ChapterPDF Available

The Evolutionary Etiologies of Autism Spectrum and Psychotic Affective Spectrum Disorders

Authors:

Abstract

Risks of human psychiatric conditions have evolved, and their symptoms represent perturbations to adaptive cognitive and affective systems. Evolutionary considerations are useful in this context because they direct us to the identification of specific human adaptations that become dysregulated in disease, though either underdevelopment or overdevelopment. Autism is thus conceptualized in terms of underdeveloped social cognition, such that the highly elaborated human social brain does not complete its usual developmental trajectory. Psychotic affective conditions, mainly schizophrenia, bipolar disorder, and depression, are, in contrast to autism, conceptualized in terms of dysfunctionally overdeveloped aspects of social cognition, such that they are caused by opposite neural system alterations to those producing autism. The hypothesis that autism and psychotic affective conditions represent diametric disorders is supported by a wide range of convergent evidence from genetics, development, neuroscience, psychology, and cognitive science. The diametric model provides for reciprocal illumination of the causes of these conditions and makes specific recommendations for research strategies and the development of novel treatments.
"Evolutionary thinking in medicine: from research to policy and practice", Oxford University Press,
edited by A. Alvergne, C. Jenkinson and C. Faurie. April 2016.
The Evolutionary Etiologies of Autism Spectrum and Psychotic-Affective Spectrum Disorders
Bernard J Crespi
Department of Biological Sciences
Simon Fraser University
Burnaby British Columbia, Canada V5A 1S6
crespi@sfu.ca
Abstract
Risks of human psychiatric conditions have evolved, and their symptoms represent perturbations to
adaptive cognitive and affective systems. Evolutionary considerations are useful in this context
because they direct us to the identification of specific human adaptations that become dysregulated
in disease, though either underdevelopment or overdevelopment. Autism is thus conceptualized in
terms of underdeveloped social cognition, such that the highly-elaborated human social brain does
not complete its usual developmental trajectory. Psychotic-affective conditions, mainly schizophrenia,
bipolar disorder, and depression, are, in contrast to autism, conceptualized in terms of dysfunctionally
over-developed aspects of social cognition, such that they are caused by opposite neural-system
alterations to those producing autism. The hypothesis that autism and psychotic-affective conditions
represent diametric disorders is supported by a wide range of convergent evidence from genetics,
development, neuroscience, psychology, and cognitive science. The diametric model provides for
reciprocal illumination of the causes of these conditions, and makes specific recommendations for
research strategies and the development of novel treatments.
Lay Summary
The mental and behavioral traits that have evolved in humans, such as complex sociality, make us
vulnerable to corresponding mental disorders, such as disorders that involve too little, or too much,
social thinking. Autism can be considered as a disorder where complex sociality does not develop,
while schizophrenia, bipolar disorder, and depression can be considered as the opposite:
pathologically over-developed social thought and behavior, as seen for example in paranoia and
hearing voices. Evolutionary biology is fundamentally important in understanding, defining and
treating mental disorders because it helps us to determine what the brain has evolved to do, which
informs us about the different ways that brain functions can become dysregulated in disease.
I. Introduction
The standard medical model and the reification of psychiatric disorders
The standard medical model for understanding and treating disease focuses on determining its
proximate physiological and developmental causes, in terms of how functional systems have become
dysregulated (Nesse and Stein 2012). High blood glucose levels, for example, may be due to type 1
diabetes, which results from specific, well-characterized physiological and molecular-biological
1
causes, and, as a result, can be unambiguously diagnosed. Understanding the normal functioning of
blood glucose regulation, or any other physiological system, thus represents a key precondition to
determining etiology and effective treatments.
How can the standard medical model be applied to psychiatric disorders? Diabetes is real, meaning
here that its presence can be objectively and unequivocally quantified. By contrast, psychiatric
disorders like autism and schizophrenia are not real, because the causes and patterns of brain
functions that underlie them are only dimly understood. Psychiatric disorders are, instead, abstract,
heuristic, descriptive constructs that are more or less useful for guiding research, diagnoses, and
treatments. The clearest evidence for such artificiality is the Diagnostic and Statistical Manual of
Mental Disorders (DSM) criteria for diagnosing psychiatric disorders, which comprise detailed lists of
symptoms, some set of which are considered necessary and sufficient to infer the presence of
disease.
Despite these considerations, it is commonplace for psychiatric conditions to be reified - that is -
considered as real, for research, medical and societal purposes (Crespi 2011). Such pragmatic
reification can be considered as innocuous, but it is not: it constrains and biases how researchers
think about mental disorders, and their associated research agendas, and leads to misconceptions of
psychiatric disorders as objectively-defined, purely-pathological 'diseases' that people 'have',
comparable in some fundamental way to diseases like diabetes, cancer, or atherosclerosis that can
be objectively and physiologically quantified in terms of their causes and effects.
Under current paradigms, determining the 'causes' of mental disorders often becomes conflated with
characterizing mental pathologies or deficits, at levels from genes, to neurodevelopment and
function, to cognitive functions, and to deleterious environments. By contrast, according to the
standard medical model, mental disorders should instead be conceptualized, and analyzed, in terms
of what functional mental systems have become dysregulated, and what forms such dysregulations
take. In this regard, for example, to better-understand autism we must also better-understand the
development of neurotypical social cognition, and to understand bipolar disorder and depression, we
must also understand the adaptive functions of normal, contextual variation in mood.
The evolution of mental adaptations
Adaptive functions of the human mind and brain, like those of glucose regulation, have, of course,
evolved. Most generally, this meaning of 'adaptive' means that such systems show, and have for
many, many past generations shown, genetically-based variation among individuals that has
influenced survival and reproduction. Such variation has thus been subject to natural selection,
which leads, across generations, to increases in, or maintenance of, the adaptive 'fit' or 'match'
between organismal phenotypes and aspects of their environments. For example, the beaks of
Darwin’s finches are ‘fit’ in their sizes and shapes for different food sources. Similarly, specific
regions of the human neocortex adaptively function to recognize individual faces (the fusiform gyrus),
or to infer the thoughts and intentions of other humans (the medial prefrontal cortex). Specific mental
adaptations, like the insulin pathway, are real and quantifiable, and the subject of intense interest in
disciplines such as cognitive neuroscience.
Natural selection of human physiology and morphology is expected, under basic evolutionary
considerations, to have led to maximization of functional robustness, homeostatic ability, and
efficiency, as well as optimal flexibility under variable circumstances, all in the service of survival and
reproduction. But what, then, is natural selection – the driver of adaptation – expected to maximize
with regard to human cognition, emotion and behavior? We usually think of mental disorders as
centrally involving unhappiness, of the subject as well as their social circle, which motivates the
2
seeking of help from the medical community. However, natural selection is by no means expected to
maximize happiness, simply because increased happiness is by no means a primary means or route
to increased survival and reproduction (Nesse 2004). Instead, natural selection is predicted, by basic
theory, to maximize condition-dependent human striving for the goals that have led, across many
past generations in relevant environments, to high survival and reproduction, relative to other
humans.
In the context of striving, human emotional systems have evolved to motivate and modulate goal
seeking, dynamically across different circumstances. Such motivation is mediated by the human
‘liking’ and ‘wanting’ reward systems, as well as by unhappiness or dissatisfaction with current
situations. Human cognitive systems, by contrast, represent sets of evolved mechanisms for
information processing, causal thinking, and decision-making that subserve identification of
appropriate goals, and tactics for reaching them. Both emotional and cognitive systems develop
across infancy, childhood and adolescence, whereby genes, environments, and gene by environment
interactions mediate neurodevelopment. To understand human psychiatric disorders from an
evolutionary perspective, it thus becomes necessary to connect these psychological trajectories and
adaptations with their corresponding maladaptations (lacks of fit of phenotypes to the environment),
expressed as developmental, emotional, and cognitive dysfunctions that revolve around human
striving and cognition. What adaptations, then, are dysregulated in major human mental disorders,
and how?
Evolutionary biology is useful in medicine for two main reasons: (1) it teaches us how to think about
human medically-relevant phenotypes, and diagnoses, in novel, productive ways, and (2) it indicates
specific new data to collect, and new approaches for therapies. In this chapter, I focus on the
evolutionary biology of psychiatric disorders centrally involving social cognition, affect, and
development. I first describe the primary types of causes of mental disorders, from evolutionary-
medical thinking. Next, I describe autism spectrum disorders, and psychotic-affective spectrum
disorders, in the context of these causes, with reference to recent findings in genetics, neuroscience,
and psychology, and in the contexts of which human-evolved adaptations have been subject to what
forms of alteration in each case. Third, I describe and evaluate hypotheses for the relationships of
these disorders with one another - relationships that define evolved axes of human development,
affect and cognition that structure variation in adaptive and maladaptive human mental functioning.
Finally, I make specific suggestions for research and clinical therapies that follow directly from these
considerations.
II. Research Findings
The evolutionary causes of psychiatric disorders represent the 'ultimate' sources of these conditions,
which indicate why, given their evolutionary history, humans exhibit particular forms of mental
disorders with particular symptoms and severities. Each of the six main causes described below
centers on explanations for deviations from mental adaptation and health, in the context of how
maladaptations can arise, and be maintained, in populations.
(1) Deleterious alleles. Mutations generate novel alleles that usually cause reduced genetic function,
because the perturbations randomly alter a system that would otherwise develop reasonably well.
Highly-penetrant mutations, with large effects, are especially likely to be highly deleterious, and
considerable evidence attests to important roles for de novo, deleterious mutations, such as copy
number variants or changes to highly conserve amino acid residues, in the causes of mental illness
(e. g., Malhotra and Sebat 2012). Highly deleterious alleles that are associated with relatively-severe
mental illnesses include, for example, monogenic causes of autism or schizophrenia, that evolve
under mutation-selection balance: rare mutations arise, and are selected against because their
3
bearers exhibit greatly-reduced reproduction.
Rare, deleterious alleles such as copy number variants have been estimated to account for a small
percentage of cases of major mental illness (Escudero and Johnstone 2014). Most inferred 'risk
alleles' for mental disorders, such as those identified with genome-wide association studies are,
however, relatively common (at frequencies above 1% or 5%), and have small effects on risk through
one dimension of their multifaceted impacts on neurodevelopment, neuronal function, and other
systems. The degree to which such alleles can be considered as deleterious to health overall - given
all of their effects - remains an open question; for example, neurodegenerative disease risk trades off
with cancer risk, such that higher risks in one domain of disease may commonly entail lower risks in
another (Plun-Favreau et al. 2010). Presumably, if psychiatric risk alleles were purely deleterious,
they would indeed not be common in populations. Risk alleles may also exhibit positive effects, on
health and reproduction, when expressed in genetic relatives of individuals with mental illness (Power
et al. 2013); these findings indicate that 'risk' alleles do not simply confer increased risk of disease,
but may, depending on context, confer benefits as well. Such considerations can help to explain the
high heritabilities of psychiatric conditions including autism, bipolar disorder and schizophrenia, on
the order of 50-80% (e. g., Singh et al. 2014).
(2) Mismatched environments. Populations and individuals are always adapted to past
environments, and if environments change more rapidly than they can be tracked by selection and
genetic response to selection, then populations will be maladapted. Human environments have
changed radically over the past few hundred years, which is expected to lead to higher risk of
psychiatric disorders to the extent that the novel environments include risk factors such as increased
social stress and isolation, or toxins such as lead and mercury that degrade neurodevelopment. For
example, some of the highest rates of schizophrenia are found among visible-minority (e. g., different
skin color) immigrants, who appear to be subject to relatively-severe psychosocial stresses due to
their novel, challenging environments (Bourque et al. 2012).
(3) Extremes of adaptations. Some psychiatric conditions, such as Generalized Anxiety Disorder, or
some manifestations of Obsessive-Compulsive Disorder such as excessive hygienic behavior, clearly
represent extremes of normally-adaptive behavior: anxiety functions to modulate arousal and
attention under challenging conditions (Stein 2013), and hygiene reduces risks of infection (Curtis
2014). This conceptual framework has been generalized to connect normal personality variation
along a spectrum to personality disorders and to severe psychiatric disorders, by demonstrating
which aspects of personality are amplified, reduced, or otherwise distorted to generate mental
dysfunction (Trull and Widiger 2013). This approach has successfully described continua in
personality traits from normal to maladaptive extremes, although the adaptive significance, in terms
of fitness-related benefits and costs of personality variation among normal individuals, remains
largely unstudied. Maladaptive extremes can also be considered more directly in the context of
human evolutionary history, in that the evolution of human-specific traits, such as large brain size and
language, has generated potential and scope for loss of these specific traits, as in microcephaly and
Specific Language Impairment, as well as potential and scope for dysfunctional over-development,
as in macrocephaly and the disordered and exaggerated components of speech in schizophrenia
(Crespi 2008; Crespi and Leach 2014).
(4) Tradeoffs. Tradeoffs have been well-characterized for developmental and physiological
phenotypes, whereby, for example, increased resource allocation in one domain takes away from
another. For neurological and psychological phenotypes, however, conceptual paradigms based on
tradeoffs have yet to be developed, despite evidence for tradeoffs of verbal-social with visual-spatial
skills (Johnson and Bouchard 2007), empathic with systemizing (rule-based) interests and abilities
(Nettle 2007), neural flexibility with stability (Liljenstrom 2003), as well as tradeoffs between neural
4
activation of the internally, self-directed default mode network, and the outwardly-focussed task-
positive network (Jack et al. 2013). Cognitive and emotional tradeoffs are important because they
structure the brain's functional architecture, and generate coincidences of relative strengths with
relative deficits; for example, Kravariti et al. (2006) found that having closer relatives with
schizophrenia was strongly associated with better verbal skills relative to visual-spatial skills.
Tradeoffs are stronger under resource-related constraints, which may commonly follow from
dysfunctional neurodevelopment, and their extremes are expected to characterize some psychiatric
conditions. Autism, for example, has been strongly associated with a combination of high systemizing
and low empathizing, whereas some combination of dysfunctionally high empathizing and low
systemizing appears to characterize some psychotic-affective conditions (Brosnan et al. 2010),
especially borderline personality disorder and depression (Dinsdale and Crespi 2013).
(5) Conflicts. Genetically-based conflicts, whereby two parties exhibit different optima for some
genetically-based phenotype, generate risk of maladaptation because one party may more or less
'lose' the conflict, resources are wasted on conflictual interactions, and conflict mechanisms generate
novel targets for dysregulation and disease (Crespi et al. 2014). The forms of evolutionary-genetic
conflict most salient to psychiatric conditions include parent-offspring conflict (e. g., Crespi 2010),
genomic imprinting conflict (Crespi and Badcock 2008; Haig 2014), and sexual conflict (Haig et al.
2014). Dysregulated genomic imprinting, for example, underlies the expression of Prader-Willi
syndrome, one of the strongest genetic causes of psychosis (Soni et al. 2008), and this syndrome
represents only an extreme case of such psychiatric effects (Crespi 2008). Similarly, a recent
epidemiological study of over two million individuals demonstrated that unaffected sisters (but not
brothers) of individuals with schizophrenia and bipolar disorder exhibit higher fertility than controls, a
pattern that is uniquely predicted by a hypothesis of 'sexually-antagonistic' alleles that impose costs
on males but benefit females (Power et al. 2013).
(6) Defenses mistaken as symptoms. This last 'cause' of disease is only apparent: some psychiatric
symptoms represent conditionally-adaptive defenses for alleviating problematic conditions, rather
than deleterious manifestations of disease. Thus, in the same way that fever represents a
conditionally-adaptive bodily response to infection, with health benefits that usually outweigh its
costs, some psychiatric symptoms can be interpreted as conferring benefits, relative to their absence
or reduction. Examples of such phenomena include: (a) repetitive behavior in autism, which serves to
dampen excessively-high levels of autonomic and sensory arousal (Hirstein et al. 2001), (b)
dissociation, as a psychological mechanism to reduce deleterious effects of trauma (Russo et al.
2014), (c) delusion formation in psychosis, as a means to mentally cope with the exaggerated and
disordered perceptions of salience (causal meaning) (Kapur 2003), and (d) mild depression (low
mood), as a conditionally-adaptive response to circumstances that favor disengagement from failing
or unreachable goals - which escalates to full depression if useless goal-seeking persists (Nesse and
Jackson 2011). The danger of conceptualizing defenses, like fever, as purely-deleterious symptoms
is that treating them is expected to make the situation specifically worse unless the underlying cause
of the disorder (and defense) is addressed, such as the sensory hyper-sensitivity in autism, the
trauma in dissociation, or the challenging life-events and personal motivational structure that underlie
liability to low mood and depression.
These six causes of etiology and symptoms of psychiatric conditions converge in their emphases on
determining what evolved genetic, developmental, neural, cognitive and emotional systems are
altered, and how they are altered, in psychiatric conditions. These causes also provide our
framework for determining how nominal, DSM-designated psychiatric conditions are related to one
another in their causes, as independent and separate, partially overlapping, or diametric to one
another in the same general way as the development or activity of any biological system or pathway
can be altered in two opposite directions.
5
Autism spectrum conditions
Autism is defined, and commonly reified, as a combination of deficits in social reciprocity and
communication with high levels of restricted interests and repetitive behavior (Figure 1). The degree
to which this combination represents a cohesive syndrome, with causally-shared rather than
independent symptoms and causal factors, remains unclear (Brunsdon and Happé 2014). Beyond
these two commonalities, autism presents diverse features, with overall intellectual abilities varying
from very low to above average, cognitive enhancements (above neurotypical) in sensory and visual-
spatial abilities in a substantial fraction of individuals, and a sex ratio that is highly male-biased
overall but much less so among more severely-affected individuals (Baron-Cohen et al. 2011).
The most straightforward connection between the major features of autism, and human evolution, is
that our evolutionary history has been characterized by elaboration of the 'social brain': the
distributed, integrated set of neural systems that subserve the acquisition, processing, and use of
social information. It is these social brain phenotypes that are specifically underdeveloped in autism.
As such, autism can be conceptualized as the expression of maladaptive extremes of social-brain
under-development, which, in principle, may be caused in a proximate way by reduction or loss of
any of the myriad systems that is necessary or sufficient for human social brain development. Autism
thus exhibits many single-gene, syndromic causes due to deleterious mutation, but it is also
commonly underlain by combined effects from the hundreds or thousands of genes bearing alleles
that affect social-brain development (Heil and Schaaf 2013). As such, there can be no primary,
proximate physiologically-based cause of autism (as there may be, for example, for type 1 diabetes),
and the search for causes becomes a differential characterization, subdivision, and prioritization of
the diverse genetic, epigenetic, and environmental influences that converge on under-development of
the social brain.
As social cognition is under-developed under all psychologically-based theories for autism, it can also
be conceptualized, and studied, in terms of developmental heterochrony, whereby child cognitive
development is not completed in autism, and childhood characteristics, including reduced social
cognition, are retained into adulthood (Woodard and van Reet 2011; Crespi 2013). In this context,
other human-elaborated traits including highly-developed, regulated social striving and goal-seeking,
guided by perceived reward-associated or cost-associated (aversive) salience (inferred, causative
meaning) of social stimuli, remain underdeveloped as well on the autism spectrum. External stimuli
may thus have salience predominantly in terms of perceived sensations, or specific, highly-restricted
non-social interests, especially foci of highly selective attention (Ploog 2010). Frith (2012) indeed
sees a weak drive to discern meaning in the world as epitomizing the weak central coherence theory
of autism, which has been supported by a wide range of evidence.
A central, unresolved question in the study of autism is whether a single, central psychological or
cognitive-level factor can explain the apparently inexplicable combination of reduced sociality with
restricted interests and repetitive behavior. In the context of social brain under-development,
increased restricted interests and repetitive behaviors, and sensory, visual-spatial, and mechanistic-
cognition enhancements in autism, can be explained by several hypotheses.
First, increases in asocial phenotypes may pre-empt the development of social phenotypes, such as
by directing perceived salience, interests, and brain specializations along asocial paths. Such
effects, which are notably represented by a theory for autism etiology based on enhanced perceptual
functioning (Mottron et al. 2006) may be mediated by over-developments of sensory perception and
mechanistic, systemizing cognition (Baron-Cohen et al. 2011).
6
Second, increased asocial cognition may itself be a direct result of reduced social cognition, as a
compensatory or tradeoff-based neurodevelopmental mechanism akin to the over-development of
non-visual senses among the blind.
Third, some such asocial cognition and behavior, and phenotypes such as insistence on sameness
and stimulus over-selectivity, may, as noted above, represent defenses that aid in coping with
challenging symptoms such as increased perceptual sensitivity or avoidance of stress from dealing
with inexplicable social-cognitive tasks.
Finally, one possible resolution, based on reduced expression of a phenotype virtually unique to
humans, is that autism is, in part, underpinned psychologically by under-developed imagination,
defined as 'the faculty or action of forming new ideas, or images or concepts of external objects not
present to the senses'. This hypothesis, originally described by Rutter (1972), and Wing and Gould
(1979) can, in principle, jointly explain social and asocial alterations in autism, including reduced
pretend play, reduced social imagination as expressed in theory of mind, restriction of interests and
repetition of behavior, and insistence on sameness.
Determining the degree to which of these hypotheses is correct, in general or for any particular
individual, is crucially important to autism therapy, especially to prevent enhancements or
conditionally-adaptive defenses of autistic individuals from being treated as deleterious symptoms.
Psychotic-affective spectrum conditions
Psychotic-affective spectrum conditions include a set of DSM disorders, mainly schizophrenia,
bipolar disorder, and depression, that broadly overlap in their symptoms, neurological and
psychological correlates, and genetic and environmental risk factors (Doherty and Owen 2014)
(Figure 1). All of these conditions exhibit substantial genetic components, and mediation in part by
rare, penetrant risk factors, although most genetic risk appears to be underlain by many alleles each
of small effect.
Schizophrenia, as well as other conditions that involve psychosis, can be understood most directly
and simply in terms of dysfunction of the human adaptive system for assigning salience (causal
meaning) to external, as well as internally-generated, stimuli (Kapur 2003; Winton-Brown et al. 2014).
Salience assignment, which is underpinned by a dedicated neural system involving the anterior
cingulate cortex and insula, is fundamental to cognition, behavior, and goal-seeking, in that it
mediates subjective causal understanding of perceptual inputs. Psychosis thus involves over-
developed and inappropriate salience, usually in the contexts of social interactions, agency,
intentionality, self-other associations, and other aspects of mentalistic (social and mind-related)
thought, apparently due to the primacy of social cognition in human goal-directed behavior (Crespi
and Badcock 2008). Paradigmatic manifestations of psychosis thus involve paranoia, other social
delusions, megalomania, belief that events always refer to the self, alterations to self-other
distinctions, and assignment of mind, agency and intentions to inappropriate subjects and inanimate
objects. Such reality distortions are mediated by top-down cognitive processes, and they can be
considered as attempts to 'make sense' of the excessively-high and inappropriate salience
assignment, for external stimuli, that is driven by hyper-dopaminergic neurotransmission (Kapur
2003; Cook et al. 2012; Howes and Murray 2014). Hallucinations, in turn, can be understood as
misinterpreted and exaggerated internal perceptions, mediated by over-developed salience of
internal representations, such that given certain neurophysiological alterations, thought, inner
speech, and imagination come to be considered as external percepts. Like delusions, hallucinations
are usually expressed as social phenomena, especially auditory hallucinations.
7
Schizophrenia is predominantly considered as a disorder of cognition, whereby the causal meanings
that guide striving become over-developed and dysfunctionally over-mentalistic. Bipolar disorder and
depression, by contrast, represent mainly disorders of emotion, the set of neural and hormonal
systems that motivate and modulate striving and goal-seeking across different contexts.
Understanding such mood disorders requires consideration of the adaptive significance of condition-
dependent variation in human emotions, especially with regard to the social interactions that
permeate human thought and behavior (Nesse and Ellsworth 2009). In this context, considerable
evidence indicates that low mood is normally adaptive in situations where individuals benefit by
disengaging from unreachable or unprofitable goals, as it facilitates such disengagement and
motivates alternative behavioral patterns of goal-seeking that should be more advantageous (Keller
and Nesse 2005). High, positive mood, in comparison, represents an emotional mechanism whereby
human reward systems motivate continuation of beneficial behavior, because one's goals are being
reached. Depression, then, can be conceptualized and studied as overly-low and overly-stable mood,
a maladaptive extreme of an adaptation, whereby individuals fail to disengage from deleterious
thought patterns and striving (Nesse 2004; Keller and Nesse 2005). Conversely, mania represents an
emotional opposite to depression, as the expression of inability to emotionally restrain high mood and
intensity of striving, even if and when its consequences become detrimental (Johnson 2005; Johnson
et al. 2012a). Behaviors associated with mania and hypomania can, moreover, be directly
interpreted in the context of extreme striving for social dominance, power, and influence, which, if
successful, leads to substantial benefits (Johnson and Carver 2012; Johnson et al. 2012b). This
evolutionary perspective can explain shifts between mania and depression in bipolar disorder, in that
mania is expected to foster pursuit of goals that become more and more risky, unreachable or
unsuccessful, eventually prompting the generation of mixed states, and descent into depression.
In bipolar disorder, then, cognitive salience systems, and choices of goals, commonly remain
functional, but the homeostatic regulation of the emotions that underlie goal pursuit becomes
dysregulated, towards overly low or overly high moods and their sequelae. Moreover, like
schizophrenia, mania and depression both centrally involve extremes of social, mentalistic thought
and behavior, here in the contexts of guilt, shame, embarrassment, perceived social defeat, and
social rumination in depression, and social-dominance pursuit and pride in mania. Affective
psychoses, which comprise psychosis with incongruence of mood, may thus be mediated by self-
punishment driven, or reward-driven, over-attributions of social salience, in the context of
emotionality that becomes sufficiently strong to dysregulate salience. These considerations can help
to explain well-documented, otherwise-inexplicable associations of bipolar disorder with high social
motivation and achievement (Coryell et al. 1989; Johnson et al. 2012a; Higler et al. 2014). Moreover,
bipolar disorder, as well as schizophrenia and schizotypy, have been associated across a wide
diversity of studies with increased social imagination, divergent thinking, creativity, and goal
attainment, especially in the arts and humanities (Nettle 2002, 2006; Burns 2004; Simeonova et al.
2005; Carson 2011; Bilder and Knudsen 2014). Imagination can indeed be considered, under
Bayesian models of cognition and learning, as directly associated with causal cognition and inference
of meaning, such that salience, causal thinking, and imagination should tend to increase, or
decrease, in concert with one another (Walker and Gropnik 2013).
The relationship between autism spectrum and psychotic-affective spectrum disorders
Bleuler invented the term 'autism' to describe withdrawal from reality and social interactions in
schizophrenia, but Kanner was careful to point out that his conceptualization of autism referred to
children who had never participated in social life (Kanner 1965). The relationship between autism
and schizophrenia, and psychotic-affective disorders more generally, has since been considered in
terms of two main hypotheses, (1) partial overlap, with some degree of shared social-cognitive
deficits and genetic risk factors; (2) a diametric (opposite) relationship, based, at a psychological
8
level, on under-development of social cognition and affect in autism, normality at the center, and
dysfunctional forms of their over-development in psychotic-affective conditions (Crespi et al. 2010)
(Figure 2). The partial overlap hypothesis is data-driven and motivated primarily by the prominence of
social deficits especially in autism and schizophrenia. By contrast, the diametric hypothesis follows
directly from evolutionary and neurodevelopmental considerations, under the premises that human
evolution has been characterized primarily by elaboration of social cognition (generating increased
scope for altered development of specific phenotypes), and that the neurodevelopmental systems
that underlie it, like all biological systems, can vary and be perturbed in two opposite directions
towards lower or higher expression (Figure 2).
A central prediction of the diametric hypothesis is that autism, and psychotic-affective conditions
(especially schizophrenia, for which most of the relevant data is available) should exhibit opposite
phenotypes and genetic risk factors. A suite of such evidence is described in Table 1, which provides
support for the diametric hypothesis from diverse and independent sources of data. The partial-
overlap hypothesis is consistent with the sharing of deficits, especially in social cognition, between
autism and schizophrenia, but such deficits can also be considered as deriving from opposite
alterations both of which reduce performance on standard tests. Genetic risk factors, such as some
genomic copy-number variants and some SNPs, have also been associated with both autism
spectrum disorders and schizophrenia (Crespi et al. 2010). Such findings, however, are subject to the
caveat that premorbidity to schizophrenia in children and young adolescents, in the form of social
deficits and associated developmental problems, can be realistically diagnosed only as autism
spectrum since there is not (and never has been) a diagnostic category for schizophrenia
premorbidity (Crespi and Crofts 2012). This structural limitation in the DSM is expected to lead to a
non-negligible incidence of false-positive diagnoses of autism among children who are actually
premorbid for schizophrenia, especially among individuals harboring relatively-penetrant genetic risk
factors such as copy number variants. Patterns of diagnoses for well-studied CNVs indeed fit with
expectations from such false positive diagnoses (Crespi and Crofts 2012).
The diametric hypothesis for autism and psychotic-affective disorders is novel and controversial, and
has just begun to be subject to systematic, large-scale testing of its predictions (e. g., Byars et al.
2014). However, to the extent that it is correct, the study of human disorders involving social
cognition should be revolutionized, and provided its first solid grounding in basic evolutionary
principles.
III. Implications for policy and practice
Risks for human mental disorders have evolved. Evolutionary conceptualizations of autism and
psychotic-affective disorders lead directly to novel, specific implications for understanding, studying
and treating these conditions.
First, autism, schizophrenia, bipolar disorder, and depression cannot justifiably be considered as
'diseases' under standard medical models of disease, because the neural-system adaptations
subject to maladaptive alteration in each case remain inadequately understood. Instead, these
conditions currently represent broad-scale, heuristic descriptions for suites of related psychological
and behavioral problems, none of which has currently-specifiable genetic or neurological causes in
the same way as do diseases like cancer or diabetes, and all of which grade smoothly in their
symptoms into normality. As such, schizophrenia and related psychotic and affective disorders can
best be considered as 'syndromes': groups of symptom dimensions that cluster in different
combinations across different individuals (van Os 2009). Risks and symptoms for these psychiatric
conditions have, however, evolved in close conjunction with the evolution of complex human social
cognition, affect and behavior, which provides the basis for an ultimate understanding, and nosology,
9
of psychiatric maladaptations. In this context, DSM descriptions of autism or a psychotic-affective
condition should represent starting points for differential diagnosis of their genetic, social and
environmental causes, for each specific individual. Such causes are expected to involve some
combination of effects from deleterious mutations, evolutionarily-novel environments, extremes of
adaptations, tradeoffs, genomic conflicts, and evolved defenses.
Second, autism can be considered, from an evolutionary perspective, in terms of under-development
of social cognition and affect, centrally involving some combination, and causal conjunction, of
reduced social development with increased non-social perception, attention, and cognition. Such
social and non-social alterations may have diverse proximate causes, but they appear to commonly
converge, psychologically, on reductions in imagination, which can explain both lower levels of
sociality and increases in restricted interests and repetitive behavior. This conceptualization of autism
is fully compatible with previously-developed psychological models founded on reduced central
coherence (Happé and Frith 2006), lower empathizing and higher systemizing (Baron-Cohen et al.
2011), and enhanced perceptual function (Mottron et al. 2006).
Third, psychotic-affective disorders can be considered as centrally involving dysfunctionally over-
developed social cognition, affect, and behavior, expressed as social hyper-salience in aspects of
psychosis, dysregulated social goal motivation and dominance-seeking in mania, and extremes of
negative social emotionality in depression. Each of these disorders, which grade into one another,
can best be understood in the individual-level contexts of the developmental causes of negatively-
valenced and imaginative social salience, and the motivational structure of one's past, current and
future imagined life goals, especially regarding regulation of, and impediments to, success in striving.
This framework is fully compatible with current psychological, neurological, cognitive-science-level
accounts of psychotic-affective conditions (e. g., Kapur 2003; Johnson et al. 2012a,b; Winton-Brown
et al. 2014), but grounds them in evolutionary considerations, and in their relationship to the autism
spectrum.
Fourth, autism and psychotic-affective conditions can be considered and analyzed as diametric
(opposite) disorders with regard to social development, cognition, affect, and behavior. This
diametric model provides for comprehensive, reciprocal illumination of the diagnoses, causes and
treatments of these disorders, such that insights derived from studying one set of disorders can be
applied directly to the other. Most generally, cognitive-behavioral treatments for autism should
especially focus on enhancing phenotypes that are over-developed in psychotic-affective conditions,
including social imagination, flexible and social salience, and social motivation and goal-seeking. By
contrast, treatments for psychotic-affective conditions, in addition to focusing more directly on the
adaptive, dynamic regulation of social-cognitive salience and mood-directed striving, should involve
therapies to make perception, cognition, affect and behavior relatively 'more autistic'. Similar
considerations apply to pharmacological effects: for example, valproate during fetal development
represents a well-established human cause, and animal model, of autism (Markram and Markram
2010), but valproate is also used to treat bipolar disorder and schizophrenia (Haddad et al. 2009);
comparably, mGlur5 pathway antagonists are being used to treat fragile X syndrome and autism
(Lozano et al. 2014), whereas mGlur5 agonists are being developed to treat schizophrenia (Matosin
and Newell 2013).
The findings and inferences described here emphasize that evolutionary approaches in medicine,
and psychiatry, can offer specific, well-rationalized hypotheses, and can help to direct research and
treatments along novel and promising paths. Such progress should lead, eventually, to the
integration of psychiatry with the standard medical model of disease, as dovetailing evolutionary and
proximate approaches to the study of brain development and function uncover the adaptive
significance of psychological, cognitive, and affective phenotypes, and their neurological and genetic
10
foundations.
References
Adams NC, Jarrold C (2009) Inhibition and the validity of the Stroop task for children with autism. J
Autism Dev Disord 39(8):1112–1121
Adida M, Maurel M, Kaladjian A, Fakra E, Lazerges P, Da Fonseca D, Belzeaux R, Cermolacce M,
Azorin JM. (2011) Decision-making and schizophrenia. Encephale 37 Suppl 2:S110-6.
Almeida RA, Dickinson JE, Maybery MT et al (2013) Visual search targeting either local or global
perceptual processes differs as a function of autistic-like traits in the typically developing
population. J Autism Dev Disord 43(6):1272–1286
Arnott W, Sali L, Copland D (2011) Impaired reading comprehension in schizophrenia: evidence for
underlying phonological processing deficits. Psychiatry Res 187(1-2):6–10
Backasch B, Straube B, Pyka M et al (2013) Hyperintentionality during automatic perception of
naturalistic cooperative behavior in patients with schizophrenia. Soc Neurosci 8(5):489–504.
doi:10.1080/17470919.2013.820666
Baez S, Herrera E, Villarin L et al (2013) Contextual social cognition impairments in schizophrenia
and bipolar disorder. PLoS One 8(3):e57664
Bara BG, Ciaramidaro A, Walter H et al (2011) Intentional minds: a philosophical analysis of intention
tested through fMRI experiments involving people with schizophrenia, people with autism, and
healthy individuals. Front Hum Neurosci 5:7
Baribeau DA, Anagnostou E (2013) A comparison of neuroimaging findings in childhood onset
schizophrenia and autism spectrum disorder: a review of the literature. Front Psychiatry 4:175.
doi:10.3389/fpsyt.2013.00175
Baron-Cohen S (2010) Empathizing, systemizing, and the extreme male brain theory of autism. Prog
Brain Res 186:167–175. doi:10.1016/B978-0-444-53630-3.00011-7
Baron-Cohen S, Lombardo MV, Auyeung B et al (2011) Why are autism spectrum conditions more
prevalent in males? PLoS Biol 9(6):e1001081. doi:10.1371/journal.pbio.1001081
Bates TC (2005) The panmodal sensory imprecision hypothesis of schizophrenia: reduced auditory
precision in schizotypy. Pers Indiv Diff 38(2):437–449
Beversdorf DQ, Smith BW, Crucian GP et al (2000) Increased discrimination of "false memories" in
autism spectrum disorder. Proc Natl Acad Sci U S A 97(15):8734–8737
Bilder RM, Knudsen KS (2014) Creative cognition and systems biology on the edge of chaos. Front
Psychol 5:1104. doi:10.3389/fpsyg.2014.01104
Bird G, Catmur C, Silani G et al (2006) Attention does not modulate neural responses to social stimuli
in autism spectrum disorders. Neuroimage 31(4):1614–1624
Blijd-Hoogewys EM, van Geert PL, Serra M et al (2008) Measuring theory of mind in children.
Psychometric properties of the ToM Storybooks. J Autism Dev Disord 38(10):19071930
11
Bonne O, Canetti L, Bachar E et al (1999) Childhood imaginary companionship and mental health in
adolescence. Child Psychiatry Hum Dev 29(4):277–286
Bourque F, van der Ven E, Fusar-Poli P et al (2012) Immigration, social environment and onset of
psychotic disorders. Curr Pharm Des 18(4):518–526
Brosnan M, Ashwin C, Walker I et al (2010) Can an ‘Extreme Female Brain’ be characterised in terms
of psychosis? Pers Indiv Diff 49(7):738–742
Brosnan M, Chapman E, Ashwin C (2014) Adolescents with autism spectrum disorder show a
circumspect reasoning bias rather than 'jumping-to-conclusions'. J Autism Dev Disord
44(3):513–520. doi:10.1007/s10803-013-1897-5
Brown WA, Cammuso K, Sachs H et al (2003) Autism-related language, personality, and cognition in
people with absolute pitch: results of a preliminary study. J Autism Dev Disord 33(2):163–167
Brunetti-Pierri N, Berg JS, Scaglia F et al (2008) Recurrent reciprocal 1q21.1 deletions and
duplications associated with microcephaly or macrocephaly and developmental and behavioral
abnormalities. Nat Genet 40(12):1466–1471
Brunsdon VE, Happé F (2014) Exploring the 'fractionation' of autism at the cognitive level. Autism
18(1):17–30. doi:10.1177/1362361313499456
Buckner RL, Andrews-Hanna JR, Schacter DL (2008) The brain's default network: anatomy, function,
and relevance to disease. In: Kingstone A, Miller MB (eds) The year in cognitive neuroscience.
Ann NY Acad Sci, New York, p 1–38. doi:10.1196/annals.1440.011
Burns JK (2004) An evolutionary theory of schizophrenia: cortical connectivity, metarepresentation,
and the social brain. Behav Brain Sci 27(6):831–855; discussion 855–885
Byars SG, Stearns SC, Boomsma JJ (2014) Opposite risk patterns for autism and schizophrenia are
associated with normal variation in birth size: phenotypic support for hypothesized diametric
gene-dosage effects. Proc Biol Sci 281(1794) pii: 20140604
Campbell BC, Wang SS-H (2012) Familial linkage between neuropsychiatric disorders and
intellectual interests. PLoS One 7(1):e30405
Cardoso-Martins C, da Silva JR (2010) Cognitive and language correlates of hyperlexia: evidence
from children with autism spectrum disorders. Read Writ 23(2):129–145
Carson SH (2011) Creativity and psychopathology: a shared vulnerability model. Can J Psychiatry
56(3):144–153
Cascio CJ, Foss-Feig JH, Burnette CP et al (2012) The rubber hand illusion in children with autism
spectrum disorders: delayed influence of combined tactile and visual input on proprioception.
Autism 16(4):406–419. doi:10.1177/1362361311430404
Chance SA (2014) The cortical microstructural basis of lateralized cognition: a review. Front Psychol
5:820
Chaste P, Sanders SJ, Mohan KN et al (2014) Modest impact on risk for autism spectrum disorder of
rare copy number variants at 15q11.2, specifically breakpoints 1 to 2. Autism Res 7(3):355–
362. doi:10.1002/aur.1378
12
Chevallier C, Kohls G, Troiani V et al (2012) The social motivation theory of autism. Trends Cogn Sci
16(4):231–239
Ciaramidaro A, Bölte S, Schlitt S et al (2014) Schizophrenia and autism as contrasting minds: neural
evidence for the hypo-hyper-intentionality hypothesis. Schizophr Bull pii: sbu124. [Epub ahead
of print]
Claridge G, McDonald A (2009) An investigation into the relationships between convergent and
divergent thinking, schizotypy, and autistic traits. Pers Indiv Diff 46(8):794–799
Clemmensen L, van Os J, Skovgaard AM et al (2014) Hyper-theory-of-mind in children with Psychotic
Experiences. PLoS One 9(11):e113082
Cook J, Barbalat G, Blakemore SJ (2012) Top-down modulation of the perception of other people in
schizophrenia and autism. Front Hum Neurosci 6:175. doi:10.3389/fnhum.2012.00175
Corlett PR, Simons JS, Pigott JS et al (2009) Illusions and delusions: relating experimentally-induced
false memories to anomalous experiences and ideas. Front Behav Neurosci 3:53.
doi:10.3389/neuro.08.053.2009
Coryell W, Endicott J, Keller M et al (1989) Bipolar affective disorder and high achievement: a familial
association. Am J Psychiatry 146(8):983–988
Courchesne E, Mouton PR, Calhoun ME et al (2011) Neuron number and size in prefrontal cortex of
children with autism. JAMA 306(18):2001–2010
Craig J, Baron-Cohen S (1999) Creativity and imagination in autism and Asperger syndrome. J
Autism Dev Disord 29(4):319–326
Crespi B (2008) Language unbound: genomic conflict and psychosis in the origin of modern humans.
In: Hughes D, D’Ettorre P (eds) Sociobiology of communication: an interdisciplinary
perspective. Oxford Univ Pr, Oxford UK, p 225248
Crespi B (2010) The strategies of the genes: genomic conflicts, attachment theory and development
of the social brain. In: Petronis A, Mill J (eds) Brain, behavior and epigenetics. Springer, Berlin,
p 143–167
Crespi B (2011) One hundred years of insanity: genomic, psychological and evolutionary models of
autism in relation to schizophrenia. In: Ritsner M (ed) Handbook of schizophrenia-spectrum
disorders, vol I. Springer, Netherlands, p 163–185
Crespi B (2013) Developmental heterochrony and the evolution of autistic perception, cognition and
behavior. BMC Med 11:119
Crespi B, Badcock C (2008) Psychosis and autism as diametrical disorders of the social brain. Behav
Brain Sci 31(3):241–261; discussion 261–320
Crespi B, Crofts HJ (2012) Association testing of copy number variants in schizophrenia and autism
spectrum disorders. J Neurodev Disord 4(1):15. doi:10.1186/1866-1955-4-15
Crespi B, Foster K, Úbeda F (2014) First principles of Hamiltonian medicine. Philos Trans R Soc
Lond B Biol Sci 369:20130366
13
Crespi B, Leach E (2014) The evolutionary biology of human neurodevelopment: evo-neuro-devo
comes of age. In: Boughner J, Rolian C (eds) Evolutionary developmental anthropology. Wiley
(in press)
Crespi B, Stead P, Elliot M (2010) Comparative genomics of autism and schizophrenia. Proc Natl
Acad Sci U S A 107 Suppl 1:1736–1741
Curtis VA (2014) Infection-avoidance behaviour in humans and other animals. Trends Immunol
35(10):457–464. doi:10.1016/j.it.2014.08.006
Daniel C, Lovatt A, Mason OJ (2014) Psychotic-like experiences and their cognitive appraisal under
short-term sensory deprivation. Front Psychiatry 5:106. doi:10.3389/fpsyt.2014.00106
Dickerson AS, Pearson DA, Loveland KA et al (2014) Role of parental occupation in autism spectrum
disorder diagnosis and severity. Res Autism Spectr Disord 8(9):997–1007
Dinsdale N, Crespi BJ (2013) The borderline empathy paradox: evidence and conceptual models for
empathic enhancements in borderline personality disorder. J Pers Disord 27(2):172–195
Doherty JL, Owen MJ (2014) Genomic insights into the overlap between psychiatric disorders:
implications for research and clinical practice. Genome Med 6(4):29. doi:10.1186/gm546
Dohn A, Garza-Villarreal EA, Heaton P et al (2012) Do musicians with perfect pitch have more autism
traits than musicians without perfect pitch? An empirical study. PloS One 7(5):e37961
Durkin MS, Maenner MJ, Meaney FJ et al (2010) Socioeconomic inequality in the prevalence of
autism spectrum disorder: evidence from a U.S. cross-sectional study. PLoS One 5(7):e11551
Dziobek I, Fleck S, Kalbe E et al (2006) Introducing MASC: a Movie for the Assessment of Social
Cognition. J Autism Dev Disord 36(5):623–636
Ek U, Fernell E, Jacobson L (2005) Cognitive and behavioural characteristics in blind children with
bilateral optic nerve hypoplasia. Acta Paediatr 94(10):1421–1426
Escudero I, Johnstone M (2014) Genetics of schizophrenia. Curr Psychiatry Rep 16(11):502.
doi:10.1007/s11920-014-0502-8
Falter CM, Braeutigam S, Nathan R et al (2013) Enhanced access to early visual processing of
perceptual simultaneity in autism spectrum disorders. J Autism Dev Disord 43(8):1857–1866.
doi:10.1007/s10803-012-1735-1
Faust M, Kenett YN (2014) Rigidity, chaos and integration: hemispheric interaction and individual
differences in metaphor comprehension. Front Hum Neurosci 8:511
Fernyhough C, Bland K, Meins E et al (2007) Imaginary companions and young children's responses
to ambiguous auditory stimuli: implications for typical and atypical development. J Child
Psychol Psychiatry 48(11):1094–1101
Force RB, Venables NC, Sponheim SR (2008) An auditory processing abnormality specific to liability
for schizophrenia. Schizophr Res 103(1):298–310
Fretland RA, Andersson S, Sundet K et al (2015) Theory of mind in schizophrenia: Error types and
associations with symptoms. Schizophr Res pii: S0920-9964(15)00028-6.
14
doi:10.1016/j.schres.2015.01.024
Frith U (2012) Why we need cognitive explanations of autism. Q J Exp Psychol (Hove) 65(11):2073–
2092. doi:10.1080/17470218.2012.697178
Gleason TR, Jarudi RN, Cheek JM (2003) Imagination, personality, and imaginary companions. Soc
Behav Pers 31(7):721–737
Granger KT, Prados J, Young AM (2012) Disruption of overshadowing and latent inhibition in high
schizotypy individuals. Behav Brain Res 233(1):201–208. doi:10.1016/j.bbr.2012.05.003
Grant P, Balser M, Munk AJ et al (2014) A false-positive detection bias as a function of state and trait
schizotypy in interaction with intelligence. Front Psychiatry 5:135
Haddad PM, Das A, Ashfaq M et al (2009) A review of valproate in psychiatric practice. Expert Opin
Drug Metab Toxicol 5(5):539–551. doi:10.1517/17425250902911455
Haig D (2014) Coadaptation and conflict, misconception and muddle, in the evolution of genomic
imprinting. Heredity (Edinb) 113(2):96–103
Haig D, Ubeda F, Patten MM (2014) Specialists and generalists: the sexual ecology of the genome.
Cold Spring Harb Perspect Biol 6(9):pii:a017525. doi:10.1101/cshperspect.a017525
Haijma SV, Van Haren N, Cahn W et al (2013) Brain volumes in schizophrenia: a meta-analysis in
over 18 000 subjects. Schizophr Bull 39(5):1129–1138. doi:10.1093/schbul/sbs118
Hanslmayr S, Backes H, Straub S et al (2012) Enhanced resting-state oscillations in schizophrenia
are associated with decreased synchronization during inattentional blindness. Hum Brain
Mapp 34(9):2266–2275. doi:10.1002/hbm.22064
Happé F, Frith U (2006) The weak coherence account: detail-focused cognitive style in autism
spectrum disorders. J Autism Dev Disord 36(1):5–25
Harkness KL, Washburn D, Theriault JE et al (2011) Maternal history of depression is associated
with enhanced theory of mind in depressed and nondepressed adult women. Psychiatry Res
189(1):91–96
Heaton P, Hudry K, Ludlow A et al (2008a) Superior discrimination of speech picth and its
relationship to verbal ability in autism spectrum disorders. Cogn Neuropsychol 25:771–782
Heaton P, Williams K, Cummins O et al (2008b) Autism and pitch processing splinter skills: a group
and subgroup analysis. Autism 12(2):203–219
Heil KM, Schaaf CP (2013) The genetics of Autism Spectrum Disorders--a guide for clinicians. Curr
Psychiatry Rep 15(1):334. doi:10.1007/s11920-012-0334-3
Higier RG, Jimenez AM, Hultman CM et al (2014) Enhanced neurocognitive functioning and positive
temperament in twins discordant for bipolar disorder. Am J Psychiatry.
doi:10.1176/appi.ajp.2014.13121683
Hillier A, Campbell H, Keillor J et al (2007) Decreased false memory for visually presented shapes
and symbols among adults on the autism spectrum. J Clin Exp Neuropsychol 29(6):610–616
15
Hirstein W, Iversen P, Ramachandran VS (2001) Autonomic responses of autistic children to people
and objects. Proc Biol Sci 268(1479):1883–1888
Hobson JA, Hobson RP, Malik S et al (2013) The relation between social engagement and pretend
play in autism. Br J Dev Psychol 31(Pt 1):114–127. doi:10.1111/j.2044-835X.2012.02083.x
Hobson RP, Bishop M (2003) The pathogenesis of autism: insights from congenital blindness. Philos
Trans R Soc Lond B Biol Sci 358(1430):335–344
Howes OD, Murray RM (2014) Schizophrenia: an integrated sociodevelopmental-cognitive model.
Lancet 383(9929):1677–1687. doi:10.1016/S0140-6736(13)62036-X
Iacoboni M (2006) Failure to deactivate in autism: the coconstitution of self and other. Trends Cogn
Sci 10:431–433
Immordino-Yang MH, Christodoulou JA, Singh V (2012) Rest is not idleness implications of the
brain’s default mode for human development and education. Perspect Psychol Sci 7(4):352–
364
Jack AI, Dawson AJ, Begany KL et al (2013) fMRI reveals reciprocal inhibition between social and
physical cognitive domains. Neuroimage 66C:385–401
Jamison KR (1993) Touched with fire: manic-depressive illness and the artistic temperament.Free
Press, New York
Jarrold C (2003) A review of research into pretend play in autism. Autism 7(4):379–390
Javitt DC (2009a) Sensory processing in schizophrenia: neither simple nor intact. Schizophr Bull
35(6):1059–1064
Javitt DC (2009b) When doors of perception close: bottom-up models of disrupted cognition in
schizophrenia. Annu Rev Clin Psychol 5:249–275
Johnson SL (2005) Mania and dysregulation in goal pursuit: a review. Clin Psychol Rev 25(2):241–
262
Johnson SL, Carver CS (2012) The dominance behavioral system and manic temperament:
motivation for dominance, self-perceptions of power, and socially dominant behaviors. J Affect
Disord 142(1-3):275–282
Johnson SL, Fulford D, Carver CS (2012a) The double-edged sword of goal engagement:
consequences of goal pursuit in bipolar disorder. Clin Psychol Psychother 19(4):352–362
Johnson SL, Leedom LJ, Muhtadie L (2012b) The dominance behavioral system and
psychopathology: evidence from self-report, observational, and biological studies. Psychol Bull
138(4):692–743. doi:10.1037/a0027503
Johnson W, Bouchard TJ Jr (2007) Sex differences in mental abilities: G masks the dimensions on
which they lie. Intelligence 35:23–39
Jung M, Kosaka H, Saito DN et al (2014) Default mode network in young male adults with autism
spectrum disorder: relationship with autism spectrum traits. Mol Autism 5:35
16
Kana RK, Libero LE, Hu CP et al (2014) Functional brain networks and white matter underlying
theory-of-mind in autism. Soc Cogn Affect Neurosci 9(1):98–105
Kana RK, Liu Y, Williams DL et al (2013) The local, global, and neural aspects of visuospatial
processing in autism spectrum disorders. Neuropsychologia 51(14):2995–3003
Kana RK, Wadsworth HM, Travers BG (2011) A systems level analysis of the mirror neuron
hypothesis and imitation impairments in autism spectrum disorders. Neurosci Biobehav Rev
35(3):894–902. doi:10.1016/j.neubiorev.2010.10.007
Kanemoto M, Asai T, Sugimori E et al (2013) External misattribution of internal thoughts and
proneness to auditory hallucinations: the effect of emotional valence in the Deese-Roediger-
McDermott paradigm. Front Hum Neurosci 7:351. doi:10.3389/fnhum.2013.00351
Kanner L (1965) Infantile autism and the schizophrenias. Behavioral Science 10(4):412–420
Kapur S (2003) Psychosis as a state of aberrant salience: a framework linking biology,
phenomenology, and pharmacology in schizophrenia. Am J Psychiatry 160(1):13–23
Kasari C, Chamberlain B, Bauminger N (2001) Social emotions and social relationships: can children
with autism compensate? In: Burack JA, Charman T, Yirmiya N, Zelazo PR (eds) The
development of autism: perspectives from theory and research. Lawrence Erlbaum Associates
Publishers, Mahwah NJ USA, p 309–323
Kaufman JC (ed) (2014) Creativity and mental illness. Cambridge Univ Press, Cambridge UK
Keller MC, Nesse RM (2005) Is low mood an adaptation? Evidence for subtypes with symptoms that
match precipitants. J Affect Disord 86(1):27–35
Kennedy DP, Courchesne E (2008) Functional abnormalities of the default network during self- and
other-reflection in autism. Soc Cogn Affect Neurosci 3(2):177–190
Kennedy DP, Redcay E, Courchesne E (2006) Failing to deactivate: resting functional abnormalities
in autism. Proc Natl Acad Sci USA 103:8275–8280
Kercood S, Grskovic JA, Banda D et al (2014) Working memory and autism: a review of the literature.
Res Autism Spectrum Dis 8(10):1316-1332.
Kim S, Thibodeau R, Jorgensen RS (2011) Shame, guilt, and depressive symptoms: a meta-analytic
review. Psychol Bull 137(1):68–96. doi:10.1037/a0021466
King D, Dockrell J, Stuart M (2014) Constructing fictional stories: a study of story narratives by
children with autistic spectrum disorder. Res Dev Disabil 35(10):2438–2449
Kohl S, Wolters C, Gruendler TO et al (2014) Prepulse inhibition of the acoustic startle reflex in high
functioning autism. PLoS One 9(3):e92372. doi:10.1371/journal.pone.0092372
Konstantakopoulos G, Oulis P, Ploumpidis D et al (2014) Self-rated and performance-based empathy
in schizophrenia: the impact of cognitive deficits. Soc Neurosci 9(6):590–600
Kravariti E, Toulopoulou T, Mapua-Filbey F et al (2006) Intellectual asymmetry and genetic liability in
first-degree relatives of probands with schizophrenia. Br J Psychiatry 188:186–187
17
Kyaga S, Landén M, Boman M et al (2013) Mental illness, suicide and creativity: 40-year prospective
total population study. J Psychiatr Res 47(1):83–90
Lahera G, Boada L, Pousa E et al (2014) Movie for the Assessment of Social Cognition (MASC):
Spanish validation. J Autism Dev Disord 44(8):1886–1896. doi:10.1007/s10803-014-2061-6
Landgraf S, Osterheider M (2013) "To see or not to see: that is the question." The "Protection-
Against-Schizophrenia" (PaSZ) model: evidence from congenital blindness and visuo-cognitive
aberrations. Front Psychol 4:352
Landin-Romero R, McKenna PJ, Salgado-Pineda P et al (2014) Failure of deactivation in the default
mode network: a trait marker for schizophrenia? Psychol Med 21:1–11
Langdon R, Still M, Connors MH et al (2014) Jumping to delusions in early psychosis. Cogn
Neuropsychiatry 19(3):241–256. doi:10.1080/13546805.2013.854198
Lawson RP, Rees G, Friston KJ (2014) An aberrant precision account of autism. Front Hum Neurosci
8:302
Leitman DI, Foxe JJ, Butler PD et al (2005) Sensory contributions to impaired prosodic processing in
schizophrenia. Biol Psychiatry 58(1):56–61
Leitman DI, Sehatpour P, Higgins BA et al (2010) Sensory deficits and distributed hierarchical
dysfunction in schizophrenia. Am J Psychiatry 167(7):818–827
Leonard H, Glasson E, Nassar N et al (2011) Autism and intellectual disability are differentially
related to sociodemographic background at birth. PLoS One 6(3):e17875
Liljenström H (2003) Neural stability and flexibility: a computational approach.
Neuropsychopharmacology 28(Suppl 1):S64–S73
Li M, Deng W, He Z, Wang Q, Huang C, Jiang L, Gong Q, Ziedonis DM, King JA, Ma X, Zhang N, Li
T. (2015) A splitting brain: Imbalanced neural networks in schizophrenia. Psychiatry Res (in
press)
Lind S, Bowler D (2008) Episodic memory and autonoetic consciousness in autism spectrum
disorders: the roles of self-awareness, representational abilities, and temporal cognition. In:
Boucher JM, Bowler DM (eds) Memory in autism: theory and evidence. Cambridge University
Press, Cambridge UK, p 166–187
Lombardo MV, Chakrabarti B, Bullmore ET et al (2011) Specialization of right temporo-parietal
junction for mentalizing and its relation to social impairments in autism. Neuroimage
56(3):1832–1838
Lozano R, Hare EB, Hagerman RJ (2014) Modulation of the GABAergic pathway for the treatment of
fragile X syndrome. Neuropsychiatr Dis Treat 10:1769–1779
Madsen GF, Bilenberg N, Cantio C et al (2014) Increased prepulse inhibition and sensitization of the
startle reflex in autistic children. Autism Res 7(1):94–103. doi:10.1002/aur.1337
Malhotra D, Sebat J (2012) CNVs: harbingers of a rare variant revolution in psychiatric genetics. Cell
148(6):1223–1241. doi:10.1016/j.cell.2012.02.039
18
Markram K, Markram H (2010) The intense world theory - a unifying theory of the neurobiology of
autism. Front Hum Neurosci 4:224
Mason OJ, Brady F (2009) The psychotomimetic effects of short-term sensory deprivation. J Nerv
Ment Dis 197(10):783–785. doi:10.1097/NMD.0b013e3181b9760b
Matosin N, Newell KA (2013) Metabotropic glutamate receptor 5 in the pathology and treatment of
schizophrenia. Neurosci Biobehav Rev 37(3):256–268
McCormick LM, Brumm MC, Beadle JN et al (2012) Mirror neuron function, psychosis, and empathy
in schizophrenia. Psychiatry Res 201(3):233–239. doi:10.1016/j.pscychresns.2012.01.004
McLewin LA, Muller RT (2006) Attachment and social support in the prediction of psychopathology
among young adults with and without a history of physical maltreatment. Child Abuse Negl
30(2):171–191
Mehta UM, Thirthalli J, Basavaraju R et al (2014) Reduced mirror neuron activity in schizophrenia
and its association with theory of mind deficits: evidence from a transcranial magnetic
stimulation study. Schizophr Bull 40(5):1083–1094. doi:10.1093/schbul/sbt155
Montag C, Dziobek I, Richter IS et al (2011) Different aspects of theory of mind in paranoid
schizophrenia: evidence from a video-based assessment. Psychiatry Res 186(2-3):203–209.
doi:10.1016/j.psychres.2010.09.006
Moore JW, Pope A (2014) The intentionality bias and schizotypy. Q J Exp Psychol (Hove)
67(11):2218–2224
Morris R, Griffiths O, Le Pelley ME et al (2013) Attention to irrelevant cues is related to positive
symptoms in schizophrenia. Schizophr Bull 39(3):575–582. doi:10.1093/schbul/sbr192
Mottron L, Bouvet L, Bonnel A et al (2013) Veridical mapping in the development of exceptional
autistic abilities. Neurosci Biobehav Rev 37(2):209–228
Mottron L, Dawson M, Soulières I et al (2006) Enhanced perceptual functioning in autism: an update,
and eight principles of autistic perception. J Autism Dev Disord 36(1):27–43
Nagai T, Tada M, Kirihara K et al (2013) Mismatch negativity as a "translatable" brain marker toward
early intervention for psychosis: a review. Front Psychiatry 4:115
Nelson B, Rawlings D (2010) Relating schizotypy and personality to the phenomenology of creativity.
Schizophr Bull 36(2):388–399
Nesse RM (2004) Natural selection and the elusiveness of happiness. Philos Trans R Soc Lond B
Biol Sci 359(1449):1333–1347
Nesse RM, Ellsworth PC (2009) Evolution, emotions, and emotional disorders. Am Psychol
64(2):129–139
Nesse RM, Jackson ED (2011) Evolutionary foundations for psychiatric diagnosis: making DSM-V
valid. In: De Block A, Adriaens P (eds) Maladapting minds: philosophy, psychiatry, and
evolutionary theory. Oxford University Press, Oxford, p 167–191
Nesse RM, Stein DJ (2012) Towards a genuinely medical model for psychiatric nosology. BMC Med
19
10:5. doi:10.1186/1741-7015-10-5
Nettle D (2001) Strong imagination: madness, creativity and human nature. Oxford University Press,
UK
Nettle D (2006) Schizotypy and mental health amongst poets, visual artists, and mathematicians. J
Res Pers 40(6):876–890
Nettle D (2007). Empathizing and systemizing: what are they, and what do they contribute to our
understanding of psychological sex differences? Br J Psychology 98:237–255
Nettle D, Clegg H (2006) Schizotypy, creativity and mating success in humans. Proc Biol Sci
273(1586):611–615
O'Connor JA, Wiffen BD, Reichenberg A et al (2012) Is deterioration of IQ a feature of first episode
psychosis and how can we measure it? Schizophr Res 137(1-3):104–109
Oberman LM, Hubbard EM, McCleery JP et al (2005) EEG evidence for mirror neuron dysfunction in
autism spectrum disorders. Brain Res Cogn Brain Res 24(2):190–198
Orekhova EV, Stroganova TA (2014) Arousal and attention re-orienting in autism spectrum disorders:
evidence from auditory event-related potentials. Front Hum Neurosci 8:34
Palmer CJ, Paton B, Hohwy J et al (2013) Movement under uncertainty: the effects of the rubber-
hand illusion vary along the nonclinical autism spectrum. Neuropsychologia 51(10):1942–1951
Park S, Nasrallah HA (2014) The varieties of anomalous self experiences in schizophrenia: splitting
of the mind at a crossroad. Schizophr Res 152(1):1–4. doi:10.1016/j.schres.2013.11.036
Pathania M, Davenport EC, Muir J et al (2014) The autism and schizophrenia associated gene
CYFIP1 is critical for the maintenance of dendritic complexity and the stabilization of mature
spines. Transl Psychiatry 4:e374. doi:10.1038/tp.2014.16
Paton B, Hohwy J, Enticott PG (2012) The rubber hand illusion reveals proprioceptive and
sensorimotor differences in autism spectrum disorders. J Autism Dev Disord 42(9):1870–1883
Pellicano E, Burr D (2012) When the world becomes 'too real': a Bayesian explanation of autistic
perception. Trends Cogn Sci 16(10):504–510
Ploog BO (2010) Stimulus overselectivity four decades later: a review of the literature and its
implications for current research in autism spectrum disorder. J Autism Dev Disord
40(11):1332–1349
Plun-Favreau H, Lewis PA, Hardy J et al (2010) Cancer and neurodegeneration: between the devil
and the deep blue sea. PLoS Genet 6(12):e1001257. doi:10.1371/journal.pgen.1001257
Power RA, Kyaga S, Uher R et al (2013) Fecundity of patients with schizophrenia, autism, bipolar
disorder, depression, anorexia nervosa, or substance abuse vs their unaffected siblings.
JAMA Psychiatry 70(1):22–30. doi:10.1001/jamapsychiatry.2013.268
Qureshi AY, Mueller S, Snyder AZ et al (2014) Opposing brain differences in 16p11.2 deletion and
duplication carriers. J Neurosci 34(34):11199–11211. doi:10.1523/JNEUROSCI.1366-14.2014
20
Rawlings D, Locarnini A (2008) Dimensional schizotypy, autism, and unusual word associations in
artists and scientists. J Res Pers 42(2):465–471
Reed P, McCarthy J (2012) Cross-modal attention-switching is impaired in autism spectrum
disorders. J Autism Dev Disord 42(6):947–953. doi:10.1007/s10803-011-1324-8
Rees E, Kirov G, Sanders A et al (2014a) Evidence that duplications of 22q11.2 protect against
schizophrenia. Mol Psychiatry 19(1):37–40
Rees E, Walters JT, Georgieva L et al (2014b) Analysis of copy number variations at 15
schizophrenia-associated loci. Br J Psychiatry 204(2):108–114
Revheim N, Butler PD, Schechter I et al (2006) Reading impairment and visual processing deficits in
schizophrenia. Schizophr Res 87(1-3):238–245
Revheim N, Corcoran CM, Dias E et al (2014) Reading deficits in schizophrenia and individuals at
high clinical risk: relationship to sensory function, course of illness, and psychosocial outcome.
Am J Psychiatry 171(9):949–959
Russell-Smith SN, Maybery MT, Bayliss DM (2010) Are the autism and positive schizotypy spectra
diametrically opposed in local versus global processing? J Autism Dev Disord 40:968–977
Russo DA, Stochl J, Painter M et al (2014) Trauma history characteristics associated with mental
states at clinical high risk for psychosis. Psychiatry Res pii:S0165-1781(14)00725-2.
doi:10.1016/j.psychres.2014.08.028
Rutter M (1972) Childhood schizophrenia reconsidered. J Autism Dev Disord 2:315–337
Samson F, Mottron L, Soulières I et al (2012) Enhanced visual functioning in autism: an ALE meta-
analysis. Hum Brain Mapp 33(7):1553–1581. doi:10.1002/hbm.21307
Sasson NJ, Touchstone EW (2014) Visual attention to competing social and object images by
preschool children with autism spectrum disorder. J Autism Dev Disord 44(3):584–592.
doi:10.1007/s10803-013-1910-z
Sharp C, Pane H, Ha C et al (2011) Theory of mind and emotion regulation difficulties in adolescents
with borderline traits. J Am Acad Child Adolesc Psychiatry 50(6):563–573.e1.
doi:10.1016/j.jaac.2011.01.017
Silverstein SM, Wang Y, Keane BP (2013) Cognitive and neuroplasticity mechanisms by which
congenital or early blindness may confer a protective effect against schizophrenia. Front
Psychol 3:624
Simeonova DI, Chang KD, Strong C et al (2005) Creativity in familial bipolar disorder. J Psychiatr Res
39(6):623–631
Singh S, Kumar A, Agarwal S et al (2014) Genetic insight of schizophrenia: past and future
perspectives. Gene 535(2):97–100. doi:10.1016/j.gene.2013.09.110
Soni S, Whittington J, Holland AJ et al (2008) The phenomenology and diagnosis of psychiatric
illness in people with Prader-Willi syndrome. Psychol Med 38(10):1505–1514
South M, Chamberlain PD, Wigham S, Newton T, Le Couteur A, McConachie H, Gray L, Freeston M,
21
Parr J, Kirwan CB, Rodgers J. (2014) Enhanced decision making and risk avoidance in high-
functioning autism spectrum disorder. Neuropsychology 28(2):222-228.
Speechley WJ, Whitman JC, Woodward TS (2010) The contribution of hypersalience to the "jumping
to conclusions" bias associated with delusions in schizophrenia. J Psychiatry Neurosci
35(1):7–17
Spek AA, Velderman E (2013) Examining the relationship between Autism spectrum disorders and
technical professions in high functioning adults. Res Autism Spectr Disord 7(5):606–612
Stein DJ (2013) What is a mental disorder? A perspective from cognitive-affective science. Can J
Psychiatry 58(12):656–662
Swerdlow NR, Light GA, Sprock J et al (2014) Deficient prepulse inhibition in schizophrenia detected
by the multi-site COGS. Schizophr Res 152(2-3):503–512
Swettenham J, Remington A, Murphy P et al (2014) Seeing the unseen: autism involves reduced
susceptibility to inattentional blindness. Neuropsychology 28(4):563–570.
doi:10.1037/neu0000042
Tang J, Liao Y, Song M et al (2013) Aberrant default mode functional connectivity in early onset
schizophrenia. PLoS One 8(7):e71061. doi:10.1371/journal.pone.0071061
Tavassoli T, Miller LJ, Schoen SA et al (2014) Sensory over-responsivity in adults with autism
spectrum conditions. Autism 18(4):428–432
Todd J, Harms L, Schall U et al (2013) Mismatch negativity: translating the potential. Front Psychiatry
4:171
Trull TJ, Widiger TA (2013) Dimensional models of personality: the five-factor model and the DSM-5.
Dialogues Clin Neurosci 15(2):135–146
van Os J (2009) 'Salience syndrome' replaces 'schizophrenia' in DSM-V and ICD-11: psychiatry's
evidence-based entry into the 21st century? Acta Psychiatr Scand 120(5):363–372.
doi:10.1111/j.1600-0447.2009.01456.x
von dem Hagen EA, Stoyanova RS, Baron-Cohen S et al (2013) Reduced functional connectivity
within and between 'social' resting state networks in autism spectrum conditions. Soc Cogn
Affect Neurosci 8(6):694–701
Walker CM, Gopnik A (2013) Causality & imagination. In: Taylor M (ed) The development of
imagination. Oxford University Press, New York, p 342–358
Werner S, Malaspina D, Rabinowitz J (2007) Socioeconomic status at birth is associated with risk of
schizophrenia: population-based multilevel study. Schizophr Bull 33(6):1373–1378
Westerhausen R, Kompus K, Hugdahl K (2011) Impaired cognitive inhibition in schizophrenia: a
meta-analysis of the Stroop interference effect. Schizophr Res 133(1-3):172–181.
doi:10.1016/j.schres.2011.08.025
Wheelwright S, Baron-Cohen S (2001) The link between autism and skills such as engineering,
maths, physics and computing: a reply to Jarrold and Routh. Autism 5(2):223–227
22
Whitfield-Gabrieli S, Thermenos HW, Milanovic S et al (2009) Hyperactivity and hyperconnectivity of
the default network in schizophrenia and in first-degree relatives of persons with
schizophrenia. Proc Natl Acad Sci U S A 106(4):1279–1284. doi:10.1073/pnas.0809141106
Wible CG (2012) Hippocampal temporal-parietal junction interaction in the production of psychotic
symptoms: a framework for understanding the schizophrenic syndrome. Front Hum Neurosci
6:180
Williams EL, Casanova MF (2010) Autism and dyslexia: a spectrum of cognitive styles as defined by
minicolumnar morphometry. Med Hypotheses 74(1):59–62
Winfield H, Kamboj SK (2010) Schizotypy and mental time travel. Conscious Cogn 19(1):321–327.
doi:10.1016/j.concog.2009.11.011
Wing L, Gould J (1979) Severe impairments of social interaction and associated abnormalities in
children: epidemiology and classification. J Autism Dev Disord 9:11–29
Winton-Brown TT, Fusar-Poli P, Ungless MA et al (2014) Dopaminergic basis of salience
dysregulation in psychosis. Trends Neurosci 37(2):85–94. doi:10.1016/j.tins.2013.11.003
Woodard CR, Van Reet J (2011) Object identification and imagination: an alternative to the meta-
representational explanation of autism. J Autism Dev Disord 41(2):213–226
23
Figure 1. Phenotypes that describe (a) the autism spectrum and (b) the psychotic-affective spectrum,
based on DSM-V diagnoses, evolutionary considerations, and the hypothesized relationships
between the two sets of disorders.
24
25
Figure 2. The autism spectrum and the psychotic-affective spectrum can be conceptualized as
diametric disorders, with regard to the direction of alterations in uniquely-human or human-elaborated
phenotypes that comprise their core features.
26
Table 1. Diametric genetic risk factors, phenotypes, and correlates of autism spectrum and psychotic affective spectrum conditions.
For phenotypes with large sets of evidence, only recent articles or reviews are cited. Crespi and Badcock (2008) present additional
evidence, from less-recent literature.
Trait Autism spectrum Psychotic-affective spectrum Comments
Copy number
variants
Duplications of 22q11.2 increase
autism risk (Crespi and Crofts
2012; Rees et al. 2014)
Duplications of 22q11.2
decrease schizophrenia risk;
deletions of 22q11.2 greatly
increase schizophrenia risk
(Rees et al. 2014)
Deletions of 22q11.2 suggested
to increased ASD risk but
pattern not found in ASD CNV
cohorts (Crespi and Crofts 2102)
Copy number
variants
Duplications of 1q21.1 increase
autism risk, increase head size
(Brunetti-Pierri et al. 2008;
Crespi and Crofts 2012)
Deletions of 1q21.1 increase
schizophrenia risk, reduce head
size (Brunetti-Pierri et al. 2008;
Rees et al. 2014b)
Deletions may increase autism
risk, or be false positive (Crespi
and Crofts 2012)
Copy number
variants
Deletions of 16p11.2 increase
autism risk, increase head size
(Qureshi et al. 2014)
Duplications of 16p11.2 increase
schizophrenia risk, reduce head
size (Rees et al. 2014b; Qureshi
et al. 2014)
Duplications may increase
autism risk, or be false positive
(Crespi and Crofts 2012)
Copy number
variants
Duplications of 15q11.2 (BP1-
BP2) increase autism risk
(Chaste et al. 2014)
Deletions of 15q11.2 (BP1-BP2)
increase schizophrenia risk
(Rees et al. 2014b)
Deletions and duplications of
CYFIP1, a key gene in this CNV
region, cause opposite
alterations to dendritic spine
complexity (Pathania et al. 2014)
27
Birth size (weight,
length)
Smaller size protects against
autism; larger size increases
autism risk (Byars et al. 2014)
Larger size protects against
schizophrenia; smaller size
increases schizophrenia risk
(Byars et al. 2014)
Each of the patterns of risk has
been replicated across many
other studies
Brain size Larger brain size in children with
autism (Courchesne et al. 2011;
Baribeau and Anagnostou 2014)
Smaller brain size in
schizophrenia (Haijma et al.
2013)
Autism involves faster brain
growth in early childhood, in
particular
Neurological
function
Congenital blindness increases
risk of autism (Hobson and
Bishop 2003; Ek et al. 2005)
Congenital blindness protects
against schizophrenia (Landgraf
and Osterheider 2013;
Silverstein et al. 2013)
Neurological
function
Sensory abilities increased in
autism (Brown et al. 2003;
Mottron et al. 2006, 2013;
Heaton et al. 2008a,b; Dohn et
al. 2012; Falter et al. 2013;
Tavassoli et al. 2014)
Sensory abilities decreased in
schizophrenia; sensory
deprivation induces features of
psychosis (Bates 2005; Leitman
et al. 2005, 2010; Force et al.
2008; Javitt 2009a,b; Mason and
Brady 2009; Daniel et al. 2014)
Strong, highly consistent pattern
in schizophrenia; substantial
although somewhat mixed
evidence in autism
Neurological
function
Prepulse inhibition increased in
autism (Kohl et al. 2014; Madsen
et al. 2014)
Prepulse inhibition decreased
(Swerdlow et al. 2014)
Findings highly consistent for
schizophrenia, variable for
autism
28
Neurological
function
Mismatch negativity increased in
autism (Orekhova and
Stroganova 2014)
Mismatch negativity decreased
in schizophrenia (Nagai et al.
2013; Todd et al. 2013)
Findings highly consistent for
schizophrenia, variable for
autism
Neurological
function
Mirror neuron system activation
decreased in autism (Oberman
et al. 2005; Kana et al. 2011)
Mirror neuron system activation
increased in actively psychotic
individuals with schizophrenia
(McCormick et al. 2012)
Same protocol used to measure
mirror neuron function, in autism
and schizophrenia (McCormick
et al. 2012); other studies of
schizophrenia usually show
reduced activation (Mehta et al.
2014) but do not involve actively-
psychotic subjects
Neurological
function
Default mode system activation
reduced in autism, in association
with reduced self-referential and
imaginative cognition (Kennedy
et al. 2006; Buckner et al. 2008;
Iocaboni 2006; Kennedy and
Courchesne 2008)
Default system over-activated in
schizophrenia, in association
with reality distortion and
increased imaginative cognition
(Buckner et al. 2008); also less
deactivation of this system
(Landin-Romero et al. 2014)
Some studies of autism show
reduced deactivations of default
system, that may be associated
with reduced activation (Buckner
et al. 2008); Immordino-Yang et
al. (2012) also contrast autism
and schizophrenia as opposite
with regard to the default
network
29
Neurological
function
Reduced connectivity within
default mode in autism (van dem
Hagen et al. 2012; Jung et al.
2014)
Increased connectivity within
default mode in schizophrenia
(Whitfield-Gabrieli et al. 2009;
Tang et al. 2013; Li et al. 2015)
Some mixed results in both
autism and schizophrenia, but
two reviews support opposite
nature of the alterations (Broyd
et al. 2009; Karbasforoushan
and Woodward 2012)
Neurological
function
Increased local brain
connectivity, decreased long-
range connectivity, in
association with early brain
overgrowth (Baribeau et al.
2013)
Decreased local brain
connectivity, increased long-
range connectivity, in
association with increased
cortical thinning, in childhood-
onset schizophrenia (Baribeau et
al. 2013)
Findings based on review of
neuroimaging findings (Baribeau
et al. 2013)
Neurological
function
Temporal-parietal junction region
shows reduced activation in
autism, underlies mentalizing
reductions (Lombardo et al.
2011; Kana et al. 2014)
Temporal-parietal junction region
shows increased activation in
schizophrenia, underlies some
psychotic symptoms (Wible
2012)
Emotionality and
motivation
Reduced social motivation in
autism (Chevallier et al. 2012)
Increased social motivation in
mania, hypomania (Johnson et
al. 2012a,b)
Motivation in general decreased
in negative symptom
schizophrenia, depression
30
Emotionality and
motivation
Cognitive empathic abilities
reduced in autism (Baron-Cohen
2010)
Some cognitive empathic
abilities enhanced in borderline
personality disorder and
subclinical depression (Harkness
et al. 2011; Dinsdale and Crespi
2013)
Cognitive empathic abilities
lower in schizophrenia, bipolar
disorder and depression, in
association with general
cognitive deficits (e. g., Baez et
al. 2013; Konstantakopoulos et
al. 2014
Emotionality and
motivation
Reduced social emotion in
autism (Kasari et al. 2001)
Increased social emotion
expression in bipolar disorder
and depression (e. g., guilt,
shame, embarrassment, pride)
(Kim et al. 2011; Johnson and
Carver 2012)
Reduced general expressed
emotionality in negative
symptom schizophrenia
Cognitive function Decreased inattentional
blindness in autism
(Swettenham et al. 2014)
Increased inattentional blindness
in schizophrenia (Hanslmayr et
al. 2012)
Cognitive function Over-selective attention (Ploog
2010; Reed and McCarthy 2012)
Reductions in selective attention
in schizophrenia and positive
schizotypy (Morris et al. 2013;
Granger et al. 2012)
Cognitive function Enhanced Stroop task
performance in autism (Adams
and Jarrold 2009)
Decreased Stroop task
performance in schizophrenia,
by meta-analysis (Westerhausen
et al. 2011)
Results mixed for autism, highly
consistent for schizophrenia
31
Cognitive function Enhanced Iowa Gambling Task
performance in high-functioning
autism (South et al. 2013)
Reduced Iowa Gambling Task
performance in schizophrenia, in
most studies (Adida et al. 2011
Results mixed for autism,
consistent for schizophrenia
Cognitive function Reduced susceptibility to rubber
hand illusion in autism and in
healthy high-ASD trait
individuals (Cascio et al. 2012;
Paton et al. 2012; Palmer et al.
2013)
Increased susceptibility to rubber
hand illusion in schizophrenia
(Park and Nasrallah 2014)
Same general pattern also found
for visual illusions, with some
inconsistencies (Crespi 2013)
Cognitive function Literal word interpretation,
under-interpretation of social
relevance, in autism (Chance
2014)
Over-interpretation of word
meaning and social relevance in
schizophrenia (Chance 2014)
Cognitive function Decreased induction of false
memories (Beversdorf et al.
2000; Hiller et al. 2007)
Increased induction of false
memories associated with
psychosis phenotypes (Corlett et
al. 2009; Kanemoto et al. 2013;
Grant et al. 2014)
Results somewhat mixed (some
non-significant) for autism
Cognitive function Semantic memory network
states overly rigid in autism
(Faust and Kenett 2014)
Semantic memory network
states chaotic in schizophrenia
(Faust and Kenett 2014)
32
Cognitive function Working memory deficits in
autism (Kercood et al. 2014);
extraordinary working memory
enhancements in child prodigies,
who score above autism range
in Attention to Detail on Autism
Quotient test, and exhibit high
rates of autism in their families
(Tuthsatz and Urbach 2012)
Large working memory deficits in
schizophrenia; highly consistent
finding (Lee and Park 2005;
Silver Feldman 2014)
Findings of Tuthsatz and Urbach
(2012) would benefit from
replication; areas of excellence
in child prodigies notably overlap
with those found in savantism in
autism (Treffert 2009)
Cognitive function Hyperlexia found predominantly
in autism (Cardoso-Martins and
da Silva 2010; Samson et al.
2012; Mottron et al. 2013)
Dyslexia associated with
schizophrenia and schizotypy
(Revheim et al. 2006, 2014;
Arnott et al. 2011)
Williams and Casanova (2010)
contrast autism and dyslexia for
cortical microstructure
Cognitive function More-deliberative decision-
related processing in autism
(Brosnan et al. 2014)
'Jumping to conclusions'
associated with delusions in
schizophrenia (Speechley et al.
2010; Langdon et al. 2014)
Cognitive function Bias towards hypo-priors in
Bayesian models of perception
and cognition (Pellicano and
Burr 2012; Lawson et al. 2014)
Bias towards hyper-priors in
Bayesian models of perception
and cognition (Cook et al. 2012;
Pellicano and Burr 2012)
33
Cognitive function Reduced inference of intentions
in autism (Ciaramidaro et al.
2014)
'Hyper-intenionality' in
schizophrenia and schizotypy
(Backasch et al. 2013; Moore
and Pope 2014
Bara et al. (2011) contrast
autism and schizophrenia
directly in this regard
Cognitive function Reduced theory of mind in
autism spectrum children by
ToM Storybooks test (Blijd-
Hoogewys et al. 2008)
'Hyper-Theory-of Mind' in
children with more psychotic
experiences by ToM Storybooks
test (Clemmensen et al. 2014)
Cognitive function Theory of mind abilities reduced
in autism, using MASC test, due
to combination of hypo-
mentalizing, lack of mentalizing,
and hyper-mentalizing (Dziobek
et al. 2006; Lahera et al. 2014)
Theory of mind abilities reduced
in association with positive
symptoms of schizophrenia,
using MASC test, due to hyper-
mentalizing (Montag et al. 2011;
Fretland et al. 2015); hyper-
mentalizing also found in
borderline personality disorder
using MASC (Sharp et al. 2011)
34
Cognitive function Reduced salience of social
stimuli, and overly-specific and
inflexible salience of primary
perceptual and non-social stimuli
(Bird et al. 2006; Sasson and
Touchstone 2014)
Over-developed and arbitrary
salience in prodrome and
psychosis, mainly involving
social phenomena (van Os
2009; Winton-Brown et al. 2013;
Howes and Murray 2014)
Cognitive function Decreased perception of
biological motion, entities, in
autism; fail to see humans who
are there (Blake Turner Smoski
2003)
Increased and false perception
of biological motion, entities, in
schizophrenia; see humans in
random dots (Kim Park Blake
2011)
Cognitive function Selectively enhanced visual-
spatial abilities in autism
(Almeida et al. 2013; Kana et al
2013).
Reduced visual-spatial skills,
relative to verbal skills, positively
associated with genetic liability
to schizophrenia (Kravariti et al.
2006; also see O'Connor et al.
2012)
Cognitive function Enhanced Embedded Figures
Test performance among
healthy individuals with more
autistic traits (Russell-Smith et
al. 2010)
Reduced Embedded Figures
Test performance among
healthy individuals with more
positively-schizotypal traits
(Russell-Smith et al. 2010)
35
Behavior Reduced imagination and
creativity in autism (Craig and
Baron-Cohen 1999; King et al.
2014)------------------------------------
-------------------
Increased imagination and
creativity, in schizophrenia,
schizotypy, and bipolar disorder
and in relatives (Jamison 1993;
Nettle 2001; Nelson and
Rawlings 2008; Claridge and
McDonald 2009; Kaufman 2014)
---------------------------
The literature relating psychotic-
affective spectrum phenotypes
and conditions to aspects of
increased imagination and
creativity is large and diverse;
reduced imagination has been
considered as a diagnostic
criterion for autism
Behavior Reduced pretend play and social
play in autism (Jarrold et al.
2003; Hobson et al. 2013)
Higher levels of dissociation,
hallucination, psychotic-affective
psychopathology associated with
presence of childhood imaginary
companions (Bonne et al.1999;
Gleason et al. 2003; McLewin
and Muller 2006; Fernyhough et
al. 2007)
-------------------------------------------
-------------------------------------------
---
36
Social correlates Autism associated with technical
professions in fathers, mothers
and grandfathers (Wheelwright
and Baron-Cohen 2001; Spek
and Velderman 2013; Dickerson
et al. 2014)
Schizophrenia, schizotypy,
bipolar disorder, and depression
associated with careers and
interests in arts, humanities and
literature (Nettle and Clegg
2006; Rawlings and Locarnini
2008)
Social correlates Autism in family associated with
technical college majors
(Campbell and Wang
2012)---------
Bipolar disorder, depression in
family associated with arts and
humanities majors (Campbell
and Wang 2012)
----------------------------------
Insufficient data on
schizophrenia for analysis, in
this study
Social correlates Autism associated with higher
socioeconomic status (Durkin et
al. 2010; Leonard et al. 2011)
Schizophrenia associated with
lower socioeconomic status
(Werner et al. 2007)
37
... The reason to select the above five mental distresses is, they all have a similar like spectrum and often get confused until a very late stage. Especially symptoms like brain fogging, acute stress or anxiety, memory-related issues, and social and emotional withdrawal are there in each spectrum of the selected disorders (Arya et al. 2018;Jankovic 2008;Crespi 2016). In this chapter, we have tried to explore and intensify the smallest difference among the selected mental disorders. ...
Chapter
Full-text available
Schizophrenia is the seventh largest cause of impairment in people aged from 15 to 44 years around the world. This is the long-term psychological disorder marked by illusions, thinking and attention difficulties, absence of desire along with that the majority will have significant and long-term social impairments. Generally, schizophrenia is associated with positive as well as negative or cognitive symptoms. The conventional antipsychotics are used to treat the positive symptoms but the negative symptoms are rarely treated. There were no effective drugs for the treatment of schizophrenia before the introduction of antidopaminergic psychotropic drugs in the 1950s. Here, we described the key meta-analytic evidence on the efficacy of antipsychotics in the acute treatment of schizophrenia, particularly clozapine in treatment-resistant patients. In this chapter, primarily we have focused on the neuropharmacological treatment options available for schizophrenia and how has the treatment changed over time. An improved understanding of how conventional antipsychotic drugs convey their therapeutic effect during the treatment and what are the cutting-edge alternatives that are available in order to mitigate the drawbacks of traditional neuroleptics.
... Second, according to the diametric model of social brain disorders (Crespi and Badcock, 2008;Crespi and Go, 2015;Crespi, 2016Crespi, , 2019Crespi, , 2020, autism and psychosis can be conceptualized as polar ends of a continuum of social cognition development, with normality at its center. Thus, autism is characterized by low mentalistic thought (i.e., social cognition) and high mechanistic thought (i.e., non-social cognition), and psychotic-affective conditions, such as schizotypy, are characterized by high mentalistic and low mechanistic thought (Crespi and Badcock, 2008). ...
Book
Full-text available
This is a collection of 21 articles published as an eBook in Frontiers in Psychology. This Research Topic aims to demonstrate that imaginative culture is an important functional part of evolved human behavior—diverse in its manifestations but unified by species-typical sets of biologically grounded motives, emotions, and cognitive dispositions. The topic encompasses four main areas of research in the evolutionary human sciences: (1) evolutionary psychology and anthropology, which have fashioned a robust model of evolved human motives organized systemically within the phases and relationships of human life history; (2) research on gene-culture coevolution, which has illuminated the mechanisms of social cognition and the transmission of cultural information; (3) the psychology of emotions and affective neuroscience, which have gained precise knowledge about the evolutionary basis and neurological character of the evolved emotions that give power to the arts, religion, and ideology; and (4) cognitive neuroscience, which has identified the Default Mode Network as the central neurological location of the human imagination. By integrating these four areas of research and by demonstrating their value in illuminating specific kinds of imaginative culture, this Research Topic aims at incorporating imaginative culture within an evolutionary conception of human nature.
... Both autistic-like traits and positive schizotypy have been associated with impaired theory-of-mind abilities (i.e., the ability to infer other people's intentions and mental states), but for very different reasons (Crespi, 2016;Crespi & Badcock, 2008;Crespi & Go, 2015). Autism and autistic-like traits are associated with hypomentalizing, meaning that there is underattribution of mental states to other people (Abu-Akel et al., 2015;Baron-Cohen, 2000). ...
Article
According to the predictive-processing framework, only prediction errors (rather than all sensory inputs) are processed by an organism’s perceptual system. Prediction errors can be weighted such that errors from more reliable sources will be more influential in updating prior beliefs. It has previously been argued that autism-spectrum conditions can be understood as resulting from a predictive-processing mechanism in which an inflexibly high weight is given to sensoryprediction errors that results in overfitting their predictive models to the world. Deficits in executive functioning, theory of mind, and central coherence are all argued to flow naturally from this core underlying mechanism. The diametric model of autism and psychosis suggests a simple extension of this hypothesis. If people on the autism spectrum give an inflexibly high weight to sensory input, could it be that people with a predisposition to psychosis (i.e., people high in positive schizotypy) give an inflexibly low weight to sensory input? In this article I argue that evidence strongly supports this hypothesis. An inflexibly low weight given to sensory input can explain such disparate features of positive schizotypy as increased exploratory behavior, apophenia, hyper theory of mind, hyperactive imagination, attentional differences, and having idiosyncratic worldviews.
... Second, according to the diametric model of social brain disorders (Crespi and Badcock, 2008;Crespi and Go, 2015;Crespi, 2016Crespi, , 2019Crespi, , 2020, autism and psychosis can be conceptualized as polar ends of a continuum of social cognition development, with normality at its center. Thus, autism is characterized by low mentalistic thought (i.e., social cognition) and high mechanistic thought (i.e., non-social cognition), and psychotic-affective conditions, such as schizotypy, are characterized by high mentalistic and low mechanistic thought (Crespi and Badcock, 2008). ...
Article
Full-text available
Video games are popular and ubiquitous aspects of human culture, but their relationships to psychological and neurophysiological traits have yet to be analyzed in social-evolutionary frameworks. We examined the relationships of video game usage, motivations, and preferences with autistic and schizotypal traits and two aspects of neurophysiology, reaction time and targeting time. Participants completed the Autism Quotient, Schizotypal Personality Questionnaire, a Video Game Usage Questionnaire, and two neurophysiological tasks. We tested in particular the hypotheses, motivated by theory and previous work, that: (1) participants with higher autism scores would play video games more, and participants with higher schizotypy scores would play video games less; and (2) autism and positive schizotypy would be associated with opposite patterns of video game use, preferences and motivations. Females, but not males, with higher autism scores played more video games, and exhibited evidence of relatively male-typical video game genre preferences and motivations. By contrast, positive schizotypy was associated with reduced video game use in both genders, for several measures of game use frequency. In line with previous findings, males played video game more than females did overall, preferred action video games, and exhibited faster reaction and targeting times. Females preferred Puzzle and Social Simulation games. Faster reaction and targeting times were associated with gaming motives related to skill development and building behavior. These findings show that gaming use and patterns reflect aspects of psychology, and gender, related to social cognition and imagination, as well as aspects of neurophysiology. More generally, the results suggest that video game use is notably affected by levels of autistic and schizotypal traits, and that video games may provide an evolutionarily novel medium for imaginative play in which immersive play experiences can be decoupled from social interaction.
... Hood, 1998), only that OCD is an extreme, yet normal, perturbation in an evolved cognitive system (Boyer & Lienard, 2006, 2008 and we all fall somewhere along this spectrum as a need for ordering a chaotic world. This is a typical characterization within cognitive science, linking psychiatric categories with normal variation in specific cognitive systems (Crespi, 2016) and has been successful in other domains, linking, for example, autism spectrum conditions and traits to normal 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 variation in the theory of mind system (ibid). To further support this premise, it has also been suggested that aspects of scientific behavior can be explained through this model of ritualized behavior (cf. ...
Article
Full-text available
The ritual handling of serpents remains an unnoticed cultural form for the explanatory aims and theoretical insights desired by cognitive scientists of religion. In the current article, we introduce the Hood and Williams archives at The University of Tennessee at Chattanooga that contains data culled from Hood's 40-plus year career of studying serpent handlers. The archives contain hundreds of hours of interviews and recordings of speaking in tongues, handling fire, drinking poison, and taking up serpents by different congregants and congregations. The archive remains a rich but untapped source of data for building, testing, and refining cognitive theories of ritual in general, and serpent handling in specific. We connect Hood's work to current cognitive theories and engage critically with research on the social functions of ritual. Finally, we discuss several further reasons to pay more attention to SHS communities and practices in cognitive theories of ritual.
... Such later delays may be foreshadowed in our nding of a plateauing of CBRs in HR infants from 9.5 to 12 months. An alternative interpretation for these ndings may fall closer in line with evo-devo informed neurodiversity frameworks suggesting that autistic characteristics may have been positively selected as evolved compensatory adaptations (Crespi, 2016). Perhaps infants with an elevated likelihood of disorder (such as the HR infants in this study) ...
Preprint
Full-text available
There is a growing body of research emphasizing the role of social and endogenous motivations in human development. The present study evaluated canonical babbling across the second-half year of life using all-day recordings of 98 children with typical or elevated likelihoods of autism i.e., at “low risk” or “high risk”, respectively. Canonical babbling ratios (CBRs) were calculated from human coding along with Likert-scale ratings on vocal turn taking and vocal play in each segment. We observed no main effect of risk on CBRs. CBRs were significantly elevated during high vocal play. High turn taking yielded a weaker significant effect. We conclude that both social and endogenous motivations may drive infants’ tendencies to produce their most advanced vocal forms.
... Differences in life history strategies are partly under genetic control but it appears that the nature and quality of the individual's early environment may also be important (Belsky et al., 1991;Ellis et al., 2011) (see Barbaro et al., 2016 for a different perspective). This renders LHT important for the understanding BOX 2.1 Pathways for the persistence of disease and disorder (Adapted from Gluckman et al. (2009) andCrespi (2016); for definition of terms see glossary: www.rcpsych. ac.uk/docs/default-source/members/sigs/evolutionary-psychiatry-epsig/evolutionary-psychiatry-glossary-2. ...
Chapter
Full-text available
In an ideal world, psychiatry would be appropriately informed by evolutionary theory and we expect that this will happen one day. In this chapter salient evolutionary contributions regarding the definition and demarcation of mental disorder are considered as well as evolutionary proposals for rethinking the classification of mental disorder. Although evolutionary analysis of mental disorders rests on a set of principles shared with the rest of medicine, these principles remain unfamiliar to the majority of psychiatrists. We examine these principles and advocate consideration of the phylogenetic or ultimate causes of disorders and how Tinbergen’s four questions may be applied to psychiatry. Leading evolutionary literature on a number of psychiatric conditions is explored including: schizophrenia spectrum disorders, depression, alcohol and drug addictions, eating disorders and others. Some preliminary evolutionary thoughts on placebo responses, pharmacological treatments and future research are also presented in brief. Mainstream psychiatry, like the rest of medicine, continues to focus on proximate causation of disease and disorder. Evolutionary science has not permeated the syllabus of most undergraduate or postgraduate medical and psychiatric education, with a few notable exceptions. Consequently, most psychiatrists have scant understanding of even the basics of evolution let alone its impact on the behavioural sciences. Therefore, the majority of psychiatrists and other mental health professionals remain largely unaware of the potential of Darwinian theory to further our understanding of human vulnerability to mental disorder. Evolutionary work consequently has not yet significantly influenced the mainstream research agenda in mental health nor has it had much impact on clinical practice. Various barriers to the incorporation and dissemination of evolutionary thinking into psychiatry are discussed, some of which are shared by the rest of medicine. Finally, we contemplate the potential contributions evolutionary science can generate for both theory and practice of psychiatry and advocate for pertinent areas of evolutionary biology to be taught at both the undergraduate and postgraduate levels as a basic science.
Chapter
Schizophrenia (SZD) is one of the chronic disorders but is not very common. The symptoms of SZD are broadly divided into two types: positive and negative symptoms. Still SZD is hard to diagnose at an initial level because it shares some similar characteristics with other disorders like Alzheimer’s (AD), Parkinson’s (PD), or chronic and manic depression. Characterizing it from other mental disorders is one of the most vital works. In this chapter, we have covered five major mental disorders which share some or the other common characteristics. We have selected SZD, AD, PD, chronic depression, and Bi-polar disorder. Starting from their definition to its severity, we have highlighted different characteristics like biological, psychological, and cognitive characteristics of SZD, and how it is different from other disorders. Wherever possible we have also linked more than two disorders and ways to differentiate them from each other. We have also discussed the clinical methods to diagnose each selected mental disorder and is there an effective way to treat it or not. This chapter sheds a limited light on each selected disorder.
Chapter
Full-text available
Evolutionary psychiatry attempts to explain and examine the development and prevalence of psychiatric disorders through the lens of evolutionary and adaptationist theories. In this edited volume, leading international evolutionary scholars present a variety of Darwinian perspectives that will encourage readers to consider 'why' as well as 'how' mental disorders arise. Using insights from comparative animal evolution, ethology, anthropology, culture, philosophy and other humanities, evolutionary thinking helps us to re-evaluate psychiatric epidemiology, genetics, biochemistry and psychology. It seeks explanations for persistent heritable traits shaped by selection and other evolutionary processes, and reviews traits and disorders using phylogenetic history and insights from the neurosciences as well as the effects of the modern environment. By bridging the gap between social and biological approaches to psychiatry, and encouraging bringing the evolutionary perspective into mainstream psychiatry, this book will help to inspire new avenues of research into the causation and treatment of mental disorders.
Article
Full-text available
Why do genes for mental disorders persist? This article argues it is because traits with a cliff-edged fitness function are shaped to a performance peak where some individuals fall off.
Article
Full-text available
We present the Spanish validation of the "Movie for the Assessment of Social Cognition" instrument (MASC-SP). We recruited 22 adolescents and young adults with Asperger syndrome and 26 participants with typical development. The MASC-SP and three other social cognition instruments (Ekman Pictures of Facial Affect test, Reading the Mind in the Eyes Test, and Happé's Strange Stories) were administered to both groups. Individuals with Asperger syndrome had significantly lower scores in all measures of social cognition. The MASC-SP showed strong correlations with all three measures and relative independence of general cognitive functions. Internal consistency was optimal (0.86) and the test-retest was good. The MASC-SP is an ecologically valid and useful tool for assessing social cognition in the Spanish population.
Article
Full-text available
Copy number variation (CNV) at the 15q11.2 region has been identified as a significant risk locus for neurological and neuropsychiatric conditions such as schizophrenia (SCZ) and autism spectrum disorder (ASD). However, the individual roles for genes at this locus in nervous system development, function and connectivity remain poorly understood. Haploinsufficiency of one gene in this region, Cyfip1, may provide a model for 15q11.2 CNV-associated neuropsychiatric phenotypes. Here we show that altering CYFIP1 expression levels in neurons both in vitro and in vivo influences dendritic complexity, spine morphology, spine actin dynamics and synaptic α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor lateral diffusion. CYFIP1 is highly enriched at synapses and its overexpression in vitro leads to increased dendritic complexity. Neurons derived from Cyfip1 heterozygous animals on the other hand, possess reduced dendritic complexity, increased mobile F-actin and enhanced GluA2-containing AMPA receptor mobility at synapses. Interestingly, Cyfip1 overexpression or haploinsufficiency increased immature spine number, whereas activity-dependent changes in spine volume were occluded in Cyfip1 haploinsufficient neurons. In vivo, Cyfip1 heterozygous animals exhibited deficits in dendritic complexity as well as an altered ratio of immature-to-mature spines in hippocampal CA1 neurons. In summary, we provide evidence that dysregulation of CYFIP1 expression levels leads to pathological changes in CNS maturation and neuronal connectivity, both of which may contribute to the development of the neurological symptoms seen in ASD and SCZ.
Article
Children are widely celebrated for their imaginations, but developmental research on this topic has often been fragmented or narrowly focused on fantasy. However, there is growing appreciation for the role that imagination plays in cognitive and emotional development, as well as its link with children’s understanding of the real world. With their imaginations, children mentally transcend time, place, and/or circumstance to think about what might have been, plan and anticipate the future, create fictional relationships and worlds, and consider alternatives to the actual experiences of their lives. The Oxford Handbook of the Development of Imagination provides a comprehensive overview of this broad new perspective by bringing together leading researchers whose findings are moving the study of imagination from the margins of mainstream psychology to a central role in current efforts to understand human thought. The topics include fantasy-reality distinctions, pretend play, magical thinking, narrative, anthropomorphism, counterfactual reasoning, mental time travel, creativity, paracosms, imaginary companions, imagination in non-human animals, the evolution of imagination, autism, dissociation, and the capacity to derive real life resilience from imaginative experiences. Many of the chapters include discussions of the educational, clinical, and legal implications of the research findings and special attention is given to suggestions for future research.
Article
Creativity is considered a positive personal trait. However, highly creative people have demonstrated elevated risk for certain forms of psychopathology, including mood disorders, schizophrenia spectrum disorders, and alcoholism. A model of shared vulnerability explains the relation between creativity and psychopathology. This model, supported by recent findings from neuroscience and molecular genetics, suggests that the biological determinants conferring risk for psychopathology interact with protective cognitive factors to enhance creative ideation. Elements of shared vulnerability include cognitive disinhibition (which allows more stimuli into conscious awareness), an attentional style driven by novelty salience, and neural hyperconnectivity that may increase associations among disparate stimuli. These vulnerabilities interact with superior meta-cognitive protective factors, such as high IQ, increased working memory capacity, and enhanced cognitive flexibility, to enlarge the range and depth of stimuli available in conscious awareness to be manipulated and combined to form novel and original ideas.
Chapter
This chapter discusses the field of evo-neuro-devo, an emerging discipline that focuses on how human neurodevelopment, and risk for psychiatric conditions mediated by neurodevelopment. It overviews segregating genetic variation, and de novo (new germline) mutations, provide novel insights into human neurodevelopmental gene functions, including effects from pleiotropy, polygenic inheritance, and developmental-genetic convergence. The chapter addresses how are the causes and phenotypes of the primary human neurodevelopmental conditions, especially autism and schizophrenia, related to recent neurodevelopmental and cognitive changes in human evolutionary history. It focuses on how genetic factors that cause changes in the rate and timing patterns of neurodevelopmental events, and associated psychological phenotypes, are involved in autism spectrum conditions and schizophrenia. The chapter describes a "developmental heterochronic" model for helping to explain the genetic bases of these two conditions, and connect the model with heterochronic change in human ancestry.
Book
Are creative people more likely to be mentally ill? This basic question has been debated for thousands of years, with the 'mad genius' concept advanced by such luminaries as Aristotle. There are many studies that argue the answer is 'yes', and several prominent scholars who argue strongly for a connection. There are also those who argue equally strongly that the core studies and scholarship underlying the mad genius myth are fundamentally flawed. This book re-examines the common view that a high level of individual creativity often correlates with a heightened risk of mental illness. It reverses conventional wisdom that links creativity with mental illness, arguing that the two traits are not associated. With contributions from some of the most exciting voices in the fields of psychology, neuroscience, physics, psychiatry, and management, this is a dynamic and cutting-edge volume that will inspire new ideas and studies on this fascinating topic.