Monoamine oxidase A

Last updated
MAOA
Monoamine oxidase A 2BXS.png
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases MAOA , MAO-A, monoamine oxidase A, BRNRS
External IDs OMIM: 309850 MGI: 96915 HomoloGene: 203 GeneCards: MAOA
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001270458
NM_000240

NM_173740

RefSeq (protein)

NP_000231
NP_001257387

NP_776101

Location (UCSC) Chr X: 43.65 – 43.75 Mb Chr X: 16.49 – 16.55 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse
MAOA gene is located on the short (p) arm of the X chromosome at position 11.3. MAOA Gene with ChrX Ideogram.svg
MAOA gene is located on the short (p) arm of the X chromosome at position 11.3.

Monoamine oxidase A, also known as MAO-A, is an enzyme (E.C. 1.4.3.4) that in humans is encoded by the MAOA gene. [5] [6] This gene is one of two neighboring gene family members that encode mitochondrial enzymes which catalyze the oxidative deamination of amines, such as dopamine, norepinephrine, and serotonin. A mutation of this gene results in Brunner syndrome. This gene has also been associated with a variety of other psychiatric disorders, including antisocial behavior. Alternatively spliced transcript variants encoding multiple isoforms have been observed. [7]

Contents

Structures

Gene

Monoamine oxidase A, also known as MAO-A, is an enzyme that in humans is encoded by the MAOA gene. [5] [6] The promoter of MAOA contains conserved binding sites for Sp1, GATA2, and TBP. [8] This gene is adjacent to a related gene ( MAOB ) on the opposite strand of the X chromosome. [9]

In humans, there is a 30-base repeat sequence repeated several different numbers of times in the promoter region of MAO-A. There are 2R (two repeats), 3R, 3.5R, 4R, and 5R variants of the repeat sequence, with the 3R and 4R variants most common in all populations. The variants of the promoter have been found to appear at different frequencies in different ethnic groups in an American sample cohort. [10]

The epigenetic modification of MAOA gene expression through methylation likely plays an important role in women. A study from 2010 found epigenetic methylation of MAOA in men to be very low and with little variability compared to women, while having higher heritability in men than women. [11] [12]

Protein

MAO-A shares 70% amino acid sequence identity with its homologue MAO-B. [13] Accordingly, both proteins have similar structures. Both MAO-A and MAO-B exhibit an N-terminal domain that binds flavin adenine dinucleotide (FAD), a central domain that binds the amine substrate, and a C-terminal α-helix that is inserted in the outer mitochondrial membrane. [13] [14] MAO-A has a slightly larger substrate-binding cavity than MAO-B, which may be the cause of slight differences in catalytic activity between the two enzymes, as shown in quantitative structure-activity relationship experiments. [15] Both enzymes are relatively large, about 60 kilodaltons in size, and are believed to function as dimers in living cells. [14]

Function

Monoamine oxidase A catalyzes O2-dependent oxidation of primary arylalkyl amines, most importantly neurotransmitters such as dopamine and serotonin. This is the initial step in the breakdown of these molecules. The products are the corresponding aldehyde, hydrogen peroxide, and ammonia:

RCH
2
-Amine + O
2
+ H
2
O
→ R-Aldehyde + H
2
O
2
+ NH
3

This reaction is believed to occur in three steps, using FAD as an electron-transferring cofactor. First, the amine is oxidized to the corresponding imine, with reduction of FAD to FADH2. Second, O2 accepts two electrons and two protons from FADH2, forming H
2
O
2
and regenerating FAD. Third, the imine is hydrolyzed by water, forming ammonia and the aldehyde. [15] [16]

Compared to MAO-B, MAO-A has a higher specificity for serotonin and norepinephrine, while the two enzymes have similar affinity for dopamine and tyramine. [17]

MAO-A is a key regulator for normal brain function. In the brain, the highest levels of transcription occur in the brain stem, hypothalamus, amygdala, habenula, and nucleus accumbens, and the lowest in the thalamus, spinal cord, pituitary gland, and cerebellum. [17] Its expression is regulated by the transcription factors SP1, GATA2, and TBP via cAMP-dependent regulation. [8] [17] MAO-A is also expressed in cardiomyocytes, where it is induced in response to stress such as ischemia and inflammation. [8]

Clinical significance

Cancer

MAO-A produces an amine oxidase, which is a class of enzyme known to affect carcinogenesis. Clorgyline, an MAO-A enzyme inhibitor, prevents apoptosis in melanoma cells, in vitro. [18] Cholangiocarcinoma suppresses MAO-A expression, and those patients with higher MAO-A expression had less adjacent organ invasion and better prognosis and survival. [19]

Cardiovascular disease

MAOA activity is linked to apoptosis and cardiac damage during cardiac injury following ischemic-reperfusion. [8]

Behavioral and neurological disorders

There is some association between low activity forms of the MAOA gene and autism. [20] Mutations in the MAOA gene results in monoamine oxidase deficiency, or Brunner syndrome. [7] Other disorders associated with MAO-A include Alzheimer's disease, aggression, panic disorder, bipolar disorder, major depressive disorder, and attention deficit hyperactivity disorder. [8] Effects of parenting on self-regulation in adolescents appear to be moderated by 'plasticity alleles', of which the 2R and 3R alleles of MAOA are two, with "the more plasticity alleles males (but not females) carried, the more and less self-regulation they manifested under, respectively, supportive and unsupportive parenting conditions." [21]

Depression

MAO-A levels in the brain as measured using positron emission tomography are elevated by an average of 34% in patients with major depressive disorder. [22] Genetic association studies examining the relationship between high-activity MAOA variants and depression have produced mixed results, with some studies linking the high-activity variants to major depression in females, [23] depressed suicide in males, [24] major depression and sleep disturbance in males [25] and major depressive disorder in both males and females. [26]

Other studies failed to find a significant relationship between high-activity variants of the MAOA gene and major depressive disorder. [27] [28] In patients with major depressive disorder, those with MAOA G/T polymorphisms (rs6323) coding for the highest-activity form of the enzyme have a significantly lower magnitude of placebo response than those with other genotypes. [29]

Antisocial behavior

In humans, an association between the 2R allele of the VNTR region of the gene and an increase in the likelihood of committing serious crime or violence has been found. The VNTR 2R allele of MAOA has been found to be a risk factor for violent delinquency, when present in association with stresses, i.e. family issues, low popularity or failing school. [30] [31] [32] [33]

A connection between the MAO-A gene 3R version and several types of anti-social behaviour has been found: Maltreated children with genes causing high levels of MAO-A were less likely to develop antisocial behavior. [34] Low MAO-A activity alleles which are overwhelmingly the 3R allele in combination with abuse experienced during childhood resulted in an increased risk of aggressive behaviour as an adult, [35] and men with the low activity MAOA allele were more genetically vulnerable even to punitive discipline as a predictor of antisocial behaviour. [36] High testosterone, maternal tobacco smoking during pregnancy, poor material living standards, dropping out of school, and low IQ predicted violent behavior are associated with men with the low-activity alleles. [37] [38] According to a large meta-analysis in 2014, the 3R allele had a small, nonsignificant effect on aggression and antisocial behavior, in the absence of other interaction factors. Owing to methodological concerns, the authors do not view this as evidence in favor of an effect. [39]

The MAO-A gene was the first candidate gene for antisocial behavior and was identified during a "molecular genetic analysis of a large, multigenerational, and notoriously violent, Dutch kindred". [40] A study of Finnish prisoners revealed that a MAOA-L (low-activity) genotype, which contributes to low dopamine turnover rate, was associated with extremely violent behavior. [41] For the purpose of the study, "extremely violent behavior" was defined as at least ten committed homicides, attempted homicides or batteries.

However, a large genome-wide association study has failed to find any large or statistically significant effects of the MAOA gene on aggression. [42] A separate GWAS on antisocial personality disorder likewise did not report a significant effect of MAOA. [43] Another study, while finding effects from a candidate gene search, failed to find any evidence in a large GWAS. [41] A separate analysis of human and rat genome wide association studies, Mandelian randomization studies, and causal pathway analyses likewise failed to reveal robust evidence of MAOA in aggression. [44] This lack of replication is predicted from the known issues of candidate gene research, which can produce many substantial false positives. [45]

Aggression and the "Warrior gene"

Low-activity variants of the VNTR promoter region of the MAO-A gene have been referred to as the warrior gene. [46] When faced with social exclusion or ostracism, individuals with the low activity MAO-A variants showed higher levels of aggression than individuals with the high activity MAO-A gene. [47] Low activity MAO-A could significantly predict aggressive behaviour in a high provocation situation: Individuals with the low activity variant of the MAO-A gene were more likely (75% as opposed to 62%, out of a sample size of 70) to retaliate, and with greater force, as compared to those with a normal MAO-A variant if the perceived loss was large. [48]

The effects of MAOA genes on aggression have also been criticized for being heavily overstated. [49] Indeed, the MAOA gene, even in conjunction with childhood adversity, is known to have a very small effect. [50] The vast majority of people with the associated alleles have not committed any violent acts. [51] [52]

In a 2009 criminal trial in the United States, an argument based on a combination of "warrior gene" and history of child abuse was successfully used to avoid a conviction of first-degree murder and the death penalty; however, the convicted murderer was sentenced to 32 years in prison. [53] [54] In a second case, an individual was convicted of second-degree murder, rather than first-degree murder, based on a genetic test that revealed he had the low-activity MAOA variant. [55] Judges in Germany are more likely to sentence offenders to involuntary psychiatric hospitalization on hearing an accused's MAOA-L genotype. [56]

Epigenetics

Studies have linked methylation of the MAOA gene with nicotine and alcohol dependence in women. [57] A second MAOA VNTR promoter, P2, influences epigenetic methylation and interacts with having experienced child abuse to influence antisocial personality disorder symptoms, only in women. [58] A study of 34 non-smoking men found that methylation of the gene may alter its expression in the brain. [59]

Animal studies

A dysfunctional MAOA gene has been correlated with increased aggression levels in mice, [60] [61] and has been correlated with heightened levels of aggression in humans. [62] In mice, a dysfunctional MAOA gene is created through insertional mutagenesis (called 'Tg8'). [60] Tg8 is a transgenic mouse strain that lacks functional MAO-A enzymatic activity. Mice that lacked a functional MAOA gene exhibited increased aggression towards intruder mice. [60] [63]

Some types of aggression exhibited by these mice were territorial aggression, predatory aggression, and isolation-induced aggression. [61] The MAO-A deficient mice that exhibited increased isolation-induced aggression reveals that an MAO-A deficiency may also contribute to a disruption in social interactions. [64] There is research in both humans and mice to support that a nonsense point mutation in the eighth exon of the MAOA gene is responsible for impulsive aggressiveness due to a complete MAO-A deficiency. [60] [62]

Interactions

Transcription factors

A number of transcription factors bind to the promoter region of MAO-A and upregulate its expression. These include:Sp1 transcription factor, GATA2, TBP. [8]

Inducers

Synthetic compounds that up-regulate the expression of MAO-A include Valproic acid (Depakote) [65]

Inhibitors

Substances that inhibit the enzymatic activity of MAO-A include:

See also

Related Research Articles

An allele, or allelomorph, is a variant of the sequence of nucleotides at a particular location, or locus, on a DNA molecule.

<span class="mw-page-title-main">Monoamine oxidase</span> Family of enzymes

Monoamine oxidases (MAO) are a family of enzymes that catalyze the oxidation of monoamines, employing oxygen to clip off their amine group. They are found bound to the outer membrane of mitochondria in most cell types of the body. The first such enzyme was discovered in 1928 by Mary Bernheim in the liver and was named tyramine oxidase. The MAOs belong to the protein family of flavin-containing amine oxidoreductases.

Antisocial personality disorder is a personality disorder characterized by a limited capacity for empathy and a long-term pattern of disregard or violation of the rights of others. Other notable symptoms include impulsivity and reckless behavior, a lack of remorse after hurting others, deceitfulness, irresponsibility, and aggressive behavior.

<span class="mw-page-title-main">Monoamine neurotransmitter</span> Monoamine that acts as a neurotransmitter or neuromodulator

Monoamine neurotransmitters are neurotransmitters and neuromodulators that contain one amino group connected to an aromatic ring by a two-carbon chain (such as -CH2-CH2-). Examples are dopamine, norepinephrine and serotonin.

<span class="mw-page-title-main">Serotonin transporter</span> Mammalian protein found in humans

The serotonin transporter also known as the sodium-dependent serotonin transporter and solute carrier family 6 member 4 is a protein that in humans is encoded by the SLC6A4 gene. SERT is a type of monoamine transporter protein that transports the neurotransmitter serotonin from the synaptic cleft back to the presynaptic neuron, in a process known as serotonin reuptake.

Catechol-<i>O</i>-methyltransferase Class of enzymes

Catechol-O-methyltransferase is one of several enzymes that degrade catecholamines, catecholestrogens, and various drugs and substances having a catechol structure. In humans, catechol-O-methyltransferase protein is encoded by the COMT gene. Two isoforms of COMT are produced: the soluble short form (S-COMT) and the membrane bound long form (MB-COMT). As the regulation of catecholamines is impaired in a number of medical conditions, several pharmaceutical drugs target COMT to alter its activity and therefore the availability of catecholamines. COMT was first discovered by the biochemist Julius Axelrod in 1957.

<span class="mw-page-title-main">Dopamine transporter</span> Mammalian protein found in Homo sapiens

The dopamine transporter (DAT) also is a membrane-spanning protein coded for in the human by the SLC6A3 gene,, that pumps the neurotransmitter dopamine out of the synaptic cleft back into cytosol. In the cytosol, other transporters sequester the dopamine into vesicles for storage and later release. Dopamine reuptake via DAT provides the primary mechanism through which dopamine is cleared from synapses, although there may be an exception in the prefrontal cortex, where evidence points to a possibly larger role of the norepinephrine transporter.

Dopamine receptor D<sub>4</sub> Protein-coding gene in the species Homo sapiens

The dopamine receptor D4 is a dopamine D2-like G protein-coupled receptor encoded by the DRD4 gene on chromosome 11 at 11p15.5.

<span class="mw-page-title-main">Gene–environment interaction</span> Response to the same environmental variation differently by different genotypes

Gene–environment interaction is when two different genotypes respond to environmental variation in different ways. A norm of reaction is a graph that shows the relationship between genes and environmental factors when phenotypic differences are continuous. They can help illustrate GxE interactions. When the norm of reaction is not parallel, as shown in the figure below, there is a gene by environment interaction. This indicates that each genotype responds to environmental variation in a different way. Environmental variation can be physical, chemical, biological, behavior patterns or life events.

<span class="mw-page-title-main">CYP2A6</span> Protein-coding gene in the species Homo sapiens

Cytochrome P450 2A6 is a member of the cytochrome P450 mixed-function oxidase system, which is involved in the metabolism of xenobiotics in the body. CYP2A6 is the primary enzyme responsible for the oxidation of nicotine and cotinine. It is also involved in the metabolism of several pharmaceuticals, carcinogens, and a number of coumarin-type alkaloids. CYP2A6 is the only enzyme in the human body that appreciably catalyzes the 7-hydroxylation of coumarin, such that the formation of the product of this reaction, 7-hydroxycoumarin, is used as a probe for CYP2A6 activity.

<span class="mw-page-title-main">CYP2C19</span> Mammalian protein found in humans

Cytochrome P450 2C19 is an enzyme protein. It is a member of the CYP2C subfamily of the cytochrome P450 mixed-function oxidase system. This subfamily includes enzymes that catalyze metabolism of xenobiotics, including some proton pump inhibitors and antiepileptic drugs. In humans, it is the CYP2C19 gene that encodes the CYP2C19 protein. CYP2C19 is a liver enzyme that acts on at least 10% of drugs in current clinical use, most notably the antiplatelet treatment clopidogrel (Plavix), drugs that treat pain associated with ulcers, such as omeprazole, antiseizure drugs such as mephenytoin, the antimalarial proguanil, and the anxiolytic diazepam.

The field of psychology has been greatly influenced by the study of genetics. Decades of research have demonstrated that both genetic and environmental factors play a role in a variety of behaviors in humans and animals. The genetic basis of aggression, however, remains poorly understood. Aggression is a multi-dimensional concept, but it can be generally defined as behavior that inflicts pain or harm on another.

<span class="mw-page-title-main">Brunner syndrome</span> X-linked recessive disorder characterised by impulsive behaviour

Brunner syndrome is a rare genetic disorder associated with a mutation in the MAOA gene. It is characterized by lower than average IQ, problematic impulsive behavior, sleep disorders and mood swings. It was identified in fourteen males from one family in 1993. It has since been discovered in additional families.

<span class="mw-page-title-main">NPAS2</span> Protein-coding gene in the species Homo sapiens

Neuronal PAS domain protein 2 (NPAS2) also known as member of PAS protein 4 (MOP4) is a transcription factor protein that in humans is encoded by the NPAS2 gene. NPAS2 is paralogous to CLOCK, and both are key proteins involved in the maintenance of circadian rhythms in mammals. In the brain, NPAS2 functions as a generator and maintainer of mammalian circadian rhythms. More specifically, NPAS2 is an activator of transcription and translation of core clock and clock-controlled genes through its role in a negative feedback loop in the suprachiasmatic nucleus (SCN), the brain region responsible for the control of circadian rhythms.

Gene–environment correlation is said to occur when exposure to environmental conditions depends on an individual's genotype.

<span class="mw-page-title-main">Monoamine oxidase B</span> Protein-coding gene in the species Homo sapiens

Monoamine oxidase B, also known as MAOB, is an enzyme that in humans is encoded by the MAOB gene.

<span class="mw-page-title-main">Proline oxidase</span> Protein-coding gene in the species Homo sapiens

Proline dehydrogenase, mitochondrial is an enzyme that in humans is encoded by the PRODH gene.

Rs6265, also called Val66Met or G196A, is a gene variation, a single nucleotide polymorphism (SNP) in the BDNF gene that codes for brain-derived neurotrophic factor.

<span class="mw-page-title-main">ANKK1</span> Protein-coding gene in the species Homo sapiens

Ankyrin repeat and kinase domain containing 1 (ANKK1) also known as protein kinase PKK2 or sugen kinase 288 (SgK288) is an enzyme that in humans is encoded by the ANKK1 gene. The ANKK1 is a member of an extensive family of the Ser/Thr protein kinase family, and protein kinase superfamily involved in signal transduction pathways.

A behaviour mutation is a genetic mutation that alters genes that control the way in which an organism behaves, causing their behavioural patterns to change.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000189221 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000025037 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. 1 2 Hotamisligil GS, Breakefield XO (August 1991). "Human monoamine oxidase A gene determines levels of enzyme activity". American Journal of Human Genetics. 49 (2): 383–92. PMC   1683299 . PMID   1678250.
  6. 1 2 Grimsby J, Chen K, Wang LJ, Lan NC, Shih JC (May 1991). "Human monoamine oxidase A and B genes exhibit identical exon-intron organization". Proceedings of the National Academy of Sciences of the United States of America. 88 (9): 3637–41. Bibcode:1991PNAS...88.3637G. doi: 10.1073/pnas.88.9.3637 . PMC   51507 . PMID   2023912.
  7. 1 2 "Entrez Gene: MAOA monoamine oxidase A".
  8. 1 2 3 4 5 6 Gupta V, Khan AA, Sasi BK, Mahapatra NR (July 2015). "Molecular mechanism of monoamine oxidase A gene regulation under inflammation and ischemia-like conditions: key roles of the transcription factors GATA2, Sp1 and TBP". Journal of Neurochemistry. 134 (1): 21–38. doi: 10.1111/jnc.13099 . PMID   25810277. S2CID   21044944.
  9. Eccles DA, Macartney-Coxson D, Chambers GK, Lea RA (10 January 2012). "A unique demographic history exists for the MAO-A gene in Polynesians". Journal of Human Genetics . 57 (5): 294–300. doi: 10.1038/jhg.2012.19 . PMID   22377710.
  10. Sabol SZ, Hu S, Hamer D (September 1998). "A functional polymorphism in the monoamine oxidase A gene promoter". Human Genetics. 103 (3): 273–9. doi:10.1007/s004390050816. PMID   9799080. S2CID   29954052.
  11. Wong CC, Caspi A, Williams B, Craig IW, Houts R, Ambler A, et al. (August 2010). "A longitudinal study of epigenetic variation in twins". Epigenetics. 5 (6): 516–26. doi:10.4161/epi.5.6.12226. PMC   3322496 . PMID   20505345.
  12. Jiang Y, Langley B, Lubin FD, Renthal W, Wood MA, Yasui DH, et al. (November 2008). "Epigenetics in the nervous system". The Journal of Neuroscience. 28 (46): 11753–9. doi:10.1523/JNEUROSCI.3797-08.2008. PMC   3844836 . PMID   19005036.
  13. 1 2 Binda C, Mattevi A, Edmondson DE (2011). "Structural properties of human monoamine oxidases a and B". Monoamine Oxidase and their Inhibitors. International Review of Neurobiology. Vol. 100. pp. 1–11. doi:10.1016/B978-0-12-386467-3.00001-7. ISBN   9780123864673. PMID   21971000.
  14. 1 2 Iacovino LG, Magnani F, Binda C (November 2018). "The structure of monoamine oxidases: past, present, and future". Journal of Neural Transmission. 125 (11): 1567–1579. doi:10.1007/s00702-018-1915-z. PMID   30167931. S2CID   52133633.
  15. 1 2 Edmondson DE, Binda C, Mattevi A (August 2007). "Structural insights into the mechanism of amine oxidation by monoamine oxidases A and B". Archives of Biochemistry and Biophysics. 464 (2): 269–276. doi:10.1016/j.abb.2007.05.006. PMC   1993809 . PMID   17573034.
  16. Binda C, Mattevi A, Edmondson DE (July 5, 2002). "Structure-function relationships in flavoenzyme-dependent amine oxidations: A comparison of polyamine oxidase and monoamine oxidase". Journal of Biological Chemistry. 277 (27): 23973–23976. doi: 10.1074/jbc.R200005200 . PMID   12015330.
  17. 1 2 3 Kolla NJ, Bortolato M (June 2020). "The role of monoamine oxidase A in the neurobiology of aggressive, antisocial, and violent behavior: A tale of mice and men". Progress in Neurobiology. 194: 101875. doi:10.1016/j.pneurobio.2020.101875. PMC   7609507 . PMID   32574581.
  18. Pietrangeli P, Mondovì B (January 2004). "Amine oxidases and tumors". Neurotoxicology. 25 (1–2): 317–24. doi:10.1016/S0161-813X(03)00109-8. PMID   14697906.
  19. Huang L, Frampton G, Rao A, Zhang KS, Chen W, Lai JM, et al. (October 2012). "Monoamine oxidase A expression is suppressed in human cholangiocarcinoma via coordinated epigenetic and IL-6-driven events". Laboratory Investigation; A Journal of Technical Methods and Pathology. 92 (10): 1451–60. doi:10.1038/labinvest.2012.110. PMC   3959781 . PMID   22906985.
  20. Cohen IL, Liu X, Lewis ME, Chudley A, Forster-Gibson C, Gonzalez M, et al. (April 2011). "Autism severity is associated with child and maternal MAOA genotypes". Clinical Genetics. 79 (4): 355–62. doi:10.1111/j.1399-0004.2010.01471.x. PMID   20573161. S2CID   24366751.
  21. Belsky J, Beaver KM (May 2011). "Cumulative-genetic plasticity, parenting and adolescent self-regulation". Journal of Child Psychology and Psychiatry, and Allied Disciplines. 52 (5): 619–26. doi:10.1111/j.1469-7610.2010.02327.x. PMC   4357655 . PMID   21039487.
  22. Meyer JH, Ginovart N, Boovariwala A, Sagrati S, Hussey D, Garcia A, et al. (November 2006). "Elevated monoamine oxidase a levels in the brain: an explanation for the monoamine imbalance of major depression". Archives of General Psychiatry. 63 (11): 1209–16. doi: 10.1001/archpsyc.63.11.1209 . PMID   17088501.
  23. Schulze TG, Müller DJ, Krauss H, Scherk H, Ohlraun S, Syagailo YV, et al. (December 2000). "Association between a functional polymorphism in the monoamine oxidase A gene promoter and major depressive disorder". American Journal of Medical Genetics. 96 (6): 801–3. doi:10.1002/1096-8628(20001204)96:6<801::AID-AJMG21>3.0.CO;2-4. PMID   11121185.
  24. Du L, Faludi G, Palkovits M, Sotonyi P, Bakish D, Hrdina PD (July 2002). "High activity-related allele of MAO-A gene associated with depressed suicide in males". NeuroReport. 13 (9): 1195–8. doi:10.1097/00001756-200207020-00025. PMID   12151768. S2CID   19874514.
  25. Du L, Bakish D, Ravindran A, Hrdina PD (September 2004). "MAO-A gene polymorphisms are associated with major depression and sleep disturbance in males". NeuroReport. 15 (13): 2097–101. doi:10.1097/00001756-200409150-00020. PMID   15486489. S2CID   39844598.
  26. Yu YW, Tsai SJ, Hong CJ, Chen TJ, Chen MC, Yang CW (September 2005). "Association study of a monoamine oxidase a gene promoter polymorphism with major depressive disorder and antidepressant response". Neuropsychopharmacology. 30 (9): 1719–23. doi: 10.1038/sj.npp.1300785 . PMID   15956990.
  27. Serretti A, Cristina S, Lilli R, Cusin C, Lattuada E, Lorenzi C, et al. (May 2002). "Family-based association study of 5-HTTLPR, TPH, MAO-A, and DRD4 polymorphisms in mood disorders". American Journal of Medical Genetics. 114 (4): 361–9. doi:10.1002/ajmg.10356. PMID   11992558.
  28. Huang SY, Lin MT, Lin WW, Huang CC, Shy MJ, Lu RB (2009). "Association of monoamine oxidase A (MAOA) polymorphisms and clinical subgroups of major depressive disorders in the Han Chinese population". The World Journal of Biological Psychiatry. 10 (4 Pt 2): 544–51. doi:10.1080/15622970701816506. PMID   19224413. S2CID   30281258.
  29. Leuchter AF, McCracken JT, Hunter AM, Cook IA, Alpert JE (August 2009). "Monoamine oxidase a and catechol-o-methyltransferase functional polymorphisms and the placebo response in major depressive disorder". Journal of Clinical Psychopharmacology. 29 (4): 372–7. doi:10.1097/JCP.0b013e3181ac4aaf. PMID   19593178. S2CID   29200403.
  30. Garcia-Arocena D. "The Genetics of Violent Behavior". The Jackson Laboratory. Retrieved 2021-03-23.
  31. Guo G, Ou XM, Roettger M, Shih JC (May 2008). "The VNTR 2 repeat in MAOA and delinquent behavior in adolescence and young adulthood: associations and MAOA promoter activity". European Journal of Human Genetics. 16 (5): 626–34. doi:10.1038/sj.ejhg.5201999. PMC   2922855 . PMID   18212819.
  32. Guo G, Roettger M, Shih JC (August 2008). "The integration of genetic propensities into social-control models of delinquency and violence among male youths". American Sociological Review. 73 (4): 543–568. doi:10.1177/000312240807300402. S2CID   30271933.
  33. Beaver KM, Wright JP, Boutwell BB, Barnes JC, DeLisi M, Vaughn MG (2012). "Exploring the association between the 2-repeat allele of the MAOA gene promoter polymorphism and psychopathic personality traits, arrests, incarceration, and lifetime antisocial behavior". Personality and Individual Differences. 54 (2): 164–168. doi:10.1016/j.paid.2012.08.014.
  34. Caspi A, McClay J, Moffitt TE, Mill J, Martin J, Craig IW, et al. (August 2002). "Role of genotype in the cycle of violence in maltreated children". Science. 297 (5582): 851–4. Bibcode:2002Sci...297..851C. doi:10.1126/science.1072290. PMID   12161658. S2CID   7882492.
  35. Frazzetto G, Di Lorenzo G, Carola V, Proietti L, Sokolowska E, Siracusano A, et al. (May 2007). "Early trauma and increased risk for physical aggression during adulthood: the moderating role of MAOA genotype". PLOS ONE. 2 (5): e486. Bibcode:2007PLoSO...2..486F. doi: 10.1371/journal.pone.0000486 . PMC   1872046 . PMID   17534436.
  36. Choe DE, Shaw DS, Hyde LW, Forbes EE (September 2014). "Interactions Between Monoamine Oxidase A and Punitive Discipline in African American and Caucasian Men's Antisocial Behavior". Clinical Psychological Science. 2 (5): 591–601. doi:10.1177/2167702613518046. PMC   4802365 . PMID   27014508.
  37. Fergusson DM, Boden JM, Horwood LJ, Miller A, Kennedy MA (February 2012). "Moderating role of the MAOA genotype in antisocial behaviour". The British Journal of Psychiatry. 200 (2): 116–23. doi:10.1192/bjp.bp.111.093328. PMC   3269651 . PMID   22297589.
  38. Sjöberg RL, Ducci F, Barr CS, Newman TK, Dell'osso L, Virkkunen M, Goldman D (January 2008). "A non-additive interaction of a functional MAO-A VNTR and testosterone predicts antisocial behavior". Neuropsychopharmacology. 33 (2): 425–30. doi:10.1038/sj.npp.1301417. PMC   2665792 . PMID   17429405.
  39. Ficks CA, Waldman ID (September 2014). "Candidate genes for aggression and antisocial behavior: a meta-analysis of association studies of the 5HTTLPR and MAOA-uVNTR". Behavior Genetics. 44 (5): 427–44. doi:10.1007/s10519-014-9661-y. PMID   24902785. S2CID   11599122.
  40. Dorfman HM, Meyer-Lindenberg A, Buckholtz JW (2014). "Neurobiological mechanisms for impulsive-aggression: the role of MAOA". Current Topics in Behavioral Neurosciences. 17: 297–313. doi:10.1007/7854_2013_272. ISBN   978-3-662-44280-7. PMID   24470068.
  41. 1 2 Tiihonen J, Rautiainen MR, Ollila HM, Repo-Tiihonen E, Virkkunen M, Palotie A, Pietiläinen O, Kristiansson K, Joukamaa M, Lauerma H, Saarela J, Tyni S, Vartiainen H, Paananen J, Goldman D, Paunio T (June 2015). "Genetic Background of Extreme Violent Behavior". Molecular Psychiatry. 20 (6): 786–92. doi:10.1038/mp.2014.130. PMC   4776744 . PMID   25349169.
  42. Vassos E, Collier DA, Fazel S (April 2014). "Systematic meta-analyses and field synopsis of genetic association studies of violence and aggression". Molecular Psychiatry. 19 (4): 471–7. doi:10.1038/mp.2013.31. PMC   3965568 . PMID   23546171. S2CID   13936647.
  43. Rautiainen MR, Paunio T, Repo-Tiihonen E, Virkkunen M, Ollila HM, Sulkava S, et al. (September 2016). "Genome-wide association study of antisocial personality disorder". Translational Psychiatry. 6 (9): e883. doi:10.1038/tp.2016.155. PMC   5048197 . PMID   27598967.
  44. Zhang-James Y, Fernàndez-Castillo N, Hess JL, Malki K, Glatt SJ, Cormand B, Faraone SV (November 2019). "An integrated analysis of genes and functional pathways for aggression in human and rodent models". Molecular Psychiatry. 24 (11): 1655–1667. doi:10.1038/s41380-018-0068-7. PMC   6274606 . PMID   29858598.
  45. Sullivan PF (May 2007). "Spurious genetic associations". Biological Psychiatry. 61 (10): 1121–6. doi:10.1016/j.biopsych.2006.11.010. PMID   17346679. S2CID   35033987.
  46. Hogenboom M (28 October 2014). "Two genes linked with violent crime". BBC News. Retrieved 2014-11-01.
  47. Gallardo-Pujol D, Andrés-Pueyo A, Maydeu-Olivares A (February 2013). "MAOA genotype, social exclusion and aggression: an experimental test of a gene-environment interaction". Genes, Brain and Behavior. 12 (1): 140–5. doi:10.1111/j.1601-183X.2012.00868.x. PMID   23067570. S2CID   4830611.
  48. McDermott R, Tingley D, Cowden J, Frazzetto G, Johnson DD (February 2009). "Monoamine oxidase A gene (MAOA) predicts behavioral aggression following provocation". Proceedings of the National Academy of Sciences of the United States of America. 106 (7): 2118–23. Bibcode:2009PNAS..106.2118M. doi: 10.1073/pnas.0808376106 . PMC   2650118 . PMID   19168625.
  49. Horgan J (26 April 2011). "Code rage: The "warrior gene" makes me mad! (Whether I have it or not)". Scientific American.
  50. Hovet K (20 February 2018). "Chasing the 'warrior gene' and why it looks like a dud so far". Genetic Literacy Project.
  51. Powledge TM (29 July 2016). "Do the MAOA and CDH13 'human warrior genes' make violent criminals—and what should society do?". Genetic Literacy Project.
  52. "MAOA and CDH13 genes linked to violent crime, but can they explain criminal behavior?". Genetic Literacy Project. 29 October 2014.
  53. Barber N (2010-07-13). "Pity the poor murderer, his genes made him do it". Psychology Today. Blog: "The Human Beast: Why we do what we do". Retrieved 2010-10-17.
  54. Hagerty BB (2010-07-01). "Can Your Genes Make You Murder?". NPR.org. National Public Radio. Retrieved 2010-10-17.
  55. Scurich N, Appelbaum PS (July 2021). "State v. Yepez: Admissibility and Relevance of Behavioral Genetic Evidence in a Criminal Trial". Psychiatric Services. 72 (7): 853–855. doi:10.1176/appi.ps.202100226. PMID   34074149. S2CID   235298342.
  56. McSwiggan S, Elger B, Appelbaum PS (January 1, 2017). "The forensic use of behavioral genetics in criminal proceedings: Case of the MAOA-L genotype". International Journal of Law and Psychiatry. 50: 17–23. doi:10.1016/j.ijlp.2016.09.005. PMC   5250535 . PMID   27823806.
  57. Philibert RA, Gunter TD, Beach SR, Brody GH, Madan A (July 2008). "MAOA methylation is associated with nicotine and alcohol dependence in women". American Journal of Medical Genetics. Part B, Neuropsychiatric Genetics. 147B (5): 565–70. doi:10.1002/ajmg.b.30778. PMC   3685146 . PMID   18454435.
  58. Philibert RA, Wernett P, Plume J, Packer H, Brody GH, Beach SR (July 2011). "Gene environment interactions with a novel variable Monoamine Oxidase A transcriptional enhancer are associated with antisocial personality disorder". Biological Psychology. 87 (3): 366–71. doi:10.1016/j.biopsycho.2011.04.007. PMC   3134149 . PMID   21554924.
  59. Shumay E, Logan J, Volkow ND, Fowler JS (October 2012). "Evidence that the methylation state of the monoamine oxidase A (MAOA) gene predicts brain activity of MAO A enzyme in healthy men". Epigenetics. 7 (10): 1151–1160. doi:10.4161/epi.21976. PMC   3469457 . PMID   22948232.
  60. 1 2 3 4 Scott AL, Bortolato M, Chen K, Shih JC (May 2008). "Novel monoamine oxidase A knock out mice with human-like spontaneous mutation". NeuroReport. 19 (7): 739–43. doi:10.1097/WNR.0b013e3282fd6e88. PMC   3435113 . PMID   18418249.
  61. 1 2 Vishnivetskaya GB, Skrinskaya JA, Seif I, Popova NK (2007). "Effect of MAO A deficiency on different kinds of aggression and social investigation in mice". Aggressive Behavior. 33 (1): 1–6. doi: 10.1002/ab.20161 . PMID   17441000.
  62. 1 2 Brunner HG, Nelen M, Breakefield XO, Ropers HH, van Oost BA (October 1993). "Abnormal behavior associated with a point mutation in the structural gene for monoamine oxidase A". Science. 262 (5133): 578–80. Bibcode:1993Sci...262..578B. doi:10.1126/science.8211186. PMID   8211186.
  63. Vishnivetskaya GB, Skrinskaya JA, Seif I, Popova NK (1 January 2007). "Effect of MAO A deficiency on different kinds of aggression and social investigation in mice". Aggressive Behavior. 33 (1): 1–6. doi: 10.1002/ab.20161 . PMID   17441000.
  64. Hebebrand J, Klug B (September 1995). "Specification of the phenotype required for men with monoamine oxidase type A deficiency". Human Genetics. 96 (3): 372–6. doi:10.1007/BF00210430. PMID   7649563. S2CID   33294633.
  65. Wu JB, Shih JC (October 2011). "Valproic acid induces monoamine oxidase A via Akt/forkhead box O1 activation". Molecular Pharmacology. 80 (4): 714–23. doi:10.1124/mol.111.072744. PMC   3187529 . PMID   21775495.
  66. Lee SA, Hong SS, Han XH, Hwang JS, Oh GJ, Lee KS, et al. (July 2005). "Piperine from the fruits of Piper longum with inhibitory effect on monoamine oxidase and antidepressant-like activity". Chemical & Pharmaceutical Bulletin. 53 (7): 832–5. doi: 10.1248/cpb.53.832 . PMID   15997146.
  67. van Diermen D, Marston A, Bravo J, Reist M, Carrupt PA, Hostettmann K (March 2009). "Monoamine oxidase inhibition by Rhodiola rosea L. roots". Journal of Ethnopharmacology. 122 (2): 397–401. doi:10.1016/j.jep.2009.01.007. PMID   19168123.

Further reading