Sunday, March 31, 2019

Neurobiological Mechanisms that Cause Aggression

neurobiological Mechanisms that Cause AggressionReview of the Neurobiological Mechanisms Which Underpin Reactive Aggression in Humans a approximate Look at Monoamine Oxidase A (MAOA)Module 1 Prosocial and antisocial doingss cross commissions the lifespanViolent acts have a significant toll on gentlemans gentleman societies annually over 1.6 million deaths worldwide occur repayable(p) to human violence (Viding Frith, 2006). Although acts of antagonism have an important adaptive purpose, when overexpressed, they whitethorn result in destructive consequences. Conventionally, antagonism has been defined as an lettered action with a purpose to inflict physical damage on another individual (Nelson Trainor, 2007). Currently two forms of encroachment have been accepted in humans controlled incursion instrumental subtype, and bring outd aggression instinctive subtype (Vatiello Stoff, 1997). Instrumental aggression, also referred to as premeditated and predatory, is agent rally planned and goal-oriented and has very oft been linked to psychopathy (Blair, 2001). This kind of aggression is thought to be regulate primarily by higher cortical systems and is characterized by diminished corpus amygdaloideum response (Nelson Trainor, 2007). Reactive aggression on the other hand, is dependent on the limbic and hypothalamic systems, and is characterized by high autonomic arousal (Siever, 2008). willing aggression generally represents a direct response to stimuli and storys for the majority of flushed acts. Individuals with borderline personality disorder, intermittent explosive disorder, or ADHD atomic number 18 particularly pr star to reactive aggression and impulsivity. foregoing question on the neurobiology of aggression stopovered out that for such individuals, repetitive acts of aggression are lots influenced by the underlying neurobiological susceptibility (Nelson Trainor, 2007). Indeed, one remarkable have got of aggression is its apparen t heritability. Twin and adoption studies suggest that genic factors eyeshade for betwixt 40% and 50% of population variance in guess of antisocial aggression (Buckholtz Meyer-Lindenberg, 2008). However, the blood amidst aggression and the underlying neurobiology is furthest from being simple (Nelson Trainor, 2007 Siever, 2008).Previous research in individuals prone to impulsivity and reactive aggression as salutary as a number of studies establish on animal models identified numerous genetic loci and neurotransmitters associated with reactive violence, including, unless not limited to, Dopaminergic genes (DRD4, DRD5, and DAT1), Serotonergic genes (5HTT, HTR1B), and genes creditworthy for encoding enzymes involved in regularization the levels of these neurotransmitters, particularly catechol-O-methyltransferase (COMT), and Monoamine Oxidase A (MAOA), often referred to as the warrior gene. However, the railroad tie between genotype and phenotype of aggression is only beg inning to be properly mum (Caspi, McClay, Moffitt, Mill, Martin, Craig, Taylor, Poulton, 2002 Nelson Trainor, 2007 Siever, 2008). While there have been umpteen studies showing the connexion between different genes and aggression, results were often mixed and inconclusive. Currently, the best candidate gene with the nearly supportive evidence appears to be Monoamine Oxidase A (see Brunner, Nelen, Breakefield, Ropers, van Oost, 1993 Byrd Manuck, 2014 Cases, Seif, Grimsby, Gaspar, Chen, Pournin, Muller, et al.,1995 Caspi, et al., 2002).The MAOA gene, located on the X chromosome, is a functional polymorphism with high use (MAOA-H) and embarrassed action at law (MAOA-L) variants, which encodes the MAOA enzyme, responsible for breaking-down neurotransmitters such as serotonin, dopamine, and noradrenaline (Shih, Chen, Ridd, 1999 Viding Frith, 2006). Previous research using animal models as well as humans exhibit an association between aggression and genetic deficiency in MAOA activity (Rowe, 2001). Transgenic mice without the gene encoding MAOA had higher amounts of brain serotonin (5-HT), dopamine (DA), and norepinephrine (NE), and displayed increase aggression (Cases, et al., 1995). After restoring MAOA activity, mice aggression was stabilized (Shih Thompson, 1999). In humans, point mutation in the MAOA gene led to MAOA deficiency and was found to be associated with reactive aggression in several men from the Dutch family. Moreover, crosswise generations, these men with MAOA knockout also showed frequent violent outbursts, particularly in response to mild provocation, and impulsive antisocial behaviour such as assault, rape, and attempted murder. (Brunner, et al., 1993). This condition, however, is quite uncommon and is unlikely to explain much rendering in human violence and aggression.Findings of the studies that only looked at levels of MAOA activity and antisocial outcomes in adulthood have been mixed and thus problematic to date since two MAOA -H and MAOA-L were linked to reactive aggression (Ficks Waldman, 2014 Nelson Trainor, 2007 Siever, 2008). The think that produced very naughty findings in that area and has later proven seminal was conducted by Caspi and colleagues (2002). This study was the first to look at the solutions of MAOA activity in compounding with puerility revilement on reactive aggression in adulthood. Indeed, perverse puerility experiences were found to affect the development and functioning of neuronal pathways involving the neurotransmitters metabolised by MAOA which can potentially result in increased aggression (Caspi, et al., 2002). Thus, Caspi and colleagues (2002) hypothesised that puerility maltreatment can predict reactive aggression in adulthood, and that this affinity is moderated by levels of MAOA expression.Results of this study demonstrated the dose-response piece of childhood maltreatment on the aggressive behaviour in adulthood, which was consistent with prior findings. Howe ver, this effect was much smaller in participants with the high-activity version of the MAOA gene as compared to men with pocket-size MAOA activity, suggesting the protective property of MAOA-H (Caspi et al, 2002). Moreover, as levels of maltreatment increased, so did the protective effect of the MAOA-H variant. A pitiful-activity MAOA gene combined with a history of childhood maltreatment increased the risk of aggressive behaviour in adulthood s neverthelessfold. These results back up the predicted hypothesis that MAOA activity would act as a moderator of the effects of childhood maltreatment on antisocial outcomes in adulthood.This study paved the way to a number of subsequent studies looking at gene and surroundings interaction. A recently published meta-analysis, which looked at 27 peer- check intoed studies on unbecoming childhood experiences, MAOA genotype, and aggressive and antisocial behaviour showed that results across 20 male person cohorts (11064 participants) were largely consistent with findings from the original study by Caspi and colleagues (2002) (Byrd Manuck, 2014). These findings remained robust even after removing each study individually.However, the question remains, how does low activity MAOA gene in combination with childhood stressors translate to antisocial behaviours in adulthood? Several theories have been proposed to answer this question. On one hand, the association between low activity MAOA gene and aggression appears paradoxical, since MAOA-L leads to increased levels of serotonin, which has been found to be positively chargedly correlated with impulse control and negatively correlated with aggression (Manuch, Flory, Ferrell, Mann, Muldoon, 2000 Siever, 2008). However, MAOA also plays a role in regulating dopamine and norepinephrine, which were shown to refuse thresholds for violent response to perceived threat (Manuch, et al., 2000). High levels of DA and NE, resulting from MAOA dysfunction, would activate a fight or fl ight response, and indirectly enhance aggression (Volavka, Bilder, Nolan, 2004). Indeed, previous studies showed a similar association between COMT gene (also responsible for breakdown of NE and DA) and aggression (Siever, 2007 Volavka, et al., 2004). Therefore the elevated levels of NE and DA, due to low expression of MAOA, would be consistent with the results of previous studies, showing an association between low activity MAOA gene and aggression. Nevertheless, this theory does not account for the role of childhood maltreatment on aggression, and as it was mentioned previously, results of studies looking whole at MAOA activity and aggression are mixed and inconclusive, showing both high and low activity MAOA gene being associated with impulsive aggression (Ficks Waldman, 2014 Nelson Trainor, 2007 Siever, 2008).Another theory that sheds more light on the mechanism by means of which MAOA deficiency in combination with childhood maltreatment influences aggression in adulthood relies on the findings that high concentrations of intracellular serotonin have been associated with increased reactivity to stress and elevated anxiety (Seif De Maeyer, 1999, Viding Frith, 2006). Therefore, it is possible that MAOA deficiency efficacy predispose individuals to spooky hyper-reactivity to a threat (maltreatment). While genetic sensitivity alone rarely results in adverse outcomes in adulthood, when combined with childhood stressors, it might potentially have consequences on brain function (Meyer-Lindenberg, Buckholtz, Kolachana, Hariri, Pezawas, Blasi, Wabnitz, et al., 2006). Previous findings in populations prone to impulsive violence demonstrated functional and geomorphological abnormalities in brain areas associated with perception and pattern of emotions, particularly in the amygdala, orbitofrontal cortex, and the merged regions (Davidson, Putnam, Larson, 2000). Neuropsychological functions associated with these brain regions were also compromised in the af orementioned populations (Blair, Peschardt, Budhani, Mitchell, Pine, 2006). Previous campaign using animal models and clinic samples seems to suggest that maltreatment negatively affects the functioning of the neural structures involved during an individuals reaction to threat (i.e., pariaquaductal gray and amygdala) and the regulation of the triggered threat response (i.e., orbitofrontal cortex and anterior cingulate). Therefore, the genetic risk (MAOA-L) along with childhood maltreatment may result in deepens to brain function, and later increases the risk of impulsive aggression (Viding Frith, 2006).These speculations imply that there is no one clear explanation for the findings we currently have regarding genetic variation and its effect on aggression. The human brain and the effects of genetic and environmental factors on its development are too complex to assume that one specialized gene, or neurotransmitter levels are responsible for aggression. More likely, it is the g radual change in neural pathways that regulate aggression. As of this moment, it hush remains unclear if aggression in adulthood that is observed in many of the aforementioned studies is due to developmental change in neural circuits or to a change in neurotransmitter function. Moreover, the effects of these changes for adults are very different than for children. Therefore, it is possible that low MAOA activity resulted in compensatory changes which transformed the organisation of the nervous system in children during the sensitive period of brain development, and was later reflected in antisocial outcomes in adulthood (Lesch Merschdorf, 2000).Attempts to replicate Caspi et. al. (2002) findings in female populations yielded importantly different results. fresh meta-analysis that looked at 11 studies with female samples produced inconclusive results even though MAOA activity had a significant association with adverse childhood events, high, as contradictory to low, MAOA activity in combination with childhood maltreatment was associated with antisocial behaviour in adulthood. Moreover, this interaction was weaker, and after removing a few individual studies, it befuddled its significance (Byrd Manuck, 2014). Replicating Caspi and colleagues study using female cohorts has proven to be significantly more complicated for two main reasons. First of all, dividing females into two groups base on MAOA activity is challenging due to uncertain inactivation of heterozygous alleles. Secondly, severe personality disorders and antisocial outcomes are quite rare in women, thus it is difficult to get a large enough sample to demonstrate dose-response relationships. In sum, taking into account findings from previous studies it is unclear what relationship MAOA activity plays in antisocial outcomes in women, and calls for further investigation (Caspi, et al., 2002 Byrd Manuck, 2014).In attempts to understand the relationship between genes, environment, and aggression, t he study by Caspi and colleagues (2002), as well as numerous studies that came out afterward, certainly advanced our understanding in the field. However, it is important to point out the limitations that characterise many of the research studies investigate the relationship between gene-environment interaction and adult antisocial behaviour. First of all, deficiency of published articles reporting null findings due to publication bias still remains a big problem in the field. As a result, published findings seem more robust than they actually are (Duncane Keller, 2011). Secondly, the samples of many studies, primarily those with female subjects, are often too small resulting in inadequate statistical power (Byrk Manuck, 2014). Indeed, negative findings had larger sample sizes compared to positive ones. These limitations make it difficult to correct for potential false-positive results (Duncan Keller, 2011). This is especially the case in replication attempts using female popula tions, and in neuroimaging studies. Finally, due to the difficulty recruiting participants for these studies, samples are often not easily comparable and consist of individuals with many comorbid psychiatrical conditions, making it difficult to tease apart effects of certain genetic variations and maltreatment on specific psychopathology (McCrory, DeBrito, Viding, 2010).While there is good evidence to suggest that genotype, particularly variants of the Monoamine Oxidase A gene, in combination with childhood maltreatment, plays an important role in reactive aggression in human adulthood, the exact underlying mechanism remains unclear. The aforementioned controversies call for discretion when making any strong conclusions regarding the effects of genetic variation on antisocial outcomes. Further research, including longitudinal studies, genome-wide association studies, gene-environment-sex and gene-gene interaction studies, and neuroimaging studies, is necessary to bettor understa nd the underlying neurobiological mechanisms which underpin reactive aggression in humans.ReferencesBlair, R. J. (2001). Neurocognitive models of aggression, the antisocial personality disorders, and psychopathy. Journal of Neurology, Neurosurgery and psychological medicine, 71, 727-731.Blair, R. J. R., Peschardt, K. S., Budhani, S., Mitchell, D. G., Pine, D. S. J. (2006). The development of psychopathy. Journal of Chid. Psychology and Psychiatry and allied disciplines, 47(3-4), 262-276.Brunner, H. G., Nelen, M., Breakefield, X. O., Ropers, H. H., van Oost, B. A. (1993). Abnormal behaviour associated with point mutation in the structural gene for monoamine oxidase A. Science, 262(5133), 578-580.Buckholtz J. W. Meyer-Lindenberg, A. (2008). MAO and the neurogenetic architecture of human aggression. Trends in Neuroscience, 31(3), 120-129.Byrd, A. L. Manuck, S. B. (2014). MAOA, childishness Maltreatment, and Antisocial Behavior Meta-analysis of a gene-Environment Interaction. Biologi cal Psychiatry, 75(1), 9-17.Cases, O., Seif, I., Grimsby, J., Gaspar, P., Chen, K., Pournin, S., Muller, U., et al. (1995). Aggressive behaviour and adapted amounts of brain serotonin and norepinephrine in mice macking MAOA. Science, 268(5218), 1763-1766.Caspi, A., McClay, J. Moffitt, T. E., Mill, J., Martin, J., Craig, I. W., Taylor, A., Poulton, R. (2002). Role of Genotype in the Cycle of Violence in Maltreated Children. Science, 297(2), 851-853.Davidson, R. J., Putnam, K. M., Larson, C. L. (2000). Dysfunction in the neural circuitry of emotion regulation- a possible prelude to violence. Science, 289(5479), 591-594.Duncan, L. Keller, M. (2011). A critical review of the first 10 years of candidate gene-by-environment interaction research in psychiatry. The American Journal of Psychiatry, 168, 1041-1049.Ficks, C. Waldman, I. (2014). Candidate Genes for Aggression and Antisocial Behavior A Meta-analysis of Association Studies of the 5HTTLPR and MAOA-uVNTR. Behavioral Genetics, 44( 5), 427-444.Lesch, K. P. Merschdorf, U. (2000). Impulsivity, aggression, and serotonin A molecular psychobiological perspective. Behavioral Sciences and the Law, 18, 581-604.Manuch, S. B., Flory, J. D., Ferrell, R. E., Mann, J. J., Muldoon, M. F. (2000). A regulatory polymorphism of the monoamine oxidase- A gene may be association with variability in aggression, impulsivity, and central nervous system serotonergic responsivity. Psychiatry search, 95, 9-23.McCrory, E., De Brito, S. A., Viding, E. (2010). Research Review The neurobiology and genetics of maltreatment and adversity. The Journal of Child Psychology and Psychiatry, 51(10), 1079-1095.Meyer-Lindenberg, A., Buckholtz, J. W., Kolachana, B., Hariri, A. R., Pezawas, L., Blasi, G., Wabnitz, A., Honea, R., Verchinski, B., Callicott, J., et al. (2006). neuronic mechanisms of genetic risk for impulsivity and violence in humans. legal proceeding of Natlional honorary society of Sciences of the linked States of America, 103(16 ), 6269-6274.Nelson, R. J., Trainor, B. C. (2007). Neural mechanisms of aggression. Nature, 8, 536-546.Shih J. C. Thompson, R. F. (1999). Monoamine oxidase in neuropsychiatry and behaviour. American Journal of Human Genetics, 65(3), 593-598.Seif, I. De Maeyer, E. (1999). Knockout command Knockout mice for monoamine oxidase A. International Journal of Neuropsychopharmacology, 12, 241-243.Shih, J. C., Chen, K., Ridd, M. J. (1999). Monoamine Oxidase from genes to behaviour. Annual Review of Neuroscience, 11, 197-217.Siever, L. J. (2008). Neurobiology of Aggression and Violence. American Journal of Psychiatry, 165, 429-442.Vatiello, B., Stoff, D. M. (1997). Subtypes of aggression and their relevance to child psychiatry. Journal of American Academy of Child and Adolescent Psychiatry, 36, 307-315.Viding, E. Frith, U. (2006). Genes for susceptibility to violence lurk in the brain. Proceedings of the National Academy of Sciences of the United States, 103(16), 6085-6086.Volavka, J., B ilder, R. J., Nolan, K. (2004). Catecholamines and aggression The role of COMT and MAO polymorphism. archives of the New York Academy of Sciences, 1036, 393-398.

No comments:

Post a Comment

Note: Only a member of this blog may post a comment.