Skip to main content
Canadian Journal of Psychiatry. Revue Canadienne de Psychiatrie logoLink to Canadian Journal of Psychiatry. Revue Canadienne de Psychiatrie
. 2016 Feb 12;61(3):159–165. doi: 10.1177/0706743716632517

Using Evolutionary Theory to Guide Mental Health Research

Utiliser la théorie évolutionniste pour guider la recherche en santé mentale

Zachary Durisko 1,2,, Benoit H Mulsant 1,3, Kwame McKenzie 1,3,4, Paul W Andrews 2
PMCID: PMC4813423  PMID: 27254091

Abstract

Evolutionary approaches to medicine can shed light on the origins and etiology of disease. Such an approach may be especially useful in psychiatry, which frequently addresses conditions with heterogeneous presentation and unknown causes. We review several previous applications of evolutionary theory that highlight the ways in which psychiatric conditions may persist despite and because of natural selection. One lesson from the evolutionary approach is that some conditions currently classified as disorders (because they cause distress and impairment) may actually be caused by functioning adaptations operating “normally” (as designed by natural selection). Such conditions suggest an alternative illness model that may generate alternative intervention strategies. Thus, the evolutionary approach suggests that psychiatry should sometimes think differently about distress and impairment. The complexity of the human brain, including normal functioning and potential for dysfunctions, has developed over evolutionary time and has been shaped by natural selection. Understanding the evolutionary origins of psychiatric conditions is therefore a crucial component to a complete understanding of etiology.

Keywords: evolution, evolutionary psychiatry, evolutionary medicine

Clinical Implications

  • Functioning adaptations can cause distress and meet current criteria for “disorder.”

  • Functioning adaptations that lead to psychiatric conditions may require alternative treatments.

  • Evolution is critical for a complete understanding of any condition’s etiology.

Limitations

  • Many researchers and clinicians have not been trained in evolutionary theory.

  • Most evolutionary hypotheses in mental health have not yet been rigorously tested.

Evolutionary theory is the most fundamental organizing principle of biology and can inform research across all areas of health.13 An evolutionary understanding of why we get sick is a framework that generates explicit, testable predictions to better direct mechanistic research regarding how we get sick. This approach has improved our understanding of many conditions (e.g., infection, cancer, morning sickness, pain, fever, genetic disorders2,3) and may be especially useful for psychiatry because the precise etiology of mental illness is often unclear.410 Indeed, because the brain is a complex integration of adaptations designed by natural selection,11 the evolutionary perspective is essential to completely understand the origin and etiology of mental illness.

Most researchers and clinicians accept that natural selection has played some role in shaping “normal” behaviour, but they are not trained to study abnormal behaviour from an evolutionary perspective.12 This article briefly summarizes evolutionary accounts of psychiatric disorders as they are currently defined. Examples are typically theoretical and remain to be rigorously tested. We then discuss the implications of the evolutionary perspective for the conceptualization of psychiatric disorder generally.

Why Do Mental Disorders Exist and Persist?

The etiology of mental disorders can be addressed at multiple interacting levels.13 Ecological and social factors certainly play a causal role, and many disorders can be triggered solely by environmental stressors or injuries. Nevertheless, estimates suggest a moderate to high degree of heritable risk (e.g., 90% of trait variation for autism can be accounted for with genetics; bipolar disorder, 85%; schizophrenia, 81%; unipolar depression, 37%1416). This indicates strong genetic components underlying vulnerability to mental disorder. Over evolutionary time, heritable risk would have been responsive to pressures of natural selection removing alleles associated with disorder from the population. In addition, age of onset is often prior to or during peak reproductive years,16 further increasing fitness costs associated with many disorders (e.g., reduced number of offspring). These facts have led some to question the so-called paradox of psychiatric disorders1620: given that these disorders are considered maladaptive and have heritable risk, why do they still exist? Explanations fall into 2 categories: despite and because of natural selection (Table 1).

Table 1.

Why do mental disorders exist and persist?a

Despite natural Selection
Mutation-selection balance Disorder-causing mutations arise faster than selection can remove them from the population. A high rate of mutation may be due to inherent trait complexity or other environmental and biological factors (e.g., paternal age).
Examples: autism21,22 or developmental disorders,23 some forms of schizophrenia,24,25 bipolar disorder26
Ancestral neutrality Alleles that were neither favoured nor disfavoured by natural selection interact with modern environments to cause disorder.
There may be no convincing examples.17
Because of natural selection
Balancing selectionb
 Antagonistic pleiotropy The same allele increases risk of disorder and improves fitness with a different trait. Selection for this secondary trait maintains the maladaptive disorder risk in the population.
Examples: some forms of schizophrenia27-29
 Stabilizing selection on continuous traits A small number of risk alleles and a moderate degree of a particular trait are beneficial, but extremes lead to disorder.
Examples: anxiety and depression,30-32 bipolar disorder33
 Alternating selection Alleles that increase risk for disorder have evolved because they also improved fitness in some environments, in only one sex, or when at a particular frequency/density in the population. They are maladaptive in some contexts and adaptive in others.
Examples: psychopathy34-36
Environmental mismatch The trait improved ancestral fitness (i.e., is an adaptation) but interacts negatively with some aspect of the modern environment.
Examples: some addictions and substance use disorders,37,38 bipolar disorder39
Functioning adaptations The trait is an adaptation that is now culturally disfavoured. Some conditions may be classified as ‘‘disorder’’ despite being adaptive and lacking any malfunction in the brain.
Examples: some instances of depressive disorders,40-42 tobacco use and nicotine addiction43,44

a This table summarizes some of the commonly used evolutionary explanations for why mental disorders exist. These explanations are not mutually exclusive, and there may be multiple mechanisms maintaining some disorders in the population. We have provided some potential examples, but note that these are hypotheses only, and although some are promising, most have not yet been convincingly tested.

b Disorders maintained as by-products of selection.

Disorder despite Natural Selection

Heritable risk for mental disorders may persist in populations despite the pressures of natural selection. First, mental and behavioural traits are complex and polygenic. With so many potential sources of error, new deleterious mutations may arise as fast or faster than natural selection can remove them (mutation-selection balance).1719,45 A greater mutation load, for example, because of inbreeding,46 maternal famine,47 or older paternal age,48 may increase an individual’s risk of schizophrenia,24,25,46 autism,21,22 other developmental disorders,23 and bipolar disorder.26

Second, preexisting genetic variation may have only recently have started contributing to disorder. Ancestral neutrality explanations propose that modern humans exist in environments critically different from those in which we evolved. In this view, ancient alleles interact with novel environmental factors to cause disorder, and too few generations have passed for natural selection to remove these alleles from the population.17,49 Theorists may cite such changes as modern diets, pollution, or decreased social connectedness. This explanation is congruent with the greater prevalence of some disorders in industrialized nations50 and among individuals using evolutionarily novel substances (e.g., cannabis and psychosis51). However, ancestral neutrality may be unlikely to solely account for the high heritability and prevalence of most disorders.17 Because successive generations consist of only a sample of prior alleles, over time random genetic drift will tend to fix or eliminate neutral alleles from the population, reducing neutral variation. Furthermore, modern fitness costs are typically great enough to remove risk alleles from the population in only a few dozen generations.17,20 Thus, if a disorder had originated because of ancestral neutrality alone, its prevalence should now be declining rapidly.

Disorder because of Natural Selection

More nuanced applications of evolutionary theory suggest many non–mutually exclusive ways that heritable risk for disorder can persist because of natural selection1618 (Table 1). A common misconception is that natural selection leads to perfectly designed traits.12 In fact, natural selection responds to trade-offs to find the best compromise for the propagation of genetic material. Selection for one trait can maintain risk for disorder as a by-product. Explanations of this kind fall under balancing selection, named for the balancing of positive and negative selective forces that maintains allelic variation. One mechanism is antagonistic pleiotropy, when a single allele increases susceptibility to disorder but also confers some benefit. Positive selection for the beneficial trait cancels out negative selection against increased disorder risk. Another mechanism of balancing selection is heterozygote advantage. In this case, individuals with one risk allele and one healthy allele have greater fitness than both homozygotes. Commonly cited examples of single-gene heterozygote advantage are sickle-cell anemia52 and cystic fibrosis,53 wherein heterozygous individuals may benefit from increased resistance to malaria and tuberculosis, respectively. For polygenic psychological traits, a moderate number of risk alleles may increase fitness while too many leads to disorder. Anxiety, depression, and bipolar disorders may all be examples in which trait extremes are maladaptive yet intermediate phenotypes are beneficial.3033 Genetic evidence suggests that there may also have been benefits to alleles that increase risk for schizophrenia,27 such as higher IQ, increased creativity, and improved mathematical reasoning.28,29 Future research should rigorously test these ideas.

Fitness benefits need not exist at all times or for all individuals. Under alternating selection, natural selection vacillates between favouring and disfavouring an allele. A special case is sexual antagonism, in which alleles confer fitness benefits in one sex but costs in the other. Some possible evidence for this exists for schizophrenia and autism, in which estimated fitness costs are greater among affected males than females, and sisters of affected individuals may have increased fitness.20 Other cases of alternating selection are frequency- or density-dependent selection, in which the strength and direction of selection depend on the relative frequency of the allele or population density. For example, some personality traits may be alternative ecological strategies that conferred fitness benefits depending on the environment and strategies expressed by others.9,34,54,55 For example, antisociality and psychopathy may represent biases toward selfishness that pay off when most others are eager to cooperate.54,34,56,57 If true, natural selection would maintain heritable variation such that a small proportion of the population expresses such traits. This perspective has generated some debate,35,58 although empirical research has largely supported the hypothesis that psychopathy is an adaptation.36,59

These examples highlight the fact that some current disorders may be the direct result of evolved adaptations interacting negatively with modern environments: environmental mismatches between modern environments and the environments for which we are designed. It has been suggested that bipolar disorder may be one example.39 The reward system of the brain is an adaptation that may be particularly susceptible to environmental mismatches, leading to addictions and substance use disorders.37,38 Furthermore, research suggests that humans may have evolved to exploit neurotoxic properties of some plant secondary compounds (e.g., nicotine) to fight parasites such as helminthic worms.43,44,60 Although it remains to be seen whether similar adaptationist accounts apply to the use of other plant-derived substances, the heritable genetic variation underlying such substance use would have been directly maintained by natural selection because of its historical adaptive value. Indeed, the “normal” functioning (as designed by natural selection) of some adaptations may still be adaptive in modern environments but nevertheless classified as “disorder” because it causes distress or is culturally disfavoured. For example, anxiety may prevent accidental death in early life,61 and depression may be due to adaptive mechanisms that reduce interest in otherwise pleasurable pursuits to conserve energy, signal social defeat, or promote focus on solving complex problems.4042,62 Indeed, controlling for comorbidities, depressive disorders may increase fitness in women.20 The implications of this perspective to our current understanding of disorder are the subject of the next section.

Psychiatric Nosology and an Evolutionary Understanding of Disorder

Some conditions or behavioural syndromes currently classified as mental disorders may have originated as adaptations (i.e., traits that historically served a particular function to increase fitness) and thus may occur as the result of “normal” functioning rather than any biological malfunction (e.g., mutation, developmental aberration, or failure of underlying mechanism). In the Diagnostic and Statistical Manual of Mental Disorders (DSM) system (III, IV, and 5), distress and impairment are required to define disorders that warrant intervention. The DSM-5 defines mental disorder as “dysfunction in the psychological, biological, or developmental processes underlying mental functioning.”63(p20) and operationalizes this dysfunction with proxies such as the inability to work, maintain interpersonal relations, and take care of one’s self. Several authors have pointed out the difficulty this approach can have in distinguishing between normal and abnormal behaviour.5,6467

Distress, impairment, and inability to function in everyday life are not necessarily indicative of biological malfunction.2,62,64,6870 The normal functioning of the body’s evolved systems can at times be unpleasant and cause suffering. Examples include feeling nauseated or having diarrhea after ingesting rotten food, running a fever when infected, and the physical pain associated with a broken limb or childbirth. The inappropriate or indiscriminate disruption of these adaptations in an attempt to relieve suffering can have negative consequences.7173 Conversely, the absence of distress can indicate disorder. For example, individuals with a congenital inability to feel pain are much more likely to suffer injury and early death.74 These insights suggest that distress is a faulty criterion, and the DSM system must therefore erroneously categorize some “normal” states as disorders and miss some instances of psychological malfunction that are not distressing.

Incorporating an evolutionary perspective may clarify the distinction between normal and disordered and improve psychiatric nosology.5,62,68,69,75 Jerome Wakefield75,76 has proposed that we consider as disorders only such conditions caused by “harmful dysfunctions:”

A condition is a disorder if and only if (a) the condition causes some harm or deprivation of benefit to the person [or others] as judged by the standards of the person’s culture (the value criterion), and (b) the condition results from the inability of some internal mechanism to perform its natural function, wherein a natural function is an effect that is part of the evolutionary explanation of the existence and structure of the mechanism.76(p384)

This definition breaks the reliance on distress inherent in the DSM approach as long as one is able to identify the “natural” function driving natural selection (i.e., identify adaptations). Identifying adaptations can be a complex and onerous process. It is impossible to directly observe the selective pressures that have led to a trait, and so the task requires historical inference. The best approach uses a thorough reverse engineering of trait design.77,78 Natural selection is the only known mechanism that can generate nonrandom biological organization. Therefore, the natural function(s) of a trait are those that can account for the trait’s complexity at all levels of analysis. For example, we conclude with confidence that the eye is an adaptation for vision because this is the only function that can account for the complexity and organization of the eye’s constituent parts.78 To demonstrate that a behavioural syndrome is an adaptation for a particular function, we must assess whether that function can account for the syndrome’s underlying complexity and organization. This requires understanding the components and mechanisms of the syndrome in increasing detail.

Applying this approach to human behaviour has been criticized for relying too heavily on speculations of adaptation (i.e., “just-so” stories).66,79 It is true that accurately identifying adaptations may require decades of research, and hypotheses may abound in the interim. However, hypotheses guide research, and such a framework allows the distinction between conditions caused by biological malfunction (true “disorder,” errors of mechanism) and those due to undesirable or out-of-context adaptations. For each supposed instance of disorder, we must propose the normal function that has gone awry. Until known and convincingly tested, the classification of a condition as disorder or nondisorder should remain tentative. This is more than semantics. Knowing whether a harmful condition is caused by the natural functioning of an adaptation or the breakdown of an adaptation leads to different models of illness, different directions of research, and, most importantly, different approaches to treatment. For the example of fever, there are very different treatments for a fever caused by the normal functioning of the immune system in response to infection and for a fever caused by hypothalamic tumor. The current approach to psychiatric nosology does not differentiate between “functional” and “dysfunctional” routes to distress and impairment.62 The incorrect assumption of brain malfunction may lead to improper and ineffective treatments that do more harm than good.80,81

As research from this perspective progresses, it leads to an ever-increasing understanding of the brain’s many adaptations. A list of brain adaptations is, in effect, a list of ways the brain may malfunction. This can improve disorder nosology by categorizing conditions according to etiology (which adaptation has malfunctioned) instead of behavioural symptoms. There may be many pathways to the same behavioural syndrome, and clustering categories of disorder based solely on observable symptoms combines cases with unique etiologies that ought to be treated by different means. Alternatively, a single underlying malfunction may present with different symptoms that can be erroneously classified as comorbid disorders. Currently, comorbidity is common. In one study, 61.8% of patients receiving a diagnosis of depressive disorder were also diagnosed with at least one other comorbid psychiatric disorder (e.g., anxiety, posttraumatic stress disorder, eating disorders, or obsessive-compulsive disorder).82 Frequent overlap between conditions may suggest that the two can in fact be caused by one underlying mechanism.5 Understanding comorbidity and parsing apart different conditions according to etiology will require more fully understanding the underlying adaptations that have gone awry.

Conclusion

Natural selection is the only scientific explanation for the origin of the brains complexity. Knowing the functions for which the brain is designed allows us to better understand whether and how mechanisms have broken down. Table 1 describes ways in which heritable risk for mental disorders (as they are currently defined) may arise and persist. Importantly, functioning adaptations can cause distress and meet current diagnostic criteria for disorder. This suggests that psychiatric nosology should sometimes think differently about distress and impairment. All distressing conditions should be treated, regardless of cause. Even the normal operation of adaptations can require medical intervention (e.g., giving birth, wisdom teeth). However, different etiological pathways require different treatment interventions. The evolutionary approach requires identifying whether a distressing condition is caused by the malfunction of an adaptation and, if so, identifying the particular malfunctioning adaptation. For functioning adaptations, more evolutionary-minded research is required to identify the best treatment for each condition given its functional etiology. For example, some conditions may respond best to alternative treatments that support rather than abolish the adaptive function (e.g., depressive symptoms brought on by an adaptation designed to promote analytical thinking may be best treated by encouraging and promoting analysis instead of abolishing rumination62). Evolutionary approaches to psychiatry are still new and do not have all the answers, but this perspective can generate hypotheses and guide research, inform treatment strategies, and shed light on the etiology of mental illness and suffering.

Acknowledgements

We thank Dr. Skye Barbic and 3 anonymous reviewers for comments on the manuscript.

Author Note: Dr. Durisko wrote the first draft of this article, and Dr. Mulsant, Dr. Andrews, and Dr. McKenzie provided critical revisions.

Declaration of Conflicting Interests: The authors declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Funding: The authors received no financial support for the research, authorship, and/or publication of this article.

References

  • 1. Williams G, Nesse R. The dawn of Darwinian medicine. Q Rev Biol. 1991;66:1–22. [DOI] [PubMed] [Google Scholar]
  • 2. Nesse RM, Williams GC. Why we get sick: the new science of Darwinian medicine. New York: Vintage; 1996. [Google Scholar]
  • 3. Gluckman P, Beedle A, Hanson M. Principles of evolutionary medicine. Oxford (UK): Oxford University Press; 2009. [Google Scholar]
  • 4. Nesse RM, Stearns SC. The great opportunity: evolutionary applications to medicine and public health. Evol Appl. 2008;1:28–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Nesse RM, Stein DJ. Towards a genuinely medical model for psychiatric nosology. BMC Med. 2012;10:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Kennair L. Evolutionary psychology and psychopathology. Curr Opin Psychiatry. 2003;16:691–699. [Google Scholar]
  • 7. Brüne M. Textbook of evolutionary psychiatry: the origins of psychopathology. New York: Oxford University Press; 2008. [Google Scholar]
  • 8. Stevens A, Price J. Evolutionary psychiatry: a new beginning. 2nd ed New York: Routledge; 2000. [Google Scholar]
  • 9. Del Giudice M. An evolutionary life history framework for psychopathology. Psychol Inq. 2014;25:261–300. [Google Scholar]
  • 10. Crespi B. An evolutionary framework for psychological maladaptations. Psychol Inq. 2014;25:322–324. [Google Scholar]
  • 11. Barkow J, Cosmides L, Tooby J. The adapted mind: evolutionary psychology and the generation of culture. Oxford (UK): Oxford University Press; 2003. [Google Scholar]
  • 12. Brüne M, Hochberg Z. Evolutionary medicine: the quest for a better understanding of health, disease and prevention. BMC Med. 2013;11:116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Shah J, Mizrahi R, McKenzie K. The four dimensions: a model for the social aetiology of psychosis. Br J Psychiatry. 2011;199:11–14. [DOI] [PubMed] [Google Scholar]
  • 14. Kendler K. Twin studies of psychiatric illness: an update. Arch Gen Psychiatry. 2001;58:1005–1014. [DOI] [PubMed] [Google Scholar]
  • 15. Cardno AG, Marshall EJ, Coid B, et al. Heritability estimates for psychotic disorders. Arch Gen Psychiatry. 1999;56:162–168. [DOI] [PubMed] [Google Scholar]
  • 16. Uher R. The role of genetic variation in the causation of mental illness: an evolution-informed framework. Mol Psychiatry. 2009;14:1072–1082. [DOI] [PubMed] [Google Scholar]
  • 17. Keller MC, Miller G. Resolving the paradox of common, harmful, heritable mental disorders: which evolutionary genetic models work best? Behav Brain Sci. 2006;29:385–404. [DOI] [PubMed] [Google Scholar]
  • 18. Keller MC. The evolutionary persistence of genes that increase mental disorders risk. Curr Dir Psychol Sci. 2008;17:395–399. [Google Scholar]
  • 19. Gangestad S, Yeo R. Behavioral genetic variation, adaptation and maladaptation: an evolutionary perspective. Trends Cogn Sci. 1997;1:103–108. [DOI] [PubMed] [Google Scholar]
  • 20. Power RA, Kyaga S, Uher R, et al. Fecundity of patients with schizophrenia, autism, bipolar disorder, depression, anorexia nervosa, or substance abuse vs their unaffected siblings. JAMA Psychiatry. 2013;70:22–30. [DOI] [PubMed] [Google Scholar]
  • 21. Durkin MS, Maenner MJ, Newschaffer CJ, et al. Advanced parental age and the risk of autism spectrum disorder. Am J Epidemiol. 2008;168:1268–1276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Reichenberg A, Gross R, Weiser M, et al. Advancing paternal age and autism. Arch Gen Psychiatry. 2006;63:1026–1032. [DOI] [PubMed] [Google Scholar]
  • 23. Vissers LELM, de Ligt J, Gilissen C, et al. A de novo paradigm for mental retardation. Nat Genet. 2010;42:1109–1112. [DOI] [PubMed] [Google Scholar]
  • 24. Rees E, Kirov G, O’Donovan MC, Owen MJ. De novo mutation in schizophrenia. Schizophr Bull. 2012;38:377–381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Xu B, Roos JL, Levy S, et al. Strong association of de novo copy number mutations with sporadic schizophrenia. Nat Genet. 2008;40:880–885. [DOI] [PubMed] [Google Scholar]
  • 26. Frans EM, Sandin S, Reichenberg A, et al. Advancing paternal age and bipolar disorder. Arch Gen Psychiatry. 2008;65:1034–1040. [DOI] [PubMed] [Google Scholar]
  • 27. Crespi B, Summers K, Dorus S. Adaptive evolution of genes underlying schizophrenia. Proc Biol Sci. 2007;274:2801–2810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Kyaga S, Landén M, Boman M, et al. Mental illness, suicide and creativity: 40-year prospective total population study. J Psychiatr Res. 2013;47:83–90. [DOI] [PubMed] [Google Scholar]
  • 29. Karlsson JL. Psychosis and academic performance. Br J Psychiatry. 2004;184:327–329. [DOI] [PubMed] [Google Scholar]
  • 30. Marks I, Nesse R. Fear and fitness: an evolutionary analysis of anxiety disorders. Ethol Sociobiol. 1994;15:247–261. [Google Scholar]
  • 31. Nesse RM. Proximate and evolutionary studies of anxiety, stress and depression: synergy at the interface. Neurosci Biobehav Rev. 1999;23:895–903. [DOI] [PubMed] [Google Scholar]
  • 32. Nettle D. Evolutionary origins of depression: a review and reformulation. J Affect Disord. 2004;81:91–102. [DOI] [PubMed] [Google Scholar]
  • 33. Akiskal KK, Akiskal HS. The theoretical underpinnings of affective temperaments: implications for evolutionary foundations of bipolar disorder and human nature. J Affect Disord. 2005;85:231–239. [DOI] [PubMed] [Google Scholar]
  • 34. Jonason PK, Duineveld JJ, Middleton JP. Pathology, pseudopathology, and the Dark Triad of personality. Pers Individ Dif. 2015;78:43–47. [Google Scholar]
  • 35. Krupp DB, Sewall LA, Lalumière ML, Sheriff C, Harris GT. Psychopathy, adaptation, and disorder. Front Psychol. 2013;4:139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Krupp DB, Sewall LA, Lalumière ML, Sheriff C, Harris GT. Nepotistic patterns of violent psychopathy: evidence for adaptation? Front Psychol. 2012;3:305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Pani L. Is there an evolutionary mismatch between the normal physiology of the human dopaminergic system and current environmental conditions in industrialized countries? Mol Psychiatry. 2000;5:467–475. [DOI] [PubMed] [Google Scholar]
  • 38. Nesse RM, Berridge KC. Psychoactive drug use in evolutionary perspective. Science. 1997;278:63–66. [DOI] [PubMed] [Google Scholar]
  • 39. Wilson DR. Evolutionary epidemiology and manic depression. Br J Med Psychol. 1998;71:375–395. [DOI] [PubMed] [Google Scholar]
  • 40. Andrews PW, Thomson JA. The bright side of being blue: depression as an adaptation for analyzing complex problems. Psychol Rev. 2009;116:620–654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Nesse RM. Is depression an adaptation? Arch Gen Psychiatry. 2000;57:14–20. [DOI] [PubMed] [Google Scholar]
  • 42. Hagen EH. Evolutionary theories of depression: a critical review. Can J Psychiatry. 2011;56:716–726. [DOI] [PubMed] [Google Scholar]
  • 43. Hagen EH, Roulette CJ, Sullivan RJ. Explaining human recreational use of ‘pesticides’: the neurotoxin regulation model of substance use vs. the hijack model and implications for age and sex differences in drug consumption. Front Psychiatry. 2013;4:142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Roulette CJ, Mann H, Kemp BM, et al. Tobacco use vs. helminths in Congo basin hunter-gatherers: self-medication in humans? Evol Hum Behav 2014;35:397–407. [Google Scholar]
  • 45. Arnheim N, Calabrese P. Understanding what determines the frequency and pattern of human germline mutations. Nat Rev Genet. 2009;10:478–488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Keller MC, Simonson MA, Ripke S, et al. Runs of homozygosity implicate autozygosity as a schizophrenia risk factor. PLoS Genet. 2012;8:1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. McClellan J, Susser E, King M. Maternal famine, de novo mutations, and schizophrenia. JAMA. 2006;296:582–584. [DOI] [PubMed] [Google Scholar]
  • 48. Kong A, Frigge ML, Masson G, et al. Rate of de novo mutations and the importance of father’s age to disease risk. Nature. 2012;488:471–475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Di Rienzo A, Hudson RR. An evolutionary framework for common diseases: the ancestral-susceptibility model. Trends Genet. 2005;21:596–601. [DOI] [PubMed] [Google Scholar]
  • 50. Demyttenaere K, Bruffaerts R, Posada-Villa J, et al. Prevalence, severity, and unmet need for treatment of mental disorders in the World Health Organization World Mental Health Surveys. JAMA. 2004;291:2581–2590. [DOI] [PubMed] [Google Scholar]
  • 51. Wilkinson ST, Radhakrishnan R, D’Souza DC. Impact of cannabis use on the development of psychotic disorders. Curr Addict Rep. 2014;1:115–128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Aidoo M, Terlouw DJ, Kolczak MS, et al. Protective effects of the sickle cell gene against malaria morbidity and mortality. Lancet. 2002;359:1311–1312. [DOI] [PubMed] [Google Scholar]
  • 53. Poolman EM, Galvani AP. Evaluating candidate agents of selective pressure for cystic fibrosis. J R Soc Interface. 2007;4:91–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Troisi A. The concept of alternative strategies and its relevance to psychiatry and clinical psychology. Neurosci Biobehav Rev. 2005;29:159–168. [DOI] [PubMed] [Google Scholar]
  • 55. Jonason PK, Baughman HM, Carter GL, Parker P. Dorian Gray without his portrait: psychological, social, and physical health costs associated with the Dark Triad. Pers Individ Dif. 2015;78:5–13. [Google Scholar]
  • 56. Colman A, Wilson J. Antisocial personality disorder: an evolutionary game theory analysis. Leg Criminol Psychol. 1997;2:23–34. [Google Scholar]
  • 57. Mealey L. The sociobiology of sociopathy: an integrated evolutionary model. Behav Brain Sci. 1995;18:523–599. [Google Scholar]
  • 58. Leedom LJ, Almas LH. Is psychopathy a disorder or an adaptation? Front Psychol. 2012;3:549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Gervais MM, Kline M, Ludmer M, et al. The strategy of psychopathy: primary psychopathic traits predict defection on low-value relationships. Proc Biol Sci. 2013;280:20122773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Sullivan RJ, Hagen EH, Hammerstein P. Revealing the paradox of drug reward in human evolution. Proc Biol Sci. 2008;275:1231–1241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Lee WE, Wadsworth MEJ, Hotopf M. The protective role of trait anxiety: a longitudinal cohort study. Psychol Med. 2006;36:345–351. [DOI] [PubMed] [Google Scholar]
  • 62. Durisko Z, Mulsant B, Andrews P. An adaptationist perspective on the etiology of depression. J Affect Disord. 2015;172:315–323. [DOI] [PubMed] [Google Scholar]
  • 63. American Psychiatric Association. Diagnostic and statistical manual of mental disorders (5th ed.). Washington (DC; ): American Psychiatric Association; 2013. [Google Scholar]
  • 64. Spitzer RL. Foreword In: The loss of sadness: how psychiatry transformed normal sorrow into depressive disorder. New York: Oxford University Press; 2007. p vi–x. [DOI] [PubMed] [Google Scholar]
  • 65. Frances A. The new crisis of confidence in psychiatric diagnosis. Ann Intern Med. 2013;159:720. [DOI] [PubMed] [Google Scholar]
  • 66. Stein DJ, Phillips KA, Bolton D, et al. What is a mental/psychiatric disorder? From DSM-IV to DSM-V. Psychol Med. 2010;40:1759–1765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Pilgrim D. Defining mental disorder: tautology in the service of sanity in British mental health legislation. J Ment Heal. 2005;14:435–443. [Google Scholar]
  • 68. Spitzer RL, Wakefield JC. DSM-IV diagnostic criterion for clinical significance: does it help solve the false positives problem? Am J Psychiatry. 1999;156:1856–1864. [DOI] [PubMed] [Google Scholar]
  • 69. Nesse R, Jackson E. Evolution: psychiatric nosology’s missing biological foundation. Clin Neuropsychiatry. 2006;3:121–131. [Google Scholar]
  • 70. Wakefield J. Disorder as harmful dysfunction: a conceptual critique of DSM-III-R’s definition of mental disorder. Psychol Rev. 1992;99:232–247. [DOI] [PubMed] [Google Scholar]
  • 71. Earn D, Andrews P, Bolker B. Population-level effects of suppressing fever. Proc R Soc London Ser B Biol Sci. 2014;281:20132571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Brandts C, Ndjave M, Graninger W, et al. Effect of paracetamol on parasite clearance time in Plasmodium falciparum malaria. Lancet. 1997;350:704–709. [DOI] [PubMed] [Google Scholar]
  • 73. Van Esch RW, Kool MM, van As S. NSAIDs can have adverse effects on bone healing. Med Hypotheses. 2013;81:343–346. [DOI] [PubMed] [Google Scholar]
  • 74. Cox JJ, Reimann F, Nicholas AK, et al. An SCN9A channelopathy causes congenital inability to experience pain. Nature. 2006;444:894–898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Wakefield JC. Evolutionary versus prototype analyses of the concept of disorder. J Abnorm Psychol. 1999;108:374–399. [DOI] [PubMed] [Google Scholar]
  • 76. Wakefield JC. The concept of mental disorder: on the boundary between biological facts and social values. Am Psychol. 1992;47:373–388. [DOI] [PubMed] [Google Scholar]
  • 77. Andrews PW, Gangestad SW, Matthews D. Adaptationism—how to carry out an exaptationist program. Behav Brain Sci. 2002;25:489–504. [DOI] [PubMed] [Google Scholar]
  • 78. Williams GC. Adaptation and natural selection: a critique of some current evolutionary thought. Princeton (NJ): Princeton University Press; 1966. [Google Scholar]
  • 79. Valles SA. Evolutionary medicine at twenty: rethinking adaptationism and disease. Biol Philos. 2012;27:241–261. [Google Scholar]
  • 80. Andrews PW, Thomson JA, Amstadter A, Neale MC. Primum non nocere: an evolutionary analysis of whether antidepressants do more harm than good. Front Psychol. 2012;3:117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. Ghaemi S. Toward a Hippocratic psychopharmacology. Can J Psychiatry. 2008;53:189–196. [DOI] [PubMed] [Google Scholar]
  • 82. Rush AJ, Zimmerman M, Wisniewski SR, et al. Comorbid psychiatric disorders in depressed outpatients: demographic and clinical features. J Affect Disord. 2005;87:43–55. [DOI] [PubMed] [Google Scholar]

Articles from Canadian Journal of Psychiatry. Revue Canadienne de Psychiatrie are provided here courtesy of SAGE Publications

RESOURCES