Articles

Revisiting the Case for ‘Feral’ Humans Under the Light of the Human Self-Domestication Hypothesis: Focusing on Language

Amy Niego1, Antonio Benítez-Burraco*2

Biolinguistics, 2022, Vol. 16, Article e9319, https://doi.org/10.5964/bioling.9319

Received: 2022-04-14. Accepted: 2022-06-23. Published (VoR): 2022-09-30.

Handling Editor: Bridget Samuels, University of Southern California, Los Angeles, CA, USA

*Corresponding author at: Área de Lingüística. Departamento de Lengua Española, Lingüística y Teoría de la Literatura. Facultad de Filología. Universidad de Sevilla. C/ Palos de la Frontera s/n. 41004-Sevilla, Spain. +34 955 420 225. E-mail: abenitez8@us.es

This is an open access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Abstract

Contemporary descriptions of ‘feral’ children generally preclude any insightful inference about the language deficits exhibited by these children, as well as the ultimate causes of their problems with language. However, they have been regularly used to support the view that language acquisition requires a proper social environment in order to occur. In this paper, we revisit the case for ‘feral’ children with the viewpoint that human evolution entailed a process of self-domestication that parallels what we find in domesticated animals. Because feralization commonly occurs in nature and because it entails a partial reversion of features of domestication, this self-domestication approach to the evolution of language reassesses the case for ‘feral’ children, particularly when compared with present-day conditions involving abnormal patterns of socialization, whether they are genetically-triggered as in autism spectrum disorder, or environmentally-triggered, as in reactive attachment disorder.

Keywords: feral children, language acquisition, language evolution, self-domestication, ASD, RAD

1 Introduction

A recent view of human evolution argues that our species went through a process similar to that experienced by domesticated mammals, which would account for many human-specific features in the physical, behavioral, and cognitive domains. This is the Human Self-Domestication (HSD) hypothesis of human evolution (Hare, 2017). Testing this hypothesis is problematic. Whereas in other instances of alleged self-domestication, such as bonobos (Hare et al., 2012), the non-domesticated counterpart is still extant (i.e. chimpanzees) and can be used for direct comparisons, the only available evidence in the case of humans is indirect, as the closest species to us became extinct. Accordingly, the HSD hypothesis builds on findings in present-day humans of traits commonly found in domesticates, as compared to other hominins (principally, Neanderthals) and archaic modern humans, but also to primate relatives. These features include physical traits, that are easier to retrieve from the fossil register (e.g. reduced cranial robusticity, reduced brain size, reduced tooth size, juvenile cranial shape retained in adulthood, smaller teeth, reduced brow ridge, smaller jaw and nasal bone projection, reduced sexual dimorphism), but also behavioral traits, which can be inferred from the fossil and archaeological registers (e.g. reduced reactive aggression, increased prosocial behavior, prolonged play behavior; see Fukase et al., 2015; Herrmann et al., 2011; Langley et al., 2019; Leach, 2003; Márquez et al., 2014; Plavcan, 2012; Shea, 1989; Stringer, 2016; Thomas & Kirby, 2018; Zollikofer & Ponce de León, 2010, among others for discussion). There is also some incidental genetic evidence to support the HSD hypothesis. Studies have found that the human genome contains a host of pseudogenes, something that is commonly associated with domestication (Deacon, 2009). More importantly, candidate genes for domestication in mammals are overrepresented in regions that show signals of positive selection in modern humans compared to Neanderthals (Theofanopoulou et al., 2017), with selection on specific genes tracing back as late as 6,000 years ago (Benítez-Burraco et al., 2021b).

The "self-" in self-domestication refers to the fact that, contrary to animal domestication, which results from an active selection by humans of more tamed individuals, the alleged domestication of humans can be instead explained by the agency of diverse external factors, like the advent of co-parenting, changes in our foraging ecology, and/or the deterioration of the human environment during the Last Glaciation (Hare et al., 2012; Nikolsky & Benítez-Burraco, 2022; Pisor & Surbeck, 2019; Spikins et al., 2021). These factors seemingly favored more prosocial and cooperative behaviors. But because the physiological management of aggression ultimately involves diverse hormones impacting on different parts of the body (particularly, the hypothalamic-pituitary-adrenal axis), this process seemingly resulted in extended changes in our body, behavior, and perhaps cognition too. This co-occurrence of modified traits after the initial selection for tameness is what we indeed observe in most domesticated animals, termed as a “domestication syndrome” (Wilkins et al., 2014). According to the proponents of the HSD hypothesis, the behavioral changes brought about by our domestication would have facilitated the emergence of many of our species-specific distinctive features, including our enhanced social cognition, increased cooperation, extended social networks, and ultimately, our sophisticated culture and advanced technology (see Hare, 2017; Hare and Woods, 2020 for details).

In the recent years, the evolution of language under the effect of HSD has been the object of particular interest. This is certainly explained by the fact that language is a hallmark of the human condition, but also because language acquisition demands both cognitive and behavioral pre-requisites that are not all found in other species, and because it certainly results from a complex interaction between our biological endowment and the social environment. The finding that in some avian species domestication increases song complexity (see Okanoya, 2017) paved the way towards claims that HSD might have played a role in the emergence of modern languages (see Thomas & Kirby, 2018 and Benítez-Burraco & Progovac, 2020 for general discussions). A common view is that HSD favored the creation of the cultural niche that enables the sophistication of language via a cultural mechanism (Thomas & Kirby, 2018), mostly through its effects on selected aspects of our behavior (e.g. increased contacts between individuals, enhanced language learning and teaching, or increased language play; Benítez-Burraco & Progovac, 2020; Progovac & Benítez-Burraco, 2019, for details), but perhaps also of our body (e.g. the development of a white sclera that facilitates eye tracking and gazing during face-to-face interactions; Tomasello et al., 2007) and our cognition (e.g. improved episodic memory; Benítez-Burraco, 2021). Ultimately, these positive-for-language effects of HSD are thought to boil down to changes in the management of aggression (mostly reactive aggression), because, as noted, selection for tameness results in multiple changes in domesticates. Unfortunately, we lack any linguistic records from the remote past. To test and eventually prove this hypothesis about language evolution, we need to rely on indirect evidence, specifically, inferred changes in our communal living, behavior, social cognition, cooperation patterns, technology and know how, social networks, culture, and the like.

Interestingly, the finding that, as noted, candidate genes for domestication have been under selection in humans until very recently, suggests that HSD might be an ongoing process. In truth, features associated with HSD do not appear suddenly in the fossil register, but gradually, with signs of domestication reaching their peak quite late, seemingly at the end of the Upper Paleolithic (Cieri et al., 2014; Leach, 2003). The presentation of HSD features in present-day human populations can also be quite variable (Gleeson & Kushnick, 2018). Overall, this evidence suggests that HSD might be not a totally fixed phenotype in humans, but rather one where features of domestication are increased or attenuated in response to environmental conditions, mostly social features. This suggests that under some specific conditions HSD may revert to a less-domesticated condition. Because we have no access to living, archaic, less self-domesticated humans, a possible way to understand how HSD took place and how it impacted on our cognition and behavior is to compare such less-domesticated phenotypes with fully domesticated phenotypes, as hypothesized for present-day human beings. In fact, animal domestication seems to be a reversible condition too. Feralization refers to the process by which once-domesticated animals change and return to a wild-like state as a result of being reintroduced to wild-type environments without human contact (Daniels & Bekoff, 1989). As noted earlier and as we discuss below in detail, cases of ‘feral’ children have been described in the literature. Comparing these ‘feral’ children with neurotypical children could thus help us understand the effects of HSD on our cognition and behavior and ultimately, on our language abilities. Some caution is in order. For one, as we also show below, the term ‘feral’ child is not uncontroversial and posits its own problems. Moreover, feralization cannot be regarded as an exact mirror condition to domestication. As discussed in detail in Niego and Benítez-Burraco (2021), not all features of domestication reverse in feralized animals and to a great extent feralization is achieved through alternative genetic routes than domestication. Still, certain distinctive traits of domestication, particularly those that are more relevant to behaviors important for language acquisition (e.g. reduced reactive aggression, prosocial behavior), are found to be reversible. Likewise, a subset of the genes involved in feralization are also candidates for domestication. Overall, we can confidently expect feralization to help us to understand domestication (and vice versa, of course), particularly if one focuses on the features and the biological mechanisms that are shared by these two conditions, that, as noted, are mostly related to response to the social environment.

In this paper we explore the possibility that the examination of cases of ‘feral’ children can shed some light, as noted, on the process of HSD and its effects on the complexification of language under HSD forces. We first review the available literature about ‘feral’ children, with a focus on their language (dis)abilities. As we will show, evidence is scarce and controversial, particularly because some of these children likely suffered from underlying cognitive and/or behavioral problems. This is why in a subsequent section of the paper we compare, vis a vis HSD features, this ‘feral’ phenotype with two conditions entailing increased reactive aggression and decreased prosocial behavior, namely, autism spectrum disorder (ASD) and reactive attachment disorder (RAD). Apart from the fact that many ‘feral’ children might suffer specifically from these exact conditions, a more important reason is that in both ASD and RAD one finds abnormally high levels of reactive aggression (internally triggered mostly, in the case of ASD; and mostly externally caused, in the case of RAD). Likewise, whereas domestication mostly results from selection for tameness, feralization essentially entails the reactivation of mechanisms triggering reactive aggression. At least in the case of ASD, there is evidence that this condition can be construed as a phenotype with attenuated features of HSD (Benítez-Burraco et al., 2016). Additionally, in the paper we delve into the genetic underpinnings of feralization, ASD and RAD, set against the genetic background of HSD, to identify candidate genes of particular relevance for explaining the observed similarities and differences at the phenotypic level. Common candidates for these conditions are expected to be involved in the features of HSD that are more sensitive to the social environment. In a subsequent section of the paper, we put the focus on the communication and language (dis)abilities of ‘feral’ children, and we compare them with children on the ASD spectrum and children with RAD, also vis-à-vis the hypothesized changes in human language/communication under the effects of HSD, but also the changes in communication patterns observed in feralized animals. We conclude with a discussion centered around the relevance of this line of inquiry for the study of language evolution and with suggestions for future research.

To finish, we wish to make it clear that throughout the paper, labels like ‘domesticated’ or ‘feral’ as applied to humans are used in the technical sense they have within this particular evolutionary framework. By no means are they intended to qualify humans as less or more similar to animals. They are used, respectively, as a shorthand for the physical, behavioral, and cognitive effects of a potentiation or an attenuation of the physiological mechanisms involved in the management of reactive aggression.

2 ‘Feral’ Children

For the reasons mentioned in the previous section, in order to achieve a better understanding of the effects of HSD on the emergence of modern languages, it is of interest to look into descriptions of ‘feral’ humans (all of them children), and what has been documented about their physical and behavioral features, and particularly, about their language. The term feral child refers to a child who has been deprived of human contact, either intentionally or through accidental circumstances, and whose development has suffered as a result. As a consequence, ‘feral’ children do not develop many of the typical human skills, particularly those necessary to become fully integrated into society, specifically, language; one reason for this is the existence of a critical period during childhood for acquiring most of these traits (Stendler, 1952; Kapoor, 1973).

A great deal of historical descriptions of ‘feral’ children are of those said to have been raised by animals. For hundreds of years, society has had a fascination with these cases, from the so-called ‘wolf boy’ to ‘savage girls’ and even one ‘gazelle boy’ (Butler, 2003; Kapoor, 1973). Over fifty cases of such ‘feral’ children have been documented from all over the world, most from centuries past, although some surfaced as late as the 1950s. Most descriptions are quite similar: after their rescue from the wild, the children initially exhibit ‘animal behaviors’ such as growling like wolves, running on all fours, or smelling food before eating it, but all of them exhibit, specifically, severe deficits in language skills (Kapoor, 1973; Steeves, 2011). A typical case is that of Kamala and Amala, known as the ‘wolf children of Midnapore’. These two girls, aged 8 and 1 ½, were found in the company of wolves on the outskirts of a small village in India, and it was assumed from their condition that they had spent most of their young lives being raised by wolves (Butler, 2003). After being taken to an orphanage, observers reported that the girls acted like wolves; they chewed on bones, walked on all fours, howled, stayed awake all night, ate raw meat, and snarled or growled at other children who approached (Gesell, 1942). Although Kamala lived for 9 years at an orphanage after her removal from the wild, she was only able to develop a vocabulary of 45 words and understand basic verbal instructions (Gesell, 1942). Some speculate that Kamala never actually learned to speak, but instead learned to signal, much like an animal would (Butler, 2003). Similar descriptions exist for many other ‘feral’ children; in fact, of the nearly 50 documented cases of ‘feral’ children, the most successful ‘recovery’ of speech was Kamala; many of the other children remained mute, or never progressed past a drastically limited vocabulary (McNeil et al., 1984). They were never able to fully integrate into society either, exhibiting high aggression levels, severe deficits in social cognition, and behavior that ultimately isolated them socially (Bettelheim, 1959).

Although many of these cases of ‘feral’ children involve them growing up alone or being raised by wild animals, the term ‘feral’ also applies to children who have suffered abuse or extreme neglect that interferes with normal social and cognitive development. One such case, perhaps the most famous one, is that of Genie, a girl who was locked in a room for most of her young life. When she was found at the age of 13, Genie was unable to speak and exhibited many ‘feral’ characteristics: she was highly antisocial, displayed high levels of reactive aggression, and avoided touch and eye contact (Curtiss et al., 1974). Sadly, Genie was never able to fully integrate into society, and never learned to speak past a rudimentary stage, prompting many linguists to claim her story as evidence for a distinct critical window for language (e.g. Fromkin et al., 1974). A similar case is that of Victor of Aveyron, a child in 16th century France who was born to alcoholic parents and severely neglected for most of his childhood (Lane, 1976). Known as ‘the Savage of Aveyron’, Victor displayed some typical behavior of ‘feral’ children in that he didn’t speak, and only learned rudimentary language, ate raw meat, had no social skills, and avoided human contact, seemingly more comfortable with living alone in the wild (Lane, 1976). It should be noted that in the case of both Genie and Victor of Aveyron, some researchers who studied the cases indicated that the cognitive and social deficits were the result of undiagnosed conditions: intellectual disability in the case of Genie (Rymer, 1994) and ASD in the case of Victor (Bettelheim, 1959). One more modern case study supports this idea; Swain and colleagues (2005) reviewed a case of an adolescent boy described as having ‘feral child syndrome’; they explained that his symptoms most likely stemmed from a condition called ‘reactive attachment disorder’(RAD), which will be discussed in more detail in a subsequent section. This child’s symptoms were typical of those described in ‘feral’ children: ‘wolf-like’ aggressive behaviors, walking on all fours, and grunting instead of speaking. The doctors who eventually diagnosed the boy pointed to his traumatic early life events and ‘pathological caregiving’--that is, a mother who was mentally unstable and unable to form a typical attachment to the son. It is easy to see that, had the child been diagnosed a century earlier, the diagnosis would most likely have been ‘feral’.

The previous case studies point to the crux of the problem in cases of ‘feral’ children: it is frequently difficult to discern whether ‘feral’ traits in children came as a result of isolation from humans per se, from the trauma involved in these cases (as in most cases the isolation was also accompanied by neglect or abuse from caregivers), or from some underlying condition favoring isolation, such as ASD, intellectual disability, RAD, or any number of other developmental disabilities. To further complicate matters, many of these cases were documented before the fields of psychology or clinical linguistics were prepared to identify symptoms as cognitive, psychological or language impairments. Still, as early as 1959, Bettelheim wrote that ‘{Feralization} seems to be the result of some persons’-usually their parents’-inhumanity and not the result, as was assumed, of animals’-particularly, wolves'-humanity.’ He also notoriously noted that, in his opinion, many cases of so-called ‘feral’ children were most likely undiagnosed cases of severely autistic or cognitively impaired children. The parallels between past descriptions of ‘feral’ children and modern descriptions of children with ASD are remarkable, e.g. children tearing up food with their hands and/or teeth, growling, little to no use of language, avoidance of human contact, irritability, aggression, social anxiety, and a host of others. Because, as noted in the introduction, ASD has been hypothesized to entail attenuated features of HSD, and because individuals with ASD exhibit notorious language and communicative deficits (Bourguignon et al., 2012; Eigsti et al., 2007; Tager-Flusberg, 2006; Tager-Flusberg et al. 2005), in the next section we will consider ASD together with cases of ‘feral humans’ in our discussion of a putative reversion of the HSD phenotype and its effects on language. For the reasons exposed above, we will also consider RAD, which is admittedly a less known and less diagnosed disorder.

Finally, it is important to note that in most cases ‘feral’ children experienced extreme isolation that does not properly parallel what we find in feralized animals, who still maintain social contact with conspecifics, and what we can infer for archaic humans (exhibiting attenuated features of HSD), who maintained some sort of social networks, albeit less complex than those of present-day humans (e.g. Sikora et al., 2017). This circumstance is worth taking into account when comparing the ‘feral’ phenotype of these children with the traits observed in feralized animals, and particularly, when hypothesizing about the presentation of the HSD phenotype in the past from what we can observe in these children.

3 ‘Feral’ Children, ASD, RAD and HSD Vis-A-Vis

In this section we will explore the parallels and differences between ‘feral’ children and children with ASD or RAD, through the framework of the HSD hypothesis and with the process of animal feralization in mind. As noted in the introduction, HSD is hypothesized to have entailed some genetic selection in several genes that contributed to fixing physical and behavioral (and perhaps cognitive) features, resulting in a new social environment that favored the cultural evolution of languages, among other features. Feralization is also purported to entail some genetic changes that contribute to the readjustment to the wild environment (with only some overlap of the involved genes with genes involved in domestication; see Niego & Benítez-Burraco, 2021, for a detailed discussion). ‘Feral’ children are not expected to have suffered any kind of genetic change that accounts for their condition, which is therefore only due to the extreme environment in which they grew up. This ‘feral’ phenotype is thus expected to result mostly from physiological and developmental changes. Nonetheless, the famous farm-fox experiment, involving selection for tameness in wild foxes, suggests that domestication can be achieved relatively quickly (Trut et al., 2009), and the same has been shown in feralization: in some feral pigs, feral traits such as hairiness, aggression, and tusk length were seen to revert to a wilder phenotype in as little as a generation (Bach, 2007). These changes can be fixed (epi)genetically quite quickly too (see Anastasiadi et al., 2022 for discussion). Finally, ASD is known to result from both genetic and environmental causes (Bhandari et al., 2020; Bölte et al., 2019; Manoli & State, 2021), but in general, children with ASD are reared in a normal environment (although some adaptations to their condition are expected). By contrast, RAD is generally diagnosed in children who experience extreme social adversity in early childhood, although it has been claimed to have some genetic basis too (Corval et al., 2020; Spangler et al., 2009). To put this differently, we can regard i) neurotypical individuals as those who have HSD cognition/behavior and have been raised in a HSD-resulting environment; ii) ‘feral’ children and children with RAD as those having HSD cognition/behavior, but raised in a non-HSD-conducive environment; iii) children with ASD as individuals with non-fully HSD cognition/behavior, but who have been raised in a HSD-conducive environment; and iv) feralized animals, as the closest proxy to truly ‘feral humans’ (or archaic humans), that is, humans having non-HSD cognition/behavior who were raised, as a consequence, in a non-HSD environment. Needless to say, the former characterization entails some undesirable simplification, as one should expect that both biological and environmental factors contribute to these four conditions, although at the same time, the respective contributions of internal and external factors can be also expected to differ from one condition to another.

With all these qualifications in mind, we first compared these four phenotypes (‘feral’ children, children with ASD, children with RAD, and feralized animals) looking for the presentation of features of the domesticated/HSD phenotype. To do this, we first compiled a comprehensive set of 29 features found in domesticated animals (and allegedly, in neurotypical modern humans too, according to the HSD hypothesis). We relied on the existing literature on domestication, specifically, the domestication syndrome in animals (see Sánchez-Villagra et al., 2016; Wilkins et al., 2014), but also on publications describing the HSD hypothesis (Cieri et al., 2014; Hare, 2017; Thomas & Kirby, 2018, among others). We then conducted a comprehensive literature search of the PubMed database (https://pubmed.ncbi.nlm.nih.gov/) to gain knowledge about the presentation of these 29 domestication-related traits in the four phenotypes under scrutiny. The search strategy included ‘ASD’, ‘feral animals’, ‘RAD’ or ‘feral children’ and keywords likely to capture publications reporting on the salient traits. In spite of some limitations in the case of ‘feral’ children, for which data are scarce, we found that most features of domestication/HSD are absent or attenuated in children with ASD, with RAD, or in ‘feral’ children (see Table 1 for a summary of our findings and Table 2 for a more detailed report). Whereas conflicting results are observed in the case of the brain, this is particularly true of the physiological mechanisms underlying the aggression response and particularly, social behavior. In the next section, we focus on communicative abilities, for which we also find notable parallels also in line with the predictions of the HSD hypothesis.

Table 1

Summary Table of the Presentation of Features of Domestication/HSD in the Four Conditions Under Scrutiny in the Paper

Feature of domestication/HSD Feralized animals ASD ‘Feral’ children RAD
Head/Brain
Reduced head circumference, reduced total brain volumes Mixed evidence X ? ?
Relative decrease of specific brain areas (particularly, the hippocampus, the thalamus, and the striatum) X X ?
Relative increase of specific brain areas (particularly, the amygdala) Mixed evidence ? X
Ear size and shape
Reduced ear size/changes in ear shape X ? ?
Orofacial region
Shorter nose Mixed evidence ? ?
Dentition
Reduced tooth size/length Mixed evidence ? ?
Behaviour
Reduced anxious response to non-social and social aspects of life X X X X
Enhanced attentiveness and sensitivity to eye or facial movements or gestures X X X X
Increased sociability X X X X
Enhanced playing behavior ? X ? ?
Reduced aggressive behavior X X X X
Neuroendocrine system
Decreased levels of glucocorticoids ? X ? ?
Decreased levels of basal adrenocorticotropic hormone in plasma ? X ? X
Decreased stress response of the HPA axis X X ? X
Increased basal levels of oxytocin (and more marked release patterns) ? X ? X
Increased basal levels of vasopressin (and more marked release patterns) ? X ? X
Skin, related features
Hypopigmentation (skin, eyes) Mixed evidence ? ?
Higher levels of vitamin D in blood ? X ? ?
Vital cycle
General undergrowth X X ? X
Neoteny ? X ? ?
Reproductive cycle
Reduced levels of androgens (with a focus on testosterone) ? X ? ?
Earlier age of menarche X X ? ?

Note. X = absent feature; ✓ = attested feature; ? = not enough data available.

Table 2

A Detailed Characterization of the Presentation of Features of Domestication/HSD in the Four Conditions Under Scrutiny in the Paper

Features of domestication/HSD Feralized animals ASD ‘Feral’ children RAD
Head/Brain
Reduced head circumference, reduced total brain volumes Smaller head circumference compared to wild counterparts (Birks & Kitchener, 1999; Kruska, 2005; Kruska & Röhrs, 1974; Röhrs & Ebinger, 1999); dingoes show a larger cranium than dogs (Smith et al., 2018) Larger head circumference and/or brain size than TD subjects (Aldridge et al., 2011; Sacco et al., 2015) Not enough data available Not enough data available
Relative decrease of specific brain areas (particularly, the hippocampus, the thalamus, and the striatum) Larger hippocampus found in feral pigs (Kruska & Röhrs, 1974) Larger relative hippocampal volumes in children; atypical coupling between hippocampal volumes and brain size (Reinhardt et al. 2020) Not enough data available Decreased size of hippocampus (Corbin, 2007; Teicher, 2002); significant decrease of activity in the striatum (Takiguchi et al., 2015); greater fiber density, axonal diameter, and myelination in selected thalamic pathways (Makita et al., 2020); larger volume in right thalamus (Jung et al, 2020)
Relative increase of specific brain areas (particularly, the amygdala) Higher density of the amygdala in some species (Kruska, 2005, 2014; Kruska & Röhrs 1974); no changes in feral cats, dogs, pigs, goats, and donkeys (Röhrs & Ebinger, 1999; Smith et al., 2018) Larger amygdala than TD peers (Mosconi et al., 2009; Murphy et al., 2012); atypical development during growth (Courchesne et al., 2011; Schumann et al., 2004) Not enough data available Decreased size of the amygdala (Corbin, 2007; Teicher, 2002)
Ear size and shape
Reduced ear size/changes in ear shape Dingoes exhibit wolf-like prominent ears, contrary to most dog breeds (Schweizer et al., 2017; Smith et al., 2019) Abnormal ear shape (Manouilenko et al., 2014) Not enough data available Not enough data available
Orofacial region
Shorter nose Some breeds of pigs retain shorter snouts after feralization (Physical Characteristics of Feral Hogs, 2019), others revert to a longer, wild type snout (Ballard & Wilson, 2019; Kruska & Röhrs,1974; Smith et al., 2018) Flatter nose bridge documented in patients with ASD (Aldridge et al., 2011) Not enough data available Not enough data available
Dentition
Reduced tooth size/length Smaller teeth documented in feral pigs, but maintain ‘wild-type’ morphology (Evin et al., 2015); dingoes show larger carnassial teeth and longer canine teeth (Smith & Litchfield, 2010) Number of tooth irregularities documented in ASD, including missing teeth and diastemas (Luppanapornlarp et al., 2010) Not enough data available Not enough data available
Behaviour
Reduced anxious response to non-social and social aspects of life Increased ‘shoaling’ behavior in feral fish suggestive of higher anxiety levels (Swaney et al., 2015) Increased anxiety (Park et al., 2016) Increased social/non-social anxiety (Bettelheim, 1959; Butler, 2003) Unexplained fearful reactions during non-threatening interactions with caregivers (Ellis et al., 2020), general increased anxiety reported (Cuyvers et al., 2020; Hinshaw-Fuselier et al., 1999)
Enhanced attentiveness and sensitivity to eye or facial movements or gestures Eye contact is avoided (Johnston et al., 2017; Park et al., 2016) Eye contact is avoided (Johnston et al., 2017; Park et al., 2016) Eye contact is avoided (Butler, 2003) Eye contact is avoided (Ellis et al., 2020; Miellet et al., 2014)
Increased sociability Avoidance of contact with humans (Johnston et al., 2017; Owens et al., 2017; Rose et al., 1985) Reduced sociability commonly (Park et al., 2016) Avoidance of human contact (Butler, 2003) Muted positive response to socially stimulating situations, general social dysfunction (Ellis et al., 2020); minimal response to social and emotional stimuli (Ellis et al., 2020; Miellet et al., 2014)
Enhanced playing behavior Not enough data available Children with ASD lack ‘pretend play’ behavior (Bettelheim, 1959; Cuyvers et al., 2020; Ellis et al., 2020; Naber et al., 2008) Not enough data available Not enough data available
Reduced aggressive behavior Increased levels of aggression (Johnston et al., 2017; Owens et al., 2017; Rose et al., 1985) Reactive aggression/irritability commonly documented (Farmer et al., 2015, Mikita et al., 2015) High prevalence of increased aggression (Butler, 2003) Episodes of aggression and/or irritability widely documented (Ellis et al., 2020)
Neuroendocrine system
Decreased levels of glucocorticoids Not enough data available Higher levels of glucocorticoids (Hamza et al., 2010; Hollocks et al., 2014; Spratt et al., 2012) Not enough data available Not enough data available
Decreased levels of basal adrenocorticotropic hormone in plasma Not enough data available Higher levels of adrenocorticotropic hormone (Curin et al., 2003; Hamza et al., 2010; Hollocks et al., 2014; Spratt et al., 2012) Not enough data available Higher levels of adrenocorticotropic hormone correlated to parental neglect/abuse (Kemph & Voeller, 2008)
Decreased stress response of the HPA axis Heightened defense responses to predators linked to changes in the HPA axis (Swaney et al., 2015) HPA axis is hyper-responsive in benign social environments (Taylor & Corbett, 2014), slower in situations of social threat (Jacobson, 2014; Spratt et al., 2012); higher serum cortisol responses in children with ASD (Spratt et al., 2012); disrupted patterns of cortisol levels in children (Tomarken et al., 2015) Not enough data available HPA axis hyper reactivity reported in response to stress (Kemph & Voeller, 2008); lower levels of cortisol documented (Kočovská et al., 2013)
Increased basal levels of oxytocin (and more marked release patterns). Not enough data available Lower blood oxytocin levels reported (John & Jaeggi, 2021) Not enough data available Lower salivary oxytocin/atypical secretion reported (Fries et al., 2005; Suzuki et al., 2020)
Increased basal levels of vasopressin (and more marked release patterns). Not enough data available Decreased vasopressin levels in children (Oztan et al., 2018; Parker et al., 2018) Not enough data available Abnormal binding of vasopressin receptor (Carter, 2003; Carter et al., 2008; Heinrichs et al., 2009); significantly lower basal levels of plasma vasopressin (Fries et al., 2005)
Skin, related features
Hypopigmentation (skin, eyes) Great deal of variation in pigmentation is present, certain colorations are conserved when advantageous in the feral environment (Anderson et al., 2009; Derelle et al., 2013; Feulner et al., 2013; Gering et al., 2019; Linderholm et al., 2016; Roulin, 2004) Various hypomelanotic diseases present autistic symptoms (Akefeldt & Gillberg, 1991; Gómez-Lado et al., 2004; von Aster et al., 1997) Not enough data available Not enough data available
Higher levels of vitamin D in blood Not enough data available Lower serum levels of vitamin D reported in children with ASD (Cannell & Grant, 2013) Not enough data available Not enough data available
Vital cycle
General undergrowth Feral canids and pigs reported to be larger than their closest domesticated relatives (Gering et al., 2019; Kruska & Röhrs, 1974; Smith et al., 2018) General overgrowth reported during childhood (Campbell et al., 2014; Chawarska et al., 2011) Not enough data available General undergrowth reported (Dobrova-Krol et al., 2008; Stinehart et al., 2012)
Neoteny Not enough data available Neotenous features uncommon due to increased body size, higher levels of androgens, and earlier puberty than TD subjects (Campbell et al., 2014; Chawarska et al., 2011; El-Baz et al., 2014) Not enough data available Not enough data available
Reproductive cycle
Reduced levels of androgens (with a focus on testosterone) Not enough data available Higher levels of androgens reported (El-Baz et al., 2014; Hauth et al., 2014) Not enough data available Higher levels of androgens linked to higher stress in early life (Hauth et al., 2014), and to disrupted HPA axis function (Goozen et al., 2000).
Earlier age of menarche Delayed onset of sexual maturation (Abbott et al., 1997; Lord et al., 2013). General delay in the age of menarche (Knickmeyer et al., 2006) Not enough data available Not enough data available

Additionally, we compared the genetic basis of these conditions against the background of the genetics of domestication. Domestication has been hypothesized to result from changes in various genes involved in the development and function of different body parts. Wilkins and colleagues (2014) have highlighted a set of core candidate genes for domestication in mammals, whereas Niego and Benítez-Burraco (2019) have compiled an expanded list of candidates based on several recent analyses of the signatures of domestication in selected domesticates (see Supplementary Materials File 1; column A). That said, as noted by Jensen (2006) or Wilkins and colleagues (2014), and discussed as well in a recent paper by Anastasiadi and colleagues (2022), domestication can be expected to also result in part from pre-existing genetic variation, and even from epigenetic changes. The genetic scenario for feralization is pretty similar, but with the additional circumstance that some genetic differences between feralized and domesticated animals can be due to interbreeding with wild populations, and not to feralization per se (see Zhang et al., 2020 for discussion). Niego and Benítez-Burraco (2021) have also compiled an enlarged list of candidates for feralization relying on available data from a variety of studies (see Supplementary Materials File 1; column B). An important conclusion of Zhang and colleagues’ study is that the overlap between the genetic signatures of domestication and feralization is not as complete and one would expect, suggesting that the (partial) reversion of the traits of domestication as found in feralized animals is achieved through modifications in genes not directly involved in domestication; at the same time, some common candidates can be found between these two conditions, which can be expected to account for the reversion of specific traits of domestication observed in feralized animals. Genetic studies on ‘feral’ children have not been done for obvious reasons. Regarding the genetic basis of RAD, there is evidence that it is genetically influenced (Minnis et al., 2007). Several studies point to the involvement of genes related to dopaminergic and serotonergic activity in abnormal changes in attachment behavior by children (Bakermans-Kranenburg & van Ijzendoorn, 2007; Caspers et al., 2009; Lakatos et al., 2000; Luijk et al., 2011; Spangler et al., 2009; Wazana et al., 2015). Animal models suggest that changes in receptors for oxytocin and endogenous opioid could also result in RAD (Winslow & Insel, 2002; Moles et al., 2004). However, our knowledge about the genes that might predispose RAD is more limited than in the case of ASD, domestication, or even feralization. In spite of these limitations, from the comparison of what we know about the genetic basis of ASD, feralization, and RAD, we can gain some insights about the genetics of human domestication/feralization. Table 3 summarizes our findings. The abnormal expression of these genes in response to internal factors (e.g. mutations) or external factors (e.g. adverse rearing conditions) are expected to account, at least in part, for the shared features between these three conditions in the domains of social cognition and behavior, and ultimately, communication (dis)abilities, to which we dedicate the last section of the paper. These common candidates also emerge as promising candidates for explaining, at least in part, the changes in social cognition, behavior, and language which occurred during our evolution under the forces of HSD. One reason is the deep link that exists between evolution and abnormal development. Hence, as noted by e.g. Pattabiraman et al. (2020) with regard to cognitive diseases, our species-specific distinctive cognitive abilities resulted to a great extent from changes in the organization and function of preexisting neural devices (in response to specific mutations, but also to selective pressures favoring a more prosocial behavior, as claimed by the HSD hypothesis), but these changes also account for our predisposition to suffer from neurodegenerative diseases, because of the reduced resilience and resistance to damage of recently evolved neuronal networks (see Toro et al. 2010 for ASD).

Table 3

Structural and Functional Features of Common Candidates for ASD, RAD, and Feralization

Gene Protein name Protein function Relation to feralization Relation to ASD Relation to RAD
TRHDE Thyrotropin Releasing Hormone Degrading Enzyme Encodes a member of the peptidase M1 family, which inactivates the neuropeptide thyrotropin releasing hormone A candidate for stress-response differences observed between feralized animals and their domesticated counterparts (Nillni & Sevarino, 1999) Linked to enhanced dopaminergic and/or reduced serotonergic activity in children; linked to hypothalamic dysfunction (Hashimoto et al., 1991) Not enough data available
OXTR Oxytocin Receptor Plays a key role in socialization and regulation of the HPA axis’ response to stress Candidate for feralization (Pan et al., 2018) Hypermethylated in subjects on the spectrum; hypermethylation correlates to severity of social cognitive deficits (Andari et al., 2020; Maud et al., 2018), as well as to reduced attention to social information (Puglia et al., 2018) Not enough data available
MAOA Monoamine Oxidase A Encodes monoamine oxidase A, involved in degradation of amine neurotransmitters Not enough data available Alleles associated with increased severity of ASD (Cohen et al., 2003; Cohen et al., 2011; Yoo et al., 2009); alleles correlated with IQ levels, adaptive skills, and language skills (Cohen et al., 2003; Cohen et al., 2011) Some polymorphisms moderate detrimental effects of childhood abuse (Caspi et al., 2002); high levels of MAOA correlate to less social problems, including disrupted attachment (Caspi et al., 2002); selected epigenetic mutations correlate to high rates of aggression (Márquez et al., 2013)
5HTT Serotonin transporter Transports the neurotransmitter serotonin from synapses to presynaptic neurons Polymorphisms associated to fear response in hens (Krause et al., 2019) and defensive behaviors in animals (Blanchard et al., 1998); enhancement of serotonin activity inhibits predatory aggression in several species (Nikulina, 1991) Lower 5-HTT availability in various brain regions correlated to increased levels of serotonin in individuals with ASD (Andersson et al., 2021; Coutinho et al., 2004; Veenstra-VanderWeele et al., 2012) Significant correlation between one allele of 5HTT and maltreatment, stress, and adversity in childhood resulting in violent/aggressive behavior later in life (Reif et al., 2007)
TH Tyrosine Hydroxylase Involved in conversion of tyrosine to dopamine; implicated in the physiology of adrenergic neurons. Affects behavior and temperament in cattle (Lourenco-Jaramillo et al., 2012) Not enough data available Not enough data available
DRD4 Dopamine Receptor D4 Inhibits adenylyl cyclase. Variants in DRD4 correlate with levels of aggression linked to domestication in chickens (Komiyama et al., 2014) An allele of the gene may increase risk for clinically elevated ASD symptoms in children and adolescents with ADHD (Reiersen & Todorov, 2011) Polymorphisms linked to disordered attachment such as that found in RAD (Bakermans-Kranenburg & van Ijzendoorn, 2007; Spangler et al., 2009)

Finally, because evidence about the genetics of ASD is more abundant than in the case of RAD, we examined in more detail the overlap that exists, specifically, between the genetics of feralization and ASD, with the aim of uncovering additional candidates that can account for the changes in human socialization and communication patterns under the effects of environmental changes, in line with the HSD hypothesis. Among the candidate genes for feralization (see Supplementary Materials File 1; column B for an updated list) one finds many candidates or risk factors for ASD. Our findings are summarized in Table 4. Although the genes discussed in this section are by no means a comprehensive list of the genes implicated in feralization, RAD, and/or ASD, they emerge, as noted, as promising candidates for the changes in human behavior under (self-)domestication forces. Further research is undoubtedly needed in order to draw conclusions about the genetic basis of ‘feralization’ in humans, but the fact that many of these genes are related to neurotransmitter function points in a useful direction.

Table 4

Structural and Functional Features of Common Candidates for ASD and Feralization

Gene Protein name Protein function Relation to feralization Relation to ASD
CACNA1A Calcium Voltage-Gated Channel Subunit Alpha1 A Encodes the alpha 1C subunit of the Cav1.2 voltage-dependent L-type calcium channel, which contributes to the production of β and γ waves during wakefulness and REM sleep (Kumar et al. 2015) Candidate for feralization (Pan et al., 2018) Linked to a complex condition entailing intellectual disability, executive dysfunction, Attention-Deficit/Hyperactivity Disorder (ADHD) and/or ASD, co-occurring with childhood-onset epilepsy (Damaj et al. 2015); risk alleles of CACNA1C also correlate to lower performance in semantic tasks in people with schizophrenia (Krug et al. 2010)
FABP4 Fatty Acid binding protein 4 Encodes a fatty acid binding protein involved in fatty acid uptake, transport, and metabolism (Furuhashi, 2019) Candidate for feralization (Pan et al., 2018) FABP4 levels are significantly lower in children on the ASD spectrum (Maekawa et al., 2020). Mice bearing defective copies of Fabp4 show ASD-like features, including abnormal behavior and abnormal morphology of pyramidal neurons (Maekawa et al., 2020)
KIT KIT Proto-Oncogene, Receptor Tyrosine Kinase The protein encoded by KIT influences the proliferation, differentiation, migration and apoptosis of various cell types, playing a key role in hematopoiesis, stem cell maintenance, gametogenesis, melanogenesis, and in mast cell development, migration and function, among other things (Naumann et al., 2021) A strong candidate for domestication due to its connection with the neural crest (Wilkins et al., 2014), and also linked to feralization (Pan et al., 2018) Involved in symptoms of ASD, possibly due to its connection with the neural crest (Kilsby et al., 2013; Rothschild et al., 2003)
MBD4 Methyl-CpG Binding Domain 4, DNA Glycosylase Encodes a protein that binds to methylated DNA and that plays a role in the epigenetic regulation of gene expression and DNA repair (Bellacosa et al., 1999) Candidate for feralization (Pan et al., 2018) There is evidence of an association of the gene to some cohorts of people with ASD (Cukier et al., 2010)
MFRP Membrane Frizzled-Related Protein Encodes a protein associated to certain retinal conditions (Katoh, 2001) Candidate for feralization (Pan et al., 2018). Strongly enriched for variants likely to affect ASD risk (Sanders et al., 2015)
MIB1 MIB E3 Ubiquitin Protein Ligase 1 Encodes an ubiquitin-protein ligase that regulates apoptosis (Itoh et al., 2003) Candidate for feralization (not domestication; Zhang et al., 2020) Strongly enriched for variants likely to affect ASD risk (De Rubeis et al., 2014)
MITF Melanocyte Inducing Transcription Factor Encodes a melanocyte inducing transcription factor, responsible for pigment cell specific transcription of the melanogenesis enzyme genes. Mutations of MITF result in auditory-pigmentary syndromes, such as Waardenburg syndrome type 2 and Tietz syndrome (Goding & Arnheiter, 2019) Found among candidates for feralization in sheep (Pan et al., 2018), and it is one core candidate as well for domestication processes (Wilkins et al., 2014) Associated to gastrointestinal problems commonly found in people with ASD (Rodríguez-Fontenla & Carracedo, 2021)
NCOR1 Nuclear Receptor Corepressor 1 Involved in chromatin condensation and specifically regulates transcription repression of thyroid-hormone and retinoic acid receptors (Zhou et al., 2019). It is found among Zhang et al. (2020)’s candidates for feralization Candidate for Mitochondrial Complex III Deficiency, Nuclear Type 2, a condition which features motor problems, including apraxia and dysarthria, as well as cognitive impairment (Ghezzi et al., 2011); it is also a candidate for Rett Syndrome, which is characterized by loss of acquired motor and language skills, stereotypic movements, and ASD-like features (Lyst et al., 2013). Pathogenic variants of NCOR1 have been found in people on the ASD spectrum (Iossifov et al., 2012; Wang et al., 2016)
NF1 Neurofibromin Encodes a GTPase involved in the negative regulation of cell proliferation (Sabbagh et al., 2013) Related to altitude adaptation and response to hypoxia by feralized sheep (Pan et al., 2018) Occasionally associated with ASD (Marui et al., 2004; Sanders et al., 2012)
NUP155 Nucleoporin Encodes a protein related to the transportation of macromolecules across the nuclear envelope, which has been related to cardiac physiology (Zhang et al., 1999) Associated with dog feralization, but not domestication (Zhang et al., 2020) Protein-truncating variants in this gene have been found in people with ASD (Satterstrom et al., 2020)
PACS1 Phosphofurin Acidic Cluster Sorting Protein Encodes a protein involved in the localization of trans-Golgi network (TGN) membrane proteins (Hinners et al., 2003) It is found among Pan et al. (2018)’s candidates for feralization Related to a syndromic form of intellectual disability also encompassing ASD features (Schuurs-Hoeijmakers et al., 2016)
PRKCA Protein Kinase C Alpha Encodes a protein kinase involved in cell proliferation, differentiation, migration and adhesion (Haughian & Bradford, 2009) It is found among Pan et al. (2018)’s candidates for feralization Variants of the gene have been identified in three ASD probands (De Rubeis et al., 2014; Iossifov et al., 2014)
PSMD12 Proteasome 26S Subunit, Non-ATPase 12 Encodes a component of a proteasome involved peptide cleavage (Saito et al., 1997) It is found among Pan et al. (2018)’s candidates for feralization Deletions and loss-of-function point mutations in this gene have been identified in people with a syndromic neurodevelopmental disorder characterized by intellectual disability and ASD features (Küry et al., 2017)
ROBO2 Roundabout Guidance Receptor 2 Encodes a protein involved in axonogenesis and which plays a key role in many aspects of neurodevelopment and cognition (Brose et al., 1999; Kidd et al., 1998) A candidate for feralization in animals, it has also been related to the olfactory system and food seeking (Bates et al., 2011; Zhang et al. 2020) Candidate for ASD (Prasad et al., 2012; Suda et al., 2011), as well as dyslexia (Fisher et al., 2002), speech-sound disorder (Stein et al., 2004), and expressive vocabulary growth in the normal population (St Pourcain et al. 2014)
SLC25A27 Solute Carrier Family 25 Member 27 Encodes a component of an ion transporter across the mitochondrial membrane (Anitha et al., 2012) Associated to feralization, but not domestication (Zhang et al., 2020) Downregulated in various brain regions of subjects on the ASD spectrum (Anitha et al., 2012)
SOX6 SRY-Box Transcription Factor 6 Encodes a transcription factor involved in the regulation of many body functions (Cantù et al., 2011) Affects both feralization and domestication due to its association with the Wnt signalling pathway, which plays a key role in initial neural crest cell differentiation (Wang et al., 2018) Mutations have been associated to a complex condition encompassing neurodevelopmental problems and behavioral abnormalities, including ADHD and/or ASD features (Tolchin et al., 2020)
TTN Connectin Encodes connectin, a protein important for muscle elasticity (Chauveau et al., 2014) It is found among Pan et al. (2018)’s candidates for feralization Rare mutations in this gene have been identified in people with ASD (O’Roak et al., 2011, 2012)

4 Human ‘Feralization’ and the Evolution of Language

In the previous section, we have shown that changes in our biological endowment (as found in ASD) or in the external environment (as found in RAD) can impact negatively on our socialization and result in ‘feral’ phenotypes in our species that resemble the features exhibited by truly feral animals compared with their wild conspecifics. Phenotypes like ASD and RAD, as well as those exhibited by ‘feral’ children, can be thus construed as hypodomesticated human phenotypes. This is seemingly explained, at least partially, by the impact of such internal and external factors on the physiological mechanisms controlling aggression, which result in turn in cognitive and behavioral changes that impede the normal acquisition of social abilities. This impact is mediated by genes involved in domestication (and presumably, in HSD too). This circumstance provides additional support for the long tradition of using of these ‘feral’ phenotypes in humans as confident proxies for previous stages in the evolution of human language(s), but mostly, we wish to add, under the umbrella of the HSD hypothesis.

As we advanced in the introduction, the HSD hypothesis of language evolution has recently emerged as a promising view of how the features exhibited by present-day human languages might have appeared during our history. In brief, the behavioral (but also cognitive and physical) changes brought about by HSD would have favored the creation of the niche that enables the complexification of language through a cultural mechanism; in turn these increasingly sophisticated forms of language would have favored HSD, as they contributed to replacing physical confrontation with verbal confrontation (see Benítez-Burraco & Progovac, 2020; Progovac & Benítez-Burraco, 2019 for details). HSD can be expected to have favored the emergence of modern uses of language (pragmatics) as well, via this feedback loop between the control of aggression and language sophistication. Accordingly, the reduction in reactive aggression brought about by HSD would have enabled the full exploitation of our cognitive and interactional potential as it applied to linguistic exchanges, and ultimately, favored the evolution of a specific form of communication governed by persuasive reciprocity (see Benítez-Burraco et al., 2021a for details). In this section, we compare the language deficits exhibited by ‘feral’ children with the language problems observed in children with ASD or RAD, vis-a-vis the changes in communication patterns observed in feralized animals compared to their domesticated conspecifics. Such a broad characterization of the linguistic and interactional profiles of people with these three conditions aims to i) check whether the parallels observed in the physical, cognitive, and behavioral domains can be also observed in the domain of language/communication; ii) discuss whether these putative shared features parallel the reversion of communication abilities observed in feralized animals; iii) provide additional support for the view that changes impacting on our socialization make language/communication simpler and less efficient; and ultimately, iv) provide additional support for the HSD hypothesis of language evolution.

4.1 Communication Deficits in ASD, RAD, and ‘Feral’ Children

With regards to ‘feral’ children, as mentioned in Section 2, the most common trait observed in historical descriptions of these children is their lack of language skills (Fromkin et al., 1974; Kapoor, 1973; McNeil et al., 1984; Vyshedskiy et al., 2017). As also observed previously, very few ‘feral’ children have recovered language skills, and although some progress was noted in certain cases (see Rymer, 1994), most of them ended up with serious deficits, including drastically reduced vocabularies, and little to no syntactic ability (Bettelheim, 1959; Kapoor, 1973; McNeil et al., 1984, Vyshedskiy et al., 2017). Because many descriptions of ‘feral’ children were written when the fields of clinical linguistics and cognitive sciences were in their infancy, solid evidence is scarce. Much of the data referred to in this subsection was gleaned from the in- depth linguistic analysis of Genie, briefly discussed in Section 2 above, which is to date the most detailed account of language in ‘feral’ children that exists (Fromkin et al., 1974). Additional data on language development in ‘feral’ children has been collected from descriptions children raised in physical isolation because of neglect, adverse circumstances, or abuse. Overall, parallels with ASD and RAD are notable, as we show below.

In the domain of phonology, some authors (e.g. Lindgren et al., 2009; Wolk & Edwards, 1993; Wolk et al., 2016) support the view that the language of people on the ASD spectrum exhibits phonetic and phonological deficits, including consonant deletions, cluster reductions, changes in voicing, and syllable coalescence (but see Bartolucci et al., 1976; Kjelgaard & Tager-Flusberg, 2001; or McCann et al., 2007 for an opposing view). Many of these features correlate to findings on ‘feral’ children (Curtiss et al., 1974; Vyshedskiy et al., 2017). Problems with prosody have also been documented in both groups, particularly with regards to paralinguistic uses of prosody.

Regarding the domain of syntax, the general consensus from studies of ‘feral’ children is that they experience marked difficulties in grasping many types of syntactic constructions, like passives, wh-questions, or movement transformations, but also other simpler aspects like constructions involving spatial prepositions (Curtiss et al., 1974; Rymer, 1994; Vyshedskiy et al., 2017). Research on ASD shows related problems with structural aspects of language (see Tager-Flusberg et al., 2005 and Tager-Flusberg, 2006 for discussion), including difficulties with binding, relative clauses, wh-questions, raising and passives (Perovic & Janke, 2013; see Benítez-Burraco & Murphy, 2016 for review). Regarding RAD, a study by Raaska and colleagues (2012) showed that children with this condition also exhibit marked difficulties with language compared to their typically-developing (TD) peers—most notably in expressive language skills, communication, and comprehension (see also Zangl & Mills, 2007).

As one would imagine, pragmatics is an area that poses a great deal of difficulty for ‘feral’ children, people on the ASD spectrum, and people with RAD. To begin with, people with ASD and ‘feral’ children have shown difficulty inferring meaning from changes in voice inflection, as well as producing appropriate pitch/prosodic inflection with their utterances or to match their interlocutor (Curtiss et al., 1974; Fromkin et al., 1974; Fusaroli et al., 2017; Vyshedskiy et al., 2017). Other pragmatic deficits commonly found in people on the ASD spectrum include problems with understanding figurative use of language, including metaphor, irony, jokes, and sense of humor (see Baron-Cohen, 1988, 1997; Kissine, 2012). Additionally, individuals with ASD show impairments in conversational skills. For instance, individuals on the spectrum generally exhibit difficulty taking turns in conversation, detecting faux pas such as interruptions, or constructing a coherent narrative discourse (Baron-Cohen, 2000; Happé, 1993; Kaland et al., 2002; Kissine, 2012; Lord & Paul, 1997; Surian et al., 1996; Tager-Flusberg, 1992, 1993, 2000). Similar results have been found for individuals with RAD and for ‘feral’ children. Children with RAD show significant deficits when asked to interpret social cues, or to understand social hierarchies, on par with difficulties reported in ASD (Bennett et al., 2009; Green & Goldwyn, 2002). According to Sadiq and colleagues’ (2012) study, appropriate use of expressive language also proves to be a difficulty for those with RAD (Smyke et al., 2002), which is in line with their failure to develop relevant social relationships (Rutter et al., 2009). In fact, Sadiq and colleagues found that children with RAD had more difficulty with pragmatics than their counterparts with ASD, demonstrating lower scores in areas such as use of context, social relationships, and rapport. Vyshedskiy et al. (2017) performed an analysis of language structure and function on linguistically deprived individuals, including so-called ‘feral’ children, and found significant deficits in the category of mental synthesis, which is a prerequisite for pragmatic functions. Further research on children who have been subject to neglect by parents or caregivers corroborates the idea that pragmatic language is one of the most dramatically affected aspects of language in these cases. For example, Di Sante and colleagues (2019) found that nearly half (44.4%) of neglected children in their study presented pragmatic problems in their language, from simple tasks such as asking for help to more complex ones like adapting the conversation towards one’s speaking partner. In a more recent study, Di Sante and colleagues (2020) found a significant link between parental neglect and children’s pragmatic abilities in areas like responsiveness, conversation support, and affect. In his work about the famous ‘feral’ child Victor of Aveyron, Lebrun (1980) pointed out that many so called ‘feral’ children, as well as those who experience speech deprivation due to neglect or other factors, have noted difficulty with pragmatic aspects of language such as turn taking, interpersonal turn taking, and understanding inference.

Since humans require social interaction during formative years for language to develop, it has been suggested that the communicative deficits displayed by ‘feral’ children are the result of this missed window of opportunity to interact with their conspecifics (Curtiss et al., 1974; Rymer, 1994), although, as noted above, the high incidence of abuse and neglect as well as the high probability of undiagnosed conditions makes it difficult to exclude other possible causes. With regards to RAD, delays in language acquisition commonly observed in this condition have been hypothesized to mostly stem from the deprived social environment that most children with RAD experience early in life, similarly to what has been suggested for ‘feral’ children. However, poor language performance in RAD could also result from some neurodevelopmental deficits such as impulsiveness and poor executive function, both associated with this condition (see Gleason et al., 2011; Raaska et al., 2012 for discussion). In the case of ASD, pragmatic deficits, specifically, have been also attributed to deficient socialization patterns (Bailey et al., 1996), resulting from an internal deficit. This is specifically the case with mindreading, which impedes the detection and recognition of others’ mental states based on external cues, such as people’s facial expressions. Typically, individuals on the ASD spectrum show difficulties with the classic ‘false belief’ task, where the participant is required to imagine the mental state of another (see Baron-Cohen, 2000; Happé, 1993, for overview). A well-known hypothesis supports the view that these problems stem from a deficit in the Theory of Mind system (ToM; Happé, 1995). ToM grants us the ability to mentalize and carry out tasks such as inferring meaning, imagining mental states of others, and interpreting communicative intentions of our interlocutors. The brain regions supporting ToM (the bilateral temporal-parietal junction, the posterior superior temporal sulcus, and the medial prefrontal cortex) show atypical activation patterns in people with ASD (Castelli et al., 2002; Colich et al., 2012; Schultz et al., 2003; Wang et al., 2006; Wang et al., 2007). Similar findings have been observed for children with RAD (and potentially, for ‘feral’ children). For example, research by Teicher and Samson (2016) found that childhood neglect (a common element in both conditions) results in atypical size, density, and function of the prefrontal cortex (see also Baker et al., 2013; Gupta et al, 2016; Heim et al., 2013).

Our contention here is that structural and functional problems with language exhibited by ‘feral’ children, children with RAD, and children on the ASD spectrum might result, at least partially, from the abnormal presentation of HSD features. On the physiological side, the hyper-responsiveness of the HPA axis in response to stress certainly impedes normal socialization and accordingly, the acquisition of language through the interaction with their peers and caregivers. Also, the brain changes found in these groups, particularly in the amygdala, the hippocampus, and the thalamus, could also account for the observed language deficits. The amygdala and the hippocampus are components of the limbic system, a group of brain structures involved in the regulation of emotion, but also in functions like motivation and memory (see Rolls, 2015 for review). It has been argued that the hippocampus supports our notable episodic memory that allows us to mentally travel both forward and backward in time. According to Corballis (2018, 2019) this Mental Time Travel (MTT) ability also supports some core features of human language, including recursion and displacement. In ASD, difficulties in MTT have been linked to hippocampal abnormalities (Cooper & Ritchey, 2019), and studies on childhood neglect have also shown that lack of a caregiver affects the structure of the hippocampus by slowing down or altering synaptic development (Andersen & Teicher, 2004; Pickering et al., 2006). Regarding the thalamus, changes in the thalamus have been hypothesized to contribute to the emergence of our species-specific ability for learning and using languages (Boeckx & Benítez-Burraco, 2014).

Given the similar profiles of people with ASD, children with RAD, and ‘feral’ children in the domain of language structure and use, we end this section with an examination of the effects of feralization in the communicative behavior of domesticated animals. Our aim is to find additional evidence that these abnormal language profiles might result from an attenuation of HSD features as levels of reactive aggression increase.

4.2 Communication in Feralized Animals Compared to Wild and Domestic Animals

Animals are routinely used as a model for language evolution, because of the strong continuity found between human language and animal cognition/communication/socialization (as language fulfills so many different functions, many domains need to be explored; Okanoya, 2017; Tyack, 2020; Vernes, 2017). This is also true for the putative effects of HSD on our language abilities: as noted in the Introduction, animal communication becomes more complex because of the relaxation of selective pressure brought about by the domestic environment. Accordingly, we find it worth examining the effects of feralization on animal communication: this should enable us to better understand how language deficits in conditions like ASD or RAD result from the attenuation of HSD features.

Naturally, dogs provide an interesting testing ground. The reason is that, like modern humans, they exhibit almost the whole suite of features associated with domestication (Sánchez-Villagra et al., 2016), seemingly because they were domesticated almost entirely based on selection against aggression and not on other traits of interest. In dogs, domestication resulted in, or reinforced, cognitive abilities which support the acquisition and use of sophisticated communication devices, such as joint attention (Nagasawa et al., 2015), gaze following (Kaminski & Nitzschner, 2013; Range & Virányi, 2013), over-imitation (Huber et al., 2020), or the ability to solve problems by relying on social cues (Hernádi et al., 2012; Udell, 2015). On the genetic level, research shows that humans and dogs share common genetic determinants that are associated with key physiological processes involved in domestication (see Benítez-Burraco et al., 2021b for review). Specifically, signals of selection of specific changes in cis-regulatory regions of the oxytocin receptor gene, OXTR, have been found both in dogs (Oliva et al., 2015; Shilton et al., 2020; vonHoldt et al., 2017) and humans (Schaschl et al., 2015). In both species, the genetic variation of OXTR is further associated with differences in social behavior (Eales, 1989; Pfenning et al., 2014; Shilton et al., 2020). More generally, the genes that have been positively selected in humans compared to Neanderthals are enriched in candidates for mammal domestication, particularly dog domestication (Theofanopoulou et al., 2017). Benítez-Burraco and colleagues (2021b) have hypothesized a positive feedback loop between dog domestication and HSD, through the reduction of reactive aggression, contributing, albeit tangentially, to aspects of language evolution. Interestingly, dingoes are dogs that went feral around 5000 years ago and have since remained isolated from other dogs (Shipman, 2021; Smith & Litchfield, 2010). From the evidence gathered, it seems that dingoes fall somewhere between domesticated and wild canids in both phenotypic terms and in terms of their ability to communicate with humans (Smith & Litchfield, 2010). Thus, dingoes are able to read human cues, such as pointing, tapping, and gazing, to reach an object—usually more successfully than wolves—displaying some skills that are more on par with domestic dogs (Smith & Litchfield, 2010). However, when it comes to more nuanced tasks such as following only the experimenter’s gaze to find food or following a signal from an incorrect location, it seems that dingoes underperform relative to domesticated dogs (Bräuer et al., 2006; Hare & Tomasello, 1999; Udell et al., 2008; Virányi et al., 2008). Additionally, Smith and Litchfield (2010) found that dingoes need to be familiar with an experimenter before being able to carry out the tasks in the experiment. This contrasts with domestic dogs, who can follow cues from any human subject. Parallels can be seen with ASD and ‘feral’ children. For instance, Watkins and colleagues (2013) found that children with ASD performed significantly better on standardized tests when they were familiar with the examiner, as opposed to when the examiner was a stranger (Watkins et al., 2013). When looking at aspects like joint attention, and/or gaze following, various studies on ASD (e.g. Paul et al., 2008; Volkmar et al., 2005; Wetherby et al., 2007) show that people on the spectrum exhibit abnormal gaze patterns, and lower frequencies of joint attention, similar to the results found in feral canids. Additionally, characterizations of ‘feral’ children describe avoidance of eye contact as a common trait, especially during initial exposure (Butler, 2003). These results indicate that domestication does seem to mark a difference in canids in terms of certain precursors of language (and particularly, language use), like joint attention, or the use of social cues to solve problems, and that these components are particularly impaired in conditions like ASD or ‘feral’ phenotypes in humans.

Another area of interest is the changes in vocal communication brought about by feralization, that can be gleaned from studies on feral cats and dogs as well. Yeon and colleagues (2011) found that feral cats generally use antagonistic vocalizations such as the growl and hiss much more than communicative ones (i.e. meow), but they do meow in response to a variety of interactions—both human and non-human. In contrast, domestic cats reserve the meow exclusively for the approach of a human, indicating a more communicative intent for the vocalization (Yeon et al., 2011). These findings concur with other research, for example, in domestic cats, the meow is heard much more frequently during cat-human interaction and much more rarely during cat-cat interactions (Bradshaw & Cameron‐Beaumont, 2000; Brown, 1993). Yeon and colleagues (2011) also found that feral cats generally produce longer vocalizations at a lower frequency (pitch) when in contact with humans, and tend to use agonistic vocalizations such as growling and hissing during encounters with humans, while domestic cats are more vocal in social situations, and less likely to use agonistic vocalizations. These findings parallel those of Nicastro (2004), who compared domestic cat vocalizations with those of wild cats, finding that in general domestic cats produce calls of a shorter duration and a higher pitch frequency than wild cats. Domestic cats are also known to produce vocalizations typical of kittens, such as purring and chirping, which is not the case in feral cats (Bradshaw & Cameron‐Beaumont, 2000), which is in line with the neotenic features typically associated to domestication. Regarding dogs and dingoes, Smith and Litchfield (2010) found in their research that, similarly to wolves, dingoes are found to howl and whimper more than domestic dogs, and bark less. Moreover, their barks tend to be shorter than their domestic counterparts, and used almost exclusively for giving warnings. In contrast, domestic dogs’ barks are more versatile, and used in a variety of social situations (Corbett, 2004). Studies on domestic and feral pigs show that they have a wide and varied repertoire that differs in some respects—for example, the ‘trumpeting’ sound produced by non-domesticates has not been recorded in domestic pigs (Garcia et al., 2016; Tallet et al., 2013). Evidence on pigs is difficult to interpret, however, given the blurred lines that exist between ‘feral’ pigs and ‘wild’ pigs; since a great deal of admixture takes place between the types of swine, it is hard to attribute changes to feralization (Garcia et al., 2016). Although vocal communication of mammals is more relevant to our research because of their phylogenetic proximity, patterns from bird species can also contribute to our knowledge of changes conferred through feralization, particularly because of the abundant research on the notable effects of domestication on birdsongs. Gering and colleagues (2015) found that feral chickens on Kauai have more extreme calls, and exhibit much more variation in their song, than either wild jungle fowl or domestic chickens that share their habitat. More research is necessary to see whether this change in vocalization complexity results from genetic admixture, environmental changes, or from a host of other epigenetic factors (Gering et al., 2015).

The vocal behavior of people with ASD seems to mirror some of the characteristics found in feral animals in the sense that it doesn’t follow the normal patterns of TD peers. Vocalizations made by individuals with ASD are often described as ‘different’ (Ricks & Wing, 1976). Wallace and colleagues (2008) found that toddlers with ASD tend to produce a larger range of syllables with atypical pronunciation than their TD peers. Schoen and colleagues (2011) found that toddlers with ASD produce more atypical vocalizations than language-matched peers, with high pitched squeals being much more common in the ASD group. Kent and Murray (1982) also found that children on the ASD spectrum produce much more irregular and complex pitch patterns within the same breath than typically developing counterparts. Problems with prosody and pitch shifting are also reported by Russo and colleagues (2008), who speculated that problems with the auditory-motor pathway are to blame. Descriptions of ‘feral’ children can be seen as also support for this view, since similar characteristics (e.g. inappropriate control of pitch and intonation) are often observed in historical descriptions (Lebrun, 1980).

5 Conclusion

From our review of the literature on the physical, behavioral, and cognitive features exhibited by ‘feral’ children, people with ASD, and children with RAD, it can be concluded that many parallels exist between these conditions, and that differences with neurotypical subjects can construed, to a certain extent, as a reversion of the changes brought about by HSD. This is also true, specifically, of the communication/language abilities, with people with these conditions exhibiting certain specific communicative deficits relative to TD counterparts, which may to some extent parallel differences seen in feralized animals compared to domesticated variants.

Overall, these differences can be interpreted as a ‘hypo-domesticated’ phenotype, resulting from some genetic differences, in the case of ASD (see Benítez-Burraco et al., 2016 for a more detailed discussion), or from some environmental effects, specifically, abnormal patterns of socialization, in the case of ‘feral’ children and people with RAD. Our findings qualify these present-day phenotypes as useful proxies for studying the effects of HSD on our language abilities in the past, particularly because comparative evidence had already suggested that the HSD phenotype is not totally fixed in present-day human populations and, accordingly, can present in a quite variable fashion (Gleeson & Kushnick, 2018). In our opinion, the specific details of our research offer fertile ground for future hypothesis generation and experimental testing. This includes, first and foremost, genetic studies aimed at enlarging the set of common genetic determinants for ASD and RAD, as well as feralization. Specifically, it would be very interesting to know whether these common genes have been subject to recent selection in human populations. Likewise, neuroimaging research on feralized animals and on people with ASD or RAD during communication/language processing tasks should help identify the brain areas with a role in language processing that are more sensitive to HSD processes, and also, to provide additional support to the view that altered socialization patterns result in abnormal processing patterns of social cues at the brain level with a negative impact on language acquisition and use. Finally, because core aspects of human languages seem to result from cultural transmission, and because the two main conditions examined in the paper, namely ASD and RAD, feature abnormal socialization patterns, it would be interesting to conduct experiments with artificial languages aimed at knowing how language is learnt and transmitted in atypical populations, and how this abnormal transmission impacts language features.

Besides its relevance for improving current theories about the evolution of language in the species, the research outlined in this paper is expected to have some clinical applications too. This HSD approach to clinical conditions like ASD or RAD is in line with systems biology approaches to complex conditions, arguing for a holistic approach to diseases, which examines all the involved factors (from genes to the environment), and their complex interaction during development. (see Benítez-Burraco, 2020 for discussion). From the considerations above about how our body, cognition, and behavior can be modified in response to environmental changes, we can expect to design better learning stimuli and more suitable learning environments for children suffering from these conditions.

Funding

This research was funded by the Spanish Ministry of Science and Innovation (grant PID2020-114516GB-I00 by MCIN/AEI/10.13039/501100011033 and by the “European Union” to ABB).

Acknowledgments

The authors have no additional (i.e., non-financial) support to report.

Competing Interests

The authors have declared that no competing interests exist.

Author Contributions

ABB and AN conceived the paper, reviewed the literature, performed the data analyses and wrote and approved the final manuscript.

Ethics Statement

The research conducted for the paper relied on previously published data by others and available datasets, hence no ethics approval was required.

Data Availability

For this article, data for the comparison of the genetic basis of Domestication-Feralization are freely available (Niego & Benítez-Burraco, 2022).

Supplementary Materials

The Supplementary Materials contain data for the comparison of the genetic basis of Domestication-Feralization (for access see Index of Supplementary Materials below).

Index of Supplementary Materials

  • Niego, A., & Benítez-Burraco, A. (2022). Supplementary materials to "Revisiting the case for ‘feral’ humans under the light of the human self-domestication hypothesis: Focusing on language" [Data]. PsychOpen GOLD. https://doi.org/10.23668/psycharchives.8216

References

  • Abbott, D. H., Dumesic, D. A., Eisner, J. R., Kemnitz, J. W., & Goy, R. (1997). The prenatally androgenized female rhesus monkey as a model for PCOS. In R. Azziz, J. E. Nestler, & D. Dewailly (Eds.), Androgen excess disorders in women (pp. 369–382). Lippincott–Raven. https://doi.org/10.1002/ajp.20679

  • Akefeldt, A., & Gillberg, C. (1991). Hypomelanosis of Ito in three cases with autism and autistic-like conditions. Developmental Medicine and Child Neurology, 33(8), 737-743. https://doi.org/10.1111/j.1469-8749.1991.tb14953.x

  • Aldridge, K., George, I. D., Cole, K. K., Austin, J. R., Takahashi, T. N., Duan, Y., & Miles, J. H. (2011). Facial phenotypes in subgroups of prepubertal boys with autism spectrum disorders are correlated with clinical phenotypes. Molecular Autism, 2(1), Article 15. https://doi.org/10.1186/2040-2392-2-15

  • Anastasiadi, D., Piferrer, F., Wellenreuther, M., & Benítez Burraco, A. (2022). Fish as model systems to study epigenetic drivers in human self-domestication and neurodevelopmental cognitive disorders. Genes, 13(6), Article 987. https://doi.org/10.3390/genes13060987

  • Andari, E., Nishitani, S., Kaundinya, G., Caceres, G. A., Morrier, M. J., Ousley, O., Smith, A. K., Cubells, J. F., & Young, L. J. (2020). Epigenetic modification of the oxytocin receptor gene: Implications for autism symptom severity and brain functional connectivity. Neuropsychopharmacology, 45(7), 1150-1158. https://doi.org/10.1038/s41386-020-0610-6

  • Andersen, S. L., & Teicher, M. H. (2004). Delayed effects of early stress on hippocampal development. Neuropsychopharmacology, 29(11), 1988-1993. https://doi.org/10.1038/sj.npp.1300528

  • Anderson, T. M., vonHoldt, B. M., Candille, S. I., Musiani, M., Greco, C., Stahler, D. R., Smith, D. W., Padhukasahasram, B., Randi, E., Leonard, J. A., Bustamante, C. D., Ostrander, E. A., Tang, H., Wayne, R. K., & Barsh, G. S. (2009). Molecular and evolutionary history of melanism in North American gray wolves. Science, 323(5919), 1339-1343. https://doi.org/10.1126/science.1165448

  • Andersson, M., Tangen, Ä., Farde, L., Bölte, S., Halldin, C., Borg, J., & Lundberg, J. (2021). Serotonin transporter availability in adults with autism-a positron emission tomography study. Molecular Psychiatry, 26(5), 1647-1658. https://doi.org/10.1038/s41380-020-00868-3

  • Anitha, A., Nakamura, K., Thanseem, I., Yamada, K., Iwayama, Y., Toyota, T., Matsuzaki, H., Miyachi, T., Yamada, S., Tsujii, M., Tsuchiya, K. J., Matsumoto, K., Iwata, Y., Suzuki, K., Ichikawa, H., Sugiyama, T., Yoshikawa, T., & Mori, N. (2012). Brain region-specific altered expression and association of mitochondria-related genes in autism. Molecular Autism, 3(1), Article 12. https://doi.org/10.1186/2040-2392-3-12

  • Bach, M. (November 28, 2007). Domestic pigs quickly revert to wild roots. Retrieved from https://www.mlive.com/flintjournal/outdoors/2007/11/domestic_pigs_quickly_revert_t.html

  • Bailey, A., Phillips, W., & Rutter, M. (1996). Autism: Towards an integration of clinical, genetic, neuropsychological, and neurobiological perspectives. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 37(1), 89-126. https://doi.org/10.1111/j.1469-7610.1996.tb01381.x

  • Baker, L. M., Williams, L. M., Korgaonkar, M. S., Cohen, R. A., Heaps, J. M., & Paul, R. H. (2013). Impact of early vs. late childhood early life stress on brain morphometrics. Brain Imaging and Behavior, 7(2), 196-203. https://doi.org/10.1007/s11682-012-9215-y

  • Bakermans-Kranenburg, M. J., & van Ijzendoorn, M. H. (2007). Research review: Genetic vulnerability or differential susceptibility in child development: The case of attachment. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 48(12), 1160-1173. https://doi.org/10.1111/j.1469-7610.2007.01801.x

  • Ballard, J. W. O., & Wilson, L. (2019). The Australian dingo: Untamed or feral? Frontiers in Zoology, 16, Article 2. https://doi.org/10.1186/s12983-019-0300-6

  • Baron-Cohen, S. (1988). Social and pragmatic deficits in autism: Cognitive or affective? Journal of Autism and Developmental Disorders, 18(3), 379-402. https://doi.org/10.1007/BF02212194

  • Baron-Cohen, S. (1997). Hey! It was just a joke! Understanding propositions and propositional attitudes by normally developing children and children with autism. The Israel Journal of Psychiatry and Related Sciences, 34(3), 174-178.

  • Baron-Cohen, S. (2000). Theory of mind and autism: A fifteen year review. In S. Baron-Cohen, H. Tager-Flusberg, & D. J. Cohen (Eds.), Understanding other minds: Perspectives from developmental cognitive neuroscience (pp. 3–20). Oxford University Press.

  • Bates, T. C., Luciano, M., Medland, S. E., Montgomery, G. W., Wright, M. J., & Martin, N. G. (2011). Genetic variance in a component of the language acquisition device: ROBO1 polymorphisms associated with phonological buffer deficits. Behavior Genetics, 41(1), 50-57. https://doi.org/10.1007/s10519-010-9402-9

  • Bartolucci, G., Pierce, S., Streiner, D., & Eppel, P. T. (1976). Phonological investigation of verbal autistic and mentally retarded subjects. Journal of Autism and Childhood Schizophrenia, 6(4), 303-316. https://doi.org/10.1007/BF01537908

  • Bellacosa, A., Cicchillitti, L., Schepis, F., Riccio, A., Yeung, A. T., Matsumoto, Y., Golemis, E. A., Genuardi, M., & Neri, G. (1999). MED1, a novel human methyl-CpG-binding endonuclease, interacts with DNA mismatch repair protein MLH1. Proceedings of the National Academy of Sciences of the United States of America, 96(7), 3969-3974. https://doi.org/10.1073/pnas.96.7.3969

  • Benítez-Burraco, A. (2020). The golden mean: A systems biology approach to language disorders. Pragmalingüística, 2, 30-44. https://doi.org/10.25267/Pragmalinguistica.2020.iextra2.02

  • Benítez-Burraco, A. (2021). Mental time travel, language evolution, and human self-domestication.  Cognitive Processing22(2), 363-367. https://doi.org/10.1007/s10339-020-01005-2

  • Benítez-Burraco, A., Ferretti, F., & Progovac, L. (2021a). Human self-domestication and the evolution of pragmatics. Cognitive Science, 45(6), Article e12987. https://doi.org/10.1111/cogs.12987

  • Benítez-Burraco, A., Lattanzi, W., & Murphy, E. (2016). Language impairments in ASD resulting from a failed domestication of the human brain. Frontiers in Neuroscience, 10, Article 373. https://doi.org/10.3389/fnins.2016.00373

  • Benítez-Burraco, A., & Murphy, E. (2016). The oscillopathic nature of language deficits in autism: From genes to language evolution. Frontiers in Human Neuroscience, 10, Article 120. https://doi.org/10.3389/fnhum.2016.00120

  • Benítez-Burraco, A., Pörtl, D., & Jung, C. A. (2021b). Did dog domestication contribute to language evolution? Frontiers in Psychology, 12, Article 695116. https://doi.org/10.3389/fpsyg.2021.695116

  • Benítez-Burraco, A., & Progovac, L. (2020). A four-stage model for language evolution under the effects of human self-domestication. Language & Communication, 73, 1-17. https://doi.org/10.1016/j.langcom.2020.03.002

  • Bennett, J., Espie, C., Duncan, B., & Minnis, H. (2009). A qualitative exploration of children’s understanding of indiscriminate friendliness. Clinical Child Psychology and Psychiatry, 14(4), 595-618. https://doi.org/10.1177/1359104509339137

  • Bettelheim, B. (1959). Feral children and autistic children. American Journal of Sociology, 64(5), 455-467. https://doi.org/10.1086/222541

  • Bhandari, R., Paliwal, J. K., & Kuhad, A. (2020). Neuropsychopathology of autism spectrum disorder: Complex interplay of genetic, epigenetic, and environmental factors. Advances in Neurobiology, 24, 97-141. https://doi.org/10.1007/978-3-030-30402-7_4

  • Birks, J. D. S., & Kitchener, A. C. (1999). Ecology of the polecat in Lowland England. In J. D. S. Birks & A. C. Kitchener (Eds.), The distribution and status of the polecat Mustela putorius in Britain in the 1990s (pp. 111–130). The Vincent Wildlife Trust. https://doi.org/10.2981/0909-6396

  • Blanchard, D. C., Griebel, G., Rodgers, R. J., & Blanchard, R. J. (1998). Benzodiazepine and serotonergic modulation of antipredator and conspecific defense. Neuroscience and Biobehavioral Reviews, 22(5), 597-612. https://doi.org/10.1016/S0149-7634(97)00054-7

  • Boeckx, C., & Benítez-Burraco, A. (2014). The shape of the human language-ready brain. Frontiers in Psychology, 5, Article 282. https://doi.org/10.3389/fpsyg.2014.00282

  • Bölte, S., Mahdi, S., de Vries, P. J., Granlund, M., Robison, J. E., Shulman, C., Swedo, S., Tonge, B., Wong, V., Zwaigenbaum, L., Segerer, W., & Selb, M. (2019). The Gestalt of functioning in autism spectrum disorder: Results of the international conference to develop final consensus International Classification of Functioning, Disability and Health core sets. Autism, 23(2), 449-467. https://doi.org/10.1177/1362361318755522

  • Bourguignon, N., Nadig, A., & Valois, D. (2012). The biolinguistics of autism: Emergent perspectives. Biolinguistics, 6(2), 124-165. https://doi.org/10.5964/bioling.8897

  • Bradshaw, J. W. S., & Cameron‐Beaumont, C. L. (2000) The signalling repertoire of the domestic cat and its undomesticated relatives. In D. C. Turner & P. P. G. Bateson (Eds.), The domestic cat: The biology of its behavior (pp. 68–93). Cambridge University Press.

  • Bräuer, J., Kaminski, J., Riedel, J., Call, J., & Tomasello, M. (2006). Making inferences about the location of hidden food: Social dog, causal ape. Journal of Comparative Psychology, 120(1), 38-47. https://doi.org/10.1037/0735-7036.120.1.38

  • Brose, K., Bland, K. S., Wang, K. H., Arnott, D., Henzel, W., Goodman, C. S., Tessier-Lavigne, M., & Kidd, T. (1999). Slit proteins bind Robo receptors and have an evolutionarily conserved role in repulsive axon guidance. Cell, 96(6), 795-806. https://doi.org/10.1016/S0092-8674(00)80590-5

  • Brown, S. L. (1993). The social behaviour of neutered domestic cats [Unpublished doctoral dissertation]. University of Southhampton.

  • Butler, T. (2003). Language and consciousness: What can we learn about feral children? [Master’s thesis]. Memorial University of Newfoundland.

  • Campbell, D. J., Chang, J., & Chawarska, K. (2014). Early generalized overgrowth in autism spectrum disorder: Prevalence rates, gender effects, and clinical outcomes. Journal of the American Academy of Child and Adolescent Psychiatry, 53(10), 1063-1073. https://doi.org/10.1016/j.jaac.2014.07.008

  • Cannell, J. J., & Grant, W. B. (2013). What is the role of vitamin D in autism? Dermato-Endocrinology, 5(1), 199-204. https://doi.org/10.4161/derm.24356

  • Cantù, C., Ierardi, R., Alborelli, I., Fugazza, C., Cassinelli, L., Piconese, S., Bosè, F., Ottolenghi, S., Ferrari, G., & Ronchi, A. (2011). Sox6 enhances erythroid differentiation in human erythroid progenitors. Blood, 117(13), 3669-3679. https://doi.org/10.1182/blood-2010-04-282350

  • Carter, C. S. (2003). Developmental consequences of oxytocin. Physiology & Behavior, 79(3), 383-397. https://doi.org/10.1016/S0031-9384(03)00151-3

  • Carter, C. S., Grippo, A. J., Pournajafi-Nazarloo, H., Ruscio, M. G., & Porges, S. W. (2008). Oxytocin, vasopressin and sociality. Progress in Brain Research, 170, 331-336. https://doi.org/10.1016/S0079-6123(08)00427-5

  • Caspers, K. M., Paradiso, S., Yucuis, R., Troutman, B., Arndt, S., & Philibert, R. (2009). Association between the serotonin transporter promoter polymorphism (5-HTTLPR) and adult unresolved attachment. Developmental Psychology, 45(1), 64-76. https://doi.org/10.1037/a0014026

  • Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W., Taylor, A., & Poulton, R. (2002). Role of genotype in the cycle of violence in maltreated children. Science, 297(5582), 851-854. https://doi.org/10.1126/science.1072290

  • Castelli, F., Frith, C., Happé, F., & Frith, U. (2002). Autism, Asperger syndrome and brain mechanisms for the attribution of mental states to animated shapes. Brain, 125(8), 1839-1849. https://doi.org/10.1093/brain/awf189

  • Chauveau, C., Rowell, J., & Ferreiro, A. (2014). A rising titan: TTN review and mutation update.  Human Mutation35(9), 1046-1059. https://doi.org/10.1002/humu.22611

  • Chawarska, K., Campbell, D., Chen, L., Shic, F., Klin, A., & Chang, J. (2011). Early generalized overgrowth in boys with autism. Archives of General Psychiatry, 68(10), 1021-1031. https://doi.org/10.1001/archgenpsychiatry.2011.106

  • Cieri, R., Churchill, S., Franciscus, R., Tan, J., & Hare, B. (2014). Craniofacial feminization, social tolerance, and the origins of behavioral modernity. Current Anthropology, 55(4), 419-443. https://doi.org/10.1086/677209

  • Cohen, I. L., Liu, X., Lewis, M. E., Chudley, A., Forster-Gibson, C., Gonzalez, M., Jenkins, E. C., Brown, W. T., & Holden, J. J. (2011). Autism severity is associated with child and maternal MAOA genotypes. Clinical Genetics, 79(4), 355-362. https://doi.org/10.1111/j.1399-0004.2010.01471.x

  • Cohen, I. L., Liu, X., Schutz, C., White, B. N., Jenkins, E. C., Brown, W. T., & Holden, J. J. (2003). Association of autism severity with a monoamine oxidase A functional polymorphism. Clinical Genetics, 64(3), 190-197. https://doi.org/10.1034/j.1399-0004.2003.00115.x

  • Colich, N. L., Wang, A. T., Rudie, J. D., Hernandez, L. M., Bookheimer, S. Y., & Dapretto, M. (2012). Atypical neural processing of ironic and sincere remarks in children and adolescents with autism spectrum disorders. Metaphor and Symbol, 27(1), 70-92. https://doi.org/10.1080/10926488.2012.638856

  • Cooper, R. A., & Ritchey, M. (2019). Cortico-hippocampal network connections support the multidimensional quality of episodic memory. eLife, 8, Article e45591. https://doi.org/10.7554/eLife.45591

  • Corballis, M. C. (2018). Space, time, and language. Cognitive Processing, 19, Suppl 189-92. https://doi.org/10.1007/s10339-018-0878-1

  • Corballis, M. C. (2019). Language, memory, and mental time travel: An evolutionary perspective. Frontiers in Human Neuroscience, 13, Article 217. https://doi.org/10.3389/fnhum.2019.00217

  • Corbett, L. (2004) Dingo. In C. Sillero-Zubiri, M. Hoffmann, & D. W. Macdonald (Eds.), Canids: Foxes, wolves, jackals and dogs (pp. 223–230). International Union for Conservation of Nature and Natural Resources.

  • Corbin, J. (2007). Reactive attachment disorder: A biopsychosocial disturbance of attachment. Child & Adolescent Social Work Journal, 24, 539-552. https://doi.org/10.1007/s10560-007-0105-x

  • Corval, R., Mesquita, A., & Soares, I. (2020). Links between reactive attachment disorder, caregiving and temperament: A differential susceptibility perspective. The Journal of Genetic Psychology, 181(5), 405-412. https://doi.org/10.1080/00221325.2020.1756206

  • Courchesne, E., Campbell, K., & Solso, S. (2011). Brain growth across the life span in autism: Age-specific changes in anatomical pathology. Brain Research, 1380, 138-145. https://doi.org/10.1016/j.brainres.2010.09.101

  • Coutinho, A. M., Oliveira, G., Morgadinho, T., Fesel, C., Macedo, T. R., Bento, C., Marques, C., Ataíde, A., Miguel, T., Borges, L., & Vicente, A. M. (2004). Variants of the serotonin transporter gene (SLC6A4) significantly contribute to hyperserotonemia in autism. Molecular Psychiatry, 9(3), 264-271. https://doi.org/10.1038/sj.mp.4001409

  • Cukier, H. N., Rabionet, R., Konidari, I., Rayner-Evans, M. Y., Baltos, M. L., Wright, H. H., Abramson, R. K., Martin, E. R., Cuccaro, M. L., Pericak-Vance, M. A., & Gilbert, J. R. (2010). Novel variants identified in methyl-CpG-binding domain genes in autistic individuals. Neurogenetics, 11(3), 291-303. https://doi.org/10.1007/s10048-009-0228-7

  • Curin, J. M., Terzić, J., Petković, Z. B., Zekan, L., Terzić, I. M., & Susnjara, I. M. (2003). Lower cortisol and higher ACTH levels in individuals with autism. Journal of Autism and Developmental Disorders, 33(4), 443-448. https://doi.org/10.1023/A:1025019030121

  • Curtiss, S., Fromkin, V., Krashen, S., Rigler, M., & Rigler, D. (1974). The linguistic development of Genie. Language, 50(3), 528-554. https://doi.org/10.2307/412222

  • Cuyvers, B., Vervoort, E., & Bosmans, G. (2020). Reactive attachment disorder symptoms and prosocial behavior in middle childhood: The role of Secure Base Script knowledge. BMC Psychiatry, 20(1), Article 524. https://doi.org/10.1186/s12888-020-02931-3

  • Damaj, L., Lupien-Meilleur, A., Lortie, A., Riou, É., Ospina, L. H., Gagnon, L., Vanasse, C., & Rossignol, E. (2015). CACNA1A haploinsufficiency causes cognitive impairment, autism and epileptic encephalopathy with mild cerebellar symptoms. European Journal of Human Genetics, 23(11), 1505-1512. https://doi.org/10.1038/ejhg.2015.21

  • Daniels, T. J., & Bekoff, M. (1989). Feralization: The making of wild domestic animals. Behavioural Processes, 19(1-3), 79-94. https://doi.org/10.1016/0376-6357(89)90032-6

  • Deacon, T. W. (2009). Relaxed selection and the role of epigenesis in the evolution of language. In M. Blumberg, J. Freeman, & S. Robinson (Eds.), Oxford handbook of developmental behavioral neuroscience (pp. 730–752). Oxford University Press.

  • Derelle, R., Kondrashov, F. A., Arkhipov, V. Y., Corbel, H., Frantz, A., Gasparini, J., Jacquin, L., Jacob, G., Thibault, S., & Baudry, E. (2013). Color differences among feral pigeons (Columba livia) are not attributable to sequence variation in the coding region of the melanocortin-1 receptor gene (MC1R). BMC Research Notes, 6, Article 310. https://doi.org/10.1186/1756-0500-6-310

  • De Rubeis, S., He, X., Goldberg, A. P., Poultney, C. S., Samocha, K., Cicek, A. E., Kou, Y., Liu, L., Fromer, M., Walker, S., Singh, T., Klei, L., Kosmicki, J., Shih-Chen, F., Aleksic, B., Biscaldi, M., Bolton, P. F., Brownfeld, J. M., Cai, J., … Buxbaum, J. D. (2014). Synaptic, transcriptional and chromatin genes disrupted in autism. Nature, 515(7526), 209-215. https://doi.org/10.1038/nature13772

  • Di Sante, M., Sylvestre, A., Bouchard, C., & Leblond, J. (2019). The pragmatic language skills of severely neglected 42-month-old children: Results of the ELLAN study. Child Maltreatment, 24(3), 244-253. https://doi.org/10.1177/1077559519828838

  • Di Sante, M., Sylvestre, A., Bouchard, C., & Leblond, J. (2020). Parental behaviors associated with the level of pragmatic language ability among 42-month-old neglected children. Child Abuse & Neglect, 104, Article 104482. https://doi.org/10.1016/j.chiabu.2020.104482

  • Dobrova-Krol, N. A., van Ijzendoorn, M. H., Bakermans-Kranenburg, M. J., Cyr, C., & Juffer, F. (2008). Physical growth delays and stress dysregulation in stunted and non-stunted Ukrainian institution-reared children. Infant Behavior and Development, 31(3), 539-553. https://doi.org/10.1016/j.infbeh.2008.04.001

  • Eales, L. A. (1989). The influences of visual and vocal interaction on song learning in zebra finches. Animal Behaviour, 37(Part 3), 507-508. https://doi.org/10.1016/0003-3472(89)90097-3

  • Eigsti, I. M., Bennetto, L., & Dadlani, M. B. (2007). Beyond pragmatics: Morphosyntactic development in autism. Journal of Autism and Developmental Disorders, 37(6), 1007-1023. https://doi.org/10.1007/s10803-006-0239-2

  • El-Baz, F., Hamza, R. T., Ayad, M. S., & Mahmoud, N. H. (2014). Hyperandrogenemia in male autistic children and adolescents: Relation to disease severity. International Journal of Adolescent Medicine and Health, 26(1), 79-84. https://doi.org/10.1515/ijamh-2012-0116

  • Ellis, E. E., Yilanli, M., & Saadabadi, A. (2020). Reactive attachment disorder. StatPearls. https://www.ncbi.nlm.nih.gov/books/NBK537155/

  • Evin, A., Dobney, K., Schafberg, R., Owen, J., Vidarsdottir, U. S., Larson, G., & Cucchi, T. (2015). Phenotype and animal domestication: A study of dental variation between domestic, wild, captive, hybrid and insular Sus scrofa. BMC Evolutionary Biology, 15(1), Article 6. https://doi.org/10.1186/s12862-014-0269-x

  • Farmer, C., Butter, E., Mazurek, M. O., Cowan, C., Lainhart, J., Cook, E. H., DeWitt, M. B., & Aman, M. (2015). Aggression in children with autism spectrum disorders and a clinic-referred comparison group. Autism, 19(3), 281-291. https://doi.org/10.1177/1362361313518995

  • Feulner, P. G. D., Gratten, J., Kijas, J. W., Visscher, P. M., Pemberton, J. M., & Slate, J. (2013). Introgression and the fate of domesticated genes in a wild mammal population. Molecular Ecology, 22(16), 4210-4221. https://doi.org/10.1111/mec.12378

  • Fisher, S. E., Francks, C., Marlow, A. J., MacPhie, I. L., Newbury, D. F., Cardon, L. R., Ishikawa-Brush, Y., Richardson, A. J., Talcott, J. B., Gayán, J., Olson, R. K., Pennington, B. F., Smith, S. D., DeFries, J. C., Stein, J. F., & Monaco, A. P. (2002). Independent genome-wide scans identify a chromosome 18 quantitative-trait locus influencing dyslexia. Nature Genetics, 30(1), 86-91. https://doi.org/10.1038/ng792

  • Fries, A. B., Ziegler, T. E., Kurian, J. R., Jacoris, S., & Pollak, S. D. (2005). Early experience in humans is associated with changes in neuropeptides critical for regulating social behavior. Proceedings of the National Academy of Sciences of the United States of America, 102(47), 17237-17240. https://doi.org/10.1073/pnas.0504767102

  • Fromkin, V., Krashen, S., Curtiss, S., Rigler, D., & Rigler, M. (1974). The development of language in Genie: A case of language acquisition beyond the ‘critical period’. Brain and Language, 1(1), 81-107. https://doi.org/10.1016/0093-934X(74)90027-3

  • Fukase, H., Kondo, O., & Ishida, H. (2015). Size and placement of developing anterior teeth in immature Neanderthal mandibles from Dederiyeh Cave, Syria: Implications for emergence of the modern human chin. American Journal of Physical Anthropology, 156(3), 482-488. https://doi.org/10.1002/ajpa.22665

  • Furuhashi, M. (2019). Fatty acid-binding protein 4 in cardiovascular and metabolic diseases. Journal of Atherosclerosis and Thrombosis, 26(3), 216-232. https://doi.org/10.5551/jat.48710

  • Fusaroli, R., Lambrechts, A., Bang, D., Bowler, D. M., & Gaigg, S. B. (2017). Is voice a marker for Autism spectrum disorder? A systematic review and meta-analysis. Autism Research, 10(3), 384-407. https://doi.org/10.1002/aur.1678

  • Garcia, M., Gingras, B., Bowling, D. L., Herbst, C. T., Boeckle, M., Locatelli, Y., & Fitch, W. T. (2016). Structural classification of wild boar (sus scrofa) vocalizations. Ethology, 122(4), 329-342. https://doi.org/10.1111/eth.12472

  • Gering, E., Incorvaia, D., Henriksen, R., Conner, J., Getty, T., & Wright, D. (2019). Getting back to nature: Feralization in animals and plants. Trends in Ecology & Evolution, 34(12), 1137-1151. https://doi.org/10.1016/j.tree.2019.07.018

  • Gering, E., Johnsson, M., Willis, P., Getty, T., & Wright, D. (2015). Mixed ancestry and admixture in Kauai’s feral chickens: Invasion of domestic genes into ancient Red Junglefowl reservoirs. Molecular Ecology, 24(9), 2112-2124. https://doi.org/10.1111/mec.13096

  • Gesell, A. (1942). Wolf child and human child. The Scientific Book Club.

  • Ghezzi, D., Antognazza, M. R., Dal Maschio, M., Lanzarini, E., Benfenati, F., & Lanzani, G. (2011). A hybrid bioorganic interface for neuronal photoactivation. Nature Communications, 2, Article 166. https://doi.org/10.1038/ncomms1164

  • Gleason, M. M., Fox, N. A., Drury, S., Smyke, A., Egger, H. L., Nelson, C. A., III, Gregas, M. C., & Zeanah, C. H. (2011). Validity of evidence-derived criteria for reactive attachment disorder: Indiscriminately social/disinhibited and emotionally withdrawn/inhibited types. Journal of the American Academy of Child and Adolescent Psychiatry, 50(3), 216-231.e3. https://doi.org/10.1016/j.jaac.2010.12.012

  • Gleeson, B. T., & Kushnick, G. (2018). Female status, food security, and stature sexual dimorphism: Testing mate choice as a mechanism in human self-domestication. American Journal of Physical Anthropology, 167(3), 458-469. https://doi.org/10.1002/ajpa.23642

  • Goding, C. R., & Arnheiter, H. (2019). MITF-the first 25 years. Genes & Development, 33(15-16), 983-1007. https://doi.org/10.1101/gad.324657.119

  • Gómez-Lado, C., Eirís-Puñal, J., Blanco-Barca, O., del Río-Latorre, E., Fernández-Redondo, V., & Castro-Gago, M. (2004). Hipomelanosis de Ito. Un síndrome neurocutáneo heterogéneo y posiblemente infradiagnosticado [Hypomelanosis of Ito. A possibly under-diagnosed heterogeneous neurocutaneous syndrome]. Revista de Neurología, 38(3), 223-228. https://doi.org/10.33588/rn.3803.2003618

  • Goozen, S. H., van den Ban, E., Matthys, W., Cohen-Kettenis, P. T., Thijssen, J. H., & van Engeland, H. (2000). Increased adrenal androgen functioning in children with oppositional defiant disorder: A comparison with psychiatric and normal controls. Journal of the American Academy of Child and Adolescent Psychiatry, 39(11), 1446-1451. https://doi.org/10.1097/00004583-200011000-00020

  • Green, J., & Goldwyn, R. (2002). Annotation: attachment disorganisation and psychopathology: new findings in attachment research and their potential implications for developmental psychopathology in childhood. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 43(7), 835-846. https://doi.org/10.1111/1469-7610.00102

  • Gupta, A., Labus, J., Kilpatrick, L. A., Bonyadi, M., Ashe-McNalley, C., Heendeniya, N., Bradesi, S., Chang, L., & Mayer, E. A. (2016). Interactions of early adversity with stress-related gene polymorphisms impact regional brain structure in females. Brain Structure & Function, 221(3), 1667-1679. https://doi.org/10.1007/s00429-015-0996-9

  • Hamza, R. T., Hewedi, D. H., & Ismail, M. A. (2010). Basal and adrenocorticotropic hormone stimulated plasma cortisol levels among Egyptian autistic children: Relation to disease severity. Italian Journal of Pediatrics, 36, Article 71. https://doi.org/10.1186/1824-7288-36-71

  • Happé, F. G. (1993). Communicative competence and theory of mind in autism: A test of relevance theory. Cognition, 48(2), 101-119. https://doi.org/10.1016/0010-0277(93)90026-R

  • Happé, F. G. (1995). The role of age and verbal ability in the theory of mind task performance of subjects with autism. Child Development, 66(3), 843-855. https://doi.org/10.2307/1131954

  • Hare, B. (2017). Survival of the friendliest: Homo sapiens evolved via selection for prosociality. Annual Review of Psychology, 68, 155-186. https://doi.org/10.1146/annurev-psych-010416-044201

  • Hare, B., & Tomasello, M. (1999). Domestic dogs (Canis familiaris) use human and conspecific social cues to locate hidden food. Journal of Comprehensive Psychology, 113(2), 173-177. https://doi.org/10.1037/0735-7036.113.2.173

  • Hare, B., Wobber, V., & Wrangham, R. W. (2012). The self-domestication hypothesis: Bonobos evolved due to selection against male aggression. Animal Behaviour, 83(3), 573-585. https://doi.org/10.1016/j.anbehav.2011.12.007

  • Hare, B., & Woods, V. (2020). Survival of the friendliest: Understanding our origins and rediscovering our common humanity. Penguin Random House.

  • Hashimoto, T., Aihara, R., Tayama, M., Miyazaki, M., Shirakawa, Y., & Kuroda, Y. (1991). Reduced thyroid-stimulating hormone response to thyrotropin-releasing hormone in autistic boys. Developmental Medicine and Child Neurology, 33(4), 313-319. https://doi.org/10.1111/j.1469-8749.1991.tb14882.x

  • Haughian, J. M., & Bradford, A. P. (2009). Protein kinase C alpha (PKCalpha) regulates growth and invasion of endometrial cancer cells. Journal of Cellular Physiology, 220(1), 112-118. https://doi.org/10.1002/jcp.21741

  • Hauth, I., de Bruijn, Y. G., Staal, W., Buitelaar, J. K., & Rommelse, N. N. (2014). Testing the extreme male brain theory of autism spectrum disorder in a familial design. Autism Research7(4), 491-500. https://doi.org/10.1002/aur.1384

  • Heim, C. M., Mayberg, H. S., Mletzko, T., Nemeroff, C. B., & Pruessner, J. C. (2013). Decreased cortical representation of genital somatosensory field after childhood sexual abuse. The American Journal of Psychiatry, 170(6), 616-623. https://doi.org/10.1176/appi.ajp.2013.12070950

  • Heinrichs, M., von Dawans, B., & Domes, G. (2009). Oxytocin, vasopressin, and human social behavior. Frontiers in Neuroendocrinology, 30(4), 548-557. https://doi.org/10.1016/j.yfrne.2009.05.005

  • Hernádi, A., Kis, A., Turcsán, B., & Topál, J. (2012). Man’s underground best friend: Domestic ferrets, unlike the wild forms, show evidence of dog-like social-cognitive skills. PLoS One, 7(8), Article e43267. https://doi.org/10.1371/journal.pone.0043267

  • Herrmann, E., Hare, B., Cissewski, J., & Tomasello, M. (2011). A comparison of temperament in nonhuman apes and human infants. Developmental Science, 14(6), 1393-1405. https://doi.org/10.1111/j.1467-7687.2011.01082.x

  • Hinners, I., Wendler, F., Fei, H., Thomas, L., Thomas, G., & Tooze, S. A. (2003). AP-1 recruitment to VAMP4 is modulated by phosphorylation-dependent binding of PACS-1. EMBO Reports, 4(12), 1182-1189. https://doi.org/10.1038/sj.embor.7400018

  • Hinshaw-Fuselier, S., Boris, N. W., & Zeanah, C. Z. (1999). Reactive attachment disorder in maltreated twins.  Infant Mental Health Journal, 20(1), 42-59.

  • Hollocks, M. J., Howlin, P., Papadopoulos, A. S., Khondoker, M., & Simonoff, E. (2014). Differences in HPA-axis and heart rate responsiveness to psychosocial stress in children with autism spectrum disorders with and without co-morbid anxiety. Psychoneuroendocrinology, 46, 32-45. https://doi.org/10.1016/j.psyneuen.2014.04.004

  • Huber, L., Salobir, K., Mundry, R., & Cimarelli, G. (2020). Selective overimitation in dogs. Learning & Behavior, 48, 113-123. https://doi.org/10.3758/s13420-019-00400-w

  • Iossifov, I., O’Roak, B. J., Sanders, S. J., Ronemus, M., Krumm, N., Levy, D., Stessman, H. A., Witherspoon, K. T., Vives, L., Patterson, K. E., Smith, J. D., Paeper, B., Nickerson, D. A., Dea, J., Dong, S., Gonzalez, L. E., Mandell, J. D., Mane, S. M., Murtha, M. T., … Wigler, M. (2014). The contribution of de novo coding mutations to autism spectrum disorder. Nature, 515(7526), 216-221. https://doi.org/10.1038/nature13908

  • Iossifov, I., Ronemus, M., Levy, D., Wang, Z., Hakker, I., Rosenbaum, J., Yamrom, B., Lee, Y. H., Narzisi, G., Leotta, A., Kendall, J., Grabowska, E., Ma, B., Marks, S., Rodgers, L., Stepansky, A., Troge, J., Andrews, P., Bekritsky, M., … Wigler, M. (2012). De novo gene disruptions in children on the autistic spectrum. Neuron, 74(2), 285-299. https://doi.org/10.1016/j.neuron.2012.04.009

  • Itoh, M., Kim, C. H., Palardy, G., Oda, T., Jiang, Y. J., Maust, D., Yeo, S. Y., Lorick, K., Wright, G. J., Ariza-McNaughton, L., Weissman, A. M., Lewis, J., Chandrasekharappa, S. C., & Chitnis, A. B. (2003). Mind bomb is a ubiquitin ligase that is essential for efficient activation of Notch signaling by Delta. Developmental Cell, 4(1), 67-82. https://doi.org/10.1016/S1534-5807(02)00409-4

  • Jacobson, L. (2014). Hypothalamic-pituitary-adrenocortical axis: Neuropsychiatric aspects. Comprehensive Physiology, 4(2), 715-738. https://doi.org/10.1002/cphy.c130036

  • Jensen, P. (2006). Domestication: From behavior to genes and back again. Applied Animal Behaviour Science, 97(1), 3-15. https://doi.org/10.1016/j.applanim.2005.11.015

  • John, S., & Jaeggi, A. V. (2021). Oxytocin levels tend to be lower in autistic children: A meta-analysis of 31 studies. Autism, 25(8), 2152-2161. https://doi.org/10.1177/13623613211034375

  • Johnston, A. M., Turrin, C., Watson, L., Arre, A. M., & Santos, L. R. (2017). Uncovering the origins of dog-human eye contact: Dingoes establish eye contact more than wolves, but less than dogs. Animal Behaviour, 133, 123-129. https://doi.org/10.1016/j.anbehav.2017.09.002

  • Jung, M., Takiguchi, S., Hamamura, S., Mizuno, Y., Kosaka, H., & Tomoda, A. (2020). Thalamic volume is related to increased anterior thalamic radiations in children with reactive attachment disorder. Cerebral Cortex, 30(7), 4238-4245. https://doi.org/10.1093/cercor/bhaa051

  • Kaland, N., Møller-Nielsen, A., Callesen, K., Mortensen, E. L., Gottlieb, D., & Smith, L. (2002). A new ‘advanced’ test of theory of mind: Evidence from children and adolescents with Asperger syndrome. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 43(4), 517-528. https://doi.org/10.1111/1469-7610.00042

  • Kaminski, J., & Nitzschner, M. (2013). Do dogs get the point? A review of dog–human communication ability. Learning and Motivation, 44(4), 294-302. https://doi.org/10.1016/j.lmot.2013.05.001

  • Kapoor, S. (1973). Socialization and feral children. Revista Internacional de Sociologia, 31(5–6), 195-213.

  • Katoh, M. (2001). Molecular cloning and characterization of MFRP, a novel gene encoding a membrane-type Frizzled-related protein. Biochemical and Biophysical Research Communications, 282(1), 116-123. https://doi.org/10.1006/bbrc.2001.4551

  • Kemph, J., & Voeller, K. (2008). Reactive attachment disorder in adolescence. Adolescent Psychiatry, 30, 159-178.

  • Kent, R. D., & Murray, A. D. (1982). Acoustic features of infant vocalic utterances at 3, 6, and 9 months. The Journal of the Acoustical Society of America, 72(2), 353-365. https://doi.org/10.1121/1.388089

  • Kidd, T., Brose, K., Mitchell, K. J., Fetter, R. D., Tessier-Lavigne, M., Goodman, C. S., & Tear, G. (1998). Roundabout controls axon crossing of the CNS midline and defines a novel subfamily of evolutionarily conserved guidance receptors. Cell, 92(2), 205-215. https://doi.org/10.1016/S0092-8674(00)80915-0

  • Kilsby, A. J., Cruwys, M., Kukendrajah, C., Russell-Eggitt, I., Raglan, E., Rajput, K., Lohse, P., & Brady, A. F. (2013). Homozygosity for piebaldism with a proven KIT mutation resulting in depigmentation of the skin and hair, deafness, developmental delay and autism spectrum disorder. Clinical Dysmorphology, 22(2), 64-67. https://doi.org/10.1097/MCD.0b013e32835e8ce5

  • Kissine, M. (2012). Pragmatics, cognitive flexibility and autism spectrum disorders. Mind & Language, 27(1), 1-28. https://doi.org/10.1111/j.1468-0017.2011.01433.x

  • Kjelgaard, M. M., & Tager-Flusberg, H. (2001). An investigation of language impairment in autism: Implications for genetic subgroups. Language and Cognitive Processes, 16(2-3), 287-308. https://doi.org/10.1080/01690960042000058

  • Knickmeyer, R. C., Wheelwright, S., Hoekstra, R., & Baron-Cohen, S. (2006). Age of menarche in females with autism spectrum conditions. Developmental Medicine and Child Neurology, 48(12), 1007-1008. https://doi.org/10.1017/S0012162206222229

  • Kočovská, E., Wilson, P., Young, D., Wallace, A. M., Gorski, C., Follan, M., Smillie, M., Puckering, C., Barnes, J., Gillberg, C., & Minnis, H. (2013). Cortisol secretion in children with symptoms of reactive attachment disorder. Psychiatry Research, 209(1), 74-77. https://doi.org/10.1016/j.psychres.2012.12.011

  • Komiyama, T., Iwama, H., Osada, N., Nakamura, Y., Kobayashi, H., Tateno, Y., & Gojobori, T. (2014). Dopamine receptor genes and evolutionary differentiation in the domestication of fighting cocks and long-crowing chickens. PLoS One, 9(7), Article e101778. https://doi.org/10.1371/journal.pone.0101778

  • Krause, E. T., Kjaer, J. B., Dudde, A., Schrader, L., & Phi-van, L. (2019). Fear but not social behaviour is affected by a polymorphism in the 5′-flanking region of the serotonin transporter (5-HTT) gene in adult hens. Behavioural Brain Research, 361, 50-53. https://doi.org/10.1016/j.bbr.2018.12.029

  • Krug, A., Nieratschker, V., Markov, V., Krach, S., Jansen, A., Zerres, K., Eggermann, T., Stöcker, T., Shah, N. J., Treutlein, J., Mühleisen, T. W., & Kircher, T. (2010). Effect of CACNA1C rs1006737 on neural correlates of verbal fluency in healthy individuals. NeuroImage, 49(2), 1831-1836. https://doi.org/10.1016/j.neuroimage.2009.09.028

  • Kruska, D. C. (2005). On the evolutionary significance of encephalization in some eutherian mammals: Effects of adaptive radiation, domestication, and feralization. Brain, Behavior and Evolution, 65(2), 73-108. https://doi.org/10.1159/000082979

  • Kruska, D. C. T. (2014). Comparative quantitative investigations on brains of wild cavies (Cavia aperea) and guinea pigs (Cavia aperea f. porcellus) A contribution to size changes of CNS structures due to domestication. Mammalian Biology, 79(4), 230-239. https://doi.org/10.1016/j.mambio.2013.12.005

  • Kruska, D., & Röhrs, M. (1974). Comparative-quantitative investigations on brains of feral pigs from the Galapagos Islands and of European domestic pigs. Zeitschrift für Anatomie und Entwicklungsgeschichte, 144(1), 61-73. https://doi.org/10.1007/BF00518633

  • Kumar, D., Dedic, N., Flachskamm, C., Voulé, S., Deussing, J. M., & Kimura, M. (2015). Cacna1c (Cav1.2) modulates electroencephalographic rhythm and rapid eye movement sleep recovery. Sleep, 38(9), 1371-1380. https://doi.org/10.5665/sleep.4972

  • Küry, S., Besnard, T., Ebstein, F., Khan, T. N., Gambin, T., Douglas, J., Bacino, C. A., Craigen, W. J., Sanders, S. J., Lehmann, A., Latypova, X., Khan, K., Pacault, M., Sacharow, S., Glaser, K., Bieth, E., Perrin-Sabourin, L., Jacquemont, M. L., Cho, M. T., … Isidor, B. (2017). De novo disruption of the proteasome regulatory subunit PSMD12 causes a syndromic neurodevelopmental disorder. American Journal of Human Genetics, 100(2), 352-363. https://doi.org/10.1016/j.ajhg.2017.01.003

  • Lakatos, K., Toth, I., Nemoda, Z., Ney, K., Sasvari-Szekely, M., & Gervai, J. (2000). Dopamine D4 receptor (DRD4) gene polymorphism is associated with attachment disorganization in infants. Molecular Psychiatry, 5(6), 633-637. https://doi.org/10.1038/sj.mp.4000773

  • Lane, H. (1976) The wild boy of Aveyron. Harvard University Press.

  • Langley, S. A., Miga, K. H., Karpen, G. H., & Langley, C. H. (2019). Haplotypes spanning centromeric regions reveal persistence of large blocks of archaic DNA. eLife, 8, Article e42989. https://doi.org/10.7554/eLife.42989

  • Leach, H. M. (2003). Human domestication reconsidered. Current Anthropology, 44(3), 349-368. https://doi.org/10.1086/368119

  • Lebrun, Y. (1980). Victor of Aveyron: A reappraisal in light of more recent cases of feral speech. Language Sciences, 2(1), 32-43.

  • Linderholm, A., Spencer, D., Battista, V., Frantz, L., Barnett, R., Fleischer, R. C., James, H. F., Duffy, D., Sparks, J. P., Clements, D. R., Andersson, L., Dobney, K., Leonard, J. A., & Larson, G. (2016). A novel MC1R allele for black coat colour reveals the Polynesian ancestry and hybridization patterns of Hawaiian feral pigs. Royal Society Open Science, 3(9), Article 160304. https://doi.org/10.1098/rsos.160304

  • Lindgren, K. A., Folstein, S. E., Tomblin, J. B., & Tager-Flusberg, H. (2009). Language and reading abilities of children with autism spectrum disorders and specific language impairment and their first-degree relatives. Autism Research, 2(1), 22-38. https://doi.org/10.1002/aur.63

  • Lord, K., Feinstein, M., Smith, B., & Coppinger, R. (2013). Variation in reproductive traits of members of the genus Canis with special attention to the domestic dog (Canis familiaris). Behavioural Processes, 92, 131-142. https://doi.org/10.1016/j.beproc.2012.10.009

  • Lord, C., & Paul, R. (1997). Language and communication in autism. In D. J. Cohen and F. R. Volkmar (Eds.), Handbook of autism and pervasive developmental disorders (2nd ed., pp. 195–225). Wiley.

  • Lourenco-Jaramillo, D. L., Sifuentes-Rincón, A. M., Parra-Bracamonte, G. M., de la Rosa-Reyna, X. F., Segura-Cabrera, A., & Arellano-Vera, W. (2012). Genetic diversity of tyrosine hydroxylase (TH) and dopamine β-hydroxylase (DBH) genes in cattle breeds. Genetics and Molecular Biology, 35(2), 435-440. https://doi.org/10.1590/S1415-47572012000300009

  • Luijk, M. P., Roisman, G. I., Haltigan, J. D., Tiemeier, H., Booth-Laforce, C., van Ijzendoorn, M. H., Belsky, J., Uitterlinden, A. G., Jaddoe, V. W., Hofman, A., Verhulst, F. C., Tharner, A., & Bakermans-Kranenburg, M. J. (2011). Dopaminergic, serotonergic, and oxytonergic candidate genes associated with infant attachment security and disorganization? In search of main and interaction effects. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 52(12), 1295-1307. https://doi.org/10.1111/j.1469-7610.2011.02440.x

  • Luppanapornlarp, S., Leelataweewud, P., Putongkam, P., & Ketanont, S. (2010). Periodontal status and orthodontic treatment need of autistic children. World Journal of Orthodontics, 11(3), 256-261.

  • Lyst, M. J., Ekiert, R., Ebert, D. H., Merusi, C., Nowak, J., Selfridge, J., Guy, J., Kastan, N. R., Robinson, N. D., de Lima Alves, F., Rappsilber, J., Greenberg, M. E., & Bird, A. (2013). Rett syndrome mutations abolish the interaction of MeCP2 with the NCoR/SMRT co-repressor. Nature Neuroscience, 16(7), 898-902. https://doi.org/10.1038/nn.3434

  • Maekawa, M., Ohnishi, T., Toyoshima, M., Shimamoto-Mitsuyama, C., Hamazaki, K., Balan, S., Wada, Y., Esaki, K., Takagai, S., Tsuchiya, K. J., Nakamura, K., Iwata, Y., Nara, T., Iwayama, Y., Toyota, T., Nozaki, Y., Ohba, H., Watanabe, A., Hisano, … Yoshikawa, T. (2020). A potential role of fatty acid binding protein 4 in the pathophysiology of autism spectrum disorder. Brain Communications, 2(2), Article fcaa145. https://doi.org/10.1093/braincomms/fcaa145

  • Makita, K., Takiguchi, S., Naruse, H., Shimada, K., Morioka, S., Fujisawa, T. X., Shimoji, K., & Tomoda, A. (2020). White matter changes in children and adolescents with reactive attachment disorder: A diffusion tensor imaging study. Psychiatry Research: Neuroimaging, 303, Article 111129. https://doi.org/10.1016/j.pscychresns.2020.111129

  • Manoli, D. S., & State, M. W. (2021). Autism spectrum disorder genetics and the search for pathological mechanisms. The American Journal of Psychiatry, 178(1), 30-38. https://doi.org/10.1176/appi.ajp.2020.20111608

  • Manouilenko, I., Eriksson, J. M., Humble, M. B., & Bejerot, S. (2014). Minor physical anomalies in adults with autism spectrum disorder and healthy controls. Autism Research and Treatment, 2014, Article 743482. https://doi.org/10.1155/2014/743482

  • Márquez, S., Pagano, A. S., Delson, E., Lawson, W., & Laitman, J. T. (2014). The nasal complex of Neanderthals: An entry portal to their place in human ancestry. Anatomical Record, 297(11), 2121-2137. https://doi.org/10.1002/ar.23040

  • Márquez, C., Poirier, G. L., Cordero, M. I., Larsen, M. H., Groner, A., Marquis, J., Magistretti, P. J., Trono, D., & Sandi, C. (2013). Peripuberty stress leads to abnormal aggression, altered amygdala and orbitofrontal reactivity and increased prefrontal MAOA gene expression. Translational Psychiatry, 3(1), Article e216. https://doi.org/10.1038/tp.2012.144

  • Marui, T., Hashimoto, O., Nanba, E., Kato, C., Tochigi, M., Umekage, T., Ishijima, M., Kohda, K., Kato, N., & Sasaki, T. (2004). Association between the neurofibromatosis-1 (NF1) locus and autism in the Japanese population. American Journal of Medical Genetics. Part B, Neuropsychiatric Genetics, 131B(1), 43-47. https://doi.org/10.1002/ajmg.b.20119

  • Maud, C., Ryan, J., McIntosh, J. E., & Olsson, C. A. (2018). The role of oxytocin receptor gene (OXTR) DNA methylation (DNAm) in human social and emotional functioning: A systematic narrative review. BMC Psychiatry, 18(1), Article 154. https://doi.org/10.1186/s12888-018-1740-9

  • McCann, J., Peppé, S., Gibbon, F. E., O’Hare, A., & Rutherford, M. (2007). Prosody and its relationship to language in school-aged children with high-functioning autism. International Journal of Language & Communication Disorders, 42(6), 682-702. https://doi.org/10.1080/13682820601170102

  • McNeil, M., Polloway, E., & Smith, D. (1984). Feral and isolated children: Historical review and analysis. Education and Training of the Mentally Retarded, 19(1), 70-79.

  • Miellet, S., Caldara, R., Gillberg, C., Raju, M., & Minnis, H. (2014). Disinhibited reactive attachment disorder symptoms impair social judgements from faces. Psychiatry Research, 215(3), 747-752. https://doi.org/10.1016/j.psychres.2014.01.004

  • Mikita, N., Hollocks, M. J., Papadopoulos, A. S., Aslani, A., Harrison, S., Leibenluft, E., Simonoff, E., & Stringaris, A. (2015). Irritability in boys with autism spectrum disorders: An investigation of physiological reactivity. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 56(10), 1118-1126. https://doi.org/10.1111/jcpp.12382

  • Minnis, H., Reekie, J., Young, D., O’Connor, T., Ronald, A., Gray, A., & Plomin, R. (2007). Genetic, environmental and gender influences on attachment disorder behaviours. The British Journal of Psychiatry, 190(6), 490-495. https://doi.org/10.1192/bjp.bp.105.019745

  • Moles, A., Kieffer, B. L., & D’Amato, F. R. (2004). Deficit in attachment behavior in mice lacking the mu-opioid receptor gene. Science, 304(5679), 1983-1986. https://doi.org/10.1126/science.1095943

  • Mosconi, M. W., Cody-Hazlett, H., Poe, M. D., Gerig, G., Gimpel-Smith, R., & Piven, J. (2009). Longitudinal study of amygdala volume and joint attention in 2- to 4-year-old children with autism. Archives of General Psychiatry, 66(5), 509-516. https://doi.org/10.1001/archgenpsychiatry.2009.19

  • Murphy, C. M., Deeley, Q., Daly, E. M., Ecker, C., O'Brien, F. M., Hallahan, B., Loth, E., Toal, F., Reed, S., Hales, S., Robertson, D. M., Craig, M. C., Mullins, D., Barker, G. J., Lavender, T., Johnston, P., Murphy, K. C., & Murphy, D. G. (2012). Anatomy and aging of the amygdala and hippocampus in autism spectrum disorder: An in vivo magnetic resonance imaging study of Asperger syndrome. Autism Research, 5(1), 3-12. https://doi.org/10.1002/aur.227

  • Naber, F. B., Bakermans-Kranenburg, M. J., van Ijzendoorn, M. H., Swinkels, S. H., Buitelaar, J. K., Dietz, C., van Daalen, E., & van Engeland, H. (2008). Play behavior and attachment in toddlers with autism. Journal of Autism and Developmental Disorders, 38(5), 857-866. https://doi.org/10.1007/s10803-007-0454-5

  • Nagasawa, M., Mitsui, S., En, S., Ohtani, N., Ohta, M., Sakuma, Y., Onaka, T., Mogi, K., & Kikusui, T. (2015). Social evolution. Oxytocin-gaze positive loop and the coevolution of human-dog bonds. Science, 348(6232), 333-336. https://doi.org/10.1126/science.1261022

  • Naumann, N., Lübke, J., Shomali, W., Reiter, L., Horny, H. P., Jawhar, M., Dangelo, V., Fabarius, A., Metzgeroth, G., Kreil, S., Sotlar, K., Oni, C., Harrison, C., Hofmann, W. K., Cross, N., Valent, P., Radia, D., Gotlib, J., Reiter, A., & Schwaab, J. (2021). Clinical and histopathological features of myeloid neoplasms with concurrent Janus kinase 2 (JAK2) V617F and KIT proto-oncogene, receptor tyrosine kinase (KIT) D816V mutations. British Journal of Haematology, 194(2), 344-354. https://doi.org/10.1111/bjh.17567

  • Nicastro N. (2004). Perceptual and acoustic evidence for species-level differences in meow vocalizations by domestic cats (Felis catus) and African wild cats (Felis silvestris lybica). Journal of Comparative Psychology, 118(3), 287-296. https://doi.org/10.1037/0735-7036.118.3.287

  • Niego, A., & Benítez-Burraco, A. (2019). Williams syndrome, human self-domestication, and language evolution. Frontiers in Psychology, 10, Article 521. https://doi.org/10.3389/fpsyg.2019.00521

  • Niego, A., & Benítez-Burraco, A. (2021). Are feralization and domestication truly mirror processes? Ethology Ecology and Evolution. Advance online publication. https://doi.org/10.1080/03949370.2021.1975314

  • Nikolsky, A., & Benítez-Burraco, A. (2022). Human aggression and music evolution: A model. PsyArxiv. https://doi.org/10.31234/osf.io/a8up7

  • Nikulina, E. M. (1991). Neural control of predatory aggression in wild and domesticated animals. Neuroscience and Biobehavioral Reviews, 15(4), 545-547. https://doi.org/10.1016/S0149-7634(05)80146-0

  • Nillni, E. A., & Sevarino, K. A. (1999). The biology of pro-thyrotropin-releasing hormone-derived peptides. Endocrine Reviews, 20(5), 599-648. https://doi.org/10.1210/edrv.20.5.0379

  • Okanoya, K. (2017). Sexual communication and domestication may give rise to the signal complexity necessary for the emergence of language: An indication from songbird studies. Psychonomic Bulletin & Review, 24(1), 106-110. https://doi.org/10.3758/s13423-016-1165-8

  • Oliva, J. L., Rault, J. L., Appleton, B., & Lill, A. (2015). Oxytocin enhances the appropriate use of human social cues by the domestic dog (Canis familiaris) in an object choice task. Animal Cognition, 18(3), 767-775. https://doi.org/10.1007/s10071-015-0843-7

  • O’Roak, B. J., Deriziotis, P., Lee, C., Vives, L., Schwartz, J. J., Girirajan, S., Karakoc, E., Mackenzie, A. P., Ng, S. B., Baker, C., Rieder, M. J., Nickerson, D. A., Bernier, R., Fisher, S. E., Shendure, J., & Eichler, E. E. (2011). Exome sequencing in sporadic autism spectrum disorders identifies severe de novo mutations. Nature Genetics, 43(6), 585-589. https://doi.org/10.1038/ng.835

  • O’Roak, B. J., Vives, L., Girirajan, S., Karakoc, E., Krumm, N., Coe, B. P., Levy, R., Ko, A., Lee, C., Smith, J. D., Turner, E. H., Stanaway, I. B., Vernot, B., Malig, M., Baker, C., Reilly, B., Akey, J. M., Borenstein, E., Rieder, M. J., … Eichler, E. E. (2012). Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature, 485(7397), 246-250. https://doi.org/10.1038/nature10989

  • Owens, J. L., Olsen, M., Fontaine, A., Kloth, C., Kershenbaum, A., & Waller, S. (2017). Visual classification of feral cat Felis silvestris catus vocalizations. Current Zoology, 63(3), 331-339. https://doi.org/10.1093/cz/zox013

  • Oztan, O., Garner, J. P., Partap, S., Sherr, E. H., Hardan, A. Y., Farmer, C., Thurm, A., Swedo, S. E., & Parker, K. J. (2018). Cerebrospinal fluid vasopressin and symptom severity in children with autism. Annals of Neurology, 84(4), 611-615. https://doi.org/10.1002/ana.25314

  • Pan, Z., Li, S., Liu, Q., Wang, Z., Zhou, Z., Di, R., Miao, B., Hu, W., Wang, X., Hu, X., Xu, Z., Wei, D., He, X., Yuan, L., Guo, X., Liang, B., Wang, R., Li, X., Cao, X., … Li, Y. (2018). Whole-genome sequences of 89 Chinese sheep suggest role of RXFP2 in the development of unique horn phenotype as response to semi-feralization. GigaScience, 7(4), Article giy019. https://doi.org/10.1093/gigascience/giy019

  • Park, H. R., Lee, J. M., Moon, H. E., Lee, D. S., Kim, B. N., Kim, J., Kim, D. G., & Paek, S. H. (2016). A short review on the current understanding of autism spectrum disorders. Experimental Neurobiology, 25(1), 1-13. https://doi.org/10.5607/en.2016.25.1.1

  • Parker, K. J., Garner, J. P., Oztan, O., Tarara, E. R., Li, J., Sclafani, V., Del Rosso, L. A., Chun, K., Berquist, S. W., Chez, M. G., Partap, S., Hardan, A. Y., Sherr, E. H., & Capitanio, J. P. (2018). Arginine vasopressin in cerebrospinal fluid is a marker of sociality in nonhuman primates. Science Translational Medicine, 10(439), Article eaam9100. https://doi.org/10.1126/scitranslmed.aam9100

  • Pattabiraman, K., Muchnik, S. K., & Sestan, N. (2020). The evolution of the human brain and disease susceptibility. Current Opinion in Genetics & Development, 65, 91-97. https://doi.org/10.1016/j.gde.2020.05.004

  • Paul, R., Bianchi, N., Augustyn, A., Klin, A., & Volkmar, F. (2008). Production of syllable stress in speakers with autism spectrum disorders. Research in Autism Spectrum Disorders, 2(1), 110-124. https://doi.org/10.1016/j.rasd.2007.04.001

  • Perovic, A., & Janke, V. (2013). Issues in the acquisition of binding and control in high-functioning children with autism. UCL Working Papers in Linguistics, 25, 131-143.

  • Pfenning, A. R., Hara, E., Whitney, O., Rivas, M. V., Wang, R., Roulhac, P. L., Howard, J. T., Wirthlin, M., Lovell, P. V., Ganapathy, G., Mouncastle, J., Moseley, M. A., Thompson, J. W., Soderblom, E. J., Iriki, A., Kato, M., Gilbert, M. T., Zhang, G., Bakken, T., … Jarvis, E. D. (2014). Convergent transcriptional specializations in the brains of humans and song-learning birds. Science, 346(6215), Article 1256846. https://doi.org/10.1126/science.1256846

  • Feral Hogs. (2019, August 28). Physical characteristics of feral hogs. https://feralhogs.extension.org/physical-characteristics-of-feral-hogs/

  • Pickering, C., Gustafsson, L., Cebere, A., Nylander, I., & Liljequist, S. (2006). Repeated maternal separation of male Wistar rats alters glutamate receptor expression in the hippocampus but not the prefrontal cortex. Brain Research, 1099(1), 101-108. https://doi.org/10.1016/j.brainres.2006.04.136

  • Pisor, A. C., & Surbeck, M. (2019). The evolution of intergroup tolerance in nonhuman primates and humans. Evolutionary Anthropology, 28(4), 210-223. https://doi.org/10.1002/evan.21793

  • Plavcan, J. M. (2012). Sexual size dimorphism, canine dimorphism, and male-male competition in primates: Where do humans fit in? Human Nature, 23(1), 45-67. https://doi.org/10.1007/s12110-012-9130-3

  • Prasad, A., Merico, D., Thiruvahindrapuram, B., Wei, J., Lionel, A. C., Sato, D., Rickaby, J., Lu, C., Szatmari, P., Roberts, W., Fernandez, B. A., Marshall, C. R., Hatchwell, E., Eis, P. S., & Scherer, S. W. (2012). A discovery resource of rare copy number variations in individuals with autism spectrum disorder. G3, 2(12), 1665-1685. https://doi.org/10.1534/g3.112.004689

  • Progovac, L., & Benítez-Burraco, A. (2019). From physical aggression to verbal behavior: Language evolution and self-domestication feedback loop. Frontiers in Psychology, 10, Article 2807. https://doi.org/10.3389/fpsyg.2019.02807

  • Puglia, M. H., Connelly, J. J., & Morris, J. P. (2018). Epigenetic regulation of the oxytocin receptor is associated with neural response during selective social attention. Translational Psychiatry, 8(1), Article 116. https://doi.org/10.1038/s41398-018-0159-x

  • Raaska, H., Lapinleimu, H., Sinkkonen, J., Salmivalli, C., Matomäki, J., Mäkipää, S., & Elovainio, M. (2012). Experiences of school bullying among internationally adopted children: Results from the Finnish Adoption (FINADO) study. Child Psychiatry and Human Development, 43(4), 592-611. https://doi.org/10.1007/s10578-012-0286-1

  • Range, F., & Virányi, Z. (2013). Social learning from humans or conspecifics: Differences and similarities between wolves and dogs. Frontiers in Psychology, 4, Article 868. https://doi.org/10.3389/fpsyg.2013.00868

  • Reiersen, A. M., & Todorov, A. A. (2011). Association between DRD4 genotype and Autistic Symptoms in DSM-IV ADHD. Journal of the Canadian Academy of Child and Adolescent Psychiatry, 20(1), 15-21.

  • Reif, A., Rösler, M., Freitag, C. M., Schneider, M., Eujen, A., Kissling, C., Wenzler, D., Jacob, C. P., Retz-Junginger, P., Thome, J., Lesch, K. P., & Retz, W. (2007). Nature and nurture predispose to violent behavior: serotonergic genes and adverse childhood environment. Neuropsychopharmacology, 32(11), 2375-2383. https://doi.org/10.1038/sj.npp.1301359

  • Reinhardt, V. P., Iosif, A. M., Libero, L., Heath, B., Rogers, S. J., Ferrer, E., Nordahl, C., Ghetti, S., Amaral, D., & Solomon, M. (2020). Understanding hippocampal development in young children with autism spectrum disorder. Journal of the American Academy of Child and Adolescent Psychiatry, 59(9), 1069-1079. https://doi.org/10.1016/j.jaac.2019.08.008

  • Ricks, D. M., & Wing, L. (1976) Language, communication and use of symbols. In L. Wing (Ed.), Early childhood autism (pp. 93–134). Pergamon. https://doi.org/10.1007/BF01538152

  • Rodríguez-Fontenla, C., & Carracedo, A. (2021). UTMOST, a single and cross-tissue TWAS (Transcriptome Wide Association Study), reveals new ASD (Autism Spectrum Disorder) associated genes. Translational Psychiatry, 11(1), Article 256. https://doi.org/10.1038/s41398-021-01378-8

  • Röhrs, M., & Ebinger, P. (1999). Verwildert ist nicht gleich wild: Die Hirngewichte verwilderter Haussäugetiere [Wild is not really wild: Brain weight of wild domestic mammals]. Berliner und Munchener Tierarztliche Wochenschrift, 112(6-7), 234-238.

  • Rolls, E. T. (2015). Limbic systems for emotion and for memory, but no single limbic system. Cortex, 62, 119-157. https://doi.org/10.1016/j.cortex.2013.12.005

  • Rose, K. M., Wodzicka-Tomaszewska, M., & Cumming, R. B. (1985). Agonistic behaviour, responses to a novel object and some aspects of maintenance behaviour in feral-strain and domestic chickens. Applied Animal Behaviour Science, 13(3), 283-294. https://doi.org/10.1016/0168-1591(85)90052-8

  • Rothschild, G., Sottas, C. M., Kissel, H., Agosti, V., Manova, K., Hardy, M. P., & Besmer, P. (2003). A role for kit receptor signaling in Leydig cell steroidogenesis. Biology of Reproduction, 69(3), 925-932. https://doi.org/10.1095/biolreprod.102.014548

  • Roulin, A. (2004). The evolution, maintenance and adaptive function of genetic colour polymorphism in birds. Biological Reviews of the Cambridge Philosophical Society, 79(4), 815-848. https://doi.org/10.1017/S1464793104006487

  • Russo, N., Larson, C., & Kraus, N. (2008). Audio-vocal system regulation in children with autism spectrum disorders. Experimental Brain Research, 188(1), 111-124. https://doi.org/10.1007/s00221-008-1348-2

  • Rutter, M., Kreppner, J., & Sonuga-Barke, E. (2009). Emanuel Miller lecture: Attachment insecurity, disinhibited attachment, and attachment disorders: Where do research findings leave the concepts? Journal of Child Psychology and Psychiatry, and Allied Disciplines, 50(5), 529-543. https://doi.org/10.1111/j.1469-7610.2009.02042.x

  • Rymer, R. (1994). Genie: A scientific tragedy (2nd ed.). Harper Perennial.

  • Sabbagh, A., Pasmant, E., Imbard, A., Luscan, A., Soares, M., Blanché, H., Laurendeau, I., Ferkal, S., Vidaud, M., Pinson, S., Bellanné-Chantelot, C., Vidaud, D., Parfait, B., & Wolkenstein, P. (2013). NF1 molecular characterization and neurofibromatosis type I genotype-phenotype correlation: The French experience. Human Mutation, 34(11), 1510-1518. https://doi.org/10.1002/humu.22392

  • Sacco, R., Gabriele, S., & Persico, A. M. (2015). Head circumference and brain size in autism spectrum disorder: A systematic review and meta-analysis. Psychiatry Research, 234(2), 239-251. https://doi.org/10.1016/j.pscychresns.2015.08.016

  • Sadiq, F. A., Slator, L., Skuse, D., Law, J., Gillberg, C., & Minnis, H. (2012). Social use of language in children with reactive attachment disorder and autism spectrum disorders. European Child & Adolescent Psychiatry, 21(5), 267-276. https://doi.org/10.1007/s00787-012-0259-8

  • Saito, A., Watanabe, T. K., Shimada, Y., Fujiwara, T., Slaughter, C. A., DeMartino, G. N., Tanahashi, N., & Tanaka, K. (1997). cDNA cloning and functional analysis of p44.5 and p55, two regulatory subunits of the 26S proteasome. Gene, 203(2), 241-250. https://doi.org/10.1016/S0378-1119(97)00524-6

  • Sánchez-Villagra, M. R., Geiger, M., & Schneider, R. A. (2016). The taming of the neural crest: A developmental perspective on the origins of morphological covariation in domesticated mammals. Royal Society Open Science, 3(6), Article 160107. https://doi.org/10.1098/rsos.160107

  • Sanders, S. J., He, X., Willsey, A. J., Ercan-Sencicek, A. G., Samocha, K. E., Cicek, A. E., Murtha, M. T., Bal, V. H., Bishop, S. L., Dong, S., Goldberg, A. P., Jinlu, C., Keaney, J. F., III, Klei, L., Mandell, J. D., Moreno-De-Luca, D., Poultney, C. S., Robinson, E. B., Smith, L., … State, M. W. (2015). Insights into autism spectrum disorder genomic architecture and biology from 71 Risk Loci. Neuron, 87(6), 1215-1233. https://doi.org/10.1016/j.neuron.2015.09.016

  • Sanders, S. J., Murtha, M. T., Gupta, A. R., Murdoch, J. D., Raubeson, M. J., Willsey, A. J., Ercan-Sencicek, A. G., DiLullo, N. M., Parikshak, N. N., Stein, J. L., Walker, M. F., Ober, G. T., Teran, N. A., Song, Y., El-Fishawy, P., Murtha, R. C., Choi, M., Overton, J. D., Bjornson, R. D., … State, M. W. (2012). De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature, 485(7397), 237-241. https://doi.org/10.1038/nature10945

  • Satterstrom, F. K., Kosmicki, J. A., Wang, J., Breen, M. S., De Rubeis, S., An, J. Y., Peng, M., Collins, R., Grove, J., Klei, L., Stevens, C., Reichert, J., Mulhern, M. S., Artomov, M., Gerges, S., Sheppard, B., Xu, X., Bhaduri, A., Norman, U., … Buxbaum, J. D. (2020). Large-scale exome sequencing study implicates both developmental and functional changes in the neurobiology of autism. Cell, 180(3), 568-584.e23. https://doi.org/10.1016/j.cell.2019.12.036

  • Schaschl, H., Huber, S., Schaefer, K., Windhager, S., Wallner, B., & Fieder, M. (2015). Signatures of positive selection in the cis-regulatory sequences of the human oxytocin receptor (OXTR) and arginine vasopressin receptor 1a (AVPR1A) genes. BMC Evolutionary Biology, 15, Article 85. https://doi.org/10.1186/s12862-015-0372-7

  • Schoen, E., Paul, R., & Chawarska, K. (2011). Phonology and vocal behavior in toddlers with autism spectrum disorders. Autism Research, 4(3), 177-188. https://doi.org/10.1002/aur.183

  • Schultz, R. T., Grelotti, D. J., Klin, A., Kleinman, J., Van der Gaag, C., Marois, R., & Skudlarski, P. (2003). The role of the fusiform face area in social cognition: Implications for the pathobiology of autism. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 358(1430), 415-427. https://doi.org/10.1098/rstb.2002.1208

  • Schumann, C. M., Hamstra, J., Goodlin-Jones, B. L., Lotspeich, L. J., Kwon, H., Buonocore, M. H., Lammers, C. R., Reiss, A. L., & Amaral, D. G. (2004). The amygdala is enlarged in children but not adolescents with autism; the hippocampus is enlarged at all ages. The Journal of Neuroscience, 24(28), 6392-6401. https://doi.org/10.1523/JNEUROSCI.1297-04.2004

  • Schuurs-Hoeijmakers, J. H., Landsverk, M. L., Foulds, N., Kukolich, M. K., Gavrilova, R. H., Greville-Heygate, S., Hanson-Kahn, A., Bernstein, J. A., Glass, J., Chitayat, D., Burrow, T. A., Husami, A., Collins, K., Wusik, K., van der Aa, N., Kooy, F., Brown, K. T., Gadzicki, D., Kini, U., Alvarez, S., … DDD study. (2016). Clinical delineation of the PACS1-related syndrome--Report on 19 patients. American Journal of Medical Genetics. Part A, 170(3), 670-675. https://doi.org/10.1002/ajmg.a.37476

  • Schweizer, A. V., Lebrun, R., Wilson, L., Costeur, L., Schmelzle, T., & Sánchez-Villagra, M. R. (2017). Size variation under domestication: Conservatism in the inner ear shape of wolves, dogs and dingoes. Scientific Reports, 7(1), Article 13330. https://doi.org/10.1038/s41598-017-13523-9

  • Shea, B. (1989). Heterochrony in human evolution: The case for neoteny reconsidered. American Journal of Physical Anthropology, 32(S10), 69-101. https://doi.org/10.1002/ajpa.1330320505

  • Shilton, D., Breski, M., Dor, D., & Jablonka, E. (2020). Human social evolution: Self-Domestication or self-control? Frontiers in Psychology, 11, Article 134. https://doi.org/10.3389/fpsyg.2020.00134

  • Shipman, P. (2021). What the dingo says about dog domestication. Anatomical Record, 304(1), 19-30. https://doi.org/10.1002/ar.24517

  • Sikora, M., Seguin-Orlando, A., Sousa, V. C., Albrechtsen, A., Korneliussen, T., Ko, A., Rasmussen, S., Dupanloup, I., Nigst, P. R., Bosch, M. D., Renaud, G., Allentoft, M. E., Margaryan, A., Vasilyev, S. V., Veselovskaya, E. V., Borutskaya, S. B., Deviese, T., Comeskey, D., Higham, T., … Willerslev, E. (2017). Ancient genomes show social and reproductive behavior of early Upper Paleolithic foragers. Science, 358(6363), 659-662. https://doi.org/10.1126/science.aao1807

  • Smith, B. P., Cairns, K. M., Adams, J. W., Newsome, T. M., Fillios, M., Déaux, E. C., Parr, W., Letnic, M., Van Eeden, L. M., Appleby, R. G., Bradshaw, C., Savolainen, P., Ritchie, E. G., Nimmo, D. G., Archer-Lean, C., Greenville, A. C., Dickman, C. R., Watson, L., Moseby, K. E., Doherty, T. S., … Crowther, M. S. (2019). Taxonomic status of the Australian dingo: The case for Canis dingo Meyer, 1793. Zootaxa, 4564(1), Article zootaxa.4564.1.6. https://doi.org/10.11646/zootaxa.4564.1.6

  • Smith, B. P., & Litchfield, C. A. (2010). Dingoes (Canis dingo) can use human social cues to locate hidden food. Animal Cognition, 13(2), 367-376. https://doi.org/10.1007/s10071-009-0287-z

  • Smith, B. P., Lucas, T. A., Norris, R. M., & Henneberg, M. (2018). Brain size/body weight in the dingo (Canis dingo): Comparisons with domestic and wild canids. Australian Journal of Zoology, 65(5), 292-301. https://doi.org/10.1071/ZO17040

  • Smyke, A. T., Dumitrescu, A., & Zeanah, C. H. (2002). Attachment disturbances in young children. I: The continuum of caretaking casualty. Journal of the American Academy of Child and Adolescent Psychiatry, 41(8), 972-982. https://doi.org/10.1097/00004583-200208000-00016

  • Spangler, G., Johann, M., Ronai, Z., & Zimmermann, P. (2009). Genetic and environmental influence on attachment disorganization. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 50(8), 952-961. https://doi.org/10.1111/j.1469-7610.2008.02054.x

  • Spikins, P., French, J. C., John-Wood, S., & Dytham, C. (2021). Theoretical and methodological approaches to ecological changes, social behaviour and human intergroup tolerance 300,000 to 30,000 BP. Journal of Archaeological Method and Theory, 28, 53-75. https://doi.org/10.1007/s10816-020-09503-5

  • Spratt, E. G., Nicholas, J. S., Brady, K. T., Carpenter, L. A., Hatcher, C. R., Meekins, K. A., Furlanetto, R. W., & Charles, J. M. (2012). Enhanced cortisol response to stress in children in autism. Journal of Autism and Developmental Disorders, 42(1), 75-81. https://doi.org/10.1007/s10803-011-1214-0

  • Steeves, P. H. (2011). Humans and animals at the divide: The case of feral children. Between the Species, 13(3), Article 7. https://doi.org/10.15368/bts.2003v13n3.7

  • Stein, C. M., Schick, J. H., Gerry Taylor, H., Shriberg, L. D., Millard, C., Kundtz-Kluge, A., Russo, K., Minich, N., Hansen, A., Freebairn, L. A., Elston, R. C., Lewis, B. A., & Iyengar, S. K. (2004). Pleiotropic effects of a chromosome 3 locus on speech-sound disorder and reading. American Journal of Human Genetics, 74(2), 283-297. https://doi.org/10.1086/381562

  • Stendler, C. (1952). Critical periods in socialization and overdependency. Child Development, 23(1), 3-12. https://doi.org/10.2307/1125888

  • Stinehart, M. A., Scott, D. A., & Barfield, H. G. (2012). Reactive Attachment Disorder in adopted and foster care children: Implications for mental health professionals. The Family Journal, 20(4), 355-360. https://doi.org/10.1177/1066480712451229

  • St Pourcain, B., Cents, R. A., Whitehouse, A. J., Haworth, C. M., Davis, O. S., O’Reilly, P. F., Roulstone, S., Wren, Y., Ang, Q. W., Velders, F. P., Evans, D. M., Kemp, J. P., Warrington, N. M., Miller, L., Timpson, N. J., Ring, S. M., Verhulst, F. C., Hofman, A., Rivadeneira, F., … Davey Smith, G. (2014). Common variation near ROBO2 is associated with expressive vocabulary in infancy. Nature Communications, 5, Article 4831. https://doi.org/10.1038/ncomms5831

  • Stringer, C. (2016). The origin and evolution of Homo sapiens. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 371(1698), Article 20150237. https://doi.org/10.1098/rstb.2015.0237

  • Suda, S., Iwata, K., Shimmura, C., Kameno, Y., Anitha, A., Thanseem, I., Nakamura, K., Matsuzaki, H., Tsuchiya, K. J., Sugihara, G., Iwata, Y., Suzuki, K., Koizumi, K., Higashida, H., Takei, N., & Mori, N. (2011). Decreased expression of axon-guidance receptors in the anterior cingulate cortex in autism. Molecular Autism, 2(1), Article 14. https://doi.org/10.1186/2040-2392-2-14

  • Surian, L., Baron-Cohen, S., & Van der Lely, H. (1996). Are children with autism deaf to gricean maxims? Cognitive Neuropsychiatry, 1(1), 55-72. https://doi.org/10.1080/135468096396703

  • Suzuki, S., Fujisawa, T. X., Sakakibara, N., Fujioka, T., Takiguchi, S., & Tomoda, A. (2020). Development of social attention and oxytocin levels in maltreated children. Scientific Reports, 10(1), Article 7407. https://doi.org/10.1038/s41598-020-64297-6

  • Swain, J. E., Leckman, J. F., & Volkmar, F. R. (2005). The wolf boy: Reactive attachment disorder in an adolescent boy. Psychiatry, 2(11), 55-61.

  • Swaney, W. T., Cabrera-Álvarez, M. J., & Reader, S. M. (2015). Behavioural responses of feral and domestic guppies (Poecilia reticulata) to predators and their cues. Behavioural Processes, 118, 42-46. https://doi.org/10.1016/j.beproc.2015.05.010

  • Tager-Flusberg, H. (1992). Autistic children’s talk about psychological states: Deficits in the early acquisition of a theory of mind. Child Development, 63(1), 161-172. https://doi.org/10.2307/1130910

  • Tager-Flusberg, H. (1993). What language reveals about the understanding of minds in children with autism. In S. Baron-Cohen, H. Tager-Flusberg, & D. J. Cohen (Eds.), Understanding other minds. Perspectives from autism (pp. 138–57). Oxford University Press.

  • Tager-Flusberg, H. (2000). Language and the understanding minds: Connections in autism. In S. Baron-Cohen, H. Tager-Flusberg, & D. J. Cohen (Eds.), Understanding other minds. Perspectives from developmental cognitive science (2nd ed., pp. 124–149). Oxford University Press.

  • Tager-Flusberg, H. (2006). Defining language phenotypes in autism. Clinical Neuroscience Research, 6(3–4), 219-224. https://doi.org/10.1016/j.cnr.2006.06.007

  • Tager-Flusberg, H., Paul, R., & Lord, C. (2005). Language and communication in autism. In F. R. Volkmar, R. Paul, A. Klin, & D. Cohen (Eds.), Handbook of autism and pervasive developmental disorders: Diagnosis, development, neurobiology, and behavior (pp. 335–364). John Wiley & Sons.

  • Takiguchi, S., Fujisawa, T. X., Mizushima, S., Saito, D. N., Okamoto, Y., Shimada, K., Koizumi, M., Kumazaki, H., Jung, M., Kosaka, H., Hiratani, M., Ohshima, Y., Teicher, M. H., & Tomoda, A. (2015). Ventral striatum dysfunction in children and adolescents with reactive attachment disorder: Functional MRI study. BJPsych Open, 1(2), 121-128. https://doi.org/10.1192/bjpo.bp.115.001586

  • Tallet, C., Linhart, P., Policht, R., Hammerschmidt, K., Šimeček, P., Kratinova, P., & Špinka, M. (2013). Encoding of situations in the vocal repertoire of piglets (Sus scrofa): A comparison of discrete and graded classifications. PLoS One, 8(8), Article e71841. https://doi.org/10.1371/journal.pone.0071841

  • Taylor, J. L., & Corbett, B. A. (2014). A review of rhythm and responsiveness of cortisol in individuals with autism spectrum disorders. Psychoneuroendocrinology, 49, 207-228. https://doi.org/10.1016/j.psyneuen.2014.07.015

  • Teicher, M. H. (2002). Scars that won't heal: The neurobiology of child abuse.  Scientific American286(3), 68-75. https://doi.org/10.1038/scientificamerican0302-68

  • Teicher, M. H., & Samson, J. A. (2016). Annual research review: Enduring neurobiological effects of childhood abuse and neglect. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 57(3), 241-266. https://doi.org/10.1111/jcpp.12507

  • Theofanopoulou, C., Gastaldon, S., O’Rourke, T., Samuels, B. D., Martins, P. T., Delogu, F., Alamri, S., & Boeckx, C. (2017). Self-domestication in Homo sapiens: Insights from comparative genomics. PLoS One, 12(10), Article e0185306. https://doi.org/10.1371/journal.pone.0185306

  • Thomas, J., & Kirby, S. (2018). Self domestication and the evolution of language. Biology & Philosophy, 33, Article 9. https://doi.org/10.1007/s10539-018-9612-8

  • Tolchin, D., Yeager, J. P., Prasad, P., Dorrani, N., Russi, A. S., Martinez-Agosto, J. A., Haseeb, A., Angelozzi, M., Santen, G., Ruivenkamp, C., Mercimek-Andrews, S., Depienne, C., Kuechler, A., Mikat, B., Ludecke, H. J., Bilan, F., Le Guyader, G., Gilbert-Dussardier, B., Keren, B., … Le Caignec, C. (2020). De Novo SOX6 variants cause a neurodevelopmental syndrome associated with ADHD, craniosynostosis, and osteochondromas. American Journal of Human Genetics, 106(6), 830-845. https://doi.org/10.1016/j.ajhg.2020.04.015

  • Tomarken, A. J., Han, G. T., & Corbett, B. A. (2015). Temporal patterns, heterogeneity, and stability of diurnal cortisol rhythms in children with autism spectrum disorder. Psychoneuroendocrinology, 62, 217-226. https://doi.org/10.1016/j.psyneuen.2015.08.016

  • Tomasello, M., Hare, B., Lehmann, H., & Call, J. (2007). Reliance on head versus eyes in the gaze following of great apes and human infants: The cooperative eye hypothesis. Journal of Human Evolution, 52(3), 314-320. https://doi.org/10.1016/j.jhevol.2006.10.001

  • Toro, R., Konyukh, M., Delorme, R., Leblond, C., Chaste, P., Fauchereau, F., Coleman, M., Leboyer, M., Gillberg, C., & Bourgeron, T. (2010). Key role for gene dosage and synaptic homeostasis in autism spectrum disorders. Trends in Genetics, 26(8), 363-372. https://doi.org/10.1016/j.tig.2010.05.007

  • Trut, L., Oskina, I., & Kharlamova, A. (2009). Animal evolution during domestication: The domesticated fox as a model. BioEssays, 31(3), 349-360. https://doi.org/10.1002/bies.200800070

  • Tyack P. L. (2020). A taxonomy for vocal learning. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 375(1789), Article 20180406. https://doi.org/10.1098/rstb.2018.0406

  • Udell, M. A. (2015). When dogs look back: Inhibition of independent problem-solving behaviour in domestic dogs (Canis lupus familiaris) compared with wolves (Canis lupus). Biology Letters, 11(9), Article 20150489. https://doi.org/10.1098/rsbl.2015.0489

  • Udell, M., Dorey, N., & Wynne, C. (2008). Wolves outperform dogs in following human social cues. Animal Behaviour, 76(6), 1767-1773. https://doi.org/10.1016/j.anbehav.2008.07.028

  • Veenstra-VanderWeele, J., Muller, C. L., Iwamoto, H., Sauer, J. E., Owens, W. A., Shah, C. R., Cohen, J., Mannangatti, P., Jessen, T., Thompson, B. J., Ye, R., Kerr, T. M., Carneiro, A. M., Crawley, J. N., Sanders-Bush, E., McMahon, D. G., Ramamoorthy, S., Daws, L. C., Sutcliffe, J. S., & Blakely, R. D. (2012). Autism gene variant causes hyperserotonemia, serotonin receptor hypersensitivity, social impairment and repetitive behavior. Proceedings of the National Academy of Sciences of the United States of America, 109(14), 5469-5474. https://doi.org/10.1073/pnas.1112345109

  • Vernes, S. C. (2017). What bats have to say about speech and language. Psychonomic Bulletin & Review, 24(1), 111-117. https://doi.org/10.3758/s13423-016-1060-3

  • Virányi, Z., Gácsi, M., Kubinyi, E., Topál, J., Belényi, B., Ujfalussy, D., & Miklósi, A. (2008). Comprehension of human pointing gestures in young human-reared wolves (Canis lupus) and dogs (Canis familiaris). Animal Cognition, 11(3), 373-387. https://doi.org/10.1007/s10071-007-0127-y

  • Volkmar, F., Chawarska, K., & Klin, A. (2005). Autism in infancy and early childhood. Annual Review of Psychology, 56, 315-336. https://doi.org/10.1146/annurev.psych.56.091103.070159

  • von Aster, M., Zachmann, M., Brandeis, D., Wohlrab, G., Richner, M., & Steinhausen, H. C. (1997). Psychiatric, neuropediatric, and neuropsychological symptoms in a case of hypomelanosis of Ito. European Child & Adolescent Psychiatry, 6(4), 227-233. https://doi.org/10.1007/BF00539930

  • vonHoldt, B. M., Shuldiner, E., Koch, I. J., Kartzinel, R. Y., Hogan, A., Brubaker, L., Wanser, S., Stahler, D., Wynne, C., Ostrander, E. A., Sinsheimer, J. S., & Udell, M. (2017). Structural variants in genes associated with human Williams-Beuren syndrome underlie stereotypical hypersociability in domestic dogs. Science Advances, 3(7), Article e1700398. https://doi.org/10.1126/sciadv.1700398

  • Vyshedskiy, A., Mahapatra, S., & Dunn, R. (2017). Linguistically deprived children: Meta-analysis of published research underlines the importance of early syntactic language use for normal brain development. Research Ideas and Outcomes, 3, Article e20696. https://doi.org/10.3897/rio.3.e20696

  • Wallace, M., Cleary, J., Buder, E., Oller, D., Sheinkopf, S., & Mundy, P. (2008, May). An acoustic inspection of vocalizations in young children with ASD. Poster presented at the International Meeting for Autism Research, London, United Kingdom.

  • Wang, T., Guo, H., Xiong, B., Stessman, H. A., Wu, H., Coe, B. P., Turner, T. N., Liu, Y., Zhao, W., Hoekzema, K., Vives, L., Xia, L., Tang, M., Ou, J., Chen, B., Shen, Y., Xun, G., Long, M., Lin, J., … Eichler, E. E. (2016). De novo genic mutations among a Chinese autism spectrum disorder cohort. Nature Communications, 7, Article 13316. https://doi.org/10.1038/ncomms13316

  • Wang, A. T., Lee, S. S., Sigman, M., & Dapretto, M. (2006). Neural basis of irony comprehension in children with autism: The role of prosody and context. Brain, 129(4), 932-943. https://doi.org/10.1093/brain/awl032

  • Wang, A. T., Lee, S. S., Sigman, M., & Dapretto, M. (2007). Reading affect in the face and voice: Neural correlates of interpreting communicative intent in children and adolescents with autism spectrum disorders. Archives of General Psychiatry, 64(6), 698-708. https://doi.org/10.1001/archpsyc.64.6.698

  • Wang, Z., Li, J., Li, K., & Xu, J. (2018). SOX6 is downregulated in osteosarcoma and suppresses the migration, invasion and epithelial-mesenchymal transition via TWIST1 regulation. Molecular Medicine Reports, 17(5), 6803-6811. https://doi.org/10.3892/mmr.2018.8681

  • Watkins, M., Szarko, J., & Brown, A. (2013). Examiner familiarity effects for children with autism spectrum disorders. Journal of Applied School Psychology, 29(1), 37-51. https://doi.org/10.1080/15377903.2013.751475

  • Wazana, A., Moss, E., Jolicoeur-Martineau, A., Graffi, J., Tsabari, G., Lecompte, V., Pascuzzo, K., Babineau, V., Gordon-Green, C., Mileva, V., Atkinson, L., Minde, K., Bouvette-Turcot, A. A., Sassi, R., St-André, M., Carrey, N., Matthews, S., Sokolowski, M., Lydon, J., … Meaney, M. J. (2015). The interplay of birth weight, dopamine receptor D4 gene (DRD4), and early maternal care in the prediction of disorganized attachment at 36 months of age. Development and Psychopathology, 27(4pt1), 1145-1161. https://doi.org/10.1017/S0954579415000735

  • Wetherby, A. M., Watt, N., Morgan, L., & Shumway, S. (2007). Social communication profiles of children with autism spectrum disorders late in the second year of life. Journal of Autism and Developmental Disorders, 37(5), 960-975. https://doi.org/10.1007/s10803-006-0237-4

  • Wilkins, A. S., Wrangham, R. W., & Fitch, W. T. (2014). The “domestication syndrome” in mammals: A unified explanation based on neural crest cell behavior and genetics. Genetics, 197(3), 795-808. https://doi.org/10.1534/genetics.114.165423

  • Winslow, J. T., & Insel, T. R. (2002). The social deficits of the oxytocin knockout mouse. Neuropeptides, 36(2-3), 221-229. https://doi.org/10.1054/npep.2002.0909

  • Wolk, L., & Edwards, M. L. (1993). The emerging phonological system of an autistic child. Journal of Communication Disorders, 26(3), 161-177. https://doi.org/10.1016/0021-9924(93)90006-V

  • Wolk, L., Edwards, M., & Brennan, C. (2016). Phonological difficulties in children with autism: An overview. Speech, Language and Hearing, 19(2), 121-129. https://doi.org/10.1080/2050571X.2015.1133488

  • Yeon, S. C., Kim, Y. K., Park, S. J., Lee, S. S., Lee, S. Y., Suh, E. H., Houpt, K. A., Chang, H. H., Lee, H. C., Yang, B. G., & Lee, H. J. (2011). Differences between vocalization evoked by social stimuli in feral cats and house cats. Behavioural Processes, 87(2), 183-189. https://doi.org/10.1016/j.beproc.2011.03.003

  • Yoo, H. J., Lee, S. K., Park, M., Cho, I. H., Hyun, S. H., Lee, J. C., Yang, S. Y., & Kim, S. A. (2009). Family- and population-based association studies of monoamine oxidase A and autism spectrum disorders in Korean. Neuroscience Research, 63(3), 172-176. https://doi.org/10.1016/j.neures.2008.11.007

  • Zangl, R., & Mills, D. L. (2007). Increased brain activity to infant-directed speech in 6- and 13-month-old infants. Infancy, 11(1), 31-62. https://doi.org/10.1207/s15327078in1101_2

  • Zhang, S. J., Wang, G. D., Ma, P., Zhang, L. L., Yin, T. T., Liu, Y. H., Otecko, N. O., Wang, M., Ma, Y. P., Wang, L., Mao, B., Savolainen, P., & Zhang, Y. P. (2020). Genomic regions under selection in the feralization of the dingoes. Nature Communications, 11(1), Article 671. https://doi.org/10.1038/s41467-020-14515-6

  • Zhang, X., Yang, H., Corydon, M. J., Zhang, X., Pedersen, S., Korenberg, J. R., Chen, X. N., Laporte, J., Gregersen, N., Niebuhr, E., Liu, G., & Bolund, L. (1999). Localization of a human nucleoporin 155 gene (NUP155) to the 5p13 region and cloning of its cDNA. Genomics, 57(1), 144-151. https://doi.org/10.1006/geno.1999.5741

  • Zhou, W., He, Y., Rehman, A. U., Kong, Y., Hong, S., Ding, G., Yalamanchili, H. K., Wan, Y. W., Paul, B., Wang, C., Gong, Y., Zhou, W., Liu, H., Dean, J., Scalais, E., O’Driscoll, M., Morton, J., DDD study, Hou, X., Wu, Q., … Sun, Z. (2019). Loss of function of NCOR1 and NCOR2 impairs memory through a novel GABAergic hypothalamus-CA3 projection. Nature Neuroscience, 22(2), 205-217. https://doi.org/10.1038/s41593-018-0311-1

  • Zollikofer, C. P., & Ponce de León, M. S. (2010). The evolution of hominin ontogenies. Seminars in Cell & Developmental Biology, 21(4), 441-452. https://doi.org/10.1016/j.semcdb.2009.10.012