Hostname: page-component-7c8c6479df-27gpq Total loading time: 0 Render date: 2024-03-18T18:34:57.249Z Has data issue: false hasContentIssue false

The development of antisocial behavior: What can we learn from functional neuroimaging studies?

Published online by Cambridge University Press:  07 October 2008

S. L. Crowe
Affiliation:
National Institute of Mental Health
R. J. R. Blair*
Affiliation:
National Institute of Mental Health
*
Address correspondence and reprint requests to: James Blair, Mood & Anxiety Program, National Institute of Mental Health, National Institutes of Health, 15 k North Drive, Bethesda, MD 20892; E-mail: blairj@intra.nimh.nih.gov.

Abstract

The recent development of low-risk imaging technologies, such as functional magnetic resonance imaging (fMRI), have had a significant impact on the investigation of psychopathologies in children and adolescents. This review considers what we can infer from fMRI work regarding the development of conduct disorder (CD) and oppositional defiant disorder (ODD). We make two central assumptions that are grounded in the empirical literature. First, the diagnoses of CD and ODD identify individuals with heterogeneous pathologies; that is, different developmental pathologies can receive a CDD or ODD diagnosis. This is indicated by the comorbidities associated with CD/ODD, some of which appear to be mutually exclusive at the biological level (e.g., posttraumatic stress disorder [PTSD] and psychopathic tendencies). Second, two populations of antisocial individuals can be identified: those that show an increased risk for only reactive aggression and those that show an increased risk for both reactive and instrumental aggression. We review the fMRI data indicating that particular comorbidities of CD/ODD (i.e., mood and anxiety conditions such as childhood bipolar disorder and PTSD) are associated with either increased responsiveness of neural regions implicated in the basic response to threat (e.g., the amygdala) or decreased responsiveness in regions of frontal cortex (e.g., ventromedial frontal cortex) that are implicated in the regulation of the basic threat response. We suggest why such pathology would increase the risk for reactive aggression and, in turn, lead to the association with a CD/ODD diagnosis. We also review the literature on psychopathic tendencies, a condition where the individual is at significantly elevated risk for both reactive and instrumental aggression. We show that in individuals with psychopathic tendencies, the functioning of the amygdala in stimulus-reinforcement learning and of the ventromedial frontal cortex in the representation of reinforcement expectancies is impaired. We suggest why such pathology would increase the risk for reactive and instrumental aggression and thus also lead to the association with a CD/ODD diagnosis.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2008

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

This research was supported by the Intramural Research Program of the NIH NIMH.

References

Adler, C. M., DelBello, M. P., Jarvis, K., Levine, A., Adams, J., & Strakowski, S. M. (2007). Voxel-based study of structural changes in first-episode patients with bipolar disorder. Biological Psychiatry, 61, 776781.CrossRefGoogle ScholarPubMed
Adolphs, R. (2002). Neural systems for recognizing emotion. Current Opinion in Neurobiology, 12, 169177.CrossRefGoogle ScholarPubMed
Altshuler, L., Bookheimer, S., Proenza, M. A., Townsend, J., Sabb, F., Firestine, A., et al. (2005). Increased amygdala activation during mania: A functional magnetic resonance imaging study. American Journal of Psychiatry, 162, 12111213.CrossRefGoogle Scholar
American Psychiatric Association. (1994). Diagnostic and statistical manual of mental disorders (4th ed.). Washington, DC: Author.Google Scholar
Armony, J. L., Corbo, V., Clement, M. H., & Brunet, A. (2005). Amygdala response in patients with acute PTSD to masked and unmasked emotional facial expressions. American Journal of Psychiatry, 162, 19611963.CrossRefGoogle ScholarPubMed
Barratt, E. S., Stanford, M. S., Dowdy, L., Liebman, M. J., & Kent, T. A. (1999). Impulsive and premeditated aggression: A factor analysis of self-reported acts. Psychiatry Research, 86, 163173.CrossRefGoogle ScholarPubMed
Baxter, M. G., & Murray, E. A. (2002). The amygdala and reward. Nature Reviews: Neuroscience, 3, 563573.CrossRefGoogle ScholarPubMed
Berkowitz, L. (1993). Aggression: Its causes, consequences, and control. Philadelphia, PA: Temple University Press.Google Scholar
Best, M., Williams, J. M., & Coccaro, E. F. (2002). Evidence for a dysfunctional prefrontal circuit in patients with an impulsive aggressive disorder. Proceedings of the National Academy of Sciences of the United States of America, 99, 84488453.CrossRefGoogle ScholarPubMed
Birbaumer, N., Veit, R., Lotze, M., Erb, M., Hermann, C., Grodd, W., et al. (2005). Deficient fear conditioning in psychopathy: A functional magnetic resonance imaging study. Archives of General Psychiatry, 62, 799805.CrossRefGoogle ScholarPubMed
Blair, R. J. (2003). Facial expressions, their communicatory functions and neuro-cognitive substrates. Philosophical Transactions of the Royal Society B: Biological Sciences, 358, 561572.CrossRefGoogle ScholarPubMed
Blair, R. J. (2004a). The roles of orbital frontal cortex in the modulation of antisocial behavior. Brain and Cognition, 55, 198208.CrossRefGoogle ScholarPubMed
Blair, R. J. (2007). The amygdala and ventromedial prefrontal cortex in morality and psychopathy. Trends in Cognitive Sciences, 11, 387392.CrossRefGoogle ScholarPubMed
Blair, R. J., Colledge, E., Murray, L., & Mitchell, D. G. (2001). A selective impairment in the processing of sad and fearful expressions in children with psychopathic tendencies. Journal of Abnormal Child Psychology, 29, 491498.CrossRefGoogle ScholarPubMed
Blair, R. J. R. (2004b). The neurobiology of aggression. In Charney, D. S. & Nestler, E. J. (Eds.), Neurobiology of mental illness (2nd ed., pp. 10761077). New York: Oxford University Press.Google Scholar
Blair, R. J. R., Mitchell, D. G. V., & Blair, K. S. (2005). The psychopath: Emotion and the brain. Oxford: Blackwell.Google Scholar
Blaise, J. H., Koranda, J. L., Chow, U., Haines, K. E., & Dorward, E. C. (2008). Neonatal isolation stress alters bidirectional long-term synaptic plasticity in amygdalo-hippocampal synapses in freely behaving adult rats. Brain Research, 1193, 2533.CrossRefGoogle ScholarPubMed
Blanchard, R. J., Blanchard, D. C., Takahashi, T., & Kelley, M. J. (1977). Attack and defensive behaviour in the albino rat. Animal Behaviour, 25, 622634.CrossRefGoogle ScholarPubMed
Blonigen, D. M., Hicks, B. M., Krueger, R. F., Patrick, C. J. & Iacono, W. G. (2005). Psychopathic personality traits: Heritability and genetic overlap with internalizing and externalizing psychopathology. Psychological Medicine, 35, 637648.CrossRefGoogle ScholarPubMed
Blumberg, H. P., Fredericks, C., Wang, F., Kalmar, J. H., Spencer, L., Papademetris, X., et al. (2005). Preliminary evidence for persistent abnormalities in amygdala volumes in adolescents and young adults with bipolar disorder. Bipolar Disorder, 7, 570576.CrossRefGoogle Scholar
Blumberg, H. P., Kaufman, J., Martin, A., Charney, D. S., Krystal, J. H., & Peterson, B. S. (2004). Significance of adolescent neurodevelopment for the neural circuitry of bipolar disorder. Annals of the New York Academy of Sciences, 1021, 376383.CrossRefGoogle ScholarPubMed
Blumberg, H. P., Kaufman, J., Martin, A., Whiteman, R., Zhang, J. H., Gore, J. C., et al. (2003). Amygdala and hippocampal volumes in adolescents and adults with bipolar disorder. Archives of General Psychiatry, 60, 12011208.CrossRefGoogle ScholarPubMed
Bourgeois, J. P., Goldman-Rakic, P. S., & Rakic, P. (1994). Synaptogenesis in the prefrontal cortex of rhesus monkeys. Cerebral Cortex, 4, 7896.CrossRefGoogle ScholarPubMed
Bremner, J. D., Narayan, M., Staib, L. H., Southwick, S. M., McGlashan, T., & Charney, D. S. (1999). Neural correlates of memories of childhood sexual abuse in women with and without posttraumatic stress disorder. American Journal of Psychiatry, 156, 17871795.CrossRefGoogle ScholarPubMed
Bremner, J. D., Vermetten, E., Vythilingam, M., Afzal, N., Schmahl, C., Elzinga, B., et al. (2004). Neural correlates of the classic color and emotional Stroop in women with abuse-related posttraumatic stress disorder. Biological Psychiatry, 55, 612620.CrossRefGoogle ScholarPubMed
Bryant, R. A., Felmingham, K. L., Kemp, A. H., Barton, M., Peduto, A. S., Rennie, C., et al. (2005). Neural networks of information processing in posttraumatic stress disorder: A functional magnetic resonance imaging study. Biological Psychiatry, 58, 111118.CrossRefGoogle ScholarPubMed
Bryant, R. A., Kemp, A. H., Felmingham, K. L., Liddell, B., Olivieri, G., Peduto, A., et al. (2008). Enhanced amygdala and medial prefrontal activation during nonconscious processing of fear in posttraumatic stress disorder: An fMRI study. Human Brain Mapping, 29, 517523.CrossRefGoogle ScholarPubMed
Budhani, S., & Blair, R. J. (2005). Response reversal and children with psychopathic tendencies: Success is a function of salience of contingency change. Journal of Child Psychology and Psychiatry, 46, 972981.CrossRefGoogle ScholarPubMed
Budhani, S., Marsh, A. A., Pine, D. S. & Blair, R. J. (2007). Neural correlates of response reversal: Considering acquisition. Neuroimage, 34, 17541765.CrossRefGoogle ScholarPubMed
Budhani, S., Richell, R. A., & Blair, R. J. (2006). Impaired reversal but intact acquisition: Probabilistic response reversal deficits in adult individuals with psychopathy. Journal of Abnormal Psychology, 115, 552558.CrossRefGoogle ScholarPubMed
Burke, J. D., Loeber, R., & Lahey, B. B. (2007). Adolescent conduct disorder and interpersonal callousness as predictors of psychopathy in young adults. Journal of Clinical Child and Adolescent Psychology, 36, 334346.CrossRefGoogle ScholarPubMed
Chen, B. K., Sassi, R., Axelson, D., Hatch, J. P., Sanches, M., Nicoletti, M., et al. (2004). Cross-sectional study of abnormal amygdala development in adolescents and young adults with bipolar disorder. Biological Psychiatry, 56, 399405.CrossRefGoogle Scholar
Coccaro, E. F., McCloskey, M. S., Fitzgerald, D. A., & Phan, K. L. (2007). Amygdala and orbitofrontal reactivity to social threat in individuals with impulsive aggression. Biological Psychiatry, 62, 168178.CrossRefGoogle ScholarPubMed
Coccaro, E. F., Posternak, M. A., & Zimmerman, M. (2005). Prevalence and features of intermittent explosive disorder in a clinical setting. Journal of Clinical Psychiatry, 66, 12211227.CrossRefGoogle ScholarPubMed
Coccaro, E. F., Schmidt, C. A., Samuels, J. F., & Nestadt, G. (2004). Lifetime and 1-month prevalence rates of intermittent explosive disorder in a community sample. Journal of Clinical Psychiatry, 65, 820824.CrossRefGoogle Scholar
Cohen, H., Matar, M. A., Richter-Levin, G., & Zohar, J. (2006). The contribution of an animal model toward uncovering biological risk factors for PTSD. Annals of the New York Academy of Sciences, 1071, 335350.CrossRefGoogle ScholarPubMed
Colledge, E., & Blair, R. J. R. (2001). The relationship in children between the inattention and impulsivity components of attention deficit and hyperactivity disorder and psychopathic tendencies. Personality and Individual Differences, 30, 11751187.CrossRefGoogle Scholar
Connor, D. F. (2002). Aggression and anti-social behaviour in children and adolescents: Research and treatment. New York: Guilford Press.Google Scholar
Coplan, J. D., Andrews, M. W., Rosenblum, L. A., Owens, M. J., Friedman, S., Gorman, J. M., et al. (1996). Persistent elevations of cerebrospinal fluid concentrations of corticotropin-releasing factor in adult nonhuman primates exposed to early-life stressors: Implications for the pathophysiology of mood and anxiety disorders. Proceedings of the National Academy of Sciences of the United States of America, 93, 16191623.CrossRefGoogle ScholarPubMed
Cornell, D. G., Warren, J., Hawk, G., Stafford, E., Oram, G., & Pine, D. (1996). Psychopathy in instrumental and reactive violent offenders. Journal of Consulting and Clinical Psychology, 64, 783790.CrossRefGoogle ScholarPubMed
Crick, N. R., & Dodge, K. A. (1996). Social information-processing mechanisms on reactive and proactive aggression. Child Development, 67, 9931002.CrossRefGoogle ScholarPubMed
Critchley, H. D., Simmons, A., Daly, E. M., Russell, A., Van, A. T., Robertson, D. M., et al. (2000). Prefrontal and medial temporal correlates of repetitive violence to self and others. Biological Psychiatry, 47, 928934.CrossRefGoogle ScholarPubMed
Dadds, M. R., Perry, Y., Hawes, D. J., Merz, S., Riddell, A. C., Haines, D. J., et al. (2006). Attention to the eyes and fear-recognition deficits in child psychopathy. British Journal of Psychiatry, 189, 280281.CrossRefGoogle Scholar
Danielyan, A., Pathak, S., Kowatch, R. A., Arszman, S. P., & Johns, E. S. (2007). Clinical characteristics of bipolar disorder in very young children. Journal of Affective Disorders, 97, 5159.CrossRefGoogle ScholarPubMed
DelBello, M. P., Zimmerman, M. E., Mills, N. P., Getz, G. E., & Strakowski, S. M. (2004). Magnetic resonance imaging analysis of amygdala and other subcortical brain regions in adolescents with bipolar disorder. Bipolar Disorders, 6, 4352.CrossRefGoogle ScholarPubMed
Dixon, A., Howie, P., Starling, J., & Franzcp, M. P. H. (2005). Trauma exposure, posttraumatic stress, and psychiatric comorbidity in female juvenile offenders. Journal of the American Academy of Child & Adolescent Psychiatry, 44, 798806.CrossRefGoogle ScholarPubMed
Dolan, M., & Fullam, R. (2006). Face affect recognition deficits in personality-disordered offenders: Association with psychopathy. Psychological Medicine, 36, 15631569.CrossRefGoogle ScholarPubMed
Eisenberg, N. (2002). Empathy-related emotional responses, altruism, and their socialization. In Davidson, R. J. & Harrington, A. (Eds.), Visions of compassion: Western scientists and Tibetan Buddhists examine human nature (pp. 131164). New York: Oxford University Press.CrossRefGoogle Scholar
Etkin, A., & Wager, T. D. (2007). Functional neuroimaging of anxiety: A meta-analysis of emotional processing in PTSD, social anxiety disorder, and specific phobia. American Journal of Psychiatry, 164, 14761488.CrossRefGoogle ScholarPubMed
Everitt, B. J., Cardinal, R. N., Parkinson, J. A., & Robbins, T. W. (2003). Appetitive behavior: Impact of amygdala-dependent mechanisms of emotional learning. Annals of the New York Academy of Sciences, 985, 233250.CrossRefGoogle ScholarPubMed
Faedda, G. L., Baldessarini, R. J., Glovinsky, I. P., & Austin, N. B. (2004). Pediatric bipolar disorder: Phenomenology and course of illness. Bipolar Disorder, 6, 305313.CrossRefGoogle ScholarPubMed
Finger, E. C., Marsh, A. A., Mitchell, D. G. V., Reid, M., Sims, C., Budhani, S., et al. (2008). Abnormal ventromedial prefrontal cortex function in children with callous and unemotional traits during reversal learning. Archives of General Psychiatry, 65, 586594.CrossRefGoogle Scholar
Flor, H., Birbaumer, N., Hermann, C., Ziegler, S., & Patrick, C. J. (2002). Aversive Pavlovian conditioning in psychopaths: Peripheral and central correlates. Psychophysiology, 39, 505518.CrossRefGoogle ScholarPubMed
Frazzetto, G., Di, L. G., Carola, V., Proietti, L., Sokolowska, E., Siracusano, A., et al. (2007). Early trauma and increased risk for physical aggression during adulthood: The moderating role of MAOA genotype. PLoS ONE, 2, e486.CrossRefGoogle ScholarPubMed
Frick, P. J., O'Brien, B. S., Wootton, J. M., & McBurnett, K. (1994). Psychopathy and conduct problems in children. Journal of Abnormal Psychology, 103, 700707.CrossRefGoogle ScholarPubMed
Frick, P. J., Stickle, T. R., Dandreaux, D. M., Farrell, J. M., & Kimonis, E. R. (2005). Callous-unemotional traits in predicting the severity and stability of conduct problems and delinquency. Journal of Abnormal Child Psychology, 33, 471487.CrossRefGoogle ScholarPubMed
Frick, P. J., & White, S. F. (2008). Research review: The importance of callous-unemotional traits for developmental models of aggressive and antisocial behavior. Journal of Child Psychology and Psychiatry, 49, 359375.CrossRefGoogle ScholarPubMed
Giedd, J. N. (2004). Adolescent brain development: Vulnerabilities and opportunities. Annals of the New York Academy of Sciences, 1021, 7785.CrossRefGoogle Scholar
Giedd, J. N., Blumenthal, J., & Jeffries, N. O. (1999). Brain development during childhood and adolescence: A longitudinal MRI study. Nature Neuroscience, 2, 861863.CrossRefGoogle ScholarPubMed
Gordon, H. L., Baird, A. A., & End, A. (2004). Functional differences among those high and low on a trait measure of psychopathy. Biological Psychiatry, 56, 516521.CrossRefGoogle ScholarPubMed
Goyer, P. F., Andreason, P. J., Semple, W. E., Clayton, A. H., King, A. C., Compton-Toth, B. A., et al. (1994). Positron-emission tomography and personality disorders. Neuropsychopharmacology, 10, 2128.CrossRefGoogle ScholarPubMed
Grafman, J., Schwab, K., Warden, D., Pridgen, A., Brown, H. R., & Salazar, A. M. (1996). Frontal lobe injuries, violence, and aggression: A report of the Vietnam Head Injury Study. Neurology, 46, 12311238.CrossRefGoogle ScholarPubMed
Graham, Y. P., Heim, C., Goodman, S. H., Miller, A. H., & Nemeroff, C. B. (1999). The effects of neonatal stress on brain development: Implications for psychopathology. Development and Psychopathology, 11, 545565.CrossRefGoogle ScholarPubMed
Gregg, T. R., & Siegel, A. (2001). Brain structures and neurotransmitters regulating aggression in cats: Implications for human aggression. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 25, 91140.CrossRefGoogle ScholarPubMed
Hare, R., Glass, S. J., & Newman, J. P. (2006). Current perspectives on psychopathy. Annual Review of Clinical Psychology, 4, 217246.CrossRefGoogle Scholar
Hare, R. D., & Neumann, C. S. (2006). The PCL-R assessment of psychopathy: Development, structural properties, and new directions. In Patrick, C. J. (Ed.), Handbook of psychopathy (pp. 5890). New York: Guilford Press.Google Scholar
Hariri, A. R., Drabant, E. M., Munoz, K. E., Kolachana, B. S., Mattay, V. S., Egan, M. F., et al. (2005). A susceptibility gene for affective disorders and the response of the human amygdala. Archives of General Psychiatry, 62, 146152.CrossRefGoogle ScholarPubMed
Harpur, T. J., & Hare, R. D. (1994). Assessment of psychopathy as a function of age. Journal of Abnormal Psychology, 103, 604609.CrossRefGoogle ScholarPubMed
Hoffman, M. L. (1994). Discipline and internalisation. Developmental Psychology, 30, 2628.CrossRefGoogle Scholar
Hooker, C. I., Germine, L. T., Knight, R. T., & D'Esposito, M. (2006). Amygdala response to facial expressions reflects emotional learning. Journal of Neuroscience, 26, 89158922.CrossRefGoogle ScholarPubMed
Kiehl, K. A., Smith, A. M., Hare, R. D., Mendrek, A., Forster, B. B., Brink, J., et al. (2001). Limbic abnormalities in affective processing by criminal psychopaths as revealed by functional magnetic resonance imaging. Biological Psychiatry, 50, 677684.CrossRefGoogle ScholarPubMed
Klinnert, M., Emde, R., Butterfield, P., & Campos, J. (1987). Social referencing: The infant's use of emotional signals from a friendly adult with mother present. Annual Progress in Child Psychiatry and Child Development, 22, 427432.Google Scholar
Kochanska, G. (1997). Multiple pathways to conscience for children with different temperaments: From toddlerhood to age 5. Developmental Psychology, 33, 228240.CrossRefGoogle ScholarPubMed
Kosson, D. S., Suchy, Y., Mayer, A. R., & Libby, J. (2002). Facial affect recognition in criminal psychopaths. Emotion, 2, 398411.CrossRefGoogle ScholarPubMed
Kraemer, G. W., Ebert, M. H., Schmidt, D. E., & McKinney, W. T. (1989). A longitudinal study of the effect of different social rearing conditions on cerebrospinal fluid norepinephrine and biogenic amine metabolites in rhesus monkeys. Neuropsychopharmacology, 2, 175189.CrossRefGoogle ScholarPubMed
Kraszpulski, M., Dickerson, P. A., & Salm, A. K. (2006). Prenatal stress affects the developmental trajectory of the rat amygdala. Stress, 9, 8595.CrossRefGoogle ScholarPubMed
Lanius, R. A., Williamson, P. C., Densmore, M., Boksman, K., Gupta, M. A., Neufeld, R. W., et al. (2001). Neural correlates of traumatic memories in posttraumatic stress disorder: A functional MRI investigation. American Journal of Psychiatry, 158, 19201922.CrossRefGoogle ScholarPubMed
Leibenluft, E., Blair, R. J., Charney, D. S., & Pine, D. S. (2003). Irritability in pediatric mania and other childhood psychopathology. Annals of the New York Academy of Sciences, 1008, 201218.CrossRefGoogle ScholarPubMed
Levenston, G. K., Patrick, C. J., Bradley, M. M., & Lang, P. J. (2000). The psychopath as observer: Emotion and attention in picture processing. Journal of Abnormal Psychology, 109, 373385.CrossRefGoogle ScholarPubMed
Lewis, D. A. (1997). Development of the prefrontal cortex during adolescence: Insights into vulnerable neural circuits in schizophrenia. Neuropsychopharmacology, 16, 385398.CrossRefGoogle ScholarPubMed
Liberzon, I., Taylor, S. F., Amdur, R., Jung, T. D., Chamberlain, K. R., Minoshima, S., et al. (1999). Brain activation in PTSD in response to trauma-related stimuli. Biological Psychiatry, 45, 817826.CrossRefGoogle ScholarPubMed
Linnoila, M., Virkkunen, M., Scheinin, M., Nuutila, A., Rimon, R., & Goodwin, F. (1983). Low cerebrospinal fluid 5-hydroxyindoleacetic acid concentration differentiates impulsive from nonimpulsive violent behavior. Life Sciences, 33, 26092614.CrossRefGoogle ScholarPubMed
Lochhead, R. A., Parsey, R. V., Oquendo, M. A., & Mann, J. J. (2004). Regional brain gray matter volume differences in patients with bipolar disorder as assessed by optimized voxel-based morphometry. Biological Psychiatry, 55, 11541162.CrossRefGoogle ScholarPubMed
Lynam, D. R., Caspi, A., Moffitt, T. E., Loeber, R., & Stouthamer-Loeber, M. (2007). Longitudinal evidence that psychopathy scores in early adolescence predict adult psychopathy. Journal of Abnormal Psychology, 116, 155165.CrossRefGoogle ScholarPubMed
Lyoo, I. K., Kim, M. J., Stoll, A. L., Demopulos, C. M., Parow, A. M., Dager, S. R., et al. (2004). Frontal lobe gray matter density decreases in bipolar I disorder. Biological Psychiatry, 55, 648651.CrossRefGoogle ScholarPubMed
Machado, C. J., & Bachevalier, J. (2003). Non-human primate models of childhood psychopathology. Journal of Child Psychology and Psychiatry and Allied Disciplines, 44, 6487.CrossRefGoogle ScholarPubMed
Maestripieri, D., Higley, J. D., Lindell, S. G., Newman, T. K., McCormack, K. M., & Sanchez, M. M. (2006). Early maternal rejection affects the development of monoaminergic systems and adult abusive parenting in rhesus macaques (Macaca mulatta). Behavioral Neuroscience, 120, 10171024.CrossRefGoogle ScholarPubMed
Maestripieri, D., McCormack, K., Lindell, S. G., Higley, J. D., & Sanchez, M. M. (2006). Influence of parenting style on the offspring's behaviour and CSF monoamine metabolite levels in crossfostered and noncrossfostered female rhesus macaques. Behavioural Brain Research, 175, 9095.CrossRefGoogle ScholarPubMed
Marsh, A. A., Finger, E. C., Mitchell, D. G. V., Reid, M. E., Sims, C., Kosson, D. S., et al. (2008). Reduced amygdala response to fearful expressions in children and adolescents with callous-unemotional traits and disruptive behavior disorders. American Journal of Psychiatry, 165, 712720.CrossRefGoogle ScholarPubMed
McElroy, S. L. (1999). Recognition and treatment of DSM-IV intermittent explosive disorder. Journal of Clinical Psychiatry, 60(Suppl. 15), 1216.Google ScholarPubMed
McElroy, S. L., Soutullo, C. A., Beckman, D. A., Taylor, P. Jr., & Keck, P. E. Jr. (1998). DSM-IV intermittent explosive disorder: A report of 27 cases. Journal of Clinical Psychiatry, 59, 203210.CrossRefGoogle ScholarPubMed
McIntosh, A. M., Job, D. E., Moorhead, T. W., Harrison, L. K., Forrester, K., Lawrie, S. M., et al. (2004). Voxel-based morphometry of patients with schizophrenia or bipolar disorder and their unaffected relatives. Biological Psychiatry, 56, 544552.CrossRefGoogle ScholarPubMed
Meyer-Lindenberg, A., Buckholtz, J. W., Kolachana, B., R Harri, A., Pezawas, L., Blasi, G., et al. (2006). Neural mechanisms of genetic risk for impulsivity and violence in humans. Proceedings of the National Academy of Sciences of the United States of America of America, 103, 62696274.Google ScholarPubMed
Mineka, S., & Cook, M. (1993). Mechanisms involved in the observational conditioning of fear. Journal of Experimental Psychology: General, 122, 2338.CrossRefGoogle ScholarPubMed
Montague, P. R., & Berns, G. S. (2002). Neural economics and the biological substrates of valuation. Neuron, 36, 265284.CrossRefGoogle ScholarPubMed
Munoz, L. C., & Frick, P. J. (2007). The reliability, stability, and predictive utility of the self-report version of the Antisocial Process Screening Device. Scandinavian Journal of Psychology, 48, 299312.CrossRefGoogle ScholarPubMed
Neller, D. J., Denney, R. L., Pietz, C. A., & Thomlinson, R. P. (2006). The relationship between trauma and violence in a jail inmate sample. Journal of Interpersonal Violence, 21, 12341241.CrossRefGoogle Scholar
Nemeroff, C. B. (2004). Neurobiological consequences of childhood trauma. Journal of Clinical Psychiatry, 65(Suppl 1), 1828.Google ScholarPubMed
Newman, J. P., & Kosson, D. S. (1986). Passive avoidance learning in psychopathic and nonpsychopathic offenders. Journal of Abnormal Psychology, 95, 252256.CrossRefGoogle ScholarPubMed
Obradović, J., Pardini, D., Long, J. D., & Loeber, R. (2007). Measuring interpersonal callousness in boys from childhood to adolescence: An examination of longitudinal invariance and temporal stability. Journal of Clinical Child and Adolescent Psychology, 36, 276292.CrossRefGoogle ScholarPubMed
Oxford, M., Cavell, T. A., & Hughes, J. N. (2003). Callous/unemotional traits moderate the relation between ineffective parenting and child externalizing problems: A partial replication and extension. Journal of Clinical Child and Adolescent Psychology, 32, 577585.CrossRefGoogle ScholarPubMed
Padoa-Schioppa, C., & Assad, J. A. (2006). Neurons in the orbitofrontal cortex encode economic value. Nature, 441, 223226.CrossRefGoogle ScholarPubMed
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. New York: Oxford University Press.CrossRefGoogle Scholar
Pavuluri, M. N., O'Connor, M. M., Harral, E., & Sweeney, J. A. (2007). Affective neural circuitry during facial emotion processing in pediatric bipolar disorder. Biological Psychiatry, 62, 158167.CrossRefGoogle ScholarPubMed
Pennington, B. F., & Bennetto, L. (1993). Main effects or transaction in the neuropsychology of conduct disorder? Commentary on “The neuropsychology of conduct disorder.” Development and Psychopathology, 5, 153164.CrossRefGoogle Scholar
Protopopescu, X., Pan, H., Tuescher, O., Cloitre, M., Goldstein, M., Engelien, W., et al. (2005). Differential time courses and specificity of amygdala activity in posttraumatic stress disorder subjects and normal control subjects. Biological Psychiatry, 57, 464473.CrossRefGoogle ScholarPubMed
Raine, A. (2002). Annotation: The role of prefrontal deficits, low autonomic arousal, and early health factors in the development of antisocial and aggressive behavior in children. Journal of Child Psychology and Psychiatry, 43, 417434.CrossRefGoogle ScholarPubMed
Raine, A., Buchsbaum, M., & Lacasse, L. (1997). Brain abnormalities in murderers indicated by positron emission tomography. Biological Psychiatry, 42, 495508.CrossRefGoogle ScholarPubMed
Raine, A., Buchsbaum, M. S., Stanley, J., Lottenberg, S., Abel, L., & Stoddard, J. (1994). Selective reductions in prefrontal glucose metabolism in murderers. Biological Psychiatry, 36, 365373.CrossRefGoogle ScholarPubMed
Raine, A., Phil, D., Stoddard, J., Bihrle, S., & Buchsbaum, M. (1998). Prefrontal glucose deficits in murderers lacking psychosocial deprivation. Neuropsychiatry, Neuropsychology, and Behavioral Neurology, 11, 17.Google ScholarPubMed
Rauch, S. L., van der Kolk, B. A., Fisler, R. E., Alpert, N. M., Orr, S. P., Savage, C. R., et al. (1996). A symptom provocation study of posttraumatic stress disorder using positron emission tomography and script-driven imagery. Archives of General Psychiatry, 53, 380387.CrossRefGoogle ScholarPubMed
Rauch, S. L., Whalen, P. J., Shin, L. M., McInerney, S. C., Macklin, M. L., Lasko, N. B., et al. (2000). Exaggerated amygdala response to masked facial stimuli in posttraumatic stress disorder: A functional MRI study. Biological Psychiatry, 47, 769776.CrossRefGoogle ScholarPubMed
Rilling, J. K., Glenn, A. L., Jairam, M. R., Pagnoni, G., Goldsmith, D. R., Elfenbein, H. A., et al. (2007). Neural correlates of social cooperation and non-cooperation as a function of psychopathy. Biological Psychiatry, 61, 12601271.CrossRefGoogle ScholarPubMed
Ruchkin, V. V., Schwab-Stone, M., Koposov, R., Vermeiren, R., & Steiner, H. (2002). Violence exposure, posttraumatic stress, and personality in juvenile delinquents. Journal of the American Academy of Child & Adolescent Psychiatry, 41, 322329.CrossRefGoogle ScholarPubMed
Sabatini, M. J., Ebert, P., Lewis, D. A., Levitt, P., Cameron, J. L., & Mirnics, K. (2007). Amygdala gene expression correlates of social behavior in monkeys experiencing maternal separation. Journal of Neuroscience, 27, 32953304.CrossRefGoogle ScholarPubMed
Salm, A. K., Pavelko, M., Krouse, E. M., Webster, W., Kraszpulski, M., & Birkle, D. L. (2004). Lateral amygdaloid nucleus expansion in adult rats is associated with exposure to prenatal stress. Brain Research: Developmental Brain Research, 148, 159167.CrossRefGoogle ScholarPubMed
Schneider, F., Habel, U., Kessler, C., Posse, S., Grodd, W., & Muller-Gartner, H. W. (2000). Functional imaging of conditioned aversive emotional responses in antisocial personality disorder. Neuropsychobiology, 42, 192201.CrossRefGoogle ScholarPubMed
Schoenbaum, G., & Roesch, M. (2005). Orbitofrontal cortex, associative learning, and expectancies. Neuron, 47, 633636.CrossRefGoogle ScholarPubMed
Shin, L. M., Kosslyn, S. M., McNally, R. J., Alpert, N. M., Thompson, W. L., Rauch, S. L., et al. (1997). Visual imagery and perception in posttraumatic stress disorder. A positron emission tomographic investigation. Archives of General Psychiatry, 54, 233241.CrossRefGoogle ScholarPubMed
Shin, L. M., McNally, R. J., Kosslyn, S. M., Thompson, W. L., Rauch, S. L., Alpert, N. M., et al. (1999). Regional cerebral blood flow during script-driven imagery in childhood sexual abuse-related PTSD: A PET investigation. American Journal of Psychiatry, 156, 575584.CrossRefGoogle ScholarPubMed
Shin, L. M., Orr, S. P., Carson, M. A., Rauch, S. L., Macklin, M. L., Lasko, N. B., et al. (2004). Regional cerebral blood flow in the amygdala and medial prefrontal cortex during traumatic imagery in male and female Vietnam veterans with PTSD. Archives of General Psychiatry, 61, 168176.CrossRefGoogle ScholarPubMed
Shin, L. M., Whalen, P. J., Pitman, R. K., Bush, G., Macklin, M. L., Lasko, N. B., et al. (2001). An fMRI study of anterior cingulate function in posttraumatic stress disorder. Biological Psychiatry, 50, 932942.CrossRefGoogle ScholarPubMed
Shin, L. M., Wright, C. I., Cannistraro, P. A., Wedig, M. M., McMullin, K., Martis, B., et al. (2005). A functional magnetic resonance imaging study of amygdala and medial prefrontal cortex responses to overtly presented fearful faces in posttraumatic stress disorder. Archives of General Psychiatry, 62, 273281.CrossRefGoogle ScholarPubMed
Silva, J. A., Derecho, D. V., Leong, G. B., Weinstock, R., & Ferrari, M. M. (2001). A classification of psychological factors leading to violent behavior in posttraumatic stress disorder. Journal of Forensic Science, 46, 309316.CrossRefGoogle ScholarPubMed
Silverthorn, P. & Frick, P. J. (1999). Developmental pathways to antisocial behavior: The delayed-onset pathway in girls. Development and Psychopathology, 11, 101126.CrossRefGoogle ScholarPubMed
Soderstrom, H., Hultin, L., Tullberg, M., Wikkelso, C., Ekholm, S., & Forsman, A. (2002). Reduced frontotemporal perfusion in psychopathic personality. Psychiatry Research, 114, 8194.CrossRefGoogle ScholarPubMed
Soderstrom, H., Tullberg, M., Wikkelso, C., Ekholm, S., & Forsman, A. (2000). Reduced regional cerebral blood flow in non-psychotic violent offenders. Psychiatry Research, 98, 2941.CrossRefGoogle ScholarPubMed
Steiner, H., & Remsing, L. (2007). Practice parameter for the assessment and treatment of children and adolescents with oppositional defiant disorder. Journal of the American Academy of Child & Adolescent Psychiatry, 46, 126141.CrossRefGoogle ScholarPubMed
Steiner, H., Saxena, K., & Chang, K. (2003). Psychopharmacologic strategies for the treatment of aggression in juveniles. CNS Spectrums, 8, 298308.CrossRefGoogle ScholarPubMed
Sterzer, P., Stadler, C., Krebs, A., Kleinschmidt, A., & Poustka, F. (2005). Abnormal neural responses to emotional visual stimuli in adolescents with conduct disorder. Biological Psychiatry, 57, 715.CrossRefGoogle ScholarPubMed
Sterzer, P., Stadler, C., Poustka, F., & Kleinschmidt, A. (2007). A structural neural deficit in adolescents with conduct disorder and its association with lack of empathy. NeuroImage, 37, 335342.CrossRefGoogle ScholarPubMed
Taylor, E., Schachar, R., Thorley, G., & Wieselberg, M. (1986). Conduct disorder and hyperactivity: I. Separation of hyperactivity and antisocial conduct in British child psychiatric patients. British Journal of Psychiatry, 149, 760767.CrossRefGoogle ScholarPubMed
van Elst, L. T., Woermann, F. G., Lemieux, L., Thompson, P. J., & Trimble, M. R. (2000). Affective aggression in patients with temporal lobe epilepsy: A quantitative MRI study of the amygdala. Brain, 123, 234243.CrossRefGoogle ScholarPubMed
Verona, E., Patrick, C. J., & Joiner, T. E. (2001). Psychopathy, antisocial personality, and suicide risk. Journal of Abnormal Psychology, 110, 462470.CrossRefGoogle ScholarPubMed
Viding, E., Blair, R. J. R., Moffitt, T. E., & Plomin, R. (2005). Evidence for substantial genetic risk for psychopathy in 7-year-olds. Journal of Child Psychology and Psychiatry, 46, 592597.CrossRefGoogle ScholarPubMed
Volkow, N. D., & Tancredi, L. (1987). Neural substrates of violent behaviour. A preliminary study with positron emission tomography. British Journal of Psychiatry, 151, 668673.CrossRefGoogle ScholarPubMed
Williams, L. M., Kemp, A. H., Felmingham, K., Barton, M., Olivieri, G., Peduto, A., et al. (2006). Trauma modulates amygdala and medial prefrontal responses to consciously attended fear. NeuroImage, 29, 347357.CrossRefGoogle ScholarPubMed
Woermann, F. G., van Elst, L. T., Koepp, M. J., Free, S. L., Thompson, P. J., Trimble, M. R., et al. (2000). Reduction of frontal neocortical grey matter associated with affective aggression in patients with temporal lobe epilepsy: An objective voxel by voxel analysis of automatically segmented MRI. Journal of Neurology, Neurosurgery, and Psychiatry, 68, 162169.CrossRefGoogle ScholarPubMed
Wong, M., Fenwick, P., Fenton, G., Lumsden, J., Maisey, M., & Stevens, J. (1997). Repetitive and non-repetitive violent offending behaviour in male patients in a maximum security mental hospital—Clinical and neuroimaging findings. Medicine, Science, and the Law, 37, 150160.CrossRefGoogle Scholar
Wootton, J. M., Frick, P. J., Shelton, K. K., & Silverthorn, P. (1997). Ineffective parenting and childhood conduct problems: The moderating role of callous-unemotional traits. Journal of Consulting and Clinical Psychology, 65, 301308.CrossRefGoogle ScholarPubMed
Yakovlev, P. I., & Lecours, A. (1967). The myelogenetic cycles of regional maturation in the brain. In Minowski, A. (Ed.), Regional development of the brain in early life (pp. 365). Oxford: Blackwell.Google Scholar
Zoccolillo, M. (1992). Co-occurrence of conduct disorder and its adult outcomes with depressive and anxiety disorders: A review. Journal of the American Academy of Child & Adolescent Psychiatry, 31, 547556.CrossRefGoogle ScholarPubMed