, it is still the preferred currency of social science. This is doubtless why Tittle believes that all of the auxiliary arguments attached to the crimino- logical theories he discussed such as impulsiveness, intelligence, crimino- genic peers, and inadequate socialization are evenly distributed across all class levels. This being a “fact” derived from a moral judgment of should, it follows that if claims are made that different levels of negative traits and practices are found at different SES levels, they are morally repugnant assumptions, and therefore factually wrong. 7 Social Class and Socialization

THE CONCEPT OF SOCIALIZATION

To state that something is related to criminal behavior is to specify a loca- tion where crime is frequently found. Theories offer guidance as to where we should look; that is, they place the descriptor in relation to other factors in an attempt to make the relationship intelligible. Searching for interven- ing factors is what we do when we introduce control variables into our models. With properly specifi ed models, the correlation between SES and criminal behavior should statistically disappear given that SES is a higher- order umbrella concept covering a large number of factors that contribute to it. This chapter examines intervening factors discussed in the theories presented in the previous chapter that have been proposed to explain the negative class-crime relationship. It is primarily about class-based social- ization practices related to those intervening factors such as poor parent- ing, low level of attachment, poor school performance, and negative peer infl uences. Given the moral concerns voiced by Tittle (1983), I want to emphasize that none of the material presented in this chapter should be considered an indictment of lower-class parents; they have enough adver- sity to contend with. To adapt and actively participate in one’s culture, one must learn its contents in a continuous process of socialization across the lifespan. Social- ization has been defi ned as the process of internalizing “the rules, roles, attitudes, standards, and values across the social, emotional, cognitive, and personal domains” (Grusec and Hastings 2007: 1). Most researchers in the past saw socialization as a top-down parent-to-child process, but modern researchers realize that children are not passive blank slates onto whom proscriptions and prescriptions are written and that they are active partici- pants in the cultivation of their natures. Parents infl uence their children’s development in multiple ways, but children’s infl uence on parents and the fact that this infl uence depends to some extent on children’s genes cannot be ignored. The socialization process is a reciprocal one that parents and children engage in together in which the children’s traits and abilities inter- act with the traits and abilities of socializing agents. The role of genes and of child effects on parents was almost com- pletely ignored by socialization researchers for most of the 20th century. 78 Social Class and Crime Researchers typically looked at only one child per family (more than 99% of studies according to one estimate [Plomin, Ashbury, and Dunn 2001]) when they needed to look at multiple children in the same family. This strategy alerts social scientists to what every parent who has more than one child knows—there are different parenting styles for different children. The same parent who is permissive with a warm and compliant child may be authoritarian with a bad-tempered and resistant child, while all the time trying to be the authoritative parent that child psychologists tell us that all parents should be to all offspring. Contemporary developmental scientists are now in full agreement that genes infl uence children’s behavior, and that parents calibrate their parent- ing to that behavior, but it took a long time after the fi rst cracks appeared in the reigning paradigm for them to come to this agreement. It is heartening to see that every chapter in the massive Handbook of Socialization: Theory and Research (Grusec and Hastings 2007) that dealt with childhood social- ization makes it a point to recognize the important role of genetics in social- ization research. Socialization agents other than the family—the schools and the peer group—are also important sources of behavioral output, and they also must be examined in terms of rGE and GxE. We begin with the family, the source that has fi rst shot at molding the incipient human being.

THE FAMILY AND SOCIALIZATION

The family is the adaptive feature of Homo sapiens by which parents take the raw biological material their offspring bring with them into the world and assist them to turn that material into civilized human beings capable of living in harmony with others and contributing to the general well-being. Mother and infant are “biologically prepared to be attracted to and remain in close proximity to each other” (Grusec and Davidov 2007: 285). Mater- nal empathy, underlain by oxytocin, is the foundation of this biological preparedness to accurately perceive and positively respond to children’s emotional and cognitive needs. Infants who do not receive the behavioral manifestations (holding, kissing, snuggling, etc.) of mother/infant attach- ment show signifi cantly lower base levels of oxytocin, the neurotransmit- ter/hormone that produces feelings of peace and calm (Wismer Fries et al. 2005). The bonds formed during this period are experience-expected and in their absence the relevant synaptic connections are pruned (Glaser 2000). There are many negative outcomes associated with the failure to form these bonds, including diffi culty in forming lasting bonds in adulthood (Lee and Hoaken 2007). According to Conger and Dogan (2007) there are three family-based models of how social class infl uences the socialization tactics and strate- gies of parents: the Family Stress Model (FSM), the Extended Investment Model (EIM), and the Model (SSM). The fi rst two are Social Class and Socialization 79 social causation perspectives that essentially ignore genetics but can easily be placed in biosocial perspective; the third is distinctly biosocial at the onset. The focus of each of these theories is the exploration of how the dis- advantages suffered by lower SES families compromise the ability of par- ents to socialize their children in accordance with middle-class standards of behavior. The essence of the FSM is that economic hardships such as low income and job instability lead to economic pressures such as debts, lack of health insurance, and the inability to provide adequate food, shelter, and clothing. The accumulation of these pressures may lead to emotional distress mani- fested as depression, anxiety, or anger, which may then lead to parental confl ict/and or withdrawal and child abuse/neglect. This, in turn, leads to childhood and adolescent adjustment problems, including weak attachment to parents, poor school performance, substance abuse, and delinquency. Disrupted parenting may have more to do with children’s temperamental irritability and non-compliance than parental temperament because it is yet another stressor to be dealt with. Longitudinal studies typically fi nd that temperamentally diffi cult children are most often found in low SES fami- lies (Wasserman et al. 1990). Another review of childhood temperament and social development pointed out that children from lower SES families are over-represented at the “problematic end of temperament dimensions, especially those relating to child diffi culty” (Sanson, Hemphill, and Smart 2004: 158). Infants and toddlers with diffi cult temperaments require more responsive and patient parenting, but studies have shown that lower SES mothers are relatively unresponsive to irritable infants relative to higher SES mothers or to lower SES mothers of non-irritable infants (Bates and Pettit 2007). Absent genetically informed studies, we cannot know if infants are irritable for genetic or parenting reasons or whether unresponsive mother- ing is a function of mother’s temperament or their responses to infant’s evocative behavior. The problem with the FSM as Conger and Dogan see it is that it con- centrates only on the economic dimension of SES, although it is perhaps the most important dimension. They maintain that the problem with this singular focus is that family income can fl uctuate greatly, with a signifi - cant number of households entering or leaving poverty in any given year. Thus, the FSM would seem to be most applicable to children from families mired in long-term poverty, or children from poor families who were most deprived during their most formative years. Whereas the FSM focuses on low-income families and how the stress- inducing nature of low SES has negative infl uences on the parenting of chil- dren, EIM focuses on higher-status families and how the greater resources enjoyed by them promote better parenting and greater cultivation of their children’s talents and abilities. In many respects, the EIM is also mainly an economic model because it focuses strongly on the investment in resources that higher SES families provide their children. Higher SES families live in 80 Social Class and Crime better neighborhoods where their children will play and become friendly with prosocial children and will enjoy many prosocial neighborhood com- munity activities, all of which, consciously or not, aid parent’s prosocial socialization strategies. In addition to the economic investment made possible by family income, higher-status parents enjoy higher educational level and greater occupation prestige than parents in the lower classes. These additional advantages aid in greater parent-child attachment and more successful socialization for reasons quite independent of income. Their greater education, for instance, would make higher SES parents more aware of the benefi ts of education for their children, and also of a variety of practices that support the overall health and well-being of their children (not drinking and smoking dur- ing pregnancy, breastfeeding, the importance of tactile stimulation, and so forth). Higher-level occupations also imply more workplace autonomy and are thus more likely to lead parents to value and promote self-direction in their children. Conger and Dogan also maintain that because lower SES par- ents are more often under strict supervision at work, they are more likely to demand unquestioning obedience from their children and to employ physi- cal rather than alternative disciplinary measures when it is not forthcoming. They also aver that lower-class parents are more likely to punish children for what they have done regardless of intent, while higher- status parents are more likely to focus on intent (accident vs. maliciousness, justifi able vs. unjustifi able, etc.). A large-scale study by Dodge, Pettit, and Bates (1994: 650) succinctly summarizes much of what the FSM and EIM empirically expect to fi nd:

Socioeconomic status was signifi cantly negatively correlated with 8 factors in the child’s socialization and social context, including harsh discipline, lack of maternal warmth, exposure to aggressive adult mod- els, maternal aggressive values, family life stressors, mother’s lack of social support, peer group instability, and lack of cognitive stimula- tion. . . . These fi ndings suggest that part of the effect of socioeconomic status on children’s aggressive development may be mediated by status- related socialization experiences.

The fi nal socialization model that Conger and Dogan (2007) discuss is the SSM. The SSM does not contend with the other two models about the importance of the factors they emphasize, but rather it augments them. This model asserts that parents not only provide their children with envi- ronments but also a range of other non-economic resources such as genes, attitudes, values, and life priorities. In other words, the SSM contends that individual differences among parents that have infl uenced their own status levels are passed on to children both genetically and environmentally (pas- sive rGE). These genetic and environmental advantages and disadvantages will slot their children into a life trajectory that will infl uence the children’s Social Class and Socialization 81 eventual SES. Conger and Dogan (2007: 451) go on to write that “once these individual differences among caregivers are taken into account, the theory proposes that there will be no direct relationship between SES and either socialization strategies or child adjustment.” One of the best examples of the model is the behavioral genetic lon- gitudinal study of 1,116 fi ve-year-old twin pairs by Kim-Cohen and her colleagues (2004). Their purpose was to discover what contributed to chil- dren’s resilience in the face of SES deprivation. The research team found that lower SES was positively associated with antisocial behaviors and negatively associated with IQ, but some children were resilient to the pro- cesses that increase antisocial behavior and depress IQ. It was found that 70.6% of the variance in behavioral resilience and 45.8% of the variance in cognitive resilience was accounted for by genetic factors. The behavioral measures contributing to resilience were children’s sociable temperament and, consistent with both EIM and FSM, parental warmth and the provi- sion of stimulating activities. Passive and evocative rGE are obvious explanations for resilience given these heritability estimates. Children with outgoing temperaments prob- ably have parents with the same kind of temperament, and such parents are likely to respond with warmth and thus reinforce their children’s personali- ties. Cognitive resilience is likely to also be the result of rGE in terms of parental responsiveness to, and cognitive stimulation of, their children. In other words, parents supply the genetic and environmental ingredients for the development of resilience, which further indicates that positive adaptive capacities run in families. Molecular genetic studies are also looking at resilience in the face of life stresses. Caspi and his colleagues’ (2003) longitudinal birth cohort study assessed subjects at age 26 for the effects of lifetime stressful events and maltreatment on major and minor depressive episodes. The researchers were concerned with a polymorphism in the 5-HTT (sero- tonin transporter) gene called the 5-HTTLPR, which has short (s) and long (l) versions. Low 5-HTTLPR expression is assumed to lead to less- effi cient serotonin functioning. In terms of maltreatment (categorized as “none,” “probably,” and “severe”), subjects who were homozygous for the short allele (s/s) showed a dramatic increase in the probability of a major depressive episode from the “none” to “severe” category. The prob- ability did not increase at all for subjects homozygous for the long allele (l/l) across the maltreatment categories, and heterozygotes (s/l) were inter- mediate in terms of probability increase. The same pattern emerged for all analyses (i.e., the l/l allele confers protection against adversity, the s/l confers some protection, and s/s individuals are particularly vulnerable to adversity). This is another example of GxE in that the main effects of genotype were not signifi cant absent stressful events (the main effects for stressful events were weakly signifi cant, however). That is, genotype only mattered in the face of environmental insults. 82 Social Class and Crime THE SCHOOLS AND SOCIALIZATION

The social life of children changes at different stages of their development, with their entry into school probably being the most dramatic. School is the socialization agent intermediate between the passive rGE processes of infancy and toddlerhood and active rGE processes of adulthood. It is that period in which others outside of the family (teachers, neighbors, and peers) react to individuals on the basis of their behavior (evocative rGE), thus magnifying differences between individuals in terms of personality and cognitive style even as they strive to make them more alike in confor- mity to the norms and goals of the general culture. It is in the school that children must come to terms with having to simultaneously adapt to the formal structural expectations of the school and to the informal structural expectations of their peer group culture. Children also face for the fi rst time the prospect of having to achieve a status rather than having one simply ascribed, such as son, daughter, brother or sister. They now have to compete with others to gain status, and they must do it coming from different social backgrounds and with different levels of ability and motivation. Albert Cohen long ago remarked on the potential role of the school experience on children from different socials classes in producing future delinquents. Cohen noted: “One of the situations in which children of all social levels come together and compete for status in terms of the same set of middle-class criteria and in which working-class children are most likely to be found wanting is the school” (1955: 112). Even if children attending a school are mostly from the working and lower class, school personnel are middle class and will evaluate students with their middle-class measuring rods. All children of all classes will be evaluated not only on their schoolwork but according to how well they exhibit middle-class values and behaviors such as ambition, self-control, rational forethought, courteous and respectful behavior, and a sense of responsibility. In Cohen’s terms, failure to meet the cognitive and behav- ioral standards required to live up to the middle-class standards of the school breeds “status frustration” and the formation of groups of similar individuals who devise an alternative system of values that will eventually foster delinquency and criminal behavior. Of course, many children from the lower and working classes will indeed assimilate the set of middle-class values and attitudes that will see them in good stead in adulthood in school and will have them reinforced if they are already encouraged at home. In attachment to school is second only to attach- ment to parents in importance to later adult commitments to a prosocial lifestyle. Academic competence increases teachers’ positive evaluations of students, which reinforces student motivation and contributes to prosocial self-esteem (as opposed to the “street juice” self-esteem that seeks valida- tion in violence and promiscuity), which further increases competence in Social Class and Socialization 83 a felicitous achievement-self-esteem spiral (Wentzel and Looney 2007). Academically competent students are thus likely to have positive feeling about school (develop bonds) that will greatly aid in their preparation for high-status occupations in the future. Conversely, academic failure weak- ens social bonds with the school and results in poor occupational oppor- tunities, which in turn may result in alienation from any sort of prosocial lifestyle (Elrod, Soderstrom, and May, 2008). As mentioned previously, it is not only academic competence but also middle-class attitudes and behaviors that the school attempts to impart and/or reinforce. Economic success can only be partially explained by the cognitive skills fostered in schools. Economists Bowles and Gintis (1976, 2002) have long argued that the formal and informal teaching of social relationships and the “proper” attitude is as much a central role of the schools as academic learning. Bowles and Gintis are openly hostile to capi- talism and claim that schools are designed to prepare students to adapt to the hierarchical and authoritarian world of the capitalist workforce. I agree completely with this, but without the negative connotations Bowles and Gintis place on it. Nevertheless, it is true that any competent agent of socialization under any form of economic system is supposed to impart the necessary skills and attitudes toward the culture its neophytes will eventually enter. Contrary to Bowles and Gintis, Durkheim wrote positively about the schools as fash- ioning individuals to society’s needs, and “for the new being that collective infl uence, through education, thus build up in each of us, represents what is best in us” (1956: 76). Schooling does affect labor market success, and if theories such as anomie/strain have any utility, it also affects the probability of criminal behavior. Bowles and Gintis agree that ambition, self-control, courteous and respectful behavior, and conscientiousness are the values schools try to impart and that they are the values necessary for success in a capitalist economy. Nowhere in their writings have I found that they distain these values even as they disdain the system that demands them. For our purposes, however, the point is that Bowles and Gintis frequently show in their works that these non-cognitive personality traits are as important to achieving economic success as is cognitive ability, and are more so in the less cognitively demanding occupations.

PEER GROUPS AND SOCIALIZATION

While it is debatable whether or not peers are the primary socialization agents from the moment children enter school (Harris 1998), few would dispute the primacy of peers in that role during adolescence. For evolution- ary reasons, all adolescent primates, including humans, are predisposed to affi liate with and adapt to a peer group and to slowly distance themselves from the family group (Rosenfeld and Nicodemus 2003). There is much to 84 Social Class and Crime learn about being an adult, and adolescence is a time to experiment with a variety of social skills before putting them into practice in earnest, and a time of sorting out what aspects of the previous generation to retain and what to discard. Adolescents are thus re-socialized to a somewhat different culture because they need to experiment with skills needed to live in their own generation, not that of their parents. The adolescent peer group con- sequently becomes an extremely important source of behavioral learning. As with parent-child socialization, however, we must again be careful of positing a passive group-to-individual socialization process shorn of indi- vidual agency. There are many different age peer groups to which individuals can attach themselves, and the groups chosen will have implications for their future success in life. For obvious reasons, individuals tend to attach themselves to the groups existing in their immediate neighborhoods, and this is where the human composition of the neighborhood matters. As ecological theory tells us, some neighborhoods have long traditions of spawning antisocial members. Although practically all adolescent peer groups engage in behaviors decried by their elders, some peer groups (gangs) engage in serious delinquent behavior to the extent that members seriously compromise their ability to climb the class ladder and to become productive members of society. Regardless of neighborhood characteristics, there are peer group choices, although they are more limited in some neighborhoods than in others. Various reviews of friendship patterns (e.g., Rodkin et al. 2000) have shown that a propensity for a given pattern of activity precedes association with individuals similarly disposed. “Birds of a feather fl ock together,” and fl ocking together makes them more alike. A longitudinal study found that as children matured and created their own mini-worlds genes played an increasingly larger role in peer choice (Kendler et al. 2007). Whatever antisocial propensities children may start out with, those pro- pensities are exacerbated during adolescence. We cannot understand why this is so without delving into the biology of puberty. Natural selection has provided adolescents with the necessary tools to engage in experimentation in preparation for adulthood, such as the huge increase in testosterone (T) accompanying puberty (Felson and Haynie 2002). The male brain is orga- nized by T in utero after which there is little difference between the sexes in levels of T until puberty when the second surge of T activates the male brain to engage in male-typical behavior (Ellis 2005). The T surge facili- tates the adolescent increase in experimental behaviors such as risk tak- ing, sexual and dominance contests, and self-assertiveness. These behaviors may take prosocial or antisocial directions according to social contexts. The interplay between T and social context is illustrated in a longitudinal study of 1,400 boys which found that T levels were unrelated to antisocial behavior for boys with “non-deviant” or “possibly deviant” friends, but such behavior was greatly elevated among boys with high T who associated Social Class and Socialization 85 with “defi nitely deviant” peers (Maughan 2005). This GxE interaction shows that T has a facilitative, not causal, role for antisocial behavior. The pubertal hormonal surge also prompts the increase of gene expres- sion in the brain to begin the process of slowly refi ning neural circuitry to adult form (Sisk and Zehr 2005). During adolescence there are changes in the ratios of excitatory to inhibitory neurotransmitters, with the former increasing and the latter decreasing (Collins 2004; Walker 2002). Ernst, Pine, and Hardin (2006: 299) explain risky adolescent behavior in terms reminiscent of reward dominance theory: “The propensity during adoles- cence for reward/novelty seeking in the face of uncertainly or potential harm might be explained by a strong reward system (nucleus accumbens), a weak harm-avoidance system (amygdala), and/or an ineffi cient supervisory system (medial/ventral prefrontal cortex).” Functional magnetic resonance imaging studies have shown that relative to children and adults, adolescents have exaggerated nucleus accumbens activity feeding into the prefrontal cortex (Eshel et al. 2007; Galvan et al. 2006). Further exacerbating the adolescent tendency toward risky behavior is the fact that the PFC undergoes a wave of synaptic overproduction just prior to puberty followed by a period of pruning during adolescence and early adulthood (Giedd 2004; Sowell et al. 2004). These changes are par- ticularly signifi cant in the PFC: “Signifi cant changes in multiple regions of the prefrontal cortex [occur] throughout the course of adolescence, espe- cially with respect to the processes of myelination and synaptic pruning” (Steinberg 2005: 70). The incompleteness of the myelination process in the adolescent PFC accounts for the “time lapse” between the perception of an emotional event in the limbic system and the rational judgment of that event in the PFC. Adolescents are thus provided with all the biological tools (increased T and excitatory chemicals and decreased inhibitory chemicals and delayed myelination) that favor increased novelty seeking, status seek- ing, and competitiveness. All these changes strongly suggest that the bio- logical mechanisms activated during adolescence are adaptations forged by natural selection precisely to assist the young in pursuing those behaviors (Spear 2000; White 2004). This brief survey of the neurobiology of adolescence informs us that there are physical reasons for the immature and sometimes antisocial behavior of adolescents. The immaturity of the teen brain combined with a teen physiology on fast-forward facilitates the tendency to assign faulty attributions to the intentions of others. In other words, a brain on “go slow superimposed on a physiology on fast forward” explains why many teen- agers “fi nd it diffi cult to accurately gauge the meanings and intentions of others and to experience more stimuli as aversive during adolescence than they did as children and will do so when they are adults” (Walsh 2002: 143). As neurologist Richard Restak (2001: 76) put it: “The immaturity of the adolescent’s behavior is perfectly mirrored by the immaturity of the adolescent’s brain.” 86 Social Class and Crime AN INTEGRATIVE LONGITUDINAL COHORT STUDY

An interesting study that ties together most of what we have been saying about SES and other correlates of antisocial behavior was conducted by Fergusson, Swain-Campbell, and Horwood (2004). This study is particu- larly useful given that it is a large (n = 1,265) longitudinal study of a birth cohort, which is the sampling “gold standard” for social science. It is also particularly informative in that the majority of the information obtained from subjects, parents, and teachers was verifi ed via multiple sources such as police records and the records of other state agencies. Children’s parental SES was assessed three times (at birth, age 3, and age 6), and children were classifi ed into one of six categories ranging from professional (class 1) to the lowest income category, consisting primarily of children from father-absent homes (class 6). Although Fergusson and his colleagues presented multiple comparisons across all six class categories, I only present the comparisons between the highest and lowest class categories. Class-based outcomes were assessed when the cohort age was 10, 16, 18, and 21. Examining Table 7.1 we fi rst note that the percentage of children from the lowest class with a parental history of offending was 10 times greater than the percentage of children from the highest class with such a parent. We could certainly be seeing the effect of passive rGE at work here, but do also take note that this also means that 65.5% of the children from the lowest class did not have a parent with a history of criminal offending (it would have been most interesting if the researchers compared class 6 sub- jects who did and who did not have a parent with a criminal history, but alas they did not). The incidence rate ratios (IRR) are particularly interesting. The IRR for self-reported offending shows that subjects from class 6 self-reported a rate of offending 3.21 times greater than subjects from class 1. However, the IRR for offi cially recorded convictions (at age 21) for class 6 subjects was 25.82 times greater than for class 1. The comparison of these two IRRs (3.21 and 25.82) supports the contention mentioned in Chapter 5 that those most seriously involved in offending are those most likely to under report their offending. Although data from classes 2, 3, 4, and 5 are omitted here, all variables showed almost perfect monotonic increases in negative outcomes moving from class 1 through class 6. The unequivocal pattern presented by these data tells us that with decreasing SES there are increases in family adver- sity (abuse, low attachment and maternal care, family changes, and paren- tal criminality), childhood adjustment problems (conduct disorder and attention problems), school problems (truancy, suspension, low scholastic achievement), and affi liation with peers with similar delinquent/criminal problems. We thus observe an accumulation of stress and strains that often result in antisocial behavior and failure in the workforce. It is in this way that SES reproduces itself across the generations in the worst families in which all negative infl uences appear to consolidate. Social Class and Socialization 87 Table 7.1 Comparison of Crime-Related and Individual, Family, School, and Peer-Related Factors by SES

Highest Lowest Variable Class (1) Class (6)

Crime-Related Factors % with parent with history of offending 3.3 34.5 % in highest quartile of deviant peer affi liation* 17.9 41.4 % in highest quartile conduct problems** 7.4 45.8 IRR self-reported offending* 1.0 3.21 IRR offi cial record of conviction 1.0 25.82 Individual, Family, School, and Peer-Related Factors % regular or severe physical punishment 3.6 15.9 % lowest quartile maternal care 13.5 40.3 % lowest quartile parental attachment 11.4 36.9 % highest quartile family change 8.8 44.7 % highest quartile school truancy 10.4 54.4 % ever suspended from school 4.2 22.8 % lowest quartile scholastic ability 5.3 49.3

Adapted from Tables 2 and 4. in Fergusson, Swain-Campbell, and Howard (2004). How Does Childhood Economic Disadvantage Lead to Crime? Classes 2, 3, 4, and 5 omitted in compari- sons. All comparisons signifi cant at < .0001. *Assessed at 21 years of age. **Assessed at 10 years of age.

The most salient point of this study is that SES per se contributes essen- tially no independent variance to the prediction of criminal behavior once the effects of all other predictor variables are entered into negative bio- nomial regression models. The IRR of 3.21 for self-reported offending reduces to 1.23, and the 25.82 for offi cially recorded convictions fell to 1.93 with all family, individual, school, and peer group factors included in the equation. This study shows once again that SES is a label for a host of advantages and disadvantages. These advantages and disadvantages are both genetic and environmental and multiply and aggregate via the pro- cesses of rGE and GxE. There are those who interpret a statistical relationship that becomes non-signifi cant after control variables are introduced as indicating that the initial relationship was spurious. Controlling for variables that reduce the direct effects of class to zero undermines direct causation, but not neces- sarily indirect causation. After all, class differences in student grade dis- tributions would statistically disappear if we controlled for intelligence, motivation, and time spent studying, as would the difference between cooked and uncooked eggs if we controlled for heat. While these factors 88 Social Class and Crime explain the different outcomes to the satisfaction of most, students still get the grades they get, eggs remain cooked or uncooked, and lower-class indi- viduals still commit more crimes regardless of the intervening variables that explain why they do. Spuriousness exists if the dependent variable (crime) and the independent variable (SES) are only coincidentally connected which is not the case with the SES-crime relationship. SES may be thought of as the sum of all the predictors in the preceding study plus others not included (X), just as Grades = intelligence + motiva- tion + time spent studying + X. SES can be considered an indirect “cause” of criminal behavior if we consider SES to be the principle component derived from a factor analysis of all component variables that contribute to it. Nevertheless, analysis of the component parts is still a far more useful strategy because without exploring further we expose ourselves to accusa- tions of essentialism, another favorite sociological boo word (is there really anyone today who could reasonably be called an essentialist?). In other words, if we make statement such as “SES explains X% of the variance in criminal behavior” without additional explanation, we imply that SES per se is a suffi cient explanation, with the additional implication that members of a given class are essentially the same and that members of other classes are simply different from them.

CONCLUSION

Socialization is the process in which biological potential is exposed to con- ditions that will affect the way it is expressed. Because SES is both the product and the producer of differential socialization, parental SES has a major, but not determining, impact on the SES level children may achieve in their adulthood. Warm, caring, nurturing parents who are sensitive to their children’s needs provide them with a wonderful start in life regardless of parental social class. The genetic endowments of both parent and child infl uence the nature of the parent-child relationship, and these endowments are naturally related. Sensitive and knowledgeable parenting is positively related to social class, but this does not doom lower-class children to per- petuate the class position of their parents. For genetic reasons, many chil- dren are resilient to the stresses they suffer in childhood if those stresses are not overly traumatic. When children enter school, they are exposed to a formal socialization agent whose task it is to provide children with the academic and attitudinal skills they need to become productive adults. Again, social class goes a long way in determining how well these lessons stick. For many middle- class children the schools reinforce what they have presumably already been exposed to, but for many, but certainly not all, children from the lower classes such learning may be entirely new and entirely unwelcome. Children who do not do well in school will become alienated from it, drop Social Class and Socialization 89 out, and congregate with others with like experiences. This is as true for middle-class children as it is for lower-class children. On the other hand, lower-class children who fi nd school congenial to their natures will bond with it, do well, and climb the class ladder as adults. Thus, the school has a mediating effect between parental social class and the social class children will achieve themselves. When children reach adolescence their behavior generates strains on all forms of social bonds. This is a normal and natural process preparing a new generation to procreate and to take its place in the adult world. For well-socialized children, antisocial behavior will be of the less serious sort and mostly limited to adolescence, while for the many of the less well- socialized it will be more serious and continue well into adulthood (Moffi tt 1993). The take-home lesson is that socialization matters, but also that the process of socialization is an interactive one that depends upon the inher- ited and acquired traits of both parents and child, the ability of children to accept and appreciate the methodological socialization of the school, and the depth and breadth of the pro- and antisocial traits youths bring with them into adolescence. 8 Poverty, Crime and Developmental Neurobiology

WHAT IS POVERTY AND WHAT CAUSES IT?

SES is viewed by most sociologists as a depoliticized smooth continuum, whereas radicals tend to think in terms of a class as a very politicized dichotomy of haves and have nots. This is the two-class model—“the over- whelming majority of respectable people on the one hand and the lumpen- proletariat on the other”—that Travis Hirschi (1969: 71) believes constitutes the true class–crime relationship implicit in most theories of crime. In this chapter we will examine crime and the “truly disadvantaged”; those unfor- tunates at the very bottom of the socioeconomic heap. Poverty is operationally defi ned by the U.S. government as 50% or less of the median family income in a particular year (Laderchi, Saith, and Stewart 2003). According to Hacker (2006), most people in the United States (about 58%) will experience poverty for at least a period of one year during their lives, with only about 12% of the population living at or below the poverty line at any one time. According to the U.S. Census Bureau (Ice- land 2003) in the years 1996 through 1999 only 2% of the population was in poverty for two or more years. These fi gures point out that poverty for most Americans is a short-term blip in their lives and that the American economy offers many avenues from which to escape it. When asking what causes poverty, we must specify in which country, at what time period, and for whom? Different contexts require different answers: geographic, structural, cultural, and individual. Countries in which almost everyone is in poverty tend to lack natural resources and have limited access to fertile land and fresh water and may be burdened with disease caused by organisms that thrive in climates found in many poor countries. Some countries may have caste or rigid class systems in which individuals are locked into their birth status, and still others are stifl ed by institutional- ized corruption or religious practices that frown on earthly status striving. No one would appeal to individual differences to explain why people are in poverty in countries where peoples’ lot in life is almost completely out of their hands due to ecological, structural, and/or cultural factors. In such societies the heritability coeffi cients of traits that are linked to success in more open societies are close to zero, indicating that conditions beyond the control of individuals trump individual traits in determining SES. Poverty, Crime and Developmental Neurobiology 91 Asking what causes poverty in a developed country such as the United States is a different matter. Constrained visionaries consider such a ques- tion strange because for them the real question is: What causes wealth? After all, one has to do something to gain wealth, but sitting for days on end in a run-down trailer park gulping Budweisers and puffi ng Marlboros while waiting for a “break” virtually assures poverty. Poverty in a nation in which very few people are in poverty for an extended period is the default option of those who do nothing to improve their standing, assuming of course that they possess the physical and psychological capability to do so. Able-bodied adults mired in long-term poverty have presumably made a series of bad choices such as to not apply themselves to school, to drop out, to have children out of wedlock, to avoid stable work, and to commit crimes. These choices have had the cumulative effect of foreclosing avenues to legitimate success. While it is true that not doing what must be done to avoid poverty will surely result in it, doing other things will also virtually assure poverty.

POVERTY: IS IT SOCIETY’S FAULT?

Rank, Yoon, and Hirschl (2003: 6) list a number of “characteristics closely associated with the risk of poverty” including “giving birth outside of mar- riage, families with large numbers of children, and having children at an early age.” This list of very private choices does not prevent them from claiming that poverty is the result of structural failure; that is, “society” is responsible for individual poverty. They go on to list low educational attainment, divorce, poor work skills, and long-term welfare dependency as failures of the economic, social, and political systems. Schwalbe and his colleagues (2000: 428) make plain that these things are subculturally chan- neled personal choices, writing that: “[T]hose who do well by the standards of the street acquire habits and create situations (drug addiction, lack of education, multiple dependents, criminal records) that are debilitating and risky, and diminish chances for mainstream success, even in the form of stable working-class employment.” Rank, Yoon, and Hirschl neglect to tell us how the “characteristics they list are societal rather than individual failures except to point out that the economy provides jobs that adequately support only about 8 out of 10 households. Why this would infl uence the poor choices they list, many of which were made before the age that people enter the job market, is any- one’s guess. If good jobs are scarce (and they are), then logically scarcity should promote precisely the opposite of the listed characteristics. To the extent that Rank, Yoon, and Hirschl’s argument has merit, they need to inform us how the economy can pull “good jobs” out of the air, and how we are to prepare folks living in poverty for such jobs when there is abundant evidence that so many of them actively resist such preparation (McWorter 92 Social Class and Crime 2000; Paterson 1998). Furthermore, their laundry list of “demographics” covers behaviors that the system spends multiple billions of dollars every year to trying to prevent, such as universal free education, family planning clinics, and job training schemes (Sawhill and Morton 2007).

RACE AND POVERTY IN THE UNITED STATES

What about the race-poverty relationship? According to the U.S. Census Bureau (2007), the median family income in 2006 for Asians was $74, 612; for whites it was $61,280; for Hispanics it was $40,000, and for blacks it was $38, 269. Correlating these fi gures with crime rates for these racial/eth- nic groups (blacks the highest, Asians the lowest) we have a perfect inverse relationship, which may lead some to assert that this provides incontrovert- ible evidence that poverty causes crime. It also provides evidence for those who assert that race and SES are excellent proxies for one another. The bite of racial discrimination certainly had a lot to do with pre-World War II black poverty when blacks were confronted with “No Negroes” signs out- side places of employment and lodging and other de jure and de facto prac- tices effectively combined to deny them access to the legitimate means of achieving upward mobility regardless of personal talents and ambitions. But such is no longer the case; barring major accidents or illnesses, any able-bodied male of any race in the United States is virtually guaranteed to avoid poverty if he does just two things: fi nishes high school and secures and maintains employment (McWorter 2002; Thernstrom and Thernstrom 1997). For females of any race, avoiding unwed motherhood is an added proviso. In 1995, less than 3% of white or black males who worked full- time year-round were in poverty. Among black females who worked full- time year-round, 7.5% were in poverty compared with 2.2% of similarly situated white females, a difference that refl ects the greater likelihood of black females having to support young children without the benefi t of male resources (Thernstrom and Thernstrom 1997: 242). Studies show that about 25% of black males between the ages of 16 and 24, and 50% of black males between the ages of 25 to 34, are non-custodial fathers who have abandoned their partners and their offspring (Holzer and Offner 2004). State pressure to force deadbeat dads to pay child support has led many of these men to abandon legitimate employment for fear of wage garnishment (Holzer and Offner 2004). Although the high black poverty rate relative to other races and eth- nicities is still often attributed to white racism, the data indicate that it is largely accounted for by the higher rate of black single-parent families. The U.S. Census Bureau’s (2002) breakdown of family types by race and income shows that a black two-parent family is less than half as likely to be poor as a white female-headed single-parent family (18.8% versus 41.1%). Stated in reverse, white female single-parent households were more than twice as likely as black two-parent households to have an annual income Poverty, Crime and Developmental Neurobiology 93 of less than $25,000 (slightly above the poverty line in 2001). These fi g- ures constitute powerful evidence that single parenting, whether the result of illegitimacy or family breakup, is a major cause of poverty across all racial lines because median family income is also perfectly correlated with illegitimacy rates among the races (i.e., Asians have the lowest illegitimacy rates and the highest median family income and blacks have the highest illegitimacy rate and the lowest income [Ventura 2009]).

CRIME AS A CAUSE OF POVERTY

In the 19th century, Belgian cartographer Lambert-Adolphe-Jacques Quetelet compared crime rates in France by age, sex, and season, and saw the same refl ections in the data that we see today in the American Uniform Crime Reports (i.e., crime is overwhelmingly committed by young males living in poor neighborhoods). Nevertheless, Quetelet discounted the idea that crime is caused by poverty, noting that the wealthiest regions of France had the greatest level of property crimes. He also noted that the level of wealth in a region does not necessarily correspond to the level of wealth among all its citizens, and that being poor amid riches as in urban France, not being poor per se as in rural France, is the condition that produces the most “misery” (in Rennie 1978). “Misery” (we call it “relative depriva- tion” today) often leads to feelings of envy generated by observing others having and enjoying what those feeling the misery would also like to have and enjoy. Envying what others have is not all that problematic if it moti- vates the envious to strive to legitimately obtain that which is desired. It is problematic, however, when it is experienced by those who believe that the world owes them a living and who subscribe to the cult of victimhood, believing they do not have what others have because of societal discrimina- tion. Intensely felt envy is often followed by anger and then by aggressive attempts to obtain that which is envied by illicit means; it is not for nothing that envy is considered one of the seven deadly sins. The link between SES and crime often boils down to the assertion that poverty causes crime. There is no doubt that poverty and crime are strongly related, but which is the cause and which is the effect? Many criminologi- cal theories implicitly or explicitly assume that poverty, or “poverty in the midst of plenty,” as one sociologist put it (Farley 1990: 217), is the major cause of crime. Poverty is not a pleasant state of affairs, which is a truism to which, if Hacker (2006) is correct, more than half of Americans can attest. While most of us endured poverty for only a brief time, most criminals tend to be mired in it, leading criminologists to make the causal link from poverty to crime. For the liberal thinker, those who commit crimes are fi nancially rather than morally bankrupt, and would settle down, get mar- ried and stay that way, get and keep a job, and even join the local church and sing hallelujah if only they could have a little more of the plenty they are supposedly in the midst of. 94 Social Class and Crime Robert Sampson (2000: 711) notes the mantra that “poverty causes crime” is pervasive in criminology: “Everyone believes that ‘poverty causes crime’ it seems; in fact, I have heard many a senior sociologist express frustration as to why criminologists would waste time with theo- ries outside the poverty paradigm. The reason we do . . . is that the facts demand it.” Frank Schmalleger (2004: 223) also notes that underlying assumption of all structural theories is that the “root causes” of crime are poverty and various social injustices. But as he also notes: “Some now argue the inverse of the ‘root causes’ argument, saying that poverty and what appear to be social injustices are produced by crime, rather than the other way around.” Sampson and Schmalleger have committed sociologi- cal heresy, but there is much truth in what they say; moral poverty not economic poverty (although they are surely a tightly knit bundle) is the prime mover of criminal behavior. Most of us who have experienced pov- erty pulled up our socks and got to work and never stooped to victimizing others to meet our needs, which may be part of the reason we ourselves are no longer in that unpleasant state. Let us perform a thought experiment in which we follow two large national samples of males from all racial/ethnic groups from birth to the age of 65. The only two criteria for inclusion in the fi rst sample is that (1) boys had to have been born into the lower classes and (2) they had never to have been convicted of any serious delinquent or criminal offense. The criteria for inclusion in the second sample is that (1) boys had to have been born to middle-class parents, and (2) they had to have been convicted of a serious delinquent or criminal offense by the age of 21. Few of us would bet against those in the fi rst sample having the highest average levels of fi nan- cial success in their adult years as opposed to those in the second sample despite their respective classes of origin. As Terrie Moffi tt (1993) and many others have pointed out, individuals who embark on a life of seriously victimizing others tend to begin before puberty, do poorly in school and drop out, do time in a juvenile detention center, and acquire a criminal record. Such a trajectory severely compro- mises their opportunities to gain meaningful employment, to form proso- cial networks, and to become attractive as a marriage partner to prosocial females. Although there are a perhaps few exceptionally talented career criminals who enter middle- and old-age fi nancially secure, and a few who manage a decent life despite their bad starts, the great majority live out their lives in poverty (Raine 1993). Economic studies have found that incar- ceration reduces employment opportunities by about 40%, reduces wages by about 15%, and wage growth by about 33% (Western 2003). Extensive interviews of burglars in the United States by Wright and Decker (1994) and Mawby’s (2001) multiple British samples provide contemporary support for Walter Miller’s (1958) focal concerns theory. Whether in the United States or Britain, burglars typically come from a Poverty, Crime and Developmental Neurobiology 95 poor run-down and socially disorganized neighborhoods where unemploy- ment is rife and the residents are poorly educated, unreliable, and resistant to taking orders and come mostly from single-parent households. Among their interviewed burglars, Wright and Decker (1994) report a strong sense of toughness, masculine independence, and fate (“I had little choice but to burgle”), excitement (sexual activity, drugs, and alcohol), autonomy (“As a burglar I am my own man”), and smartness (outwitting the law and getting something for nothing). Wright and Decker and Mawby are dubious about the possibility that job creation programs will change burglars into law-abiding citizens, fi nd- ing that most burglars chose burglary over jobs as being more suitable to their needs and lifestyles. Rengert and Wasilchick (2000: 47) also reject the notion that burglary is a default option of the jobless, claiming that many burglars who had jobs gave them up to concentrate on burglary: “Unem- ployment is not what caused crime. Crime caused the unemployment. Time confl icts forced individuals to choose burglary or their jobs. The resolution of the confl ict led directly to unemployment.” Employment means taking orders from bosses, and “Not taking orders from anyone . . . is a bedrock value on which male streetcorner culture rests; to be regarded as hip one must always do as he pleases” (Wright and Decker 1994: 49). Legitimate employment simply would not fi t into the lives of these people because for them party time is all the time. Most of the proceeds of Wright and Deck- er’s (1994) burglars were spent on drugs, alcohol, and sex; legitimate jobs provide neither adequate money nor time to engage in these pleasures at the level they want to indulge in them. As well as foreclosing on personal opportunities for success, crime also causes poverty by stripping communities where crime is most prevalent of its businesses and its jobs. Business owners do not relish remaining in areas where robberies, thefts, break-ins, and muggings are an everyday occur- rence. In addition to issues of personal safety, the costs of doing business in such areas are sharply increased (insurance premiums and expenses for security guards and alarm systems, etc.). After building costs and space constraints, crime is the most important reason for companies either not moving into inner-city areas or moving out of them if they are already there (Porter 1995). When teachers and children are more concerned about rapes, dope dealing, and assaults, how can any meaningful learning take place in schools serving high-crime area? How can people be proud of their homes and hope to accrue equity when all other houses around them are run down and bullet marked? In a causal sense the statement “Poverty is endemic where crime is endemic” is a more plausible proposition than “Crime is endemic where poverty is endemic.” Democratic Congressman John Lewis sums up this argument well: “In a very real sense it is crime that has caused poverty, and is the most powerful cause of poverty today” (in Walinsky 1997: 11). 96 Social Class and Crime THE NEUROBIOLOGICAL PERILS OF AN UNDERCLASS UPBRINGING

I must reverse my “pull-up-your-socks-and-stop-whining” stance here and acknowledge that severe deprivation suffered by innocent children in hor- rendous family environments prevents them from realizing their potential for normal development. Those who rise above poverty to become produc- tive members of society are assumed to come from family environments within the normal range, but some children are born into family environ- ments that are way beyond the pale of species-expected environments and cannot justly be evaluated in terms of human agency. Children in these environments are subjected to more noise, crowding, family and neigh- bor confl ict and instability, ambient pollution, abuse and neglects, and numerous other stress-inducing problems than are the vast majority of children. Such stressors are endemic in many inner-city environments in which children simply fi nd themselves and have done nothing to create. Relentless stress can alter neurobiological mechanisms in such a way as to put those who experience them at an elevated risk for all kinds of antiso- cial behavior. Stress is a state of psychophysiological arousal experienced when and organism perceives a challenge to its physical or mental well-being. Stress is an inevitable part of life; it energizes and focuses us, and without it we would be seriously handicapped in our ability to meet life’s challenges. Adults who experienced average levels of stress during childhood most likely possess brains calibrated to better navigate the travails of life than those who were been protected from almost all stress. People differ in their ability to manage stress for genetic and experiential reasons that have been captured in their brain’s circuitry (Perry 2002). While stress is functional, toxic and protracted stress does damage to vital brain areas responsible for memory storage and behavioral regulation such as the amygdala and hippocampus. Childhood stressors that lead to frequent activation of stress response mechanisms may lead to the dysregulation of these mechanisms, leading to a number of psychological, emotional, and behavioral problems (Gunnar and Quevedo 2007). The stress response is mediated by two separate but interrelated sys- tems controlled by the hypothalamus: the autonomic nervous system and the hypothalamic-pituitary-adrenal (HPA) axis. As previously discussed, when an organism perceives a threat to its well-being, the hypothalamus directs the ANS’s sympathetic system to mobilize the body for vigorous action aided by pumping out the hormone epinephrine (adrenaline). When the organism perceives that the threat is over, the parasympathetic system restores the body to homeostasis. Because the HPA axis response occurs through changes in gene expres- sion, it is slower than the ANS response and lasts longer (Gunnar and Quevedo 2007). The HPA axis is activated in situations that call for a Poverty, Crime and Developmental Neurobiology 97 prolonged rumination rather than the visceral immediacy of the ANS’s preparation of the body for fi ght or fl ight. The HPA axis response begins with the hypothalamus feeding various chemical messages to the pituitary gland which leads to further chemical products that stimulate the adre- nal glands to release the hormone cortisol. The brain is a major target for cortisol which, unlike epinephrine, is able to cross the blood-brain barrier (van Voorhees and Scarpa 2004). Cortisol is the fuel that energizes our coping mechanisms by increasing vigilance and activity, and is therefore functional within the normal range. Frequent HPA axis arousal may lead to upward or downward dysregula- tion of arousal mechanisms. Upward dysregulation means the overproduc- tion of cortisol, or hypercortisolism. Hypercortisolism leads to anxiety and depressive disorders and is most likely to be found in females who have been maltreated because females activate signifi cantly more neural systems asso- ciated with emotional stress and with encoding it into long-term memory than males (van Voorhees and Scarpa 2004). Females will thus have emo- tional experiences more readily available for rumination than males, and constant rumination increases their valence over time. Hypercortisolism suggests a failure of the system to adjust to chronic environmental stressors and leads to internalizing problems such as chronic depression and post- traumatic stress disorder. Hypocortisolism, on the other hand, suggests an adaptive downward adjustment to chronic stress and leads to externalizing problems. It is adap- tive because frequent stressful encounters habituate the organism to them, and as a consequence the organism does not react to further encounters as it had previously. Habituation means that both HPA axis and ANS response mechanisms have become blunted. Hypocortisolism has been linked to early onset of aggressive antisocial behavior (McBurnett et al. 2000), to criminal behavior in general (Ellis 2005), and is more likely to be found in maltreated males than in females (van Goozen et al. 2007). A study by O’Leary, Loney, and Eckel (2007) found that males high in psychopathic traits lacked stress-induced increases in cortisol displayed by males low in psychopathic traits. As noted when we discussed the formation of a con- science, blunted arousal means a low level of anxiety and fear, something quite useful for those committing or contemplating committing a crime. We should note that while chronically high levels of cortisol are associated with aggression inhibition, acutely elevated levels of cortisol and/or chronic extreme low levels of cortisol tend to promote aggression (van Goozen et al. 2007). The development of hypocortisolism is an example of the process of allostasis, which literally means to achieve physiological stability through change. Allostasis describes the body’s attainment of equilibrium by alter- ing the acceptable range of physiological set-points to adapt to extreme acute stress or chronic stress rather than returning them to their previous state as in homeostasis. According to Goldstein and Kopin (2007: 111): 98 Social Class and Crime “Adaptations involving allostasis to cope with real, simulated, or imagined challenges are determined by genetic, developmental and previous experi- ential factors. While they may be effective for a short interval, over time the alterations may have cumulative adverse effects.” Frequent allostatic responses is termed “allostatic load,” which Goldstein and Kopin (2007) liken to having the heating and cooling systems running simultaneously, a situation guaranteed to hasten the wear and tear of both systems and eventual breakdown. Likewise, Massey (2004) points out that chronically elevated cortisol also interferes with the normal operation of glutamate, a common excitatory neurotransmitter that is vital in the formation of syn- aptic connections and learning. Children exposed to chronic stress will thus have fewer cognitive resources regardless of their genetic potential.

MATERNAL SUBSTANCE ABUSE AND TERATOGENIC FETAL EFFECTS

During the early process of building a brain, the immature neurons must migrate from their birthplace to their eventual home guided by molecu- lar messengers. The nascent brain is most vulnerable to insults during the migratory phase of maturation because these molecular guides are suscep- tible to attacks by teratogenic chemicals, which may send neurons to the wrong area or even direct them to self-destruct (Prayer et al. 2008). The most common teratogens are those associated with alcohol. When pregnant women drink they introduce their fetuses to neurotoxins that produce a number of neurological disorders, the most serious of which is fetal alcohol syndrome. For criminologists the primary concern related to FAS is how it effects behavior via is effects on the frontal lobes, amygdala, hippocam- pus, hypothalamus, serotonergic system, and myelination process (Noble, Mayer-Proschel, and Miller 2005). FAS is the foremost preventable cause of intellectual impairment in the world (O’Leary 2004). The prevalence of fetal alcohol disorders (not all are full-blown FAS) in the United States is around 1% of live births (Manning and Hoyme 2007). Because heavy drinking is most prevalent among low SES individu- als (Casswell, Pledger, and Hooper 2003), FAS rates are higher in the lower classes. A review of numerous studies by the National Institute of Alcohol Abuse and Alcoholism (May and Gossage 2008) found an average rate of FAS of about 0.26 per thousand for the middle class and one of about 3.4 per thousand for the lowest classes, which is about 13 times greater. Besides low IQ, other developmental defi cits associated with FAS that are linked to high levels of antisocial and criminal behavior independent of FAS are hyperactivity, impulsivity, poor social, emotional, and moral development, and a highly elevated probability of alcoholism independent of any genetic transmission (Jacobson and Jacobson 2002). FAS is found in all racial groups, but African American children are fi ve times more likely to exhibit Poverty, Crime and Developmental Neurobiology 99 FAS than white children (Sokol, Delaney-Black, and Nordstrom 2003), and Asian American children have the lowest rate (Meaney and Miller 2003). Other voluntary use of substances more common among lower SES women that have teratogenic effects on neuron migration and development include smoking and illicit drug use. It has been estimated that as many as 375,000 infants born each year have been exposed to cocaine in utero, and that in some inner-city hospitals as many as 50% of pregnant women test positive for crack or cocaine (Mayes et al. 1995). A large study of neonates who were born in inner-city hospitals in four large American cities and were both exposed and not exposed to cocaine in utero found that while cocaine exposure does not result in the anatomical problems associated with FAS, it does result in serious central and peripheral nervous system symptoms (Bauer et al. 2005). Cocaine-exposed infants were also 42 times more likely to have a diagnosis of hepatitis and 15 times more likely to have a diagnosis of syphilis. Comparing the mothers of cocaine-exposed infants (n = 717) to non-exposed infants (n = 7442) it was found that only 10.1% of the mothers of cocaine-exposed infants were married ver- sus 40.4% of mothers of non-exposed infants, 50.2% had less than a high school education (versus 28.7%), and only 22.7% worked in the previous year (versus 53.6%). Maternal smoking has its primary risk in fetal intermittent hypoxia, which reduces the oxygen available to the fetus that may lead to cell death (Zechel et al. 2005). Cohort studies (e.g., Brennan, Grekin, and Sarnoff 1999) consistently fi nd that maternal smoking during pregnancy predicts criminal behavior in fetally exposed offspring independent of other risk factors. A review of a number of such studies found odds ratios ranging between 1.5 and 4.0 for fetal tobacco-exposed individuals versus non- exposed individuals for various forms of antisocial behavior across diverse contexts and independent of other factors such as maternal SES and IQ (Wakschlag et al. 2002). In the Bauer et al. (2005) study discussed earlier, 81.8% of the mothers of cocaine-exposed infants also smoked during preg- nancy versus 19.7% of the mothers of non-cocaine-exposed infants.

LEAD EXPOSURE

Exposure to noxious substances in the environment outside the womb such as lead over which the mother has no control has negative effects on children’s brain, manifested most clearly in the reduction of IQ. The IQ decrement per one unit increase in micrograms per deciliter of blood (µg/dl) of lead is an average of 0.50 points (Koller et al. 2004). Poverty status means having to reside in the poorest neighborhoods and living in the oldest houses, and the main source of lead exposure today is lead dust from paint in older houses. Toxic levels of lead (>40 µg/dl) distort enzymes, interfere with the development of the endogenous opiate system, disrupt 100 Social Class and Crime the dopamine system, and reduce serotonin and MAOA levels (Wright et al. 2008). An fMRI study found that brain gray matter was inversely correlated with mean childhood lead concentrations in mostly black young adults (n = 157) taken from the longitudinal Cincinnati Lead Study (Cecil et al. 2008). The mean childhood blood lead concentration of this sample was 13.3µg/ dl, which is far in excess of the 2006 average of 1.5µg/dl for the general population (Bellinger 2008). Although the gray matter lost to lead expo- sure was relatively small (about 1.2%), it was concentrated in the frontal lobes and the anterior cingulate cortex, which are vital behavior moderat- ing areas responsible for executive functioning and mood regulation. A larger sample (n = 250; 90% black; mean age 22.5) from the same Cin- cinnati Lead Study was examined for the relationship between childhood blood lead and verifi ed criminal arrests. The mean number of arrests for the 136 males in the sample who were ever arrested was 5.2. The main fi nd- ing of this study was that after adjusting for relevant covariates for every 5 µg/dl lead increase there was an increase in the probability of arrest for a violent crime of about 50% (Wright et al. 2008). There are a number of other studies that have found these weak to moderate effects of lead expo- sure on crime, particularly violent crime (reviewed in Needleman 2004). Lead competes with calcium for absorption in the body, so if children do not receive adequate dietary calcium, lead can trick the body into absorb- ing it instead. This is important because blood lead levels are associated with a calcium absorption gene called the vitamin D receptor gene (VDR) (Chakraborty et al. 2008). Certain polymorphisms of the VDR (an enzyme known as Fok1) make the absorption of calcium more effi cient for their carriers, and these polymorphisms are more prevalent among blacks than among whites or Asians (Chakraborty et al. 2008; Haynes et al. 2003). If these polymorphisms render calcium absorption easier, given lead’s ability to mimic calcium lead is also more easily absorbed. This creates a gene x environment interaction in which African Americans absorb more lead than similarly exposed members of other racial groups.

BREASTFEEDING

While drinking, smoking, and drug use during pregnancy is a sign of inex- cusable irresponsibility, prolonged breastfeeding refl ects a desire to care for and be close to one’s infant. One of the positive effects of prolonged breastfeeding was demonstrated in a large study of 13,889 Belarusian breastfeeding mothers. A random half of these mothers were given incen- tives that encouraged prolonged breastfeeding while the remaining half continued their usual maternity hospital and outpatient care. The children were assessed six years later, and the researchers found that the children breastfed for a prolonged period of time (>6 months) had a mean IQ almost Poverty, Crime and Developmental Neurobiology 101 six points higher than the control group children and received higher aca- demic ratings from teachers (Kramer et al. 2008). The experimental design allowed researchers to measure breastfeeding effects on IQ without any biasing confounds such as the positive relationship between mothers’ IQ and the probability of prolonged breastfeeding. Kramer and his colleagues could not determine if the breastfeeding-IQ relationship was due to the constituents of breast milk, mother-child inter- actions and the warm kin-to-skin contact experienced during breastfeeding, or both. Tactile stimulation of human infants confers enormous benefi ts on the infant and is recommended by physicians for the optimal brain develop- ment of low-birth-weight infants and infants who have suffered some kind of head injury. Breastfeeding and tactile stimulation results in the release of oxytocin, a polypeptide known popularly as “the cuddle chemical,” the proximate molecular mechanism of mother-infant bonding. Oxytocin released during breastfeeding reduces the mother’s sensitivity to environ- mental stressors, which allows for greater sensitivity to the infant’s needs. It has been repeatedly shown via skin conductance and cardiac-response measures that lactating mothers show signifi cantly fewer stress responses to infant stimuli than non-lactating mothers, and show signifi cantly greater desire to pick up and cuddle their infants in response to all infant-presented stimuli (Hiller 2004). Nutritionally, mothers’ milk confers many benefi ts, not the least of which is its immunological benefi ts. In terms of IQ enhancement, a study of 3,000 British and New Zealand children born in the early 1970s and IQ-tested in the 1990s found that subjects who were breastfed scored on average six to seven points higher on IQ tests than non-breastfed subjects (Caspi et al. 2007). This was only true for breastfed subjects who had at least one copy of the C (cytosine) base variant of the fatty acid desaturase (FADS2) gene (90% of the sample) that codes for an enzyme that converts fatty acids into two polyunsaturated acids that accumulate in the infant brain in the early months after birth. There was no IQ increase noted for breastfed subjects homogenous for the less common G (guanine) base allele of the FADS2 gene. This is another example gene-environment interaction. That is, there was no difference in the average scores of the C versus G allele individuals who were not breastfed (the gene made no difference alone), and breast- feeding made no difference for individuals with two copies of the G allele. Unfortunately, the literature consistently shows a marked downward gradient in the rates of breastfeeding as IQ, income, education, and occu- pational status fall, and children of lower-class women are thus more likely to be deprived of important evolutionarily experience-expected input (Gla- ser 2000). A random sample of 10,519 mothers in California found that the odds of breastfeeding for the women in the highest income category was 3.65 times the odds of the women in the lowest income category (Heck et al. 2006). Data from the Department of Health and Human Service’s National Immunization Program survey (2004) reveal that in 2001 only 102 Social Class and Crime 29.3% of black infants were breastfeeding at six months versus 43.2% of whites and 53.7% of Asians.

CONCLUSION

This chapter has taken a skeptical look at the crime-poverty relationship, concluding that it was more plausible to assume that crime is more a cause of poverty than poverty is a cause of crime. It was noted that most of us have endured poverty for a short time at some point in our lives, and that very few people are mired in poverty for more than a few years. Poverty is “caused” by individuals doing nothing constructive to prevent it, and by doing almost everything guaranteed to assure it. Committing crimes is cer- tainly one way of precipitating the slide into poverty and/or if one is already in that sad state, assuring that one will remain there. The emphasis on human agency must be downgraded considerably when confronted with the truly horrendous experiences of children born to the deepest depths of the class hierarchy. Many of these children suffer devel- opmental defi cits that are the result of their mothers ingesting toxic sub- stances while pregnant with them that resulted in many brain defi ciencies. These children are also less likely to be breastfed and to be raised in homes and neighborhoods in which there are high levels of lead. Children born into intergenerational poverty families suffer many additional environmen- tal stressors such as abuse and neglect that may lead to hyper- or hypocor- tisolism that lead to internalizing and externalizing problems, respectively. Putting all these problems together, it not diffi cult to see how the products of such environments come to perceive the world as a hostile place where no one is to be trusted or valued, and therefore all are fair game for criminal victimization. 9 Social Stratifi cation, the Genome, and Social Structure

SOCIAL STRATIFICATION

This chapter explores the evolutionary origins of stratifi cation and its neu- rohormonal underpinnings. The status striving that eventually leads to SES stratifi cation is a hard-wired behavior because high status equals increased fi tness. However, I want to allay concerns that some may have that all this talk of hard-wiring means that social class is fi xed (biologically inher- ited) across generations and that social structure does not matter. I show that social structure matters immensely because it determines the extent to which genes matter for the attainment of social status. Personal traits have a large impact on achieving social class level in some societies, but hardly at all in others. I also want to show that social class cannot be genetically fi xed by brief discussions of regression to the mean, the Hardy-Weinberg equilibrium, and decreased assortative mating, all of which contribute to mixing-up the classes in modern societies. In sociology, stratifi cation refers to the social ranking or level of people based on their possession of socially valued traits and/or resources such as education and income. The importance of stratifi cation in noted by Theo- dore Kemper: “Perhaps the fundamental social structure of society is the system of stratifi cation. It so bluntly determines individual conduct, belief, and value preferences, on the one hand, and sheer biological fate on the other, including pathology, morbidity, and longevity, that for many pur- poses it is . . . the social structure par excellence worthy of close attention” (1994: 47–8). SES is the central concept of macro-sociology akin to the atom in phys- ics and the gene in biology, although it is not as neatly conceptualized and measured because it is not a tangible “thing” that can be directly observed or measured with scientifi c accuracy. The existence of both atoms and genes were once doubted, but they were still considered useful heuristic concepts because they explained numerous observations in physics, chemistry, and biology. SES is a similarly useful heuristic that serves sociologists to explain many facets of human behavioral variation. No one, of course, expects that we will ever see SES under a microscope because it is more like gravity than atoms or genes, but like gravity we know it by its consequences. 104 Social Class and Crime When people of my generation entered the workforce in the 1950s, the class landscape was crystal clear, or at least we though it was. The work- ing class car-pooled or took the bus to the factory, paid union dues, voted Democrat, took vacations at grandpa’s, and worshiped a Catholic or Bap- tist God. The middle class drove to the offi ce in their Buicks (next stop, Cadillac), paid professional dues, voted Republican, vacationed in Florida, and worshiped an Episcopalian God. The upper crust were chauffeured to the golf course, paid bribes, voted for whomever they had so blessed, vacationed in the Mediterranean, and worshiped Mammon. At the very lowest level of the class structure were the “dangerous classes” who never worked, never paid anyone anything if they could help it, never voted, never vacationed, and worshiped nothing transcendental. Except for the top and bottom rungs of the class ladder, today’s class landscape has been muddied considerably. Perhaps with a little humility at one end and hubris at the other, high court judges and high school janitors are apt to describe themselves as middle class. And indeed there may not be much in terms of possessions and lifestyle that separates individuals whose social prestige is poles apart. So many of the goods (entertainment devices, home appliances, automobiles) and services (cruises, air travel, nice restau- rants) only affordable for the comparatively few in the 1950s are available to almost everyone today.

KARL MARX AND MAX WEBER ON CLASS

We begin our exploration of class stratifi cation by examining the works of two theorists who arguably set the intellectual framework for the study of class for generations of sociologists: Karl Marx and Max Weber. Both men were scholars of stupendous intellect who agreed on many things, but disagreed more. Philosophically, Marx was a materialist, a position that seeks the causes of human action in material things such as economic fac- tors and social structure. Weber was more of an idealist, a philosophical principle that posits that ideas create reality and cause human action, the structure of society, and how individuals and groups behave. This differ- ence in philosophical positions led them to view, dissect, and evaluate a number of nominally identical concepts such as class and confl ict in differ- ent ways. Because the works of Marx and Weber are so vast, I discuss them here only briefl y to note where their ideas cohere with biosocial thought on the topic of status striving and social stratifi cation. For Marx social classes are defi ned by their relationship to the means of production, and only two classes really mattered—the owners of the means of production (the bourgeoisie) and those who must sell their labor (the workers or the proletariat). He recognized gradations within these classes, and even prestige classes above (the aristocracy) and below (the lumpenproletariat) them, but they were largely irrelevant to his theory since Social Stratifi cation, the Genome, and Social Structure 105 he envisioned society evolving via ruthless capitalist competition into two homogeneous warring classes of owners and workers. Marx saw social life as a struggle for existence and that social class was both the cause and consequence of that struggle. This view has a dim evolu- tionary ring to it, and indeed many concepts central to Marxism such as con- fl ict, exploitation, self-interest, dominance, and alliance building are also central to evolutionary biology. There is much evidence throughout Marx’s work that he saw humans as primarily motivated by self-interest (Jost and Jost 2007) and that he abjured the naturalistic fallacy. For instance, Marx and Engels wrote that good communists: “do not put the moral demand: love one another, do not be egoists, etc; on the contrary, they are well aware that egoism, just as much as selfi shness, is in defi nite circumstances a neces- sary form of self-assertion for individuals” (in Replogle 1990: 695). While Marx’s concern was with class confl ict, he knew that it is only apparently driven by group interests, and that underneath all the class soli- darity rhetoric is the quest for individual power. As Marx put it: “The sepa- rate individuals form a class only in so far as they have to carry on a common battle against another class; otherwise they are on hostile terms with each other as competitors” (in Coser 1971: 48). In other words, cooperative rela- tions exist among humans because they best promote the self-interest of each individual, not because they promote the interests of the class to which they belong. Of course, cooperation within the group improves things for all within it who abide by the cooperative norm, which once again illus- trates how self-interest is the ultimate evolutionary key to social harmony. As the great biologist Edward O. Wilson put it: “Human beings appear to be suffi ciently selfi sh and calculating to be capable of infi nitely greater harmony and social homeostasis. This statement is not self-contradictory. True selfi shness, if obedient to the other constraints of mammalian biology, is the key to a nearly perfect social contract” (1978: 157). Max Weber saw Marx’s optimism regarding the elimination of confl ict and inequality as a biological fantasy. For Weber, social life is ultimately about the pursuit of social power, which is in turn a way to acquire social honor. Whereas Marx saw only class division in terms of economics, Weber identifi ed three separate but related spheres in which this social power is sought: class, status, and party. Class denotes economic power and is for Weber non-social or “barely social” because it is instrumental in orienta- tion and lacks any sense of shared social belonging. Status is fully social because it is communal (i.e., status groups hold common values, express a common lifestyle, and possess a sense of belonging). Party represents politi- cal power and is based on voluntary associative social relationships, which are rational-legal in outlook rather than communal and traditional and so are thus less “social” than status groups (Gane 2005). For Weber a society is composed of multiple temporary as well as per- manent alliances of self-interested individuals that coalesce and dissolve as their interests wax and wane. These coalitions exist within more permanent 106 Social Class and Crime alliances such as economic classes, religious groups, political parties, and racial/ethnic groups. Many of the confl icts between and within these groups are considered by their participants to be just as important, and often more so, than Marx’s economic class struggle (Sanderson 2001). Weber was also a selectionist in that he recognized “that in the unending competition for social advantage some individuals are better fi tted than others to succeed in occupying coveted roles because of their inherited abilities and personal characteristics” (Runciman 2001: 14). Weber frequently used selectionist logic in asserting that certain types of individuals are favored over others in the struggle for wealth and status, and that competition will always be present under any social system conditions. Responding to Marx’s optimism that would eliminate social inequality, Weber (1978: 39) stated that “even on the utopian assumption that all competition were completely eliminated, conditions would still lead to a latent process of selection, biological or social, which would favor the types best adapted to the conditions, whether their relevant qualities were mainly determined by hereditary or environment.” Thus, regardless of the social, economic, or political nature of society, confl ict will always exist. It was Weber’s position that even though those with wealth, status and power are able to impose their values and vision for society on others, class divisions in society are normal, inevitable, functional, and acceptable (Gane 2005). In agreement with Marx: “Weber quite unmistakably regards personal self-interest as innate and universal among humans. . . . personal self- interest is already fi xed by genetic inheritance in all human individuals and needs no further fi xing there by external imposition” (Wallace 1990: 209). Despite Weber’s faith in idealism, he asserts that self-interest trumps ideas in the governance of human conduct, an assertion that turns him into a materialist. In Weber’s own words: “Yet very frequently the ‘world images’ that have been created by ‘ideas’ have, like switchmen, determined the tracks along which action has been pushed by the dynamics of interest” (in Wallace 1990: 209). Weber agreed with Marx in characterizing the growing alienation he observed in society as an “iron cage.” While capitalism was the root cause of alienation for Marx, for Weber it was a remedy: “an involved worker is a type of man, bred by free association in which the individual have [sic] to prove himself before his equals, where no authoritative commands, but autonomous decisions, good sense, and responsible conduct train for citizenship” (Sarfraz 1997: 48). Weber reasoned this way because he saw growth of capitalism and state bureaucracies marching in lock-step with democratization, which implies greater liberty to make one’s own deci- sions and to track one’s own fate in fair and open competition. The point I am making here is that while both Marx and Weber had a deep interest in social class and in confl ict, neither man made any effort to plumb the depths of these concepts. This is not a criticism of these men because they did not have the technical or conceptual tools we have today to do so. Social Stratifi cation, the Genome, and Social Structure 107 THE PRIMORDIAL ORIGINS OF SOCIAL STRATIFICATION

There is little or no discussion of the underlying psychology of status or of the evolutionary processes underlying that psychology in sociological text- books on stratifi cation. This should not surprise us, for as Kemper (1994: 48) has pointed out: “If sociology and biology have not been on speaking terms in general, sociological distain for the biological reaches its apogee when it comes to social stratifi cation.” Perhaps this is so because most soci- ologists are unconstrained visionaries committed to changing the status quo, and biology conjures up the ghosts of . They fear that if they were to admit that a person’s position in the social hierarchy was even partly “biological” they would be admitting that class positions are immutable and predetermined. This position is derived from an under- standing of biology to which they are not only hostile but also “militantly and proudly ignorant” (Van den Berghe 1990: 177). The origins of social stratifi cation lie in the psychology of organisms striving for status in social groups. As Loch, Galunic, and Schneider (2006: 222) remark: “People crave general respect and recognition . . . in all cul- tures of the world. In other words, the striving for success is hard-wired, utilizing basic emotions (such as anger, sadness, happiness, pride) depend- ing on whether status is achieved or not.” This basic feature of human nature is universal and “hard-wired” because status leads to priority of access to resources and mating opportunities, the twin central concerns of all sexually reproducing animals. Of course, status can be and is valued for its own sake, and the criteria by which status is attained and conferred and the symbols that signify its holders are cultural. Nevertheless, for the biosocial criminologist the ultimate origin of social stratifi cation is the fun- damental trait of individual status striving. A biosocial analysis of common status concerns across species should help to explain the status concerns that animate much of the criminal behavior observed among lower SES individuals.

NEUROHORMONAL UNDERPINNINGS OF STATUS HIERARCHIES

Status is a ubiquitous concern of all sexually reproducing animals that expend energy and resources obtaining mates in the face of sometimes brutal competition. High standing in the group relieves those who enjoy it of having to compete with others because status simultaneously draws females to those who have it and leads lower status males to defer. Tes- tosterone (T), serotonin, and cortisol are the chemicals most related to striving status across animal species (Anestis 2006; Archer 2006). High T levels foster dominance and extraversion; low cortisol refl ects low levels of stress, anxiety, and fear; and high serotonin fosters confi dence and self- esteem. These substances operate within the context of an interacting soup 108 Social Class and Crime of other neurotransmitters, hormones, and receptor and transporter mol- ecules entailing complex feedback loops, and their secretion patterns are infl uenced by the social context. That is, although basal levels of these sub- stances are heritable, they rise and fall according to the organism’s experi- ences. In established status hierarchies leaders best assert their leadership by a subtle mix of coercion and bonhomie. Henrich and Gil-White (2001) refer to this leadership style as persuasive because it relies on the leader’s prestige and the freely conferred deference of followers. In more chaotic hierarchies more aggressive strategies may be needed to maintain control and thus require a different leadership style that Henrich and Gil-White (2001) call a dominance style characterized by the use of force or the threat of force. A shift in leadership strategies requires a dif- ferent mix of chemistry. For instance, a male who has a high T who also has a hyperactive autonomic nervous system will encounter anxiety and fear when challenged, and this will elevate cortisol. One of the effects of cortisol is that it suppresses T, so the combination of anxiety and the loss of T-driven surgency may lead such a male to withdraw from future status competitions. Thus, different genotypes rise to high rank depending on the social context. Serotonergic mechanisms modulate cortisol and T responses in sta- tus competitions among primates (Anderson and Summers 2007). When male vervet monkeys are provided with exogenous serotonin, they tend to attain high dominance status in the troop, and in naturalistic settings the highest ranking males have the highest levels of serotonin (Anestis 2006). Low-ranking males, typically low in serotonin, defer without much fuss to higher-ranking males over access to females and other resources in estab- lished hierarchies. When the hierarchy is in fl ux, however, it is the low-sta- tus males who become the most aggressive in seeking status. If a low-status male is successful in improving his status, his serotonin level rises to levels commensurate with his new status, which again emphasizes the feedback nature of behavior-related chemistry and environmental events. The same relationship between serotonin and violence are consistently found among human males, as one would expect given the common fi tness concerns among all primates. Indeed, the relationship between low serotonin and impulsive violence among humans has been called “perhaps the most reli- able fi ndings in the history of psychiatry” (Fishbein 2001: 15). Given the bidirectional relationship between social status and neuro- hormonal mechanisms, we might hypothesize that natural selection has equipped us to adjust ourselves to the social statuses we fi nd ourselves in within well-ordered groups (“Best not to challenge when the status hierar- chy is solid and I have little chance to challenge successfully”). This is not social Darwinism asserting that the status quo exists because it is naturally ordained and therefore good (the naturalistic fallacy). These same mech- anisms also equip those with little to lose with the necessary biological wherewithal to attempt to elevate their status by taking violent risks when Social Stratifi cation, the Genome, and Social Structure 109 social restraints are weak, as they are in our lowest SES neighborhoods (Anderson and Summers 2007). Given the well-established serotoninÆstatus relationship among non- human primates, it is reasonable to expect that is also exists among humans. Matthews and her colleagues (2000) subjected a sample of 270 black and white males and females to a neuropharmacological challenge that enabled them to measure serotonergic responsivity. They also measured each sub- ject’s SES, impulsiveness and hostility. Serotonin responsivity was inversely related to SES and hostility-impulsiveness overall and across both races and sexes. In a multiple regression model, SES remained signifi cantly related to serotonin after adjusting for a wide range of biological variables and for scores on the hostility-impulsiveness scales. It should be noted that the researchers hypothesized the causal direction to be low SESÆlow sero- tonin/high hostility-impulsiveness, not low serotoninÆlow SES/high hos- tility-impulsiveness, although there are arguments about causal direction. Of course, it does not have to be linear in either direction; it is most likely bidirectional as it is in other species. Thus, it is plausible to assert that the ultimate origins of social status in any society past or present lie in primate dominance struggles which afforded the winners priority of access to food, territory, and mates. These dominance struggles are not a matter of brute strength and aggression, but more a matter of coalition building in which favored members of the alliance share in the fruits of alpha’s success at the expense of those not favored. Herein is also the origin of exploitation of one class (alpha’s “rul- ing alliance”) over another. But as Lopreato and Crippen (1999: 225) point out, dominant males and their allies maintain social order in the group, and in exchange for this vital social role a certain amount of exploitation (expropriating more than their “fair share” of resources and mates) is toler- ated by the exploited.

THE CHANGING INFLUENCE OF THE GENOME ACROSS DIFFERENT SOCIAL STRUCTURES

Sociologists Daniel Adkins and Guang Guo (2008), who are rare exceptions to the dismissal of biology in sociology, provide us with a useful model of the shifting infl uence of the human genome on status attainment across dif- ferent levels of social development. Their thesis is that the extent to which genes matter for the attainment of social status depends on the structural nature of society. Genes matter most in modern open societies in which legal equity is the norm, and hardly at all in societies characterized by social closure (e.g., slave and caste societies). Adkins and Guo’s thesis is derived from the phenomenon discussed in Chapter 2 regarding heritability—genes matter least, as indexed by a low h2, in the most unequal and disadvan- taged environments, and their infl uence increases linearly as environments 110 Social Class and Crime become more equal and advantaged. This is the central maddening irony facing egalitarianians: As the environment becomes more equal, the more people are differentiated by their genes. Using data from the Standard Cross-Cultural Sample of the Ethnographic Atlas (SCCS), Adkins and Guo found that the infl uence of the genotype on status attainment describes a reverse J pattern across the expanse of human history. Hunter-gatherer societies are the most egalitarian, with only 7.4% having a stratifi cation system, 7.4% inherited leadership, 26.7% political elites, 13% inherited land ownership, and 3.7% with slavery. Hunter-gath- erer societies, particularly in the earliest evolutionary environments, lack surplus resources, so all individuals engage in subsistence activities, leaving little possibility for the emergence of elite classes to exploit others. There are, of course, differences in status among hunter-gathers, but these dif- ferences are a function of the individual’s personal characteristics—skilled hunter, fi ghter, strategist, speaker, leader, manipulator, and so forth. As Adkins and Guo (2008: 248) explain:

With each generation, status assortment begins on a fairly level play- ing fi eld and proceeds on the basis of ability and luck. Thus, the genomeÆability relationship is apt to be relatively strong as all society members develop in largely the same environment; the abilityÆstatus relationship is also likely to be relatively strong as there is generally no social closure or inequality creating disparate life chances for society members.

With the advent of food surplus and storage in horticultural societies (semi-sedentary peoples cultivating a variety of crops on a small scale), we see the beginning of social stratifi cation systems. In the SCCS sample, 75% of horticultural societies had stratifi cation, 51.6% political elites, 77.8% inherited land ownership, 37% hereditary leadership, and 56.9% slavery. These fi gures indicate a higher level of inequality and social closure and thus weaker genomeÆability and abilityÆstatus relationships. Inequality and social closure reached its zenith with the arrival of agrar- ian societies (sedentary societies cultivating crops on a large scale with a technology superior to that of the earlier horticulturalists). The SCCS sam- ple shows that 96.8% of agrarian societies had class stratifi cation, 76.9% political elites, 96.5% inherited land ownership, 52% hereditary leader- ship, and 43.3% slavery. In agrarian societies only the most talented and lucky could rise above their station of birth. Such societies are character- ized by huge environmental inequalities between the elites and common- ers, thus drastically reducing both the genomeÆability relationship and the abilityÆstatus relationship. Things began to change with the onset of industrialism. The democ- ratization of the polity marching in lockstep with the industrialization of the economy simultaneously reduced power and economic inequality. The Social Stratifi cation, the Genome, and Social Structure 111 sheer size and complexity of industrial societies demands large-scale par- ticipation of their members since it is not possible for a small cadre of elites to do everything to keep those societies viable. Modern societies require a constant fl ow of millions of ambitious, diligent, and capable individuals to fi ll positions vital to economic health and growth. The people who fi ll the most important and desirable positions naturally demand compensa- tion commensurate with their training and education. The genomeÆability relationship thus becomes strong again, as does the abilityÆstatus relation- ship, although not as strong as it was in hunter-gatherer societies. Unfortunately, as Adkins and Guo point out, the decreasing income inequality observed during industrialization began to reverse in the 1970s due to economic globalization, and this trend casts “uncertainty on the direction of future changes of the strength of genetic infl uence on status outcomes” (2008: 251). For instance, between 1979 and 2004 real (after- tax) income rose by 9% for the poorest fi fth of Americans, 69% for the richest fi fth, and 176% for the richest 1% (Sawhill and Morton 2007). The increase in economic inequality between the classes may work in two opposite directions given that it is taken place concurrent with the greater intellectual demands of the modern workplace. That is, for genotypes that confer exceptional phenotypic traits necessary for success in the mod- ern workforce, the genomeÆability and abilityÆstatus relationships may increase in strength due to increased competition for the professional jobs, while reducing the relationships for those not so genetically well equipped. As long as we do not experience social closure, there will always be room for talent to exert itself.

MIXING UP THE CLASSES: REGRESSION TO THE MEAN AND HARDY-WEINBERG EQUILIBRIUM

The U-turn in income inequality beginning in the 1970s coupled with our interest in the human genome suggests a disturbing question: Can social class evolve to genetic fi xation? Although social hierarchies are ubiquitous in social species, in no sense can social class be considered an adaptation. Although there are genetic factors at play that result in a person’s attainment of a position in the social hierarchy, there are no genes subject to selection that can place some members of the human species on a trajectory for class fi xation. Social class is not like morphological or physiological features that are functional requirements, and therefore at genetic fi xation. Rather it is a macro phenotype far from the genotype. Social class is a complex compos- ite of a large number of endophenotypes such as intelligence, conscientious- ness, agreeableness, and ambition that are themselves the results of a maze of genetic, epigenetic, and environmental inputs, often including sheer luck arising from random events. Thus, there are many processes on the jour- ney from genes to SES that militate against the hardening of social class 112 Social Class and Crime into caste-like systems. Classes do not naturally beget themselves; only very strong socio-cultural practices such as those present in agrarian cultures can maintain artifi cial social classes across the generations. As will be shown in Chapter 11, class mobility studies plainly show that genetic endowment contributes immensely to an individual’s SES, but this does not mean that the offspring of these individuals will possess the same endowment. Whatever that particular suite of genes is, it will be composed of a number of alleles at different chromosomal loci or at distant loci on the same chromosome, ensuring that very few of which will be passed on to the next generation in toto. This is because during the process of meiosis (gamete formation) the chromosomes separate and the alleles are segre- gated into two different gametes (the law of segregation). Additionally, the alleles coding for the proteins underlying a particular trait will probably reside on separate chromosomes or at distant locations of the same chro- mosome. In the process of gamete formation, the alleles most generally assort independently of one another (i.e., they are not necessarily linked together in transmission [the law of independent assortment]). Segregation and independent assortment ensure that whatever constellation of proteins may have contributed to parental social class will be inherited piecemeal rather than intact, and that the larger the number of alleles contributing to the phenotype, the more likely this is. Regression to the mean is a mechanism that may be refl ecting the effects of independent assortment and segregation. Although the genetic shuffl ing of meiosis ensures a great deal of variability in the traits being passed from parents to offspring, we know that typically bright parents give birth to bright children and that dull parents give birth to dull children. Thus, a sig- nifi cant number of parental traits are passed on. But children born to excep- tionally high IQ parents, despite all the environmental advantages afforded them, are usually less bright than their parents, although they are typi- cally brighter than average. Conversely, children of low IQ parents, despite all the disadvantages this entails, are usually brighter than their parents, although they are still below average. For instance, Sandra Scarr (1996) points out that parents in the professional classes have an average IQ of 120 and their children an average 115, and that unskilled worker parents have an average IQ of 90 and their children an average of 95. Regression to the mean (of the population) is most pronounced at parental extremes. Regression to the mean is common for a number of polygenic traits. Some view the phenomenon as a statistical artifact arising from imper- fect correlations (in this case, between mid-parent IQ and offspring IQ) due to factors that may affect one set of scores but not the other. But to the extent that it refl ects some real underlying phenomenon, it may operate in the following way: Extremely bright and extremely dull parents are, respec- tively, the recipients of a larger than average number of IQ-enhancing or IQ- depressing alleles. Because parents pass on a random half of their genes, not their entire genotype, the odds are greatly against their offspring receiving Social Stratifi cation, the Genome, and Social Structure 113 the same large number of IQ-enhancing or -depressing alleles contributing to the phenotypic trait. The larger the difference between mid-parent IQ and the population mean, the less the odds are, and thus the greater the offspring regression to the population mean. Rushton and Jensen (2005: 263) liken this to rolling a pair of dice and having them come up two sixes (high IQ) or two ones (low IQ), and say that “the odds are on the next roll [the next generation], you will get some value that is not quite as high (or low).” The laws of genetics (the principles of segregation and recombina- tion) work against the hardening of populations into caste-like systems of social stratifi cation: “regression mixes up the social classes, ensures social mobility, and favors meritocracy” (Eysenck and Kamin 1981: 64). A similar phenomenon that operates to prevent caste-like groups emerg- ing for genetic reasons is the Hardy-Weinberg equilibrium (H-WE). The H-WE a basic law of evolutionary genetics stating that allele frequencies in a mating population will remain stable (unaltered) unless evolutionary mechanisms such as , (salient only in small popula- tions), and gene fl ow (the introduction of new alleles into the population) are working. All these mechanisms are operating to some extent, but it makes little difference to the reservoir of genetic variation in large populations. To illustrate with a hypothetical example: suppose there is an autosomal recessive allele (a) with a 1% frequency in the population (probably too large for a deleterious allele in reality, but useful to keep the decimal points from multiplying promiscuously) coding for a protein that results in an average loss of 15 IQ points. If both partners in a mating relationship are homozygous (aa) for the allele, both are affl icted with an IQ signifi cantly below the population average. Since both parents are homozygous for the allele, all of their offspring will be similarly affl icted. Does this imply a cross-generational inheritance of social class (to the extent that IQ indexes SES)? No, it does not. In fact, unaffected heterozy- gous (aA) parents will contribute more of this recessive trait to the next gen- eration than will the affected homozygous parents; here is why. The H-WE informs us why genes having no selective value (at least at present) will be retained in the gene pool. The H-WE is written as a binomial expansion. When dealing with probabilities (p) of an event occurring or not occurring (q), by defi nition p + q = 1. Thus, if the probability of the recessive allele (a) in the population is .01, the probability of the dominant allele (A) is .99. If p + q = 1, then (p + q)2 = 1 also. In binomial expansion: (p + q)2 = p2 + 2pq + q2 = 1. Thus, the allelic frequencies are: a2+ 2aA + A2, = 100% of all alleles at that locus. Given the 1% frequency of the aa allele and doing the math, we have .0001 + (2)(.0099) + .9801 = .0001 + .0198 + .9801 = 1.0. Assuming random mating, the probability of two individuals homozy- gous for the recessive allele and thus certain to have affected offspring is (aa)(aa) = (.0001)(.0001) = .000001. The probability of two heterozygous individuals mating is (aA)(aA) = (.0198)(.0198) = .00039204. The laws of Mendelian genetics tell us that on average one-fourth of the offspring of 114 Social Class and Crime heterozygote matings would be homozygous for the recessive allele (i.e., .00039204/4 = .00009801). Thus, the number of affected offspring of unaffected heterozygotes will be on average 98 times (.00009801/.000001 = 98) greater than the number of affected offspring of homozygotes. Matings between AA and AA individuals will never produce offspring affected by our hypothetical allele. Neither will matings between aA and AA individuals, but an average of one-fourth of the offspring of aA-AA matings will be unaffected carriers of the deleterious allele, thus helping to maintain its frequency in the population. These are idealized calculations assuming random mating, which is almost never the case. Non-random mating is assortative mating; the ten- dency of similar phenotypes to mate. An aa phenotype who manifests the traits associated with low-level IQ is not likely to be attractive to the aA and AA phenotype. Assortative mating is thus an evolutionary process that works against the H-WE and therefore will tend to somewhat maintain social classes. Evolutionary biologists differentiate between positive sand negative assor- tative mating, but for humans the process is primarily positive (“birds of a feather”) rather than negative (“opposites attract”). In other words, we tend to mate with partners who are similar to us on many demographic and attitudinal measures (Luo and Klohnen 2005). A review of the American lit- erature on assortative mating shows that the strongest assortment is for race, religion, and ethnicity, followed by (in descending order) education level, atti- tudes, IQ, SES, height and weight, and personality traits (Hur 2003). Decreased assortative mating should lessen intellectual differences between families. Assortative mating requires assortative meeting, and the lessening institutional and fi nancial constraints on opportunities to meet individuals from other classes should have the effect of mixing up the classes more than was the case in the past. The average correlation between American spouses for SES of .30 (Hur 2003) is piddling compared to what it was 100 years ago when it was very rare to marry outside of a person’s class of origin (van Leeuwen and Maas 2002). In a comparison of assortative mating across 55 countries, Smits (2003) found that in countries where access to higher education is most open the strength of the relation- ship between spouses for SES diminishes (i.e., it becomes more random with respect to class of origin). Open societies in which individuals from diverse backgrounds are thrown together in educational, occupational, and recreational situations will reduce, but by no means eliminate, assortative mating and therefore add to the fl uidity of social class.

CONCLUSION

We cannot satisfactorily understand the dynamics of class stratifi cation without understanding the primordial reasons why the individual pursuit Social Stratifi cation, the Genome, and Social Structure 115 of status is so important. The ultimate reason that high status in one’s social group is so important for males is that status brought with it greater access to females in ancestral environments, and still does today. Of course, this is not necessarily a conscious motivation for status seeking, because status is a valuable asset quite independent of its adaptive evolutionary value. After all, the homosexual, the asexual, the chaste, and the monogamous strive for status also because the desire for status is a universal feature built into us all. The chemistry of social status should be of great interest to crimi- nologists, but it is rarely addressed. This, combined with the evolutionary importance of the family, helps us to make sense of the kinds of behavior often observed in lower-class neighborhoods where altercations are fre- quently sparked over trivial matters of street reputation. A major message of this chapter is that the social class systems of mod- ern societies are fl uid, and that individuals may move up and down the class ladder as never before. It was shown that independent assortment and segregation, regression to the mean, the Hardy-Weinberg equilibrium, and reduced assortative mating for SES are mechanisms that “mix up” the classes and stop them from stagnating. Another major message is that how society is arranged and where one is within that arrangement affects the role that the genome plays in helping one to achieve social status. Adkins and Guo (2008: 252) suggest that sociology can have an impact on future discussions of SES if only it can get over its dread of biology and require its students to learn some: “[W]e suggest that sociology may fi nd a role in the oncoming synthesis of social and biological sciences in the very maxim that has described its reluctance. Structure matters—not only to social pro- cesses, but also to the actualization of genetic potential.” 10 The Nature and Nurture of Intelligence

THE IMPORTANCE OF HUMAN INTELLIGENCE

All normally functioning humans learn the basic requirements of social participation, but the rewards of a modern complex society go to those with the ability to most quickly and effi ciently assimilate a culture forever in fl ux and forever moving in the direction of broadening scope and deepen- ing complexity. Human learning requires cognition, and cognition makes use of symbols—letters, words, pictures, formulas, and so on—that stand for various aspects of the environment and the relationships among them. Because of its increasing importance to success, the issue of differential abil- ity becomes more and more socially relevant: “Intelligence has a profound effect on the structure of society, not necessarily because it is the most highly valued of individual differences—although conceivably it is—but rather because it may have the widest and most stable distribution among all the traits that are valuable in industrialized nations” (Gottfredson 1986: 406). Intelligence is more closely tied to class mobility in modern societies than any other single factor, but having said this we must not confuse intel- ligence with moral worth. As Herrnstein and Murray (1994: 21) point out: “one of the problems of writing about intelligence is how to remind readers often enough how little IQ scores tells you about whether the human being next to you is someone you will admire or cherish.” The root word of intelligence is intelligo, which means “to select among.” Etymologically intelligence implies the ability to select from among a variety of elements (whatever they may be) and analyze, synthesize, and arrange them in such a way as to provide satisfactory, and sometimes novel, solutions to problems the elements may pose. We argue over verbal defi nitions of important concepts because they encompass a multitude of sub-concepts; they would hardly be important if they did not. Most defi ni- tions do contain elements about which there is a fairly general consensus, and while a head-count of experts is no guarantee of accuracy and truth, it does provide an authoritative place to begin. Snyderman and Rothman (1988) conducted a survey of 1,020 Ph.D.’s identifi ed as experts in the area in which they were asked to check all descrip- tors they considered to be important components of intelligence. There was virtual unanimity on some such as “abstract thinking or reasoning” The Nature and Nurture of Intelligence 117 (99.3%), “problem solving ability (97.7%), and capacity to acquire knowl- edge” 96%. More than two-thirds checked “memory,” “adaptation to one’s environment,” and “mental speed.” These elements provide a basis for a defi nition which would differ little from that offered by David Wechsler, the originator of many of today’s most widely used IQ tests, who defi ned it as: “The aggregate or global capacity of the individual to act purpose- fully, to think rationally, and to deal effectively with his environment. It is aggregate or global because it is composed of elements or abilities (features) which, although not entirely independent, are qualitatively differentiable” (in Matarazzo 1976: 79). The terms aggregate and global recognize that different mental tests measure different cognitive abilities but also that there is a factor common to them all that unites the various abstracting functions. Psychometricians call this common factor Spearman’s g, or simply g, which is the principal component derived from factor analysis of a variety of mental tests. Spear- man’s g (“g” for “general” intelligence) is the only independent factor com- mon to all “intelligences” as they are operationalized by various cognitive tests. Factor loadings (i.e., the correlation between a test and the general factor) on g across a number of mental tests are mostly in the range of .50 to .90 (Deary, Spinath, and Bates 2006); thus, there is no measure of learning ability that is independent of g. Psychometricians of all ideological persua- sions agree that g conforms to what both they and laypersons mean when they speak of intelligence (Lubinski 2004). Although Snyderman and Rothman’s (1988) experts, the National Acad- emy of Sciences (Seligman 1992), and the American Psychological Associa- tion’s task force on intelligence (Niesser et al. 1995) concluded that IQ tests are not biased, remarkably there are still some who believe that they are. The only evidence these naysayers need is that certain groups consistently show lower average scores than certain other groups. Proponents of the bias argument do not necessarily mean that the tests are biased in terms of the reliability and validity of the tests, but rather they mean that lower SES children are less likely to have the kinds of intellectual experiences that pre- pare them for some kinds of items found on IQ tests, which confuses two meanings of intelligence: crystallized and fl uid. Crystallized intelligence is heavily dependent on acquired knowledge; fl uid intelligence refl ects the ability to analyze problems in novel ways, to synthesize, and to recognize relationships between the various elements of a problem. Most items on modern tests rely heavily on items that tap into fl uid intelligence (Deary et al. 2006). We cannot develop culture free tests in any absolute sense because any test requiring the use of mental faculties depends on skills and knowledge developed in a culture. Yet lower SES chil- dren score lower on tests designed specifi cally to compensate for cultural deprivation than they do on “culturally loaded” tests (Loehlin 2000). This is because culture-free tests tap heavily into fl uid intelligence and are more g-loaded than tests of crystallized intelligence. The higher the g-load is (the 118 Social Class and Crime more it depends on the common ), the greater the cognitive com- plexity of the test is; the more g-loaded a test is, the higher its heritability is (Jensen, 1998; Lubinski 2004). The crystallized-fl uid distinction tells us why intelligence is not learned in the ordinary sense of the term. IQ is remarkably stable across the life course with correlations as high as .94 over a 10-year span (Deary et al. 2000). IQ remains quite stable even as we acquire vast storehouses of knowledge from childhood to old age. A 60-year-old Ph.D., for instance, may have acquired enough crystallized knowledge to write learned books and articles on esoteric subjects that would have mystifi ed him or her at age 13. But despite increases in knowledge of more than a thousand-fold, his or her IQ would differ at age 60 from that at age 13 by only a few points in either direction. If IQ was acquired through learning as knowledge is, it would increase dramatically across the life course, which it does not. It is fairly well agreed upon that fl uid intelligence stops developing during adolescence—around 14 to 16 years of age (Deary, Johnson, and Houli- han 2009). This is illustrated by the ease with which prepubescent children learn an accent-free foreign language and the diffi culty with which adults learn one, and never accent-free.

THE NEUROBIOLOGY OF INTELLIGENCE

Whatever intelligence is, it involves complex neurobiological processes requiring the participation of many enzymes and neurotransmitters. The amount, concentration, and metabolism of these substances vary from person to person, with variability being under appreciable genetic con- trol (Haier 2003). Variation among these factors will be refl ected in vari- ability in information processing (i.e., the speed and accuracy of nerve impulse activity), which can be accurately measured and compared with a person’s IQ. The crudest biological correlate of intelligence is brain volume. Eleven magnetic resonance imaging (MRI) studies yielded a weighted mean cor- relation with brain volume and intelligence of .381 (Vernon et al. 2000). A more sophisticated measure is reaction time (RT). RT is premised on the notion that more intelligent people possess brains that are able to quickly spot, differentiate, and respond to relevant stimuli. There are three different test of RT. Simple RT measures only the time (in milliseconds) it takes to recognize and respond to a stimuli; recognition RT involves responding to the correct stimuli and ignoring distracter stimuli; and choice RT involves matching a response corresponding to the stimulus, such as pressing a the same letter key that appears on a screen. Correlations between RT and IQ range between –.25 and –.50 (higher IQ = shorter mental reaction time), with the highest correlations being associated with the more challenging RT tasks (reviewed in Deary 2003). The Nature and Nurture of Intelligence 119 The highest correlations are found in more sophisticated measures of brain activity such as cerebral glucose metabolism. Cerebral glucose metab- olism is measured by positron emission tomography (PET) scans which reveal metabolic functioning as the brain takes up positron-emitting glu- cose applied by injection or inhalation. A computer reveals colorized bio- chemical maps of the brain identifying the parts activated while engaged in some task as the energy supplied by the glucose is metabolized. Glu- cose metabolic rates at various brain-slice levels have been correlated with scores on Raven’s Advanced Progressive Matrices, a highly g-loaded test of abstract reasoning. Depending on brain-slice level and hemisphere, cor- relations ranging between –.44 and –.84 are found, which means that high- IQ subjects expend less brain energy when performing intellectual tasks and possess brains that are speedier, more accurate, and more “energy effi - cient” than low-IQ subjects (see Haier 2003 and Gray and Thompson 2004 for reviews). Functional magnetic resonance imaging studies offer more impressive evidence for the biological substrate of intelligence. An MRI study of MZ and DZ twins found that Spearman’s g is related to the amount of gray matter in the brain, and that the amount of gray matter is highly heritable, with a mean heritability of .88 (Glahn, Thompson, and Blangero 2007). Posthuma and his colleagues (2002) showed that the association between brain volume and g is mediated by common genetic factors (i.e., the same genes that infl uence brain volume infl uence g). Such studies pin-point which specifi c areas of the brain are larger or more effi cient in more intelligent people rather than merely indicating larger overall cerebral brain mass (Anderson 2003). The increasing sophistication of brain imaging technol- ogy has allowed researchers to discover that g correlates with gray matter across the frontal, parietal, temporal, and occipital lobes (Colom, Jung, and Haier 2006; Frangou, Chitins, and Williams 2004). Richard Haier and Rex Jung (2007) formulated their Parieto-Frontal Integration Theory (P-FIT) after reviewing 37 neuroimaging studies. P-FIT suggests that intelligence is related to how effi ciently information process- ing is shunted around the brain. Most neuroscientists in the past consid- ered the seat of human intelligence to be found in the frontal lobes until it was conclusively shown that damage to the frontal lobes did not affect IQ even as it affected motivation, foresight, and emotional modulation. Haier and Jung assert that there is no single brain location where intelligence is situated, and place their emphasis on brain areas that multiple neuroim- aging studies had found to be related to IQ. Based on this information they concluded that most of these brain areas are found clustered in the frontal and parietal lobes (the parietal lobes function to integrate sensory information, process visual-spatial details, manipulate objects, determine relationships among numbers, among other things). This fi ts in nicely with the body of research showing that the parieto-frontal network is activated under confl ict situations and when response selection is required (Brass and 120 Social Class and Crime von Cramon 2004). Resolving confl icting possibilities by making the cor- rect selection is, of course, what one engages in during IQ testing. Variation in IQ is thus related to variation in the effi ciently with which these brain networks communicate dynamically with one another.

GENETICS AND INTELLIGENCE

Intelligence is among the most heritable of human traits, with estimates ranging from .50 in childhood to .80 in old age (Neubauer and Fink 2009). Heritability studies provide initial evidence of genetic involvement in a trait, and neuroimaging studies identify anatomical and physiologi- cal correlates, but scientists have still not conclusively identifi ed specifi c genes related to intelligence. Geneticists do not expect to discover genes with large effect sizes for intelligence; rather they expect to fi nd a large number with small effect sizes in quantitative trait loci studies (Craig and Plomin 2006). Early molecular genetic work by Robert Plomin’s team (1994) found a number of DNA markers that differentiated between low- IQ (M = 82) and high-IQ (M = 130) subjects, although it required this large mean IQ gap to locate signifi cant differences. Grigorenko (2000) suggests that the IQ-enhancing affects of alleles may refl ect pleiotropic affects (the multiple affects that a gene can have on an organism via the protein it produces) rather than direct affects, which if true will further complicated the search. A genome-wide study of 500,000 SNPs from 7,000 school children who had taken a battery of IQ tests illustrates just how elusive the search has been and how small the contribution of any single polymorphism is (Butcher et al. 2008). Butcher’s team used a two-stage associative QTL method to examine and eliminate SNPs. They fi nally arrived at six SNPs that were signifi cantly related to g, but their aggregate effect only signifi - cantly correlated 0.11 with g. Nevertheless, molecular geneticists have been relentless in their search for a variety of candidate genes using a wide variety of sophisticated tech- niques. A large number of SNPs and mini- and microsatellites have been signifi cantly or “suggestively” linked to various IQ measures, very few of which are familiar to biosocial criminologists. One of the most familiar is the val158met polymorphism of the COMT gene, which is vital for dop- amine regulation in the prefrontal cortex (Ohnishi et al. 2006). Another candidate is the cholinergic muscarinic 2 receptor gene (CHRM2). Cer- tain SNPs on the CHRM2 gene have been found to signifi cantly corre- late with performance, but not verbal, IQ (Dick et al. 2007). Of course, none of these genetic polymorphisms are genes “for” intelligence. They are polymorphisms involved in certain kinds of brain processing, and then perhaps only in the presence of other polymorphisms and environ- mental instigation. The Nature and Nurture of Intelligence 121 ENVIRONMENTAL INFLUENCES ON INTELLIGENCE

The best and most solid evidence for environmental infl uence on intelligence comes from genetic research: “The fi ndings of greatest social signifi cance to emerge from human behavioral genetic research to date involve nurture, not nature” (Plomin and Daniels 1987: 1). (Recall the breastfeeding x fatty acid desaturase gene interaction effects in Chapter 8). Since any pheno- typic trait is the result of genes interacting with environments, and because the best modern estimates of the percentage of IQ variance attributable to genes at adolescence are between 50 and 60% at adolescence, it follows that between 40 and 50% of the variance has to be attributed to environmental factors minus measurement error. Yet no one has been able to empirically account for as much as 10% of the variance (Plomin and DeFries 1980). This is probably more a function of our inability to nail environmental factors down rather than an underestimation of genetic effects. It is more diffi cult to get a fi rm grasp on environmental infl uences than on genetic infl uences (at least at the level of population genetics) because the “environ- ment” encompasses everything not encoded in the DNA, and these infl u- ences are numerous and interact in a myriad ways. Adoption studies provide some evidence of environmental effects, but Locurto (1990) points out that adoption studies tell us nothing about mal- leability per se absent knowledge of what the IQ’s of children would have been had they not been adopted. Children or infants who may appear to be unhealthy or lethargic or to have symptoms of retardation may be excluded from adoption consideration. However, it is possible to estimate what IQ it might have been by comparing the IQs of adopted children with those of biological siblings who were not adopted. Dumaret’s (1985) study of French adoptees found that the adopted children had a mean IQ advan- tage over their stay-at-home half-siblings of 16 points, and one of almost 29 points over their in-care half-siblings. Unfortunately, some of the children were half-siblings, so genetic differences cannot be discounted. Nor can we discount selectivity on the part of adoptive parents, or the possibility that characteristics associated with the IQ of the in-care children evoked reac- tions from mothers that led to the children being taken into care. A French cross-fostering study by Capron and Duyme (1989) merits attention because of its unique design. The study compared adoptive par- ent-adopted child pairs from both extremes of SES. Parents at the high SES extreme were senior executives and professionals, and at the low extreme they were small farm workers and “diverse unskilled.” All children were relinquished at birth, and the mean age of testing was 14 years. Chil- dren born to high-SES parents, regardless of SES of the adoptive parents, had higher IQs (M = 113.5) than children born to low-SES parents (M = 98.0), and children adopted by high-SES parents, regardless of SES of bio- logical parents, have higher IQs (M = 111.6) than children adopted by low- SES parents (M = 99.95). There was a difference of 27.2 points between 122 Social Class and Crime children both born to, and adopted by, high-SES parents (M = 119.6) and children both born to, and adopted by, low-SES parents (M = 92.4). This is another example of positive rGE. Children born to low-SES parents adopted by high-SES parents show a lower mean IQ (103.6) than children born to high-SES parents adopted by low-SES parents (107.5), despite an aver- age of 14 years exposure to high- and low-SES environments, respectively (negative rGE). This study shows both genetic and environmental effects on IQ, and it supports the view that the reaction range for an individual’s IQ appears to be an average of about 12 points (Locurto 1990). Neuroscience also provides evidence for the importance of learning on brain structure and functioning. Success in modern society requires motivation, inspiration, perspiration, and mastery of a body of acquired (crystallized) knowledge as well as fl uid intelligence. Neuroscience tells us that routinely stretching our minds develops new synapses and neural networks and channels to increase the effi ciency of information process- ing. Acquiring large amounts of abstract knowledge of a particular subject leads to measurable structural changes in the brain, as a number of studies have demonstrated (reviewed in Neubauer and Fink 2009). A study of the IQs of 146 science faculty members of prestigious Cambridge University revealed almost as many scientists with IQ’s in the “bright-average” range (IQ of 111–119) as in the “superior” range (IQ of 120–128) (Eysenck and Kamin 1981). Thus, “genius,” whatever it is, requires more than raw fl uid intelligence: if we convert an IQ of 111 to a Z score, we fi nd that 23% of the general population have a higher IQ. In addition to the g factor, perhaps we should also be looking at what Itzkoff (1987: 161) calls the P (for “post- ponement”) factor, which he describes as “the ability to concentrate, focus, plan, persevere, and above all to postpone the momentary gratifi cations with which life tempts us, all for the sake of the long-term goal.” Of course, genetic factors are also implicated in these “postponement” abilities, but the point is that success, in the workplace or elsewhere, is not determined by raw intelligence alone. My conservative temperament draws me to the oft-quoted words of Thomas Edison: “Success is 10 percent inspiration and 90 percent perspiration.”

THE FLYNN EFFECT

Perhaps the most important evidence of environmental effects on intelli- gence is the so-called Flynn Effect. Flynn (2007) has shown that the pop- ulation IQ mean increased in all countries studied by approximately 3.1 points per decade from 1932 to 2000, or about 21 points. Because the gene pool cannot possibly have changed to account for a change of this magni- tude, the effect has to be attributed to the environment. Our grandparents were not blithering idiots; they had just as much “raw” intelligence as the current generation but, according to Flynn, it was of a more concrete sort The Nature and Nurture of Intelligence 123 and anchored in everyday reality. The increase has been mostly limited to similarities (“in what way is a rabbit like a dog”), which has increased by 24 points, whereas the vocabulary, arithmetic, and information components have increased only 3 points in 55 years (Flynn 2007). Flynn attributes the increase in similarities scores to the increasing use of “the categories of science” that has freed the mind from the concrete and allows us to under- stand abstractions that have no concrete referents. The largest IQ gains have concentrated at the lower end of the distribution as the environment has improved to allow for more of us to realize our intellectual potential. Dickens and Flynn (2001) believe that the evidence of high heritabil- ity of IQ (implying strong genetic effects) with large intergenerational IQ gains (implying strong environmental effects) evidenced over the last seven decades provide us with a paradox that requires reconciliation (note that the paradox is not that more equal environments yield higher heritability coeffi cients; that is a mathematical given). Flynn (2007: 90) avers that the direct genetic effect on IQ is only about 36%, with 64% resulting from the indirect effects of genes interacting with the environment, as well as envi- ronmental effects uncorrelated with genes. The gene-environment interplay results in what Dickens and Flynn (2001: 349) call the “multiplier effect,” and state that “genes can get matched with environments of correspond- ing quality only through genetically-infl uenced traits.” That is, by a match between parents and offspring by which what may have been an initially small genetic advantage is magnifi ed (“multiplied”) into a large phenotypic advantage by passive, evocative, and active rGE. In less complex and less egalitarian times the multiplier effect suppressed the power of environmental effects, much as the effects of poverty mask the power of genetic effects. Dickens and Flynn (2001: 351) explain: “The match between genes and environment means that environmental factors, however potent, to a large extent just reinforce the advantage or disadvan- tage that genes confer. So the match masks the potency of environmental factors.” They go on to say that across the generations on which the Flynn effect has been working, the masking has slipped because of the better envi- ronments to which successive generations have been exposed and: “The potency of environmental factors stands out in bold relief” (Dickens and Flynn 2001: 351). Dickens and Flynn are not saying that rGE is not operating today and is not as individually potent; rather they are saying that the match is not as socially potent as in the past as the environmental advantages enjoyed by the wealthier classes have become more widely available to all. The Flynn effect has worked on the population as a whole and does not imply that IQ levels of individuals can be increased permanently via transient programs designed for that purpose. Dickens and Flynn’s model presup- poses a constant trajectory in the same positive direction. Low-IQ children placed in IQ-enhancing programs do tend to make initial gains, but these gains are quickly lost when the program ends and children return to less 124 Social Class and Crime cognitively demanding environments that match their disadvantaged geno- types (because low-IQ children typically have low-IQ parents) rather than with their temporarily environmentally enhanced IQs (Lubinski 2004). Many specifi c environmental explanations for the Flynn Effect have been offered ranging from increased educational opportunities to test wiseness, to changes in fertility patterns (smaller families), to the sheer complexity of modern life. For instance, computer video games have provided a mea- sure of mental stimulation to practically all children in modern societies that more cerebral children obtained from simply reading in the past, and even today’s TV shows, according to Flynn (2007), demand more cognitive involvement than the I Love Lucy- and Leave It to Beaver-type offerings of yesteryear . In the same way that molecular genetics posits a role for numer- ous genes with small effect sizes, it is probably true that the Flynn Effect is caused by a combination of all of these and other environmental factors, each having a small effect size. Richard Lynn (2009) provides the most empirically sound explanation for the Flynn Effect attributable to a single variable with a large effect size. Lynn notes that the rise in IQ has been matched by a corresponding increase in the Development Quotients (DQs) of children over the same time period. DQs are measured in infants by the Bayley Scales and the Grifi ths Test that assess motor (holding up the head, sitting, standing, walking, jumping) and mental (uttering words, displaying curiosity, naming objects, responding to requests) development. Lynn notes that the malnutrition and mineral defi ciencies that caused rickets, anemia, and many other environmentally induced problems prevalent in the 1930s had all but disappeared by the 1970s in developed societies and are virtually unknown today. Lynn also notes that height and head size have increased during the period, and that DQ is signifi cantly related to IQ (DQ at 6 to 12 months is correlated with IQ at .53 when corrected for unreliability). Lynn claims that these data rule out other proposed explanations for IQ gains such as those mentioned ear- lier since increased DQs occur prior to exposure to these things. Improve- ments in pre- and postnatal care and nutrition thus provide us with foci with which to try to alleviate cognitive defi ciencies in the current cohort of underprivileged children. The Flynn Effect appears to have ended in developed nations (as well as height gains), meaning that we seem to have wrung all the juice out of the environment that we can (Sundet, Barlaug, and Torjussen 2004). We might even be witnessing a decline in mean IQ since the high point achieved in 1998, although the Flynn Effect is still in evidence in developing countries (Teasdale and Owen 2008). Dennis Garlick (2002, 2003) makes much the same point as Dickens and Flynn from a neurodevelopmental perspective, arguing that low-IQ individuals have brains that do not adapt well to environmental condi- tions. Garlick (2002, 2003) posits an initial genetic set-point (for differen- tial brain plasticity rather than differential intelligence per se) and suggests The Nature and Nurture of Intelligence 125 that environmental factors then provide the necessary input to realize one’s intellectual potential. Garlick’s neuroplasticity model of intelligence argues that different individuals have different set-points of brain plastic- ity just as they have, for instance, different metabolic rates or different levels of natural strength. The model posits that intelligence is “created” when neural connections in the brain are changed and strengthened by environmental input, and more or less of it will be created according to the differential capacity of individuals’ neurons to adapt (due to genetic polymorphisms associated with neurotransmitters, receptors, transporters, and enzymes). Garlick argues that there is a critical period in which the developing brain fi ne-tunes its ability to adapt to novel phenomena. This period begins in infancy when the human brain is most plastic; that is, most able to make new neural connections in response to experience, and ends around puberty. The work of Garlick and Dickens and Flynn, the neurosci- ence evidence regarding structural brain changes with experience, and the diffi culty that geneticists have discovering intelligence-related genes sug- gest that intelligence may be as much an emergent trait as a latent trait, or perhaps even more so.

THE IQ-CRIME CONNECTION

As the quintessentially human characteristic, it is no surprise that intel- ligence has long been considered an important correlate of criminal behav- ior. Claims of test bias and the assumed eugenic implications of IQ testing following World War II led to the virtual disappearance of IQ from the criminological literature from the 1930s to the 1970s, even though many well-known studies continued to demonstrate a negative relationship between IQ and crime. A number of recent reviews have also characterized the relationship as ubiquitous and robust (see Ellis and Walsh 2003 for a review). The National Longitudinal Survey of Youth (NLSY) found a number of interesting relationships between IQ and a variety of lifetime outcomes, including the fact that 93% of the males in the sample who had ever been interviewed in a correctional facility over a 10-year period had IQs in the bottom half of the IQ distribution. Furthermore, of those subjects located at the bottom 20% of IQ (IQ ≤ 87), 62% had been interviewed in jail or prison at some point over the study period compared with only 2% from the top 20% (IQ ≥ 113)—a ratio of 31:1 (Herrnstein and Murray 1994: 376). Despite such evidence, many criminology textbooks still imply that there is little evidence to support the claim of average IQ difference between offenders and non-offenders (Wright and Miller 1998). IQ may be related to crime and delinquency more strongly than suggested by a simple comparison of mean IQ levels of offenders and non-offenders. Offenders’ IQs are typically compared with the general population mean 126 Social Class and Crime IQ of 100. It is often forgotten that the general population includes a fairly large number of offenders, as well as individuals with such low IQs that they are largely incapable of committing crimes. Thus, the difference in IQ between offenders and normally functioning non-offenders (the mean of which is probably around 103–105) is actually greater than reported. A related problem is that many IQ studies are conducted with delin- quents rather than with adult criminals. We know that almost all teenage boys commit some acts which could get them into trouble with the law, and also that most delinquents do not become adult criminals (Moffi tt 1993). Criminals who offend most frequently and most seriously tend to begin their antisocial careers prior to puberty and to continue them long after the typical delinquent has desisted from offending. Delinquents who desist in early adulthood have accrued enough social capital to allow them to do so, much of that by virtue of their cognitive abilities. It has been pointed out that the IQ difference between non-offenders and adolescent-onset offend- ers is typically only one IQ point, but the same comparison with life-course persistent offenders reveals about a 17-point difference (Gatzke-Kopp et al. 2002; Moffi tt 1993). Simple arithmetic tells us that pooling these two very different groups hides the magnitude of the IQ difference between non-offenders and the most persistent and serious offenders in our midst, and thus the true strength of the relationship between IQ and criminality. David Wechsler’s (1958: 176) statement—“The most outstanding feature of the sociopath’s test profi le is the systematic high score on the perfor- mance as opposed to the verbal part of the scale”—sparked another way of examining the relationship between IQ and antisocial behavior. This con- cept of intellectual imbalance is popular among correctional psychologists. IQ scores are typically rendered in terms of an individual’s full-scale IQ (FIQ). FIQ is the average score obtained by summing scores on verbal IQ (VIQ) and performance IQ (PIQ) and dividing by two. Most people have closely matched VIQ and PIQ scores, with a population average of 100 on each sub-scale. People who have either VIQ or PIQ scores signifi cantly in excess of the other (VIQ > PIQ or PIQ > VIQ) are considered intellectually imbalanced. Correlating FIQ with offending is problematic because offend- ers are almost always found to have signifi cantly lower VIQ, but not lower PIQ, than non-offenders (Walsh 2003). Miller (1987: 120) concludes the following about the PIQ > VIQ imbalance: “This PIQ > VIQ relationship was found across studies, despite variations in age, sex, race, setting, and form of the Wechsler scale administered, as well in differences in criteria for delinquency.” In the general population of American males, it has been estimated that 18% are VIQ > PIQ imbalanced, 66% are balanced (VIQ = PIQ), and 16% are PIQ > VIQ imbalanced. A discrepancy of 12 points or more is consid- ered a signifi cant imbalance at p < .01 (Kaufman 1976). It is consistently found that VIQ > PIQ imbalanced individuals are signifi cantly under- represented in criminal and delinquent populations, and that PIQ > VIQ The Nature and Nurture of Intelligence 127 individuals are signifi cantly overrepresented. Averaged across eight studies, VIQ > PIQ boys are underrepresented in delinquent populations by a factor of about 2.6, and PIQ > VIQ boys are overrepresented by a factor of about 2.2, rendering an odds ratio of about 5.7 (Walsh 2003). A VIQ > PIQ pro- fi le would appear to be a major predictor of prosocial behavior, especially among adults. Barnett et al. (1989) found that only 0.9% of prison inmates had such a profi le compared to the 18% of the general male population; a 20-fold difference. Using FIQ to assess the role of cognitive ability in explaining antisocial behavior provides another example of how research- ers can arrive at an incomplete picture of the nature of the relationship.

CONCLUSION

Intelligence, as measured by IQ, has huge impact on life outcomes via its impact on education, occupation, and income, and thus on SES. I have not looked at these outcomes in this chapter, but I have briefl y discussed the nature of intelligence as currently understood by psychologists, geneticists, and neuroscientists. I have emphasized that intelligence is a phenotypic trait developed through the constant interplay of genes and environment and have offered evidence for this. The Flynn Effect is certainly good news in that we have evidence that the social environment is more conducive today than ever before in helping us all to realize whatever genetic potential we possess. It is still unavoidably axiomatic that the more the environment provides equal opportunities the more biological factors will be responsible for differences among individuals. No single book, never mind a single chapter, can do justice to the fasci- nating topic of intelligence and its relationship to so many things in life. For the reader interested in getting to the bottom of the concept of intelligence, I would recommend Flynn’s (2007) excellent book and Jung and Haier’s (2007) more technically challenging P-FIT theory of intelligence. 11 Class Mobility Ascription or Achievement?

STATUS IN A MERITOCRATIC SOCIETY

There are two main contending positions regarding how individuals become situated in the class structure. The mainstream sociological position is that the social advantages and disadvantages of childhood SES largely determine adult education, occupation, and income, and therefore adult SES. SES is seen as socially “inherited” in this view and does not require any explana- tion beyond that. This is called the class structuralist or status ascription position. The other position is the meritocratic or status attainment posi- tion, held primarily by psychologists and economists, and avers that indi- vidual cognitive ability and motivation largely determine adult SES, and that they do so regardless of childhood SES (Nielsen 2006). Both positions agree to the extent that adult SES is achieved via educa- tion, occupation, and income, but the disagreement is the extent to which the burden of achievement or non-achievement of these things is on social structure or on individual differences in cognitive and personality attri- butes. Proponents of neither position reject all aspects of the other. That is, the meritocratic model realizes that parental SES has some effect on offspring SES via both social and genetic routes, and the ascription model realizes that abilities do infl uence success, although they insist that such abilities are themselves a function childhood advantage-disadvantage based on parental SES. The end-point of both models is class inequality regardless of the route. Class structuralists tend to view this as unfair and as leading to a host of class-related problems such as crime. For instance, Steven Messner and Rich- ard Rosenfeld (2001: 9) point out that the competition for economic success requires inequality of outcomes because “winning and losing have meaning only when rewards are distributed unequally.” They also claim that inequal- ity in the United States is not an aberration of the American Dream but an expression of it. It is because inequality is a natural outcome of competition that: “High crime rates are intrinsic to American society. . . . at all social levels, America is organized for crime” (emphasis added). From this posi- tion a meritocracy is both criminogenic and unfair. Sawhill and Morton (2007: 4) also question the fairness of the American meritocracy when they write: “people are born with different genetic endowments and are raised Class Mobility 129 in different families over which they have no control, raising fundamental questions about the fairness of even a perfectly functioning meritocracy.” Claims of unfairness are claims made by unconstrained visionaries who imagine solutions without trade-offs. For their part, constrained visionar- ies view unequal outcomes as necessary and not al all unfair, but rather desirable because they function to motivate healthy competition without which we would have economic stagnation. The alleged problems caused by inequality are unfortunate trade-offs. Those with this vision see inequal- ity as desirable because the meritocratic system ultimately confers benefi ts on everyone, including its losers, by providing a dynamic economy and a free polity. For instance, Federal Reserve chairman and one of the world’s top economists Ben Bernanke (in Sawhill and Morton 2007: 1) evaluates capitalism’s requirement for unequal outcomes quite differently from Mess- ner and Rosenfeld:

Although we Americans strive to provide equality of economic op- portunity, we do not guarantee equality of economic outcomes, nor should we. Indeed, without the possibility of unequal outcomes tied to differences in effort and skill, the economic incentives for productive behavior would be eliminated and our market-based economy—which encourages productive activity primarily through the promise of fi nan- cial reward—would function far less effectively.

INTELLIGENCE AND SES

Intelligence as measured by IQ tests is a member of a class of social science “evils” that must be excoriated and exorcised, or at least ignored in the hope that it will go away. To discuss intelligence as a determinant of a per- son’s social class is nothing less than sociological heresy. Anyone spending an hour or so perusing books on social stratifi cation would be struck by the absence of such discussions. The largest of these books that I examined, a densely packed 750 pages (Grusky 1994), does not even have the terms IQ or intelligence listed in the index. If we do encounter a discussion of IQ, we are subjected to “explanations” without evidence of why intelligence has nothing to do with anything, or ad hominem attacks against those who supply evidence that it as a lot to do with just about everything in the modern social world. Lee Ellis (1996: 28) also comments about sociology’s negative stance toward IQ and on the absence on any discussion of intelligence in soci- ological theories of class: “Someday historians of social science will be astounded to fi nd the word intelligence is usually not even mentioned in late-twentieth-century text books on social stratifi cation.” IQ tests were designed to be a benign class-neutral measure of aptitude that was sup- posed to turn schools into capacity-catching institutions by siphoning off 130 Social Class and Crime the brightest children to provide an increasingly complex economic system with competent workers (Pinker 2002), not malignant measures of exclu- sion. Whatever one’s IQ score, it is not an indicator of innate inferiority, moral value, or social worthiness, but rather a measure of the probability of being able to master certain tasks of an intellectually demanding nature. The litmus test for any assessment tool is its ability to predict outcomes (criterion-related validity). IQ tests do a particularly good job with regard to occupational success. A fi rst-order analysis of eleven meta-analyses of the relationship between IQ and occupational success found that IQ pre- dicted success better than any other variable in most occupations (Got- tfredson 1986). As one would expect, the more intellectually demanding and the higher the status of the job, the better predictor it was. The analy- ses also showed that IQ predicted occupational success equally well across all class and racial/ethnic lines. Another study of over 32,000 workers in 515 different occupations revealed that the correlations between IQ and job performance rose from .23 for “low complexity” jobs to .58 for “high complexity” jobs (Gottfredson 1997). Intelligence is particularly important in technologically advanced soci- eties in which low complexity occupations become mechanized and high complexity occupations become more prevalent. Modernization requires open competition for the most rewarding jobs with everything being sec- ondary to talent, the engine driving modern economies. Employers compete for talented employees, and IQ testing has been a major tool in all modern societies to locate them. Whether you view the role of intelligence testing in modern societies positively or negatively, the fact remains that in open societies high-IQ individuals tend to achieve higher SES than do individuals with lower IQs.

HOW MOBILE IS THE AMERICAN CLASS STRUCTURE?

We can agree that there is no such thing as a totally open society, and that being born into an upper- or middle-class family confers many advantages, and that being born into the lower class brings with it many disadvan- tages. Unlike the rigid class societies of pre-modern times when status was ascribed to individuals by others, today it is something that is attained by dint of effort. At the present time class mobility in the United States is less robust than it is in Canada or in most nations of Western Europe (Sawhill and Morton 2007), but this could be the result of America’s eco- nomic tide rising and lifting everyone’s boat earlier, and that these other nations are simply catching up. Nevertheless, according to one major study occupational mobility is still quite substantial in the United States (Hurst 1995). This study found that 51% of sons of lower-manual status fathers achieved higher status, with 22.5 achieving “upper white-collar” status, and that 48% of sons of upper white-collar status fathers had lower status Class Mobility 131 occupations than their fathers, with 17% falling all the way to “lower man- ual” status (Hurst 1995: 270). Hurst (1995: 276) concluded: “There was a great deal of movement both within and between generations. Well over half of the sons moved out of the occupational strata of their fathers and out of the strata of their own fi rst jobs.” Nevertheless, there is still a certain “stickiness” to social class since once a family has attained an advantaged class position it is not happy to see its offspring generation relinquish it. It has been calculated that in Britain the odds ratios of offspring of the highest class falling to the lowest at the same time that offspring in the lowest class rising to the highest class is 7.4 for men and 3.3 for women (Saunders 2002). This indicates a high degree of mobility, but not enough to satisfy those who believe that genuine equal- ity of opportunity means an odds ratio of 1.0 (Goldthorpe 1987). An even ratio might make sense if there were an equal percentage of highly intelli- gent and motivated individuals in the lowest class as there is in the highest, and an equal percentage of dull and unmotivated individuals in the highest class as there is in the lowest, but who really believes that to be the case? Given unequal talents and motivation, an odds ratio anywhere near 1.0 would imply that mobility occurs at random and that it is not at all tied to achievement. The fact remains that more undeserving offspring of the higher classes fail to fall down the class ladder than deserving offspring of the lower classes fail to climb up it, but surely this is a positive thing (more “losers” become “winners” than “winners” become “losers”). As Saunders (2002: 561) explains: “the main ‘blockage’ in social mobility has less to do with talented lower-class individuals failing to move up the system than with less talented higher-class individuals failing to move down it.” In other words, while there are few barriers preventing the talented children from the lower classes moving up the class ladder, less-than-talented children from the higher classes enjoy a number of safeguards against failure not enjoyed by less-than-talented children from the lower classes, such as parental sup- port and encouragement, and parental infl uence and contacts. There is, of course, no way that a society could (or should) prevent parents’ efforts on their children’s behalf—after all, nepotism runs deep in our genes.

ASCRIPTION VERSUS ACHIEVEMENT: THE DATA

When compelled to confront the correlation between IQ and SES, sociolo- gists tend to consider a child’s IQ to be an effect of his or her parents’ SES (i.e., a reproduction of parental class advantage or disadvantage in the off- spring generation). A number of studies have found the correlation between parental SES and children’s IQ to be within the .30 to .40 range (Lubinski 2004). These correlations are predictable from polygenic models since a proportion of the genes underlying the traits that contributed to parents’ 132 Social Class and Crime achieved class are transmitted to offspring. The real test is not the corre- lation between parental SES and offspring IQ, but rather the correlation between offspring IQ and their own adult SES. As Jensen remarks: “If SES were the cause of IQ [rather than the other way around], the correlations between adults’ IQ and their attained SES would not be markedly higher than the correlation between children’s IQ and their parents’ SES” (1998: 491). Correlations between offspring IQ and offspring attained SES across a number of studies are in the .50 to .70 range (Lubinski 2004). Offspring IQ is thus a considerably more powerful predictor of offspring SES than is parental SES. The weak affect of parental SES on offspring SES in the United States is almost considered a truism outside of sociology. For instance, Kingston (2006: 121) writes: “The net impact of measured family background on economic success is easy to summarize: very little. This conclusion holds across different data sets with different model specifi cations and measure- ments and applies to both occupational status and earnings.” For instance, DiRago and Viallant (2007) examined 60 years of data from males born between 1925 and 1932 who had lived in poor high-crime neighborhoods in Boston when growing up. These men were interviewed about their occupational status when they were 25, 32, 47, and 65 years of age. Attrition whittled the original 500 males down to 345 at age 65. It was found that parental SES was signifi cantly related to offspring occupa- tional status (r = .17) at age 25, but the correlation progressively dwindled to a non-signifi cant .079 by age 65, by which time none of the measured environmental factors were related to occupational status. Years of educa- tion (–.410) and IQ (–.347) were related to occupational success at age 65 (occupational status was coded 1 = professional down to 7 = unskilled). The results support the behavior genetic “law” that states that as we age the effects of shared environment (in this case, childhood SES and all the bag- gage that comes with it) on phenotypic traits fade to insignifi cance while genes and non-shared environments become more salient. Rod Bond and Peter Saunders (1999) pitted the class structuralist posi- tion and status attainment positions against one another in a longitudinal study of 4,298 British males. The study found that individual meritocratic factors (intelligence, aspirations, motivation, and possession of educational qualifi cations) that were assessed when subjects were 11 and 16 years old accounted for 48% of the variance in occupational status at age 33. All measured background variables (parental SES, type of housing, type of school, and parents’ aspirations) combined accounted for only 8% of the variance. Based on this 6-fold difference in the proportion of variance explained, Bond and Saunders concluded that: “occupational selection in Britain appears to take place largely on meritocratic principles” (1999: 217). The remaining variance would have to be attributed to unmeasured variables such as the quality of the subjects’ interpersonal relationships, other personality traits, random chance, and measurement errors. Class Mobility 133 Another British longitudinal study (Nettle 2003) followed a cohort of all children born in Britain in one week in 1958 to the age of 42. All children had their mental abilities tested at the age of 7. At the cohort age of 42 there were 11,419 remaining subjects for whom there were mental ability test scores available. The primary fi nding of this study was that childhood intellectual ability is associated with class mobility in adulthood uniformly across all social classes of origin. Nettle found an IQ difference of 24.1 points between those who attained professional status and those in the unskilled class, regardless of the class or origin. Parental SES accounted for only 3% of the variation in these individuals’ attained social class after partialling out general mental ability. Nettle concluded that “intelligence is the strongest single factor causing class mobility in contemporary societies that has been identifi ed” (2003: 560). An American behavior genetic study pitting the ascription thesis against the achievement thesis also found strongly in favor of the latter Nielsen (2006). This study consisted of 1,072 sibling pairs that included MZ and DZ twins, full siblings, half siblings, cousins, and non-genetically related pairs. The variables examined were verbal IQ (VIQ), grade-point average (GPA), and college plans (CPL). Partitioning the variance of these three measures into the usual genetic, shared environment, and non-shared envi- ronment components, the three heritability coeffi cients were VIQ = .536, GPA = .669, and CPL = .600; the shared environment coeffi cients were VIQ = .137, GPA = .002, and CPL = .030, and the non-shared environmental variances were .327, .329, and .370, respectively. Shared environment, of course, is everything shared by siblings as they grew up, including parental SES. The shared environment component also captures other factors in the pantheon to which sociologists appeal to explain important life outcomes such as neighborhood and school characteristics, and ethnic status. The proportions of variance explained in these three vital components of occu- pational success by SES of origin are miniscule compared with the propor- tions explained by genes and non-shared environment. Charles Murray (2002) tackled the SES of origin vs. IQ issue head on as it relates to individual class attainment by controlling for the entire com- plex of variables that constitute the SES environment of rearing. He created what he called a “utopian sample” of sibling pairs to compare with con- trol subsample of sibling pairs from the same data set. Murray’s utopian sample was composed of sibling pairs who were born in wedlock, who had not experienced parental divorce by age 8, and who grew up in families with a median income of $64,586 in 2000 dollars. The utopian subsample consisted of 733 sibling pairs, and the “other” subsample, consisting of all sibling pairs failing one or more of the criteria for inclusion in the utopian subsample, numbered 1,075 pairs. Murray stated that he wanted to deter- mine “How much difference would it make to income inequality if, magi- cally, every child in the country could be given the same advantages as the more fortunate of our children” (2002: 140). 134 Social Class and Crime Table 11.1 shows that the sibling pairs in the utopian subsample enjoyed a considerable family background advantage over the “other” subsample. The second panel shows that this childhood advantage survived into adulthood since in every IQ category the adult members of the utopian sample enjoyed greater median family incomes (in 2000 dollars and assessed when subjects were 30–38 years old) across all IQ categories. However, the within subsam- ple differences based on IQ levels are considerably greater than the between subsample differences. Comparisons across the rows reveal the effects of childhood advantaged or disadvantaged and comparisons down the columns show the effects of individuals’ cognitive ability (i.e., siblings with higher IQs do substantially better on all indicators of success than their siblings with lower IQs). These data show that income inequality will exist in a meritocratic society even if, as Murray says, “by some miracle” we could equalize fi rst- generation family income, convince women not to give birth out of wedlock, and prevent divorce during children’s formative years. The intellectual abili- ties of the second generation will reproduce inequality by earning less, giving birth out-of-wedlock more and divorcing more, all of which is predictable from their IQ alone. Adding heritable personality variables also linked to the outcomes of interest to the equation sheds further light on class attainment.

Table 11.1 Background Data of Family of Origin of NLSY Sibling Pairs

Family Demographics Utopian (733 pairs) Other (1,075 pairs)

Median Parental Income $64,586 $25,858 Fathers with 16+ years of 19% 10% education Offspring born out of wedlock 0 13% Parents divorced before children 0 32% were age 8 Data on Adult Sibling Pairs Broken Down by IQ Category Women Who Have Married with Given Birth Out of Median Family Income Children* Wedlock** IQ Category Utopian Other Utopian Other Utopian Other

120+ $70,700 $65,100 87% 82% 2% 10% 110–119 $60,500 $59,900 91% 83% 10% 7% 90–109 $52,700 $41,600 78% 62% 17% 30% 80–89 $39,400 $31,300 63% 48% 33% 40% < 80 $23,600 $18,400 43% 39% 44% 69%

All income data in 2000 dollars. *Utopian pairs = 469; Other pairs = 412. ** Utopian pairs = 291; Other pairs = 303. Source: Adapted from Tables 1, 2, 3, and 4 in C. Murray (2002). The American Economic Review, vol. 92. Class Mobility 135 TEMPERAMENT AND PERSONALITY

Chamorro-Premuzic and Furnham (2005: 352) inform us that tempera- ment and intelligence are “the two great pillars of differential psychology.” They add that these two constructs are vital to predicting all kinds of life outcomes, including social class attainment. As we have seen, temperamen- tal variation is important in the differential development of conscience, and temperament and intelligence are identifi ed by Agnew (1992) as impor- tant factors in determining prosocial versus antisocial responses to strain. Temperament is defi ned as: “individual differences in emotional, motor, and attentional reactivity [that are] biologically based and linked to an individual’s genetic endowment” (Rothbart 2007: 207). Temperament is the phenotypic expression of the organism’s habitual mode of emotion- ally responding to stimuli. The various components of temperament include mood (happy/sad), activity level (high/low), sociability (introverted/extra- verted), reactivity (calm/excitable), and affect (warm/cold), among others. These components are continua rather than dichotomies, and depending on where individuals are on the various measures they are either easy or diffi cult for others to like and get along with. People with temperamental diffi culties will fi nd it tough to form social bonds and to compete success- fully in life. Numerous studies have shown that children with disinhibited and irritable temperaments evoke negative responses from parents, teach- ers, and peers, and that these children fi nd acceptance only by peers with similar temperamental dispositions (reviewed in Sanson et al. 2004). Temperamental differences among individuals have their basis in heritable variation in central (primarily limbic) and autonomic nervous system arousal patterns, with estimates for the different components ranging from .20 to .80 (Kagan 2007; Saudino 2005). This pattern indicates that environmen- tal experiences impact the development of temperament weakly or strongly, depending on the component. Although temperament is fairly stable across the life course, environmental input can strengthen or weaken innate pro- pensities: “Temperament develops, that is, emotions and components of emo- tions appear at different ages” (Rothbart, Ahadi and Evans, 2000: 124). It is these differences in temperament that make children differentially responsive to socialization. Because of genetic similarity, the temperaments of parents and children are typically (but not always) positively correlated; children who are temperamentally unresponsive to socialization will have their unresponsiveness exacerbated by the fact that their parents are incon- sistent disciplinarians, irritable, impatient, and unstable. This renders them unable or unwilling to cope constructively with their children, thereby sad- dling them with both a genetic and an environmental liability (Saudino 2005). As Caspi (2000: 170) summarizes this literature:

In the early years of life, person-environment covariation occurs be- cause of the joint transmission of genes and culture from parents to offspring. Given that parents and children resemble each other in 136 Social Class and Crime temperamental qualities, children whose diffi cult temperament might be curbed by fi rm discipline will tend to have parents who are incon- sistent disciplinarians, and the converse is also true: Warm parents tend to have infants with an easy temperament. Later in life, person- environment covariation occurs because people choose situations and select partners who resemble them, reinforcing their earlier established interaction style.

Longitudinal studies typically fi nd that children from low SES families are over-represented at the “problematic end of temperament dimensions, especially those relating to child diffi culty” (Sanson et al. 2004: 158). Caspi, Bem, and Elder’s (1989) longitudinal study illustrates the hetero- geneity of negative outcomes that can arise from a single temperamental dimension. This study identifi ed males from middle-class families with or without a history of temper tantrums in childhood and traced them for 30 years investigating multiple areas of their lives. The majority of bad- tempered boys ended up in lower-status occupations than their fathers, had erratic work histories, and experienced more unemployment than other males with more tranquil temperaments. They were also more than twice as likely as other men to be divorced by age 40. Women with ill-tempered dispositions as children also experienced many negative life outcomes in the Caspi et al. (1989) study. The tended to marry men from classes lower than their own, were more than twice as likely as their more pleasant sisters to be divorced, had signifi cantly more marital confl ict, and were almost universally described by their husbands and chil- dren as ill-tempered mothers. This study illustrates that men and women with a tendency to be dis- agreeable and impulsive, and to respond more often with negative rather than positive emotionality, tend to take their temperamental “style” with them wherever they go, from home, to school, to work, to marriages, and that its cumulative effect can land one at the low end of the class hierarchy via erratic work histories and marital disruptions. This does not mean that a negative temperament is a general property of lower-class individuals; it only means that individuals with such a temperament lack the personal resources to move up the status hierarchy or to prevent themselves from sliding down it even if, like the individuals in the Caspi et al. (1989) study, they have middle-class origins. Temperament is the biological superstructure upon which personality is built. We can think of temperament as a computer with factory-installed hard wiring, and personality as the sum of all the soft-wired modifi cations its owner makes to it over his or her lifetime. Personality it therefore the relatively enduring, distinctive, integrated, and functional set of psycho- logical traits that results from an individual’s temperament interacting with his or her environment over the life course. Class Mobility 137 There are potentially hundreds of personality traits that could be enu- merated depending upon how fi ne a distinction one wants to make. Mod- ern psychology has boiled down most of these diverse traits into fi ve broad categories known as the “Big Five” or the “Five-Factor Model.” These fi ve personality traits are derived from factor analyses in which a number of specifi c traits cluster together (i.e., are highly inter-correlated) into broad categories or factors. These fi ve factors are openness, conscientiousness, extraversion, agreeableness, and neuroticism. All these traits are moder- ately to strongly heritable (Lensvelt-Mulders and Hettema 2001), are rela- tively stable across the life course, and are predictive of many patterns of behavior. Because SES is a refl ection of patterns of behavior, certain of these traits may contribute considerably to one’s level of class attainment. I will concentrate only on the two personality traits consistently and most strongly found to be related to both occupational success (and therefore SES) and to criminal and other forms of antisocial behavior—conscien- tiousness and agreeableness (Wiebe 2004).

CONSCIENTIOUSNESS AND AGREEABLENESS

Conscientiousness is the broad term applied to a number of sub-traits that range from well-organized, disciplined, scrupulous, orderly, responsible, and reliable at one end of the continuum to disorganized, careless, unreli- able, irresponsible, and unscrupulous at the other (Lodi-Smith and Roberts 2007). Conscientiousness is particularly important to success in the work- force and climbing the class ladder, and thus, if the anomie/strain argument has merit, to criminal behavior. Behavior genetic studies of conscientious- ness fi nd a mean heritability of .49 (Bouchard et al. 2003). Agreeableness is the tendency to be friendly, considerate, courteous, helpful, and cooperative with others. Agreeable persons tend to trust oth- ers, to compromise with them, to empathize with and aid them. This list of sub-traits suggests a high degree of concern for prosocial conformity and social desirability. Disagreeable persons simply display the opposite characteristics—suspicious of others, unfriendly, uncooperative, unhelpful, and lacking in empathy—which all suggest a lack of concern for proso- cial conformity and social desirability. A pooled heritability estimate of .48 from Canadian and German samples for agreeableness is reported by Jang et al. (1998). As we might expect, conscientiousness and agreeableness are positively correlated but far from perfectly (about .39 [Witt et al. 2002]). A person can be very conscientious at work but thoroughly disagreeable socially (think of the Machiavellian white-collar criminal). Conversely, one can be most agreeable to everybody but lackadaisical at work (think of the happy-go-lucky ritualist of Mertonian theory). Thus, we might intuit that 138 Social Class and Crime agreeableness is a stronger protective factor against antisocial behavior than conscientiousness, and that conscientiousness is a stronger predictor of occupational success than agreeableness, although both will be nega- tively related to antisocial behavior and positively related to occupational success. From a meta-analysis of personality and various forms of antisocial behavior, Miller and Lynham (2001: 780) describe the personality of the “typical” criminal in terms of agreeableness and conscientiousness thusly: “Individuals who commit crimes tend to be hostile, self-centered, spiteful, jealous, and indifferent to others (i.e., low in Agreeableness). They tend to lack ambition, motivation, and perseverance, have diffi culty controlling their impulses, and hold nontraditional and unconventional values and beliefs (i.e., are low in Conscientiousness).” They go on to note that these personality traits conform to those that Gottfredson and Hirschi (1990) include as components of their low self-control construct. Sociologically trained criminologists have traditionally dismissed the notion that personality contributes to antisocial behavior, but the correla- tions between antisocial behavior and various personality constructs are often considerably larger than they are between with antisocial behav- ior and most sociological variables. For instance, Miller and Lynam’s (2001) meta-analysis comparing 29 /non-prisoner samples found a weighted mean effect size of -.41 for agreeableness and antisocial behav- ior and -.25 for conscientiousness and antisocial behavior. Another meta- analysis (Saulsman and Page 2004) examining the relationship between the Big Five and each of the behavioral disorders listed in the Diagnostic and Statistical Manual of Mental Disorders (DSM-IV) found mean effect sizes for antisocial personality disorder over 15 studies of–.35 for agreeableness and–.26 for conscientiousness. Jacobwitz and Eagan’s (2006) study of a general community sample found that agreeableness was signifi cantly negatively related to all of what they call the “dark triad” of personality traits: primary psychopathy (–.43), secondary psychopathy (–.23), Machiavellianism (–.41) and narcissism (–.43). Conscientiousness was also negatively related to all four measures, but only signifi cantly so for Machiavellianism (–.27) and narcissism (–.24). Finally, Miller, Lynam, and Jones (2008) found signifi cant negative cor- relations between agreeableness and substance abuse (–.30), antisocial behavior (–.32), and aggression (–.56). They also found weaker signifi cant correlations between conscientiousness and antisocial behavior (–.19) and aggression (–.15). Employers can be expected to favor high levels of conscientiousness in their employees and perspective employees, and they do. Conscientious- ness is more important in high-autonomy jobs than in low-autonomy jobs because it “affects motivational states and stimulates goal setting and goal commitment” (Schmidt and Hunter 2004: 169). In an intergenerational study following subjects from early childhood to retirement, Judge and his Class Mobility 139 colleagues (1999) found that conscientiousness measured in childhood pre- dicted adult occupational status (r = .49) and income (r = .41) in adulthood, which were only slighter less than the correlations between “general mental ability” (GMA) and the same variables (.51 and .53, respectively). Schmidt and Hunter’s (2004:170) analysis of GMA and personality variables in attaining occupational success concluded that “the burden of prediction is borne almost entirely by GMA and conscientiousness.”

CONCLUSION

The evidence is overwhelming that in modern societies SES is achieved by dint of one’s own talents and efforts. This fact must be assimilated by ano- mie/strain theorists, subcultural theorists, and Marxist/confl ict theorists who continue to blind themselves to the body of literature cited in this chapter and continue to work under the assumption that social class is caused by the failures of social institutions. If class is the result of the fail- ure of social institutions, and if class is a cause of crime, than criminal behavior must ultimately be blamed on these same institutional failures. This is a proposition losing its cachet, even among sociologically trained criminologists, as the evidence seeping in from the more advanced sciences is becoming too strong to ignore. The whole notion of criminal behavior as the result of “strained” individuals righteously hitting back at a society that has denied them opportunities to obtain status and resources legitimately is deservedly pooh-poohed by Vold, Bernard, and Snipes (1998:177): “It is not merely a matter of talented individuals confronted with inferior schools and discriminatory hiring practices. Rather, a good deal of research indi- cates that many delinquents and criminals are untalented individuals who cannot compete effectively in complex industrial societies.” Epilogue

Having explored the class-crime relationship at some length, it is time to arrive at some overall conclusions. I hope that I have managed to convince the reader that there is nothing to fear and everything to gain from a per- spective that endeavors to incorporate genes, culture, brains, developmental twists, and evolutionary history into its thinking and from one that rejects any suggestion that nature and nurture are opponents. We are designed by nature to incorporate nurture and are only able to express our natures via the medium of nurture. Modern biology and neuroscience reveals that every animal is born prepared to easily incorporate the essential lessons of its species; we are all born pregnant with built-in assumptions that are activated by experience-expected events. As Tittle (1983) pointed out in Chapter 6, it is anomie/strain theory that most clearly draws a straight arrow from social class to crime. It is with this theory in mind that I have examined the class-crime relationship because it clearly points the way to the next step in the analysis but never itself took that step. That is, while it is unambiguously a class-based the- ory, it never asks what the precursors of social class are. In anomie/strain theory social class seems to be something that is sui generis—an uncaused fi rst-cause that requires no explanation. This naturally leads to “blame society” explanations for the criminal behavior of the lower classes. The issue of SES as ascribed or achieved is thus very important to criminol- ogy. If SES is “caused” by social structure, and if crime is “caused” by SES, then ultimately social structure “causes” crime, and this is where we should concentrate our efforts at amelioration. On the other hand, if SES is “caused” by individual cognitive and temperamental differences, and if crime is “caused” by SES, then ultimately crime is “caused” by individual choices conditioned by factors both internal and external. My persistent point is that in an open society SES is attained by the dint of one’s efforts rather than ascribed by “society.” To the extent that SES and criminal behavior are linked, both are “caused” by the individual traits and characteristics with which each of us enter the status competition. It is undeniable that some people are advantaged in the game by virtue of having been given a strong start, and others are disadvantaged by virtue Epilogue 141 of being born on the wrong side of the tracks. It is interesting to note that Robert Merton, the formulator on anomie/strain theory, was the son of Jewish immigrants and was raised in slum conditions (Lilly, Cullen, and Ball 2007). He achieved the American dream by dint of his own ability and efforts, yet his theory glosses over human agency and the power of self- reliance in favor of positing blank-slated creatures blown hither and thither by the capricious winds of an environment they have no part in creating.

FAIRNESS AND MERITOCRACY

What about the moral issue of fairness with which Messner and Rosenfeld (2001) and Sawhill and Morton (2007) were so concerned? The concept of fairness appeals to our moral sentiments because fairness is a process by which we expect to “make things right.” Who can fail to feel sorry for individuals burdened with environmental and genetic disabilities they did nothing to create, and who would not like to make it right for them? But having sympathy for such individuals does not tell us how their position in life coheres with the moral issue of fairness. Fairness is a philosophical issue saturated with contradictory notions; we all praise it but differ as to when its promise is fulfi lled. Constrained visionaries view fairness as an equal opportunity process—a non-discriminatory chance to play the game—which governments can guarantee via law. Unconstrained vision- aries tend to view fairness as equality of outcome—all participants are winners—which no power on earth can guarantee. In the constrained view different outcomes are considered fair if the pro- cess is fair, and the process is considered fair if everyone is subjected to the same rules and judged by the same standards. With the rules of the game and the standards of judgment held constant, the only things that vary are the qualities that individuals bring to it. Life is like a poker tournament; we all get dealt our hands and must play the best way we can within the rules of the game. We can hone our skills and pray to lady luck; we win some and we lose some, but we can’t all win jackpots. Some of us move up the status hierarchy, some of us move down, and some of us stay where our parents put us. All of us bring advantages and disadvantages with us to the game that help or hinder us, but they are not the fault of the game. A 5’6”, 155-lb man does not aspire to the NFL or NBA if he has any sense of reality at all, nor does anyone suggest that we change the rules of those games or devise an athletic version of affi rmative action to accommodate such men. The rules of the game virtually guarantee that the likes of Peyton Manning and Kobe Bryant (physically blessed + talent + effort) will succeed in those are- nas while “Mr. Average” will not—it’s really quite that simple. No doubt some of the dull siblings represented in Murray’s (2002) study feel it grossly unfair that their smarter siblings fare so much better in life than they do 142 Social Class and Crime simply because they inherited different snippets of their parent’s DNA. The more realistic among them will realize that fairness has nothing to do with it, because there is no being or “system” on earth upon whom they could lay blame for their misfortune. We can only talk of fairness or unfairness if some entity has the power to distribute resources to equally deserving individuals and deliberately favors some over others. Of course, no one is in charge of handing out parents, genes, or environments; we just fi nd our- selves existing with these gifts of fate and must do our best with them. The insinuation of the moral issue of fairness-unfairness into the realm of class and status attainment helps us understand why we so often see Sowell’s unconstrained visionaries viciously attacking constrained vision- aries, but rarely, if ever, see the opposite. It is the unconstrained vision’s fusion of so many issues with morality that leads holders of that vision to consider those who disagree with them as immoral supporters of inequal- ity rather than simply wrong-headed. As Thomas Sowell (1987: 227) puts it: “Implications of bad faith, venality, or other moral and or intellectual defi ciencies have been much more common in the unconstrained vision’s criticism than vice versa.” Unconstrained visionaries seem to believe that if apparently intelligent and well-educated individuals oppose programs they believe could improve the lives of the less fortunate, or if they research dif- ferences among races, sexes, and classes, they must be evil racists, sexists, classists, or some other hissing epithet. On the other hand, constrained visionaries never accuse their opponents of deliberately opposing the com- mon good, even if they see them as inadvertently doing just that. Con- strained visionaries see their opponents as well-meaning but naive, wrong, or unrealistic; rarely do they resort to ad hominem attacks that impugn their moral characters. For the unconstrained visionary, morally all people should be equal in their intellectual potential and other positive traits, so therefore they are. Those who subscribe to this moralistic fallacy are logically committed to appealing to forces outside the individual to explain observed intellectual differences since variance can only be explained by something that varies (“society’s” treatment of individuals). Even if unequal talents are recog- nized, many on the left would still like to see all the peas in the pot distrib- uted equally regardless of contributions individuals made to fi ll the pot. We all love equality, but as French philosopher wrote: “equality is at once the most natural and the most chimerical thing in the world: natural when it is limited to rights, unnatural when it attempts to level goods and power” (in Durant 1952: 245). It is unnatural in that it defi es the Aristo- telian notion of treating equals equally and unequals unequally according to relevant differences, which has long been considered the essence of dis- tributive justice by philosophers around the world. References

Adkins, D. & G. Guo. 2008. “Societal development and the shifting infl uence of the genome on status attainment.” Research in Social Stratifi cation and Mobil- ity 26:235–255. Agnew, R. 1992. “Foundations for a general strain theory of crime and delin- quency.” Criminology 30:47–87. . 2005. Why do criminals offend? A general theory of crime and delin- quency. Los Angeles: Roxbury. Alcock, J. (2005). Animal behavior: An evolutionary approach. Sunderland, MA: Sinauer Associates. Altukhov, Y., and E. Salmenkova. 2002. “DNA polymorphisms in population genetics.” Russian Journal of Genetics 38:1173–1195. Amato, P. and B. Keith. 1991. “Parental divorce and well-being of children: A meta-analysis.” Psychological Bulletin 110:26–46. Anderson, B. 2003. Brain imaging and g. Pp. 29–39 in The scientifi c study of gen- eral intelligence: Tribute to Arthur R. Jensen, edited by H. Nyborg. Oxford: Elsevier. Anderson, E. 1994. “The code of the streets.” The Atlantic Monthly, 5:81–94. . 1999. Code of the street: Decency, violence, and the moral life of the inner city. New York: W. W. Norton. Andersen, S. (2003). Trajectories of brain development: point of vulnerability or window of opportunity? Neuroscience and Biobehavioral Reviews, 27:3–18. Anderson, W. and C. Summers. 2007. “Neuroendocrine mechanisms, stress coping strategies, and social dominance: Comparative lessons about leadership potential.” Annals of the American Academy of political & Social Science 614:102–130. Anestis, S. 2006. “Testosterone in juvenile and adolescent chimpanzees (Pan trog- lodytes): Effects of dominance, rank, aggression, and behavioral style.” Ameri- can Journal of Physical Anthropology 130:536–545. Archer, J. 2006. “Testosterone and human aggression: An analysis of the challenge hypothesis.” Neuroscience and Biobehavioral Reviews, 30:319 –345. Badcock, C. 2000. Evolutionary psychology: A critical introduction. Cambridge: Polity Press. Baker, L., S. Bezdjian, and A. Raine. 2006. “Behavior genetics: The science of anti- social behavior.” Law and Contemporary Problems 69:7–46. Barak, G. 1998. Integrating criminologies. Boston: Allyn and Bacon. Barkow, J. 1989. Darwin, sex and status: Biological approaches to mind and cul- ture. Toronto: University of Toronto Press. . 1992. Beneath new culture is an old psychology: Gossip and social strati- fi cation. Pp. 627–637 in : Evolutionary psychology and the generation of culture, edited by J. Barkow, L. Cosmides, and J. Tooby. New York: Oxford University Press. 144 References

Barnett, C. 2002. The measurement of white-collar crime using uniform crime reporting (UCR) data. U.S. Department of Justice. Washington, DC: Govern- ment Printing Offi ce. Barnett, R., L. Zimmer, and J. McCormack 1989. P>V sign and personality pro- fi les. Journal of Correctional and Social Psychiatry 35:18–20. Bartels, M. and J. Hudziak 2007. “Genetically informative designs in the study of resilience in developmental psychopathology.” Child and Adolescent Psychiat- ric Clinics of North America 16:323–339. Bates, J. and G. Pettit 2007. Temperament, parenting, and socialization. Pp. 153– 177 in Handbook of socialization: Theory and research, edited by J. Grusec and P. Hastings. New York: Guilford Press. Bauer, C., J. Lange, S. Shankaran, H. Bada, B. Lester, L. Wright, H. Krause-Stein- rauf, V. Smeriglio, L. Finnegan, P. Maza, and J. Verter. 2005. “Acute neonatal effects of cocaine exposure during pregnancy.” Archives of Pediatric and Ado- lescent Medicine, 159:824–834. Beaver, K., J. Wright, M. DeLisi, A. Walsh, M. Vaughn, D. Boisvert, and J.Vaske. 2007. “A gene X gene interaction between DRD2 and DRD4 in the etiology of conduct disorder and antisocial behavior.” Behavioral and Brain Functions 30:1–8. Beaver, K., J. Wright, and A. Walsh. 2008. “A gene-based evolutionary explanation for the association between criminal involvement and number of sex partners.” Biodemography and Social Biology 54:47–55. Bell, M. and K. Deater-Deckard. 2007. “Biological systems and the development of self-regulation: Integrating behavior, genetics, and psychophysiology.” Journal of Developmental & Behavioral Pediatrics 28:409–420. Bellinger, D. 2008. “Neurological and behavioral consequences of childhood lead exposure.” PLoS Medicine 5:690–692. Benson, M. and K. Kerley. 2001. “Life course theory and white-collar crime.” Pp. 121–136 in Contemporary issues in crime and criminal justice: Essays in honor of Gil Geis, edited by H. Pontell and D. Shicor. Benson, M. & E. Moore (1992). Are white-collar and common offenders the same? An empirical and theoretical critique of a recently proposed general theory of crime. Journal of Research in Crime and Delinquency, 29:251–272. Bentham, J. 1789/1948. A fragment on government and an introduction to the principles of morals and legislation, edited by W. Harrison. Oxford: Basil Blackwell. Blumstein, A. 1995. “A LEN interview with Professor Alfred Blumstein of Carn- egie Mellon University.” Law Enforcement News, 21:10–13. Bond, R. and P. Saunders. 1999. “Routes of success: Infl uences on the occupational attainment of young British males.” British Journal of Sociology 50:217–240. Bonger, Willem. 1969. Criminality and economic conditions. Bloomington: Indi- ana University Press. Booth, A., K. Carver, and D. Granger. 2000. “A biosocial perspective on the fam- ily.” Journal of Marriage and the Family, 62:1018–1034. Bouchard, T., N, Segal, A. Tellegen, M. McGue, M. Keyes, and R. Krueger. 2003. “Evidence for the construct validity and heritability of the Wilson-Patterson scale: A reared-apart twins study of social attitudes.” Personality and Individual Differences 34:959–969. Bowles, S. and H. Gintis. 1976. Schooling in capitalist America. New York: Basic Books. . 2002. “Schooling in capitalist America revisited.” Sociology of Education 75:1–18. Braithwaite, J. 1981. “The myth of social class and criminality reconsidered.” American Sociological Review 46:36–57. References 145

Brass, M. and D. von Cramon. 2004. “Selection for cognitive control: A functional magnetic resonance imaging study on the selection of task-relevant informa- tion.” Journal of Neuroscience 24:8847–8852. Brennan, P., E. Grekin, and M. Sarnoff. 1999. “Maternal smoking during preg- nancy and adult male criminal outcomes.” Archives of General Psychiatry 56:215–219. Brennan, P., A. Raine, F. Schulsinger, L. Kirkegaard-Sorenen, J. Knop, B. Hutch- ings, R. Rosenberg, and S. Mednick. 1997. Psychophysiological protective fac- tors for male subjects at high risk for criminal behavior. American Journal of Psychiatry 154:853–855. Bronfenbrenner, U. and S. Ceci. 1994. “Heredity, environment, and the question ‘how’—A fi rst approximation.” Pp. 313–324 in Nature, nurture, and psychol- ogy, edited by R. Plomin and G. McClearn. Washington, DC: American Psy- chological Association. Brunero, J. 2002. “Evolution, altruism and internal reward explanations.” - sophical Forum 33:413–424. Buss, D. 2001. “Human nature and culture.” Journal of Personality 68:955–978. . 2005. The murderer next door: Why the mind is designed to kill. New York: Penguin. Buss, D. and J. Duntley. 2006. “The evolution of aggression.” Pp. 263–285 in Evolution and , edited by In M. Schaller, J. Simpson, and D. Kenrick. New York: Psychology Press. Butcher, L., O. Davis, I. Craig, and R. Plomin. 2008. “Genome-wide quantitative trait locus association scan of general cognitive ability using pooled DNA and 500K single nucleotide polymorphism microarrays.” Genes, Brain, and Behav- ior 7:435–446. Cadoret, R., W. Yates, E. Troughton, G. Woodworth, and M. Stewart. 1995. “Genetic-environmental interaction in the genesis of aggressivity and conduct disorders.” Archives of General Psychiatry 52:916–924. Calavita, K. and H. Pontell. 1994. “‘Head I win, tails you lose’: Deregulation, crime, and crisis in the savings and loan industry.” Pp. 460–480 in Contemporary soci- eties: Problems and prospects, edited by in D. Curran and C. Renzetti. Boston: Ally & Bacon. Campbell, A. 2006. “Sex differences in direct aggression: What are the psychologi- cal mediators?” Aggression and Violent Behavior 6:481–497. Campbell, A. 2009. Gender and crime: An evolutionary perspective. In A. Walsh & K. Beaver (Eds.), Criminology and Biology: New directions in theory and research, 117-136. New York: Routledge. Campbell, A., S. Muncer, and D. Bibel. 2001. “Women and crime: An evolutionary approach.” Aggression and Violent Behavior 6:481–497. Canli, T. 2006. Biology of personality and individual differences. New York: Guilford. Capron, C. and M. Duyme. 1989. “Assessment of effects of socioeconomic status on IQ in a full cross fostering study.” Nature 340:552–554. Carey, G. 2003. Human genetics for the social sciences. Thousand Oaks, CA: Sage. Carnagey, N., C. Anderson, and B. Bushman. 2007. “The effect of video game violence on physiological desensitization to real violence.” Journal of Experi- mental Social Psychology 43:489–496. Carrasco, X., P. Rothhammer, M. Moraga, H. Henríquez, R. Chakraborty, F. Aboitiz, and F. Rothhammer. 2006. “Genotypic interaction between DRD4 and DAT1 loci is a high risk factor for attention-defi cit/hyperactivity disorder in Chilean families.” American Journal of Medical genetics (Neuropsychiatric Genetics) 141B:51–54. 146 References

Caspi, A. 2000. The child is the father of the man: Personality continuities from Child- hood to adulthood. Journal of Personality and Social Psychology 78:158–172. Caspi, A., D. Bem, and G. Elder. 1989. “Continuities and consequences of interac- tion styles across the lifecourse.” Journal of Personality 57:375–406. Caspi, A., J. McClay, T. E. Moffi tt, J. Mill, J. Martin, I. W. Craig, A. Taylor, and R. Poulton. 2002. “Role of genotype in the cycle of violence in maltreated chil- dren.” Science 297:851–854. Caspi, A., K. Sugden, T. Moffi tt, A. Taylor, I Craig, H. Harrington, J. MacClay, J. Mill, J. Martin, A. Braithwaite, and R. Poulton. 2003. “Infl uence of life stress on depression: Moderation by a polymorphism in the 5-HTT gene.” Science 301:386–389. Caspi, A., B. Williams, J. Kim-Cohen, I. Craig, B. Milne, R. Poulton, L. Schalk- wyk, A. Taylor, H. Werts, and T. Moffi tt. 2007. “Moderation of breastfeeding effects on the IQ by genetic variation in fatty acid metabolism.” Proceedings of the National Academy of Sciences 104:1–6. Casswell, S., M. Pledger, and R. Hooper. 2003. “Socioeconomic status and drink- ing patterns in young adults.” Addiction 98:601–610. Cauffman, E., L. Steinberg, and A. Piquero, A. 2005. “Psychological, neuropsy- chological, and psychophysiological correlates of serious antisocial behavior in adolescence.” Criminology 43:133–176. Cecil K., C. Brubaker, C. Adler, K. Dietrich, M. Altaye, J.Egelhoff, S. Wessel, I. Elangovan R. Hornung, K. Jarvis, and B. Lanphear. 2008. “Decreased brain volume in adults with childhood lead exposure.” PLoS Medicine 5:742–750. Cernkovich, S., P. Giordano, and M. Pugh. 1985. “Chronic offenders: The missing cases in self-report delinquency research.” Journal of Criminal Law and Crimi- nology 76:705–732. Chakraborty, B., H. Lee, M. Wolujewicz, J. Mallik, G. Sun, K. Dietrich, A. Bhat- tacharya, R. Deka, & R. Chakraborty. 2008. Low dose effect of chronic lead exposure on neuromotor response impairment in children is moderated by genetic polymorphisms. Journal of Human Ecology, 23:183–194. Chamorro-Premuzic, T. and A. Furnham. 2005. “Intellectual competence.” The Psychologist 18:352–354. Chen, Z. and H. Kaplan. 2003. “School failure in early adolescence and status attainment in middle adulthood: A longitudinal study.” Sociology of Education 76:110–127. Cohen, A. 1955. Delinquent boys. New York: Free Press. Collins, R. 2004. “Onset and desistence in criminal careers: Neurobiology and the age-crime relationship.” Journal of Offender Rehabilitation 39:1–19. Colum, R., R. Jung, and R. Haier. 2006. “Distributed brain sites for the g-factor of intelligence.” NeuroImage 31:1359–1365. Conger, R. and S. Dogan. 2007. Social class and socialization in families. Pp. 13–41 in Handbook of socialization: Theory and research, edited by J. Grusec and P. Hastings. New York: Guilford Press. Cooley-Quille, M., R. Boyd, E. Frantz, and J. Walsh. 2001. “Emotional and behav- ioral impact of exposure to community violence in inner-city adolescents.” Jour- nal of Clinical Child Psychology 30:199–206. Cooper, A., A. Perkins, and P. Corr. 2007. A confi rmatory factor analytic study of anxiety, fear, and behavioral inhibition system measures. Journal of Individual Differences 4: 179–187. Cooper, J., A. Walsh, & L. Ellis. 2010. Is Criminology Ripe for a Paradigm Shift? Evidence from a Survey of American Criminologists. Journal of Criminal Jus- tice Education, (in press). References 147

Cornish, D., and R. Clarke, eds. 1986. The reasoning criminal. New York: Springer-Verlag. Corr, P. 2004. “Reinforcement sensitivity theory and personality.” Neuroscience and Biobehavioral Reviews 28:317–332. Coser, L. (1971). Masters of sociological thought. New York: Harcourt Brace Jovanovich. Craig, I. and R. Plomin. 2006. “Quantitative trait loci for IQ and other complex traits: Single-nucleotide polymorphism genotyping using pooled DNA and microarrays.” Genes, Brain and Behavior, 5:32–37. Crow, T. 2007. “How and why genetic linkage has not solved the problem of psy- chosis. Review and hypothesis.” American Journal of Psychiatry 164:13–21. Cullen, F. 2009. Foreword, In A. Walsh and K. Beaver, Biosocial Criminology: New directions in theory and research. New York: Routledge, xv–xvii. Cunradi, C., R. Caetano, and J. Schafer. 2002. “Socioeconomic predictors of inti- mate partner violence among white, black, and Hispanic couples in the United States.” Journal of Family Violence 17:377–389. D’Alessio, S. and L. Stolzenberg. 2003. “Race and the probability of arrest.” Social Forces 81:1381–1397. Daly, M. and M. Wilson. 1985. “Child abuse and other risks of not living with both parents.” Ethology & Sociobiology 6:197–210. . 1988. “Evolutionary social psychology and family homicide.” Science 242:519–524. . 1996. “Violence against stepchildren.” Current directions in psychological science 5: 77–81. . 2005. “Human behavior as animal behavior.” Pp. 393–408 in Behavior of animals: mechanisms, function, and evolution, edited by J. J. Bolhuis and L. A. Giraldeau. Oxford: Blackwell Publishing. Day J. and R. Carelli. 2007. “The nucleus accumbens and Pavlovian reward learn- ing.” The Neuroscientist 13:148–159. Deary, I. 2003. “Reaction time and psychometric intelligence: Jensen’s contri- butions.” Pp. 53–75 in The scientifi c study of general intelligence: Tribute to Arthur R. Jensen, edited by H. Nyborg. Oxford: Elsevier. Deary, I., W. Johnson, and L. Houlihan. 2009. “Genetic foundations of human intelligence.” Human Genetics 126:215–232. Deary, I., F. Spinath, and T. Bates. 2006. Genetics of intelligence. European Jour- nal of Human Genetics 14:690–700. Deary, I., L. Whalley, H. Lemmon, J. Crawford, and J. Starr. 2000. The stability of individual differences in mental ability from childhood to old age: Follow-up of the 1932 Scottish mental survey. Intelligence 28:49–55. Dennett, D. 1995. Darwin’s dangerous idea: Evolution and the meanings of life. New York: Simon & Schuster. Department of Health and Human Services. 2004. “Breastfeeding practices— Results from the National Immunization Survey.” http://www.cdc.gov/breast- feeding/data/NIS_2004.htm. Depue, R. and P. Collins. 1999. “Neurobiology of the structure of personality: Dopamine, facilitation of incentive motivation, and extraversion.” Behavioral and Brain Sciences 22:491–569. de Waal, F. 2002. “Evolutionary psychology: The wheat and the chaff.” Current Directions in Psychological Science 11:187–191. . 2008. “Putting the altruism back into altruism: The evolution of empathy.” Annual Review of Psychology 59:279–300. Dick, D., F. Aliev, J. Kramer, J. Wang, A. Hinricks, S. Bertlesen, S. Kuperman, M. Schuckit, J, Nurnberger, H. Edenburg, P. Porjesz, H. Begleiter, V. Hasselbrock, 148 References

A Goate, and L. Bierut. 2007. “Association of CHRM2 with IQ: Converging evidence for a gene infl uencing intelligence.” Behavior Genetics 37:265–272. Dickens, W. and J. Flynn. 2001. “Heritability estimates versus large environmental effects; The IQ Paradox resolved.” Psychological Review 108: 346–349. Ding, Y., H. Chi, D. Grady, A. Morishima, J. Kidd, K. Kidd, P. Flodman, M. Spence, S. Schuck, J. Swanson, Y. Zhang, & R. Moyziz. 2002. “Evidence of positive selection acting at the Human dopamine receptor D4 gene locus.” Pro- ceedings of the National Academy of Science 99:309–314. DiRago, A. and G. Viallant. 2007. “Resilience in inner city youth: Childhood pre- dictors of occupational status across the lifespan.” Journal of Youth Adoles- cence 36:61–70. Dodge, K., G. Pettit, and J. Bates. 1994. “Socialization mediators of the relation between Socioeconomic status and child conduct problems.” Child Develop- ment 65:649–645. Dumaret, A. 1985. “IQ, scholastic performance and behavior of sibs raised in contrasting environments.” Journal of Child Psychology and Psychiatry 26:553–580. Dunbar, R. and S. Shultz. 2007. “Evolution of the social brain.” Science 317:1344–1347. Durant, W. (1952). The story of philosophy. New York: Simon and Schuster. Durkheim, E. 1951. Suicide. Translator J. Spalding. New York: Free Press. . 1956. Education and sociology. Translator S. Fox. New York: Free Press. Edelman, G. 1992. Bright air, brilliant fi re. New York: Basic Books. Ellis, L. 1996. “A discipline in peril: Sociology’s future hinges on curing its biopho- bia.” American Sociologist 27:21–41. . 2005. A theory explaining biological correlates of criminality. European Journal of Criminology 2:287–315. Ellis, L. and J. McDonald. 2001. Crime, delinquency, and social status: A reconsid- eration. Journal of Offender Rehabilitation 32:23–52. Ellis, L. and A. Walsh. 2000. Criminology: A global perspective. Boston: Allyn & Bacon. . 2003. “Crime, delinquency and intelligence: A review of the worldwide literature.” Pp. 345–365 in The scientifi c study of general intelligence: A tribute to Arthur R. Jensen, edited by H. Nyborg. Amsterdam: Pergamon. Elrod, P., I. Soderstrom, and D. May. 2008. “Theoretical predictors of delinquency in and out of school among a sample of rural public school; youth.” Southern Rural Sociology 23:131–156. Ember, M. and C. Ember. 1998. “Facts of violence.” Anthropology Newsletter October:14–15. Ernst, M., D. Pine, & M. Hardin (2006). Triadic model of the neurobiology of motivated behavior in adolescence. Psychiatric Medicine, 36:299–312. Eshel, N., E. Nelson, R. Blair, D. Pine, and M. Ernst. 2007. “Neural substrates of choice selection in adults and adolescents: Development of the ventrolateral pre- frontal and anterior cingulated cortices.” Neuropsychologia 45:1270–1279. Evans, P., S. Gilbert, N. Mekel-Bobrov, E. Vallender, J. Anderson, L. Vaez- Azizi, S. Tishkoff, R. Hudson, and B. Lahn. 2005. “Microcephalin, a gene regulating brain size, continues to evolve adaptively in humans.” Science 309:1717–1720. Eysenck, H. and L. Kamin. 1981. The intelligence controversy. New York: John Wiley. Farley, J. 1990. Sociology. Englewood Cliffs, NJ: Prentice Hall. Felson, R. and D. Haynie. 2002. “Pubertal development, social factors, and delin- quency among adolescent boys.” Criminology 40:967–988. References 149

Fergusson, D., N. Swain-Campbell, and J. Horwood. 2004. “How does childhood economic disadvantage lead to crime?” Journal of Child Psychology and Psy- chiatry 45:956–966. Fishbein, D. 2001. Biobehavioral perspectives in criminology. Belmont, CA: Wadsworth. Fletcher, R. 1991. “Mating, the family, and marriage: A sociological view.” Pp. 111–162 in Mating and Marriage, edited by In V. Reynolds and J. Kellett. Oxford: Oxford University Press. Flynn, J. 2007. What is intelligence? Beyond the Flynn effect. Cambridge: Cam- bridge University Press. Fraga, M., E. Ballestar, M.F. Paz, S. Ropero, F. Setien, M.L. Ballestar, D. Heine- Suñer, J.C. Cigudosa, M. Urioste, J. Benitez, M. Boix-Chornet, A. Sanchez- Aguilera, C. Ling, E. Carlsson, P. Poulsen, A. Vaag, Z. Stephan, T.D. Spector, Y.Z. Wu, C. Plass, and M. Esteller. 2005. Epigenetic differences arise during the lifetime of monozygotic twins.” Proceedings of the National Academy of Sciences 102:10604–10609. Frangou, S., X. Chitins & S. Williams. 2004. Mapping IQ and gray matter density in healthy young people. NeuroImage, 23:800–805. Friedman, N., A. Miyake, S. Young, J. DeFries, R. Corely, and J. Hewitt. 2008. “Individual differences in executive functions are almost entirely genetic in ori- gin.” Journal of Experimental Psychology 137:201–225. Galvan, A., T. Hare, C. Parra, J. Penn, H. Voss, G. Glover, and B. Casey. 2006. “Earlier development of the accumbens relative to orbitofrontal cortex might underlie risk-taking behavior in adolescents.” The Journal of Neuroscience 26:6885–6892. Gane, N. 2005. “Max Weber as social theorist: ‘Class, Status, Party.’” European Journal of Social Theory 8: 211–226. Garlick, D. 2002. “Understanding the nature of the general factor of intelligence: The role of Individual differences in neural plasticity as an explanatory mecha- nism.” Psychological Review 109:116–136. . 2003. “Integrating brain science research with intelligence research.” Cur- rent Directions in Psychological Science 12:185–192. Gatzke-Kopp, L., A. Raine, R. Loeber, M. Stouthamer-Loeber, and S. Steinhauer. 2002. “Serious delinquent behavior, sensation seeking, and electrodermal arousal.” Journal of Abnormal Child Psychology 30:477–486. Geary, D. 2005. The origin of mind: Evolution of brain, cognition, and general intelligence. Washington, DC: American Psychological Association. Giedd, J. 2004. “Structural magnetic resonance imaging of the adolescent brain.” Annals of the New York Academy of Science 1021:77–85. Gintis, H. 2003. “The hitchhiker’s guide to altruism: Gene-culture coevolu- tion and the internalization of norms.” Journal of Theoretical Biology 220:407–418. Glahn, D., P. Thompson, and J. Blangero. 2007. “Neuroimaging endophenotypes: Strategies for fi nding genes infl uencing brain structure.” Human Brain Map- ping 28:461–463. Glaser, D. 2000. “Child abuse and neglect and the brain—A review.” Journal of Child Psychology and Psychiatry 41:97–116. Goldsmith, H. and R. Davidson. 2004. “Disambiguating the components of emo- tion regulation.” Child Development 75: 361–365. Goldthorpe, J. 1987. Social mobility and class structure in modern Britain. Oxford: Clarendon Press. Goldstein, D. and I. Kopin. 2007. “Evolution of the concept of stress.” Stress 10:109–120. 150 References

Gottfredson, L. 1986. “Social consequences of the g factor in employment.” Jour- nal of Vocational Behavior 29:379–410. . 1997. “Why g matters: The complexity of everyday life.” Intelligence, 24:79–132. Gottfredson, M. and T. Hirschi. 1990. A general theory of crime. Stanford, CA: Stanford University Press. Gottlieb, G. 2007. “Probabilistic epigenesis.” Developmental Science 10(1):1–11. Gottesman, I. and D. Hanson. 2005. “Human development: biological and genetic processes.” Annual Review of Psychology 56:263–286. Gray, J. and P. Thompson. 2004. “Neurobiology of intelligence: Science and eth- ics.” Nature Reviews: Neuroscience, 5:471–482. Grigorenko, E. 2000. “Heritability and intelligence.” Pp. 53–91 in Handbook of Intelligence, edited by R. Sternberg. Cambridge: Cambridge University Press. Gross, M. 2005. “The evolution of parental care.” The Quarterly Review of Biol- ogy 80:37–46. Grosvenor, P. 2002. “Evolutionary psychology and the intellectual left.” Perspec- tives in Biology and Medicine 45:433–448. Grusec, J. and M. Davidov. 2007. “Socialization in the family: The roles of par- ents.” Pp. 284–308 in Handbook of socialization: Theory and research, edited by J. Grusec and P. Hastings. New York: Guilford Press. Grusec, J. and P. Hastings, eds. 2007. Introduction. Pp. 1–9 in Handbook of social- ization: Theory and research. New York: Guilford Press. Grusky, D. 1994. Social stratifi cation: Class, race, and gender in sociological per- spective. Boulder, CO: Westview Press. Gunnar, M. and K. Quevedo. 2007. “The neurobiology of stress and development.” Annual Review of Psychology 58:145–173. Guo, G., M. Roettger, and J. Shih. 2007. “Contributions of the DAT1 and DRD2 genes to serious and violent delinquency among adolescents and young adults.” Human Genetics 121:125–136. Guo, G., Y. Tong, and T. Cai. 2008. “Gene by social context interactions for num- ber of sexual partners among white male youths: Genetics-informed sociology.” American Journal of Sociology 114:S36–66. Hacker, J. 2006. The great risk shift: The new insecurity and the decline of the American dream. New York: Oxford University Press. Hacking, I. 2006. “Genetics, biosocial groups & the future of identity.” Daedalus 135: 81–95. Haier, R. 2003. “Positron emission tomography studies of intelligence: From psy- chometrics to neurobiology.” Pp. 41–51 in The scientifi c study of general intel- ligence: Tribute to Arthur R. Jensen, edited by H. Nyborg. Oxford: Elsevier. Hall, S. 2002. “Daubing the drudges of fury: Men, violence and the piety of the ‘hegemonic masculinity thesis.” Theoretical Criminology 6:35–61. Hamilton, W. 1964. “The evolution of social behavior.” Journal of Theoretical Biology 7:1–52. Harris, A. and J. Shaw. 2001. “Looking for patterns: Race, class, and crime.” Pp. 129–163 in Criminology: A contemporary handbook , edited by J. Shelley. Bel- mont, CA: Wadsworth. Harris, J. 1998. The nurture assumption: Why children turn out the way they do. New York: Free Press. Hawks, J., E. Wang, G. Cochran, H. Harpending, and R. Moyzis. 2007. “Recent acceleration of human adaptive evolution.” Proceedings of the National Acad- emy of Science 104:20753–20758. Haynes, E., H. Kalkwarf, R. Hornung, R. Wenstrup, K. Dietrich, & P. Lanphear (2003). Vitamin receptor Fok1 polymorphism and blood lead concentration in children. Environmental Health Perspectives, 111:1665-1669. References 151

Heck, K., P. Braveman, C. Cubbin, G. Chavez, and J. Kelly. 2006. “Socioecomonic status and breastfeeding initiation among California mothers.” Public Health Reports 121:51–59. Henrich, J. and F. Gil-White. 2001. “The evolution of prestige: Freely conferred deference as a mechanism enhancing the benefi ts of cultural transmission.” Evolution and Human Behavior 22:165–196. Herrnstein, R. and C. Murray. 1994. The Bell Curve: Intelligence and class struc- ture in American society. New York: Free Press. Hiller, J. 2004. “Speculations on the links between feelings, emotions and sexual behaviour: Are vasopressin and oxytocin involved?” Sexual and Relationship Therapy 19:1468–1479. Hindelang, M., T. Hirschi, and J. Weis. 1981. Measuring delinquency. Beverly Hills, CA: Sage. Hirschi, T. 1969. The causes of delinquency. Berkeley: University of California Press. Hirschi, T. and M. Gottfredson. 1987. Causes of white-collar crime. Criminology 25:949–974. Hirschi, T., M. Hindelang, and J. Weis. 1980. “The status of self-report measures.” Pp. 473–488 in Handbook of criminal justice evaluation, edited by M. Klien and K. Teilman. Beverly Hills: Sage. Holzer, H. and P. Offner. 2004. “The puzzle of black male unemployment.” The Public Interest 154:74–84. Hosking, G. 1985. The fi rst socialist society: A history of the Soviet Union from within. Cambridge: Harvard University Press. Hrdy, S. 1999. Mother Nature: A history of mothers, infants, and natural selec- tion. New York: Pantheon. Hur, Y. 2003. “Assortative mating for personality traits, educational level, reli- gious affi liation, height, weight, and body mass index in parents of a Korean twin sample.” Twin Research 6:467–470. Hurst, C. 1995. Social inequality: Forms, causes, and consequences. Boston: Allyn and Bacon. Iceland, J. 2003. Dynamics of economic well-being: Poverty 1996–1999. U.S. Cen- sus Bureau, July. http://www.census.gov/prod/2003pubs/p70-91.pdf. Itzkoff, S. 1987. Why humans vary in intelligence. Ashfi eld, MA: Paideia. Jacobson, J. and S. Jacobson. 2002. “Effects of prenatal alcohol exposure on child development.” Alcohol Research & Health 26:282–286. Jakobwitz, S. and V. Egan. 2006. “The dark triad and normal personality traits.” Personality and Individual Differences 40:331–339. Jafee, S., T. Moffi tt, A. Caspi, and A. Taylor. 2003. “Life with (or without) father: The benefi ts of living with two biological parents depend on the father’s antiso- cial behavior.” Child Development 74:109–126. Jang, K., R. McCrae, A. Angleitner, R. Riemann, and W. Livesley. 1998. “Heri- tability of facet-level traits in a cross-cultural twin sample: support for a hier- archical model of personality.” Journal of Personality and Social Psychology 74:1556–1565. Jensen, A. 1998. The g factor. Westport, CT: Praeger. Jost, L. and J. Jost. 2007. Why Marx left philosophy for social science. Theory and Psychology, 17:297–322. Jung, R. and R. Haier. 2007. The parieto-frontal integration theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sci- ences, 30:135–187. Kagan, J. 2007. “A trio of concerns.” Perspectives on Psychological Science 2:361–376. Kanazawa, S. 2003. “A general evolutionary psychological theory of criminal- ity and related male-typical behavior.” Pp. 37–60 in Biosocial criminology: 152 References

Challenging environmentalism’s supremacy, edited by A. Walsh and L. Ellis. Hauppauge, NY: Nova Science. Kaufman, A. 1976. “Verbal-performance IQ discrepancies on the WISC-R.” Jour- nal of Counseling and Clinical Psychology 44:739–744. Kelly, R. 2005. “The evolution of lethal intergroup violence.” Proceedings of the National Academy of Sciences 102:15294–15298. Kemper, T. 1994. “Social stratifi cation, testosterone, and male sexuality.” Pp. 47–61 in Social stratifi cation and socioeconomic inequality, vol. 2: Repro- ductive and interpersonal aspects of dominance and status, edited by L. Ellis. Westport, CT: Greenwood. Kendler, K., K. Jacobson, C. Gardner, N. Gillespie, S. Aggen, and C. Prescott. 2007. “Creating a social world: A developmental of peer group devi- ance.” Archives of General Psychiatry 64:958–965. Kennair, L. 2002. “Evolutionary psychology: An emerging integrative perspective within the science and practice of psychology.” The Human Nature Review 2:17–61. Kim-Cohen, J., A. Caspi, A. Taylor, B. Williams, R. Newcombe, I. Craig, and T. Moffi tt. 2006. “MAOA, maltreatment, and gene-environment interaction pre- dicting children’s mental health: New evidence and a meta-analysis.” Molecular Psychiatry 11:903–913. Kim-Cohen, J., T. Moffi tt, A. Caspi, and A. Taylor. 2004. “Genetic and environ- mental processes in young children’s resilience and vulnerability to socioeco- nomic deprivation.” Child Development 75:651–668. Kim, J., M. Fendrich, and J. Wislar. 2000. “The validity of juvenile arrestees’ drug use reporting: A gender comparison.” Journal of Research in Crime and Delin- quency, 37, 429–432. Kingston, P. 2006. “How meritocratic is the United States?” Research in Social Stratifi cation and Mobility 24:11–130. Knight, K. 2008. “Assortative mating for antisocial behavior: A comparison of mates and partners.” Paper presented at the annual meeting of the American Sociological Association Annual Meeting, Boston, MA. Kochanska, G. and N. Aksan. 2004. “Conscience in childhood: Past, present, and future.” Merrill-Palmer Quarterly 50:299–310. Koller, K., T. Brown, A. Spurfeon, and L. Levy. 2004. “Recent developments in low-level lead exposure and intellectual impairment in children.” Environmen- tal Health Perspectives 112:987–994. Kramer, D. 2005. “Commentary: Gene-environment interplay in the context of genetics, epigenetics, and gene expression.” Journal of the American Academy of Child and Adolescent Psychiatry 44:19–27. Kramer, M., F. Aboud, E. Mironova, I. Vanilovich, R. Platt, L. Matush, S. Igumnov, E. Fombonne, N. Bogdanovich, T. Ducruet, J. P. Collet, B. Chalm- ers, E. Hodnett, S. Davidovsky, O. Skugarevsky, O. Trofi movich, L. Kozlova, and S. Shapiro, for the Promotion of Breastfeeding Intervention Trial (PROBIT) Study Group. 2008. “Breastfeeding and child cognitive develop- ment: New evidence from a large randomized trial.” Archives of General Psychiatry 65:578–584. Krueger, R., T. Moffi tt, A. Caspi, A. Bleske, and P. Silva. 1998. “Assortative mat- ing for antisocial behavior: Developmental and methodological implications.” Behavior Genetics 28:173–185. Kruger, D. 2003. “Evolution and altruism: Combining psychological mediators with naturally selected tendencies.” Evolution and Human Behavior 24:118–125. Kuhn, T. 1970. The structure of scientifi c revolutions. Chicago: University of Chi- cago Press. References 153

Lacourse, E., D. Nagin, F. Vitaro, S. Côté, L. Arseneault, and R. Tremblay. 2006. “Prediction of early-onset deviant peer group affi liation: A 12-year longitudinal study.” Archives of General Psychiatry 63:562–568. Laderchi, C., R. Saith, and F. Stewart. 2003. “Does it matter that we do not agree on the defi nition of poverty? A comparison of four approaches.” Oxford Devel- opmental Studies 31: 243–274. Lancaster, J. and C. Lancaster. 1987. “The watershed: Changes in parental invest- ment and family formation strategies in the course of .” Pp. 187–205 in Parenting across the lifespan: Biosocial perspectives, edited by J. Lancaster, J. Altman, A. Rossi, and L. Sherrod. New York: Aldine De Gruyter. Lee, V. and P. Hoaken (2007). “Cognition, emotion, and neurobiological devel- opment: Mediating the relation between maltreatment and aggression.” Child Maltreatment 12: 281–-298. Lehmann, L. and L. Keller. 2006. “The evolution of cooperation and altruism—A general framework and classifi cation of models.” European Society for Evolu- tionary Biology 19:1365–1376. Lensvelt-Mulders, G. and J. Hettema. 2001. “Analysis of the genetic infl uences on the consistency and variability of the big fi ve across different stressful situa- tions.” European Journal of Personality 15:355–371. Levine, D. 2006. “Neural modeling of the dual motive theory of economics.” The Journal of Socio-Economics 35:613–625. Lilly, J., F. Cullen, & R. Ball. 2007. Criminological theory: Context and conse- quences. Thousand Oaks, CA: Sage. Loch, C., D. Galunic, and S. Schneider. 2006. “Balancing cooperation and com- petition in human groups: The role of emotional algorithms and evolution.” Managerial and Decision Economics 27:217–233. Locurto, C. 1990. “The malleability of IQ as judged from adoption studies.” Intel- ligence 14:275–292. Lodi-Smith, J. and B. Roberts. 2007. “Social investment and personality: A meta- analytic analysis of the relationship of personality traits to investment in work, family, religion, and volunteerism.” Personality and Social Psychology Review 11:68–86. Loehlin, J. 2000. “Group differences in intelligence.” Pp. 176–193 in Handbook of intelligence, edited by R. Sternberg. Cambridge: Cambridge University Press. Lopez-Rangel, E. and M. Lewis. 2006. “Loud and clear evidence for gene silencing by epigenetic mechanisms in autism spectrum and related neurodevelopmental disorders.” Clinical Genetics 69:21–25. Lopreato, J. and T. Crippen. 1999. Crisis in sociology: The need for Darwin. New Brunswick, NJ: Transaction. Lubinski, D. 2004. “Introduction to the special section on cognitive abilities: 100 years after Spearman’s (1904) “’General intelligence,’ objectively determined and measured.” Journal of Personality and Social Psychology 86:96–111. Luo, S. and E. Klohnen. 2005. “Assortative mating and marital quality in newly- weds: A couple-centered approach.” Journal of Personality and Social Psychol- ogy 88:304–326. Lynn, R. 2009. “What has caused the Flynn effect? Secular increases in the devel- opment quotients of infants.” Intelligence 37:16–24. Maccoby, E. 2000. “Parenting and its effects on children: On reading and misread- ing behavior genetics.” Annual Review of Psychology 51:1–27. Manning, M. and H. Hoyme. 2007. “Fetal alcohol syndrome disorders: A practi- cal clinical approach to diagnosis.” Neuroscience and Biobehavioral Review 31:230–238. Marx, K. 1967. Capital, vol. 1. New York: International Publishers. 154 References

. 1978. “Economic and philosophical manuscripts of 1844.” Pp. 66–123 in The Marx-Engels reader, edited by R. Tucker. New York: W. W. Norton. Marx, K. and F. Engels. 1948. The communist manifesto. New York: International. Massey, D. S. 2002. “A brief history of human society: The origin and role of emo- tions in social life.” American Sociological Review 67:1–29. . 2004. “Segregation and stratifi cation: A biosocial perspective.” Du Bois Review 1:7–25. Matarazzo, J. 1976. Wechsler’s measurement and appraisal of adult intelligence. Baltimore: Williams and Wilkins. Matthews, K., J. Flory, M. Mulldoon, and S. Manuck. 2000. “Does socioeconomic status relate to serotonergic responsivity in humans?” Psychosomatic Medicine 62:231– 237. May, P. and P. Gossage. 2008. Estimating the prevalence of fetal alcohol syndrome: A summary. National Institute of Alcohol Abuse and Alcoholism. National Insti- tute of Health. http://pubs.niaaa.nih.gov/publications/arh25-3/159-167.hm. Maughan, B. 2005. Developmental trajectory modeling: A view from developmen- tal psychopathology. The Annals of the American Academy of Political and Social Science, 602:118–130. Maughan, B., R. Rowe, J. Messer, R. Goodman, and H. Meltzer. 2004. “Conduct disorder and oppositional defi ant disorder in a national sample: Developmental epidemiology.” Journal of Child Psychology and Psychiatry 43:609–621. Mawby, R. 2001. Burglary. Colompton, Devon: Willan Publishing. Maxfi eld, D., B. Weiler, and C. Widom. 2000. “Comparing self-report and offi cial records of arrests.” Journal of Quantitative Criminology 16:87–110. Mayes, L., M. Bornstein, K. Chawarska, O. Haynes, and R. Granger. 1995. “Infor- mational processing and developmental assessment in 3-month-old infants exposed prenatally to cocaine.” Pediatrics 95:539–545. Mazur, A. and A. Booth. 1998. “Testosterone and dominance in men.” Behavioral and Brain Sciences 21:353–397. McBurnett, K., B. Lahey, P. Rathouz, and R. Loeber. 2000. “Low salivary cortisol and persistent aggression in boys referred for disruptive behavior.” Archives of General Psychiatry 57:38–43. McWhorter, J. 2000. Losing the race: Self-sabotage in black America. New York: Free Press. Meaney, F. & L. Miller. 2003. A comparison of fetal alcohol syndrome surveillance network and birth defects surveillance methodology in determining prevalence rates of fetal alcohol syndrome. Birth Defects Research, 67:819-821. Mekel-Bobrov, N., S. Gilbert, P. Evans, E. Vallender, J. Anderson, R. Hudson, S. Tishkoff, and B. Lahn. 2005. “Ongoing adaptive evolution of ASPM, a brain size determinant in Homo sapiens.” Science 309:1720–1722. Menard, S. and S. Mihalic. 2001. “The tripartite conceptual framework in adoles- cence in adolescence and adulthood: Evidence from a national sample.” Journal of Drug Issues 31:905–940. Merton, R. 1938. “Social structure and anomie.” American Sociological Review 3:672–682. . 1968. Social theory and social structure. New York: The Free Press. Messner, S. and R. Rosenfeld. 2001. Crime and the American dream, 3rd ed. Bel- mont, CA: Wadsworth. Meyer-Lindenberg, A., J. Buckholtz, B. Kolachana, A. Hariri, L. Pezawas, G. Blasi, A. Wabnitz, R. Honea, B. Verchinski, J. Gallicott, M. Egan, V. Mattay, and D. Weinberger. 2006. “Neural mechanisms of genetic risk for impulsivity in violence in humans.” Proceedings of the National Academy of Sciences, 103:6269-6274. References 155

Miles, D. and G. Carey. 1997. “Genetic and environmental architecture of human aggression.” Journal of Personality and Social Psychology, 72:207–217. Miller, L. 1987. “ of the aggressive psychopath: An integrative review.” Aggressive Behavior 13:119–140. Miller, J. and D. Lynham. 2001. “Structural models of personality and their rela- tion to antisocial behavior: A meta-analytic review.” Criminology 39:765–798. Miller, J., D. Lynham, and S. Jones. 2008. “Externalizing behavior through the lens of the fi ve-factor model: A focus on agreeableness and conscientiousness.” Journal of Personality Assessment, 90:158–164. Miller, W. 1958. “Lower-class culture as a generating milieu of gang delinquency.” Journal of Social Issue, 14:5–19. Mitchell, K. 2007. “The genetics of brain wiring: From molecule to mind.” PLoS Biology, 4:690–692. Moffi tt, T. 1993. “Adolescent-limited and life-course-persistent antisocial behav- ior: A developmental .” Psychological Review 100:674–701. . 2005. “The new look of behavioral genetics in developmental psychopa- thology: Gene-environment interplay in antisocial behavior.” Psychological Bulletin 131:533–554. Murray, C. 2002. “IQ and income inequality in a sample of sibling pairs from advan- taged family backgrounds.” The American Economic Review 92:339–343. Murray, J., M. Liotti, H. Mayberg, Y. Pu, F. Zamarripa, and Y. Liu. 2006. “Chil- dren’s brain activations while viewing televised violence revealed by fMRI.” Media Psychology 8:25–37. Needleman, H. 2004. “Lead poisoning.” Annual Review of Medicine 55:209–222. Nettle, D. 2003. “Intelligence and class mobility in the British population.” British Journal of Psychology 94:551–561. Niehoff, D. 2003. “A vicious circle: The neurobiological foundations of violent behavior.” Modern Psychoanalysis 28:235–245. Nielsen, F. 2006. “Achievement and ascription in educational attainment: Genetic and environmental infl uences on adolescent schooling.” Social Forces 85:193–216. Neisser, U., G. Boodoo, T. Bouchard, A. Boykin, N. Brody, S. Ceci, D. Halpern, J. Loehlin, R. Perloff, R. Sternberg, and S. Urbina. 1995. Intelligence: Knowns and unknowns. Report of a task force established by the board of scientifi c affairs of the American Psychological Association. Washington, DC: American Psychological Association. Neubauer, A. and A. Fink. 2009. Intelligence and neural effi ciency. Neuroscience and Biobehavioral Reviews, 33:1004–023. Noble, M., M. Mayer-Proschel, and R. Miller. 2005. The oligodendrocyte. Pp. 151–196 in Developmental neurobiology, edited by M. Rao and M. Jacobson. New York: Kluwer/Plenum. O’Brien, R. 2001. “Crime facts: Victim and offender data.” Pp. 59–83 in Criminol- ogy: A contemporary handbook, edited by J. Sheley. Belmont: CA, Wadsworth. O’Connor, T., K. Deater-Deckard, D. Fulker, M. Rutter, and R. Plomin. 1998. “Genotype-environment correlations in late childhood and early adolescence: Antisocial behavioral problems and coercive parenting.” Developmental Psy- chology 34:970–981. Ohnishi, T., R. Hashimoto, T. Mori, K. Nemoto, Y. Moriguchi, H. Iida, H. Noguchi, T. Nakabayashi, H. Hori, M. Ohmori, R. Tsukue, K. Anami, N. Hirabayashi, S. Harada, K. Arima, O. Saitoh, and H. Kunugi. 2006. “The association between the Val158Met polymorphism of the catechol-O-methyl transferase gene and morpho- logical abnormalities of the brain in chronic schizophrenia.” Brain 129: 399–410. 156 References

O’Leary, C. 2004. “Fetal alcohol syndrome: Diagnosis, epidemiology, and devel- opmental outcomes.” Journal of Paediatrics and Child Health 40:2–7. O’Leary, M., B. Loney, and L. Eckel. 2007. “Gender differences in the associa- tion between psychopathic personality traits and cortisol response to induced stress.” Psychoneuroendocrinology 32:183–191. Orians, G. 2008. “Nature and human nature.” Daedalus, 137:39–48. Osgood, D. and J. Chamber. 2003. “Community correlates of rural youth vio- lence.” Juvenile Justice Bulletin, May. U.S. Department of Justice. Osofsky, J. 1995. “The effects of exposure to violence on young children.” Ameri- can Psychologist 50:782–788. Palmer, C. & Tilley, C. 1995. Sexual access to females as a motivation for joining gangs: An evolutionary approach. The Journal of Sex Research, 32:213–217. Patterson, O. 1998. Rituals of blood: Consequences of slavery in two American centuries. Washington, DC: Civitas Counterpoint. Penn, A. 2001. “Early brain wiring: Activity-dependent processes.” Schizophrenia Bulletin 27:337–348. Perry, B. 2002. “Childhood experience and the expression of genetic potential: What childhood neglect tells us about nature and nurture.” Brain and Mind 3:79–100. Perry, B. and R. Pollard. 1998. “Homeostasis, stress, trauma, and adaptation: A neurodevelopmental view of childhood trauma.” Child and Adolescent Psychi- atric Clinics of America 7:33–51. Pigliucci, M., C. Murren, and C. Schlichting. 2006. “Phenotypic plastic- ity and evolution by genetic assimilation.” Journal of Experimental Biology 209:2362–2367. Pillworth, E. and M. Haselton. 2005. “The evolution of coupling.” Psychological Inquiry 16:98–104. Pinker, S. 2002. : The modern denial of human nature. New York: Viking. Pitchford, I. 2001. “The origins of violence: Is psychopathy and adaptation?” Human Nature Review 1:28–38. Plato. 1960. The Republic and other works. Garden City, NY: Doubleday. Plomin, R. 2005. “Finding genes in child psychology and psychiatry: When are we going to be there.” Journal of Child Psychology and Psychology 46:1030–1038. Plomin, R., K. Ashbury, and J. Dunn. 2001. “Why are children in the same fam- ily so different? Nonshared environment a decade later.” Canadian Journal of Psychiatry 46:225–233. Plomin, R. and D. Daniels. 1987. “Why are children from the same family so dif- ferent from one another?” Behavioral and Brain Sciences 10:1–60. Plomin, R. and DeFries. 1980. “Genetics and intelligence: Recent data.” Intelli- gence 4:15–24. Plomin, R., J. Defries, I. Craig, and P. McGuffi n. 2001. Behavioral genetics, 4th ed. New York: Worth Publishers. . 2003. “Behavioral genomics.” Pp. 531–540 in Behavioral genetics in the postgenomic era, edited by R. Plomin, J. Defries, I. Craig, and P. McGuffi n. Washington, DC: American Psychological Association. Plomin, R., McClearn, G., Smith, D., Vignetti, S., Chorney, M., Chorney, K., Ven- ditti, C., Kasarda, S. Thompson, L., Detterman, D., Daniels, J., Owen, M., & McGuffi n, P. 1994. DNA markers associated with high versus low IQ: The IQ quantitative trait loci (QTL) project. Behavior Genetics, 24:107-118. Pope, C. and Snyder, H. 2003. “Race as a factor in juvenile arrests. Juvenile justice bulletin” (NCJ 189180). Washington, DC: Offi ce of Juvenile Justice and Delin- quency Prevention. References 157

Porter, M. 1995. “The competitive advantage of the inner city.” Harvard Business Review 73:63–64. Posthuma, D., E. De Geus, W. Baar, H. Pol, R. Kahn & D.Boomsma. 2002. “The association between brain volume and intelligence is of genetic origin”. Nature neuroscience, 5:83–84. Prayer, D., G. Kasprian, E. Krampl, B. Ulm, L. Witzani, L. Prayer, and P. Brugger. 2006. “MRI of normal fetal brain development.” European Journal of Radiol- ogy 57:199–216. Propper, C. and G. Moore. 2006. “The infl uence of parenting on infant emo- tionality: A multilevel psychobiological perspective.” Developmental Review 26:427–460. Quartz, S. and T. Sejnowski. 1997. “The neural basis of cognitive development: A constructivist manifesto.” Behavioral and Brain Sciences 20:537–596. Quinsey, V. 2002. “Evolutionary theory and criminal behavior.” Legal and Crimi- nological Psychology 7:1–14. Raine, A. 1993. The psychopathology of crime: Criminal behavior as a clinical disorder. San Diego: Academic Press. . 2002. “Biosocial studies of antisocial and violent behavior in children and adults: A review.” Journal of Abnormal Child Psychology 30: 311–326. Raine, A., T. Lencz, S. Bihrle, L. Lacasse, and P. Colletti. 2000. “Reduced prefron- tal gray matter volume and reduced autonomic activity in antisocial personality disorder.” Archives of General Psychiatry 57: 19–127. Raine, A., J. Meloy, S. Bihrle, J. Stoddard, L. LaCasse, and M. Buchsbaum. 1998. “Reduced prefrontal and increased subcortical brain functioning assessed using positron emission tomography in predatory and affective murderers.” Behav- ioral Sciences and the Law 16:319–332. Rank, M., H. Yoon, and T. Hirschl. 2003. “American poverty as a structural failing: Evidence and arguments.” Journal of Sociology and Social Welfare 30:3–29. Rengert, G. and J. Wasilchick. 2001. Suburban burglary: A tale of two suburbs. Springfi eld, IL: Charles C Thomas. Rennie, Y. 1978. The search for criminal man. Lexington, MA: Lexington Books. Replogle, R. 1990. “Justice as superstructure: How vulgar is vulgar materialism?” Polity 22:675–699. Restak, R. 2001. The secret life of the brain. New York: co-published by Dana Press and Joseph Henry Press. Reuter, M., A. Schmitz, P. Corr, and J. Hennig. 2006. “Molecular genetics sup- port Gray’s personality theory: The interaction of COMT and DRD2 polymor- phisms predicts the behavioural approach system.” The International Journal of Neuropsychopharmacology 9:155–166. Rhee, S. and I. Waldman. 2002. “Genetic and environmental infl uences on anti- social behavior A meta-analysis of twin and adoption studies.” Psychological Bulletin 128: 490–529. Richerson, P. and R. Boyd. 2004. Not by genes alone: How culture transformed human evolution. Chicago: University of Chicago Press. Ridley, M. 2003. Nature via nurture: Genes, experience and what makes us human. New York: Harper Collins. Riggins-Caspers, K. R., J. Cadoret, J. Knutson, and D. Langbehn. 2003. “Biol- ogy– environment interaction and evocative biology–environment correlation: Contributions of harsh discipline and parental psychopathology to problem adolescent behaviors.” Behavior Genetics 33:205–220. Robinson, M. 2004. Why crime? An integrated systems theory of antisocial behav- ior. Upper Saddle River, NJ: Prentice Hall. 158 References

Rodkin, P., T. Farmer, R. Pearl, and R. Van Acker. 2000. “Heterogeneity of popu- lar boys: antisocial and prosocial confi gurations.” Developmental Psychology 36:14–24. Rosenfeld, R. and B. Nicodemus. 2003. “The transition from adolescence to adult life: Physiology of the transition phase and its evolutionary basis.” Hormone Research 60:74–77. Rossi, P. 1990. “Theory and methods: Moving forward.” Contemporary Sociology 19:623–624. Rothbart, M. 2007. “Temperament, development, and personality.” Current Direc- tions in Psychological Science 16:207–212. Rothbart, M., A. Ahadi, and D. Evans. 2000. “Temperament and personal- ity: Origins and outcomes.” Journal of Personality and Social Psychology 78:122–135. Rowe, D. 1994. The limits of family infl uence: Genes, experience, and behavior. New York: Guilford Press. . 1996. “An adaptive strategy theory of crime and delinquency.” Pp. 268– 314 in Delinquency and crime: Current theories, edited by J. Hawkins. Cam- bridge: Cambridge University Press. Rowe, D., D. Almeida, and K. Jacobson. 1999. “School context and genetic infl u- ences on aggression in adolescence.” Psychological Science 10:277–280. Rowe D. and D. Farrington. 1997. “The familial transmission of criminal convic- tions.” Criminology 35:177–201. Rowe, D., K. Jacobson, and E. Van den Oord. 1999. “Genetic and environmen- tal infl uences on vocabulary IQ: Parents education level as moderator.” Child Development 70:1151–1162. Ruden, R. 1997. The craving brain: The biobalance approach to controlling addic- tions. New York: Harper Collins. Runciman, C. 2001. Was Max Weber a selectionist in spite of himself? Journal of Classical Sociology 1:15–32. Rushton J. and A. Jensen. 2005. Thirty years of research on race differences in cognitive ability. Public Policy and Law 11:235–294. Rutter, M. 2007. “Gene-environment interdependence.” Developmental Science 10:12–18. Sampson, R. 2000. “Whither the sociological study of crime.” Annual Review of Sociology, 26:711–714. Sampson, R. and J. Laub. 2005. “A life-course view of the development of crime.” American Academy of Political & Social Sciences,602:12–45. Sanderson, S. 2001. The evolution of human sociality: A Darwinian confl ict per- spective. Lanham, MD: Rowman & Littlefi eld. Sanson, A., S. Hemphill, and D. Smart. 2004. “Connections between temperament and social development: A review.” Social Development 13:142–170. Sarfraz, H. 1997. “Alienation: A theoretical overview.” Pakistan Journal of Psy- chological Research 12:45–60. Saudino, K. 2005. “Behavioral genetics and child temperament.” Journal of Devel- opmental and Behavioral Pediatrics 26:214–223. Saulsman, L. & A. Page. 2004. The fi ve-factor model and personality disorder empirical literature: A meta-analytic review. Clinical Psychology Review, 23:1055–1085. Saunders, P. 2002. “Refl ections on the meritocratic debate in Britain in response to Richard Breen and John Goldthorpe.” British Journal of Sociology 53:559–574. Sawhill, I. and J. Morton. 2007. Economic mobility: Is the American dream alive and well? Washington, DC: The Economic Mobility Project/Pew Charity Trusts. References 159

Sayers, S. 2005. “Why work? Marx and human nature.” Science & Society 69:606–616. Scarpa, A. and A. Raine. 2003. “The psychophysiology of antisocial behavior: Interactions with environmental experiences.” Pp. 209–226 in Biosocial crimi- nology: Challenging environmentalism’s supremacy, edited by A. Walsh and L. Ellis. Hauppauge, NY: Nova Science. Scarr, S. 1996. “How people make their own environments: Implications for par- ents and policy makers.” Psychology, Public Policy, and Law 2:204–228. Schmalleger , F. 2004. Criminology Today. Upper Saddle River, NJ: Prentice Hall. Schmidt, F. and K. Hunter. 2004. “General mental ability in the world of work: Occupational attainment snd job performance.” Journal of Personality and Social Psychology 86:162–173. Schon, R. and M. Silven. 2007. “Natural parenting—back to basics in infant care.” Evolutionary Psychology 5:102–183. Schwalbe, M., S. Goodwin, D. Holden, D. Schrock, S. Thompson, and M. Wolko- mer. 2000. “Generic processes in the reproduction of inequality: An interac- tionist analysis.” Social Forces 79:419–452. Seligman, D. 1992. A question of intelligence: The IQ debate in America. New York: Birch Lane Press. Shaw, C. and H. McKay. 1972. and urban areas. Chicago: University of Chicago Press. Shea, C. 2009. “The nature-nurture debate, redux: Genetic research fi nally makes its way into the thinking of sociologists.” Chronicle of Higher Education: Chronicle Review, 9: B6. http://chronicle.com/free/v55/i18/18b00601.htm. Shi, S., T. Cheng, L. Jan, and Y. Jan. 2004. “The immunoglobin family mem- ber dendrite arborization and synapse maturation 1 (Dasm1) controls excit- atory synapse maturation.” Proceedings of the National Academy of Sciences 101:13246–13351. Sisk, C. & J. Zehr. 2005. “Pubertal hormones organize the adolescent brain and behavior.” Frontiers in Neuroendocrinology 26:163–174. Smits, J. 2003. “Social closure among the higher educated: Trends in educational homogamy in 55 countries.” Social Science Research 32:251–277. Snyderman, M. and Rothman, S. 1988. The IQ controversy, the media and public policy. New Brunswick, NJ: Transaction. Sokol, R., V. Delaney-Black, & B. Nordstrom. 2003. Fetal alcohol spectrum disor- der. Journal of the American Medical Association, 290:2996–2999. Sowell, E., P. Thompson, and A. Toga. 2004. “Mapping changes in the human cortex throughout the span of life.” Neuroscientist 10:372–392. Sowell, T. 1987. A confl ict of visions: Ideological origins of political struggles. New York: William Morrow. Spear, L. 2000. “Neurobehavioral changes in adolescence.” Current Directions in Psychological Science 9:111–114. Stark, R. 1979. “Whose status counts? Comment on Tittle, Villemez, and Smith.” American Sociological Review 44:668–669. Steinberg, L. 2005. “Cognitive and affective development in adolescence.” Trends in Cognitive Sciences. 9:69–74. Sundet, J., D. Barlaug, and T. Torjussen. 2004. “The end of the Flynn effect? A study of secular trends in mean intelligence test scores of Norwegian conscripts during half a century.” Intelligence 32:349–362. Sutherland, E. 1940. “White collar criminality.” American Sociological Review 5:1–20. Sutherland, E. and D. Cressey. 1974. Criminology. Philadelphia: Lippincott. 160 References

Tang, T., Y. Chen, and T. Sutarso. 2008. “Bad apples in bad (business) barrels: The love of money, Machiavellianism, risk tolerance, and unethical behavior.” Management Decision 46:243–263. Teasdale, T. and D. Owen. 2007. “Secular declines in cognitive test scores: A rever- sal of the Flynn effect.” Intelligence 36:121–126. Thernstrom, S. and A. Thernstrom. 1997. America in black and white: One nation indivisible. New York: Simon and Schuster. Tibbetts, S. 2003. “Selfi shness, social control, and emotions: An integrated per- spective on criminality.” Pp. 83–101 in Biosocial criminology: Challenging environmentalism’s supremacy, edited by A. Walsh and L. Ellis. Hauppauge, NY: Nova Science. Tittle, C. 1983. “Social class and criminal behavior: A critique of the theoretical foundation.” Social Forces 62:334–358. Tittle, C., W. Villemez, and D. Smith. 1978. “The myth of social class and crimi- nality: Evidence of the relationship between social class and criminal behavior.” American Sociological Review 49:398–411. Trivers, R. 1971. “The evolution of reciprocal altruism.” Quarterly Review of Biology 46:35–57. Turkheimer, E. and I. Gottesman. 1991. “Is H2 = 0 a null hypothesis anymore?” Behavioral and Brain Sciences 14:410–411. Turkheimer, E., A. Haley, M. Waldron, B. D’Onofrio, and I. Gottesman. 2003. “Socioeconomic status modifi es heritability of IQ in young children.” Psycho- logical Science 14:623–628. Turkheimer, E. and M. Waldron. 2000. “Nonshared environment: A theoretical, methodological, and quantitative review.” Psychological Bulletin 126:78–108. Turner, H., D. Finkelhor, and R. Ormrod. 2006. “The effects of lifetime victim- ization on the mental health of children and adolescents.” Social Science and Medicine 62:13–27. Udry, J. R. 2003. The National Longitudinal Study of Adolescent Health (Add Health). Chapel Hill, NC: Carolina Population Center, University of North Carolina at Chapel Hill. Uggen, C. 2000. “Class, gender, and arrest: An intergenerational analysis of work- place power and control.” Criminology 38:835–862. U.S. Census Bureau. 2002. The Population of the United States. Washington, DC: U.S. Census Bureau. . 2007. Current Population Reports. Washington, DC: U.S. Census Bureau. Useem, Michael. 1989. Liberal Education and the Corporation. Hawthorn, NY: Aldine de Gruyter. van den Berghe, P. 1990. “Why most sociologists don’t (and won’t) think evolu- tionarily.” Sociological Forum 5:173–185. van Goozen, S., G. Fairchild, H. Snoek, and G. Harold. 2007. “The evidence for a neurobiological model of childhood antisocial behavior.” Psychological Bul- letin 133:149–182. van Leeuwen, M. and I. Maas. 2002. “Partner choice and homogamy in the nine- teenth century: Was there a sexual revolution in Europe?” Journal of Social History 36:101–123. van Voorhees, E. and A. Scarpa. 2004. “The effects of child maltreatment on the hypothalamic- pituitary-adrenal axis.” Trauma, Violence, and Abuse 5:333–352. Ventura, S. 2009. Changing patterns of nonmarital childbearing in the United States. NCHS data brief no. 18. Hyattsville, MD: National Center for Health Statistics. Vernon, P., J. Wickett, P. Bazana, and R. Stelmack. 2000. “The neuropsychology and psychophysiology of intelligence.” Pp. 245–264 in Handbook of Intelli- gence, edited by R. Sternberg. Cambridge: Cambridge University Press. References 161

Vold, G., T. Bernard, and J. Snipes. 1998. Theoretical criminology. New York: Oxford University Press. Wakschlag, L., K. Pickett, E. Cook, N. Benowitz, and B Leventhal. 2002. “Mater- nal smoking during pregnancy and severe antisocial behavior in offspring: A review.” American Journal of Public Health 92:966–974. Walinsky, A. 1997. “The crisis of public order.” Pp. 8–15 in Criminology 97/98, edited by M. Fisch. Guilford, CT: Dushkin. Walker, E. 2002. “Adolescent neurodevelopment and psychopathology.” Current Directions in Psychological Science 11:24–28. Wallace, W. 1990. “Rationality, human nature, and society in Weber’s theory.” Theory and Society 19:199–223. Walsh, A. 1997. Methodological and vertical integration in the social sciences. Behavior and Philosophy, 25:121-136. . 2002. Biosocial criminology: Introduction and integration. Cincinnati, OH: Anderson. . 2003. “Intelligence and antisocial behavior.” Pp. 105–124 in Biosocial criminology: Challenging environmentalism’s supremacy., edited by A. Walsh and L. Ellis. Huntington, NY: Nova Science. . 2009. Biology and criminology: The biosocial synthesis. New York: Routledge. Walsh, A. and L. Ellis. 2004. “Ideology: Criminology’s Achilles’ heel?” Quarterly Journal of Ideology 27:1–25. Walsh, A. and H-H. Wu. 2008. “Differentiating antisocial personality disorder, psychopathy, and sociopathy: Evolutionary, genetic, neurological, and socio- logical considerations.” Criminal Justice Studies 21:135–152. Walters, G. and M. Geyer. 2004. “Criminal thinking and identity in male white- collar offenders.” Criminal Justice and Behavior 31:263–281. Wasserman, R., C. DiBlasio, L. Bond, P. Young, and R. Colletti. 1990. “Infant temperament and school age behavior: 6-year longitudinal study in pediatric practice.” Pediatric 85:801–807. Watters, E. 2006. “DNA is not destiny.” Discover: Science, Technology and the Future. November. Weber, Max. 1978. Economy and Society: An outline of interpretative sociol- ogy, vol. 2, edited by G. Roth & C. Wittich. Berkeley: University of California Press. Wechsler, D. 1958. The measurement and appraisal of adult intelligence. Balti- more: Williams and Wilkin. Weinberger, D., B. Elvevag, and J. Giedd. 2005. The adolescent brain: A work in progress. Washington, DC: The National Campaign to Prevent Teen Pregnancy. Weinhold, B. 2006. “Epigenetics: The science of change.” Environmental Health Perspective 114:161–167. Weisburd, D., S. Wheeler, E. Waring, and N. Bode. 1991. Crimes of the middle classes: White-collar offenders in the federal courts. New Haven, CT: Yale Uni- versity Press. Wentzel, K. and L. Looney. 2007. “Socialization in school settings.” Pp. 382–403 in Handbook of socialization: Theory and research, edited by J. Grusec and P. Hastings. New York: Guilford Press. Western, B. 2003. Incarceration, employment, and public policy. New Jersey Insti- tute for Social Justice. http://www.njisj.org/reports/western_report.html. Wexler, B. 2006. Brain and culture: Neurobiology, ideology, and social change. Cambridge, MA: MIT Press. White, A. 2004. Substance use and the adolescent brain: An overview with the focus on alcohol. Durham, NC: Duke University Medical Center. 162 References

Wiebe, R. 2004. “Delinquent behavior and the fi ve-factor model: Hiding in the adaptive landscape?” Individual Differences Research 2:38–62. Widom, C. and L. Brzustowicz. 2006. “MAOA and the ‘cycle of violence’: Child- hood abuse and neglect: MAOA genotype and the risk for violent and antisocial behavior.” Biological Psychiatry 60:684–689. Wilson, E. O. 1978. . Cambridge, MA: Harvard University Press. Wilson, M. and M. Daly. 1997. “Life expectancy, economic inequality, homicide and reproductive timing in Chicago neighborhoods.” British Medical Journal 314:1271–1274. Witt, L., L. Burke, M. Barrick, and M. Mount. 2002. “The interactive effects of conscientiousness and agreeableness on job performance.” Journal of Applied Psychology 87:164–169. Wismer Fries, A., T. Ziegler, J. Kurian , S. Jacoris, and S. Pollak. 2005. “Early experience in humans is associated with changes in neuropeptides critical for regulating social behavior.” Proceedings of the National Academy of Sciences 102:17237–17240 Wright, J. & P. Boisvert. 2009. What biosocial criminology offers criminology. Criminal Justice and Behavior, 36:1228–1239. Wright, J., K. Dietrich, M. Ris, R. Hornung, S. Wessel, and B. Lanphear. 2008. “Association of prenatal and childhood blood lead concentrations with criminal arrests in early childhood.” PLoS Medicine 5:732–740. Wright, R. and S. Decker. 1994. Burglars on the job: Streetlife and residential break-ins. Boston: Northeastern University Press. . 1997. Armed robbers in action. Boston: Northeastern University Press. Wright, R. A. and J. Miller. 1998. “Taboo until today? The coverage of biological arguments in criminology textbooks, 1961 to 1970 and 1987 to 1996.” Journal of Criminal Justice 26:1–19. Yacubian, J., T. Sommer, K. Schroeder, J. Glascher, R. Kalisch, B. Leuenberger, D., F. Braus, and C. Buchel. 2007. “Gene-gene interaction associated with neural reward sensitivity.” Proceedings of the National Academy of Sciences 104:8125–8130. Yang, Y., A. Raine, T. Lencz, S. Bihrle, l. LaCasse, and P. Colletti. 2005. “Volume reduction in prefrontal gray matter in unsuccessful criminal psychopaths.” Bio- logical Psychiatry 57:1103–1108. Yoav, G., O. Man, S. Pääbo, and D. Lancet. 2003. “Human specifi c loss of olfac- tory receptor genes.” Proceedings of the National Academy of Sciences 100: 3324– 3327. Yu, O. and L. Zhang. 2006. “Does acceptance of corporate wrongdoing begin on the “training ground” of professional managers?” Journal of Criminal Justice 34:185–194. Zechel, J., J. Gamboa, A. Peterson, M. Puchowicz, W. Selman, and D. Lust. 2005. “Neuronal migration is transiently delayed by prenatal exposure to intermittent hypoxia.” Birth Defects Research 74:287–299. Index

A 85, 96–97, 99; and personality AbilityÆstatus relationship 110–111 138; and racial differences 54; abuse: child 18, 22, 37, 79, 96, 102; and SES 16, 48, 81, 86, 137; and maternal substance 49, 98; sub- socialization 89; and (T) 85 stance 138 antisocial personality disorder (ASPD) active rGE 17–19, 82, 123 49, 138 activity level 135 anxiety 46–47, 75, 79, 97, 107–108 adaptations 3, 26, 30, 36, 40, 85, 98 aristocracy 104 addiction 44, 50, 91 association analysis 21 adenine (A) 21 assortative mating 18–19, 103, Adkins, D. 109–111, 115 114–115 Administration Improvement Act (AIA) attention defi cit with hyperactivity 58 disorder (ADHD) 21–22 adrenal gland 97 autonomic nervous system (ANS) 44, affect 135 96, 108, 135 afferent nerves 13 axon 39, 41 aggression 11, 21, 34, 38, 43, 97, 109, 138 B Agnew, R. 68–69, 135 Badcock, C. 25 agrarian societies 110 Barkow, J. 27, 46 agreeableness 111, 137–138 Barnett, R. 59, 127 alcohol 44, 66, 95, 98 base pairs 21 alienation 73, 83, 106 Bates, J. 79 alleles 22, 27, 45, 112–113, 120 bayley scales 124 allostasis 97–98 Beaver, K. 22, 33, 124 allostatic load 98 behavior genetics 14, 19, 26; and crimi- altruism 4, 21, 27–30, 33, 73 nal behavior 17 American Psychological Association behavioral activating system (BAS) 44 117 behavioral analysis 26 amygdala 44–45, 50, 85, 96, 98 behavioral inhibiting system (BIS) 44 Anderson, E. 34, 70 behavioral traits 24, 26–27, 38 anomie 61, 66 benefi t/cost ratio 64 anomie/strain theory 65–68, 83, 137, Bentham, J. 74 139–141 Bernanke, B. 129 anterior cingulate cortex 100 Bernard, T. 139 antisocial behavior 5, 53; and assor- biosocial criminology 2, 5, 9 tative mating 18–19, 33; and biparental care 36 friends 84; and IQ 126; and Black>White> Asian pattern 56–67, neurology 15–18, 21–23, 37, 50, 92–93, 99–100 164 Index blood- brain barrier 97 connectionists 40 Blumstein, A. 37 conscience 39, 44–48, 63, 97, 135 Bond, R. 31, 36, 40, 78, 132 conscientiousness 83, 111, 137–139 bonding: neurological importance of constrained vision 8–9, 29, 91, 129, 101; social 83, 89; transient 33 141–142 Bonger, W. 73 constructivism 2, 4, 40 bourgeoisie 73, 104 cooperation 11, 27, 29, 31, 46, 105 bourgeoisie family 37 corporate crime 59, 63 Bowles, S. 83 correlation 16, 19, 54–56, 77, 112, brain: anatomy 11; craving 44, 50; grey 114, 117–119, 130–132, matter volume 49, 100, 119; 138–139 social 27; volume 118–119 cortisol 97–98, 102, 107–108 Braithwaite, J. 53–56 craving brain 44, 50 breastfeeding: importance of 80, 100, Cressey, R. 73 102, 121; IQ relationship 101 crime: normality 30, 61 Buss, D. 4, 30, 34 criminal families 19 Butcher, L. 120 criminal threshold 62–63 criminality 17–19, 28, 30, 32, 37; and C IQ 126; and SES 52, 55, 73; as calcium 100 misbehavior 64; parental 86 Capron, C. 121 criminogenic environment 47, 50, 62–63 Carey, G. 19 Crippen, T. 109 Caspi, A. 22, 55, 81, 101, 135–136 crystallized intelligence 117 central nervous system (CNS) 13 Cullen, F. 12, central tendency 3, 26 culture 4, 33, 37–38, 82; agrarian 112; cerebral glucose metabolism 119 and evolutionary psychology 27; Chamorro- Premuzic, T. 135 and SES 65–67, 70; and social- Carrasco, X. 22 ization 77 cheating: detection 31; theory 31, 35, cytosine (C) 21, 23, 101 46 Chicago School of Ecology 70 D children: abuse/neglect 18, 22, 37, 51, Daly, M. 35–36, 56 79, 86, 96, 102; desertion of Darwin, C.: The Decent of Man 11 37, 92; developmental factors de waal, F. 29 96–102, 112, 117, 120–121, Decker, S. 94–95 124, 130–136; exposure to vio- delinquency: patterns of 33, 52–55, 64, lence 43–43, 50 68, 71, 79, 82, 125–126 cholinergic muscarinic 2 receptor gene dendrites 39, 41 120 Dennett, D. 6 chromosomes 21, 112 determinism 10; genetic 25 Cincinnati Lead Survey 100 development quotients (DQs) 124 class: Weber on 74, 104–106 developmentalism 6 class structuralist position 128, 132 Diagnostic and Statistical Manual of class-crime relationship 52–53, 55–56, Mental Disorders (DSM-IV) 138 64–65, 67–69, 71–72, 74–77, Dickens, W. 123–125 90, 140 differential association theory (DAT) 72 classic conditioning 46 differential social organization 72 cognitive traits 19–20 DiRago, A. 132 Cohen, A. 67–70, 82 divorce 18, 36–37, 91, 133–134, 136 collective effi cacy 73 DNA 13, 21, 142; markers 120; methy- conditional cooperation 31 lation 23–25 conduct disorder (CD) 23, 46, 86 Dobzhanksy, T. 27 conformity 46–47, 66–67, 82, 137 Dodge, K. 80 Conger, R. 78–81 Dogan, S. 78–81 Index 165 dopamine (DA) 44–45, 50, 100 ization 78–83; as a life domain dopamine deaminating enzyme 68–69 (COMT) 44, 50, 120 family stress model (FSM) 78 dopamine receptor gene: DRD2 23; fatty acid desaturase gene (FADS2) 101, DRD4 21–23, 45 121 dopamine transporter gene (DAT1) 22, fear 46–47, 49–50, 97; male and female 33 differences 38, 107–108 Dumaret, A. 121 Fergusson, D. 86–87 Durkeim, E. 30, 56, 65–66, 83 fetal alcohol syndrome (FAS) 98 Duyme, M. 121 fetal intermittent hypoxia 99 fi ght/fl ight/freeze system (FFS) 44–46 E fi ve-factor model 137 Eckel, L. 97 Flynn effect 122–124, 127 ecological fallacy 55–56, 64 Flynn, J. 122–125, 127 ecological studies 55–56, 64 focal concerns theory 94 Edison, T. 122 fok1 100 Edleman, G. 41 full scale IQ (FIQ) 126 efferent nerves 13 functional magnetic resonance imaging Eagan, V. 138 (fMRI) 43, 49–50, 100 ego 44–45 Furnham, A. 135 egoism 63, 73, 105; crabbed 30 Ellis, L. 53, 55, 129 G emotion 4, 39, 41, 44; arousal 48–50; g load 117–119 and cognition 29; social 46–47 Gage, P. 49 empathy: and altruism 28–29 ;and Galunic, D. 107 criminality 17–18; male/female gangs 69, 74, 84 differences 33, 38, 46, 78, 137 Garlick, D. 124–125 endophenotypes 17–18, 36, 111 gene-culture co-evolution 26–27, 36 Engels, F. 73, 105 gene/environment correlation (rGE) epigenesis 7 16–19, 22, 42, 63, 68–69, 78, epigenetics 23–25 80–82, 68–87, 122–124 epinephrine (adrenaline) 45, 96–97 gene/environment interaction (GxE) essentialism 88 16–18, 20,22, 33, 62, 78, 81, evocative rGE 17–18, 68–69, 81–82 85, 87 evolutionary psychology 26–27 general mental ability (GMA) 139 evolutionary theory: of aggression general strain theory (GST) 68 33–34, 43; of crime 30–31; of generalized other 48 family 35–36; of mating strategy genes 13, 112; and behavior 21–22; and 32, 38; of socialization 83, 103, children 77–78; and environ- 105, 107, 110, 13–115 ment 5, 13–19, 123–125, 127, excitatory neurotransmitters 85, 98 132–133; and IQ 112, 119–120– executive functions 20, 45, 48–49 121; and social status 109–111 exons 13 genetic drift 113 experience-dependent process 40 genetics 11–12, 14–16, 113; and intel- experience-expected process 40–41, 50, ligence 120–121; behavior 19, 78, 101, 140 24; molecular 18, 20, 124 extended investment model (EIM) 78 genome 11, 14, 21, 23–37, 39, 50, 103, external deterrents 31 109–111, 115 extrafamilial non-shared factors 15 genome-> ability and ability-> status relationship 110–111 F genome->ability relationship 110–111, familial non-shared environment 15 124 family 35–38, 68; and SES 86–87, 90, genotype 16–18, 22, 50, 81, 108, 92–93, 96, 130–134; and social- 100–112 166 Index

Gil-White, F. 108 hypermasculinity 33 Gintis, H. 83 hypoarousable ANS 47 glial cells 39, 41 hypocortisolism 97, 102 glucose 119 hypothalamic-pituitary-adrenal (HPA) glutamate 98 axis 48, 96 Goldstein, D. 97–98 hypothalamus 96–98 Gottfredson, M. 59–60,116, 138 Grifi th’s test 124 I Grigorenko, E. 120 id 44–45 Grosvenor, P. 4 idealism 106 guanine (G) 21, 101 ideology 7–9, 63 guilt 31, 46–47, 75 illegitimacy 37,73, 93 Guo, G. 33,109–111, 115 impulsiveness 18, 43,76, 109; male/ female differences 30 H impulsivity 21, 35, 98 habituation 97 incidence rate ratios (IRR) 86–87 Hacker, J. 90, 93 inclusive fi tness theory 28 Hacking, I. 1 individual differences 20,23, 54, 80–81, Haier, R. 119, 127 90, 116, 128, 135 Hamilton, W. 28 industrialism 110 Handbook of Socialization: Theory and inhibitory neurotransmitters 85 Research 78 innovation 66–67 Hardy- Weinberg equilibrium (H-WE) intellectual imbalance 126 103, 111, 113, 115 intelligence 75–76, 116; and SES Harper, L. 24 88, 111,116–118, 129–130, Harris, A. 52, 56, 59 133, 135; genetic basis of 21, Harris, J. 30, 83 118–122, 125 Hawks, J. 27 internal deterrents 31 hedonism 67 IQ: and criminality 125–126; and heri- Henrich, J. 108 tability 21 heritability 14; computing 20–21, 109; IQ tests 101, 117, 120, 125, 129–130 in different environments15, 17, Itzkoff, S. 122 48, 81, 90, 118–120; of consci- entiousness 137; of divorce 36; J of IQ 123 Jacobwitz, S. 138 heritability coeffi cient 14–15, 36, 90, Jensen, A. 113, 132 123, 133 Jirtle, R. 25 hippocampus 44–45, 50, 96, 98 Jung, R. 119, 127 Hirshi,T. 59–60, 90, 138 Hirschl, T. 91 K histone acetylation 23 Kagan, J. 51 histones 23 Kemper, T. 103, 107 holism 5–6 Kim-Cohen, J. 81 homeostasis 45–48, 96–97, 105 kin selection theory 28 homicide 34–35, 43, 56, 72 Kingston, P. 132 honor subcultures 34 Kopin, I. 97–98 horticultural societies 110 Kruger, D. 19 Horwood, J. 86 Kuhn, T. 12 human nature 2–4, 8, 11, 26, 30, 35, 61, 74, 76, 107 L Hunter, K. 139 language 3, 40 Hurst, C. 131 law of independent assortment 112 hyperarousable ANS 47 law of segregation 112 hypercortisolism 97 lead exposure 99–100 Index 167

Lewis, J. 95 myth of social class and criminality 52, life-course persistent offender 118 55, 68, 70 limbic system 44, 46–47, 49, 85 linkage analysis 21 N Loch, C. 107 narcissism 138 Locurto, C. 121 National Academy of Sciences 117 Loney, B. 97 National Crime Victimization Survey Lopreato, J. 109 (NCVS) 53 love 5, 30, 35 National Immunization Program Survey low self-control 17, 138, 69 101 lumpenproletariat 73, 104 National Incident Based Reporting Lynn, R. 124 System (NIBRS) 57, 59, 64 National Institute of Alcohol Abuse and M Alcoholism 98 Machiavellianism 137–138 National Institute of Health Study 69 Madoff, B. 58 National Longitudinal Study of Adoles- magnetic resonance imaging (MRI) 49, cent Health Survey 55 118–119 National Longitudinal Survey of Youth marriage 6, 36, 69, 91, 94, 136 125 Marx, K.: Capital 3 National Youth Survey 55 Marxism 105 nativists 40 Massey, D. 2, 98 natural selection 11, 26–27, 29, 32, materialism 104, 106 35–36, 40, 84–85, 108, maternal smoking 99 naturalistic fallacy 76, 105, 108 mating effort 32–34, 38 negative assertive mating 114 Matthews, K. 109 negative class-crime relationship 53, Mawby, R. 94–95 64–65, 68, 71–72, 74–77 McKay, H. 70–72 negative emotionality 136 Mead, G. 4, 48 neighborhood effects Meaney, M. 24 Nettle, D. 133 meiosis 112 neural Darwinism 41 Mendelian genetics 113 neurons 5–6, 39, 41–42, 44, 98, 125 meritocratic position 128–129, 132, neurotransmitters 10, 13, 22, 108, 118, 134 125 Merton, R. 66–67, 137, 141 Newtonism 6 messenger RNA (mRNA) 13, 23 Nielsen, F. 133 Messner, S. 128–129, 141 nonshared environment 15 methylation 23 nuclear family 36 Meyer- Lindenberg, J. 50 nucleotides 6, 21 microsatellites 21, 120 nucleus accumbens 44, 85 Miller, W. 69–70, 94, 98, 126, 138 nurturing 24, 32, 88; and empathy 39 minisatellites 21 minority power groups 74 O molecular genetics 18, 20, 124 O’Leary, M. 97 monoamine oxidase A (MAOA) 22 occupational success: and conscientious- mood 100, 135 ness 138; and IQ 130; and per- moralistic fallacy 76, 142 sonality 139; and SES 133, 137 Morton, J. 128, 141 openness 137 multiplier effect 69, 123 operant conditioning 46, 64 Murray, C. 116, 133–134, 141 Osgood, D. 71 mutation 113 oxytocin 36, 78, 101 mutual sharing 35 myelin 41 P myelination 41, 85, 98 parasympathetic system 45, 47, 96 168 Index parietal lobes 119 Rank, M. 91 parieto-frontal integration theory rational choice theory 74 (P-FIT) 119 reaction time (RT) 118 party: Weber on 105 reactivity 46, 135 passive rGE 16, 19, 80, 82, 86 rebellion 66 Patterson, O. 9 reciprocal altruism 28 peer groups 15, 63; and socialization recognition reaction time 118 83–84 reductionism 5–6 performance IQ (PIQ) 126 regression to the mean 103, 111–112, peripheral nervous system (PNS) 44, 99 115 Perry, B. 39, 42 reinforcement sensitivity theory (RST) personality traits 14, 19, 68, 83, 114, 44 132, 137–138 Rengert, G. 95 Pettit, G. 80 reproductive success 11, 26, 28, 30, phenotype 14, 16–18, 36, 111–112, 114 32–34, 36, 39, 46, 50 phenotypic traits 14, 21, 111, 132 Restak, R. 85 pituitary gland 97 retreatism 66 Plato 4, 61, 63 reward dominance theory 44, 50, 85 Plomin, R. 26, 120 Ridley, M. 6, 24 police bias hypothesis 56–57 ritualism 66–67 polymorphisms 20–23, 27, 44, 50, 100, Robinson, M. 1, 39 120, 125 Rosenfeld, R. 128–129, 141 positive assortative mating 19 Rothman, S. 116–117 positive emotionality 136 Rowe, D. 32 positron emission tomography (PET) Rushton, J. 113 49, 119 postponement factor (P) 122 S poverty 51, 71, 79, 90–91; and race Sampson, R. 6–7, 43, 94, 92–93; and society 91; crime as Saunders, P. 131–132 cause of 73, 93–95, 102 Sawhill, I. 128, 141 pre-experiential brain organization 40 Scarr, S. 112 preformationism 7 schizophrenia 23 prefrontal cortex (PFC) 44, 48, 50, 85, Schmalleger, F. 94 120 Schmidt, F. 139 Pleistocene epoch 27–28 Schneider, S. 107 primary psychopathy 138 school 33, 54–55, 68; and socialization probability 19, 23, 32, 35–37, 42, 79, 82–89, 91–95; characteristics 56–57, 81, 83, 98, 101, 113, 130 133, 136, 139 proletariat 73, 90, 104 secondary psychopathy 138 promoter factors 13 selectionism 40, 106 prosocial behavior 28, 49, 127 selfi shness 29–30, 105–106 psychological altruism 28–29 self-report strategies 22, 49, 52–56, psychopathy 46, 49, 138 86–87 puberty 84–85, 94, 125–126 sensation-seeking 11, 17, 21 serotonin (5-HT) 44–45, 50, 81, 100, Q 107–109 qualitative differences 14, 17 serotonin transporter gene (5-HTTPR) quantitative differences 13, 17, 20- 21 45, 81 quantitative trait loci mapping 21, 120 shared environment 14, 17, 19, 132–133 Shaw, C. 70–72 R Shaw, J. 52, 56, 59 race: and antisocial behavior 54; and simple reaction time 118 crime 56–57, 59–60, 74; and single nucleotide polymorphism (SNPs) poverty 92–93, 109, 114 21 Index 169 sky hooks 6 T Smith, D. 52, 65 tactile stimulation 80, 101 Smits, J. 114 temperament 3, 46, 68, 75, 79, 81; and Snipes, J. 139 personality 135–136, 140 Snyderman, M. 116–117 teratogenic chemicals 98–99 sociability 135 testosterone (T) 84, 107 social brain 27 thymine (T) 21 social class 8, 10; and criminal behavior Tittle, C, 52–53, 55–57, 65, 67–69, 71, 52–53, 55, 57, 64, 73, 75–76, 73–77, 140 140; and genetics 111–115, 133; trait variance 20 and socialization 78, 88–89, transcription factors 13 103; Marx on 104–106 transient bonding 33 social control theories 82 turning points 6–7 social disorganization 71–72 social emotions 46–47 U social push hypothesis 48 unconstrained vision 8, 107, 129, social selection model (SSM) 78, 80 141–142 social stratifi cation 103 Uniform Crime Reports (UCR) 53, socialization 77–78, 80–81, 83, 88–89; 56–57, 59–60 and criminal behavior 44, United States Census Bureau studies 47–48; and peer groups 83–85; 90, 92 and schools 82; and SES 67; United States Department of Justice responsiveness to 135 studies 60 socioeconomic status (SES) 11–12, 15, uracil (U) 21 37, 52 ,80 utilitarian/deterrence tradition 74–75 sociology 1–2, 5, 10, 12, 103, 107, 109, 115, 129, 132 V Sowell, T. 8, 142 values 15, 19, 80; and SES 69–71, 74, Spearman’s g 117, 119 77, 82–83, 105–106, 138 spuriousness 88 variance 5, 14–15, 20–21, 48, 81, 88, status 11, 29–30; and crime 26, 58–61; 142; and environment 133; and and violence 33–36 IQ 121 status ascription position 128 ventral striatum area 49 status attainment position 128, 132 ventral tengmental area 44 status frustration 67, 69, 82 vicarious violence 43 status hierarchies 34, 107–108 Villemez, W. 52, 65 status offenses 54 violence: and brain development 42,43, status-striving 3, 38,103–104, 107 49–50, 108; and illegitimacy 70; stepparenting 37 and status 33–35, 37; and video stock-based compensation 63 games 43 strain 65–68, 86, 89, 135, 139 vitamin D receptor gene (VDR) 100 structural theories 94 Vold, G. 139 subcultural theory 67 Voltaire 142 subcultures of violence 34 sub-optimal arousal 21 W Sui generis 140 Walsh, A. 9 Sun-Tzu: Art of War 4 Wasilchick, J. 95 super traits theory 68 Weber, M.: on class 74, 104; on confl ict superego 44–45 74, 105–106 Sutherland, E. 58–59, 64, 72–73 Wechsler, D. 117, 126 Swain-Campbell, J. 86 Weisburd, D. 60–61 sympathetic system 45, 47, 96 white collar crime 57–61, 64 synapses 41, 122 Wilson, E. 105 synaptogenesis 41 Wilson, M. 35–36, 56 170 Index

Wright, R. 94–95 Yoon, H. 91 Yu, O. 63 Y Yacubian 49 Z Yanomamo 34 Zhang, L. 63