Abstract
Neurodevelopmental disorders, which emerge early in development, include a range
of neurological phenotypes and exhibit marked differences in prevalence between
sexes. A male predominance is particularly pronounced in autism spectrum
disorder (ASD). Although the precise cause of ASD is still unknown, certain
genetic variations and environmental influences have been implicated as risk
factors. Preclinical ASD models have been instrumental in shedding light on the
mechanisms behind the sexual dimorphism observed in this disorder. In this
review, we explore the potential processes contributing to sex bias by examining
both intrinsic differences in neuronal mechanisms and the influence of external
factors. We organize these mechanisms into six categories: 1) sexually dimorphic
phenotypes in mice with mutations in ASD-associated genes related to synaptic
dysfunction; 2) sex-specific microglial activity, which may disrupt neural
circuit development by excessively pruning synapses during critical periods; 3)
sex steroid hormones, such as testosterone and allopregnanolone, that
differentially influence brain structure and function; 4) escape from X
chromosome inactivation of the O-linked-N-acetylglucosamine transferase gene in
the placenta; 5) sexually dimorphic activation of the integrated stress response
pathway following maternal immune activation; and 6) immunological responses
that are differentially regulated by sex. Understanding these mechanisms is
essential for deciphering the underlying causes of ASD and may offer insights
into other disorders with notable sex disparities.
-
Keywords: Autism spectrum disorder; Genetic variation; Pregnancy; Risk factors; Sex characteristics
Introduction
Background
Neurodevelopmental disorders (NDDs) are a heterogenous group of conditions that
manifest during the developmental period and are characterized by impairments in
various aspects of neurological functioning [
1]. As defined by the Diagnostic and Statistical Manual of Mental
Disorders, Fifth Edition, NDDs include autism spectrum disorder (ASD),
intellectual disability/developmental delay, attention-deficit/hyperactivity
disorder, motor and tic disorders, and specific language disorders. These
conditions can lead to a range of developmental deficits, from specific
challenges in learning or executive function management to more extensive
impairments in social skills or intellectual abilities [
2]. This review is primarily focused on ASD, a NDD marked by
deficits in social interaction and engagement in repetitive and stereotyped
behaviors [
2].
The prevalence, age of onset, pathophysiology, and symptomatology of many NDDs
vary substantially by sex. A pronounced male bias is evident in ASD, with a
male-to-female prevalence of approximately four to one [
3,
4]. The potential
underdiagnosis of females with ASD has raised concerns, suggesting the need for
distinct diagnostic criteria in their assessment [
5,
6]. Even apart from
variations in diagnostic practices, a sex bias persists in the prevalence of
ASD, with a male-to-female ratio ranging from at least 2:1 to 3:1. This
highlights the critical need to explore the biological basis of sexual
dimorphism, which may be key to understanding the processes underlying ASD
pathogenesis [
7].
Here, we explore the potential mechanisms contributing to the sex-biased
prevalence disparity in ASD using various preclinical models (
Table 1). While preclinical rodent models
of ASD cannot fully replicate the spectrum of human ASD phenotypes [
8], potentially due to differences in brain
structures and developmental trajectories [
9], they remain invaluable for gaining mechanistic insights into the
pathogenesis of ASD and for exploring potential therapeutic strategies [
10,
11].
Table 1.Potential contributing mechanisms underlying the sex-biased
prevalence of ASD as demonstrated in preclinical rodent models
Potential contributing
mechanisms |
Preclinical models showing sexually
dimorphic ASD-like phenotypes |
Suggested mechanism(s) |
References |
Synaptic dysfunction |
Shank3 KO |
Reduced levels of mGluR5 in male
mice |
[27] |
Chd8+/N2373K
|
Sexually dimorphic changes in neuronal
activity, synaptic transmission, and transcriptomic
profiles |
[40] |
Fmr1 KO |
Sexually dimorphic upregulation of ASD
risk genes (male↑: Ctnnb1a and Grin1a, female↑:
Homer1a,
Ptgs2a,
Drd1a, Pik3cab, and Csnk1g1b) |
[44] |
Microglial
abnormalities |
Cntnap2 KO |
Activated morphology and phagocytosis
of synaptic structures in male microglia |
[58] |
DEP/MS |
Hyper-ramified phenotype in male
microglia |
[63] |
Hormones |
VPA-induced ASD mouse model |
Lower levels of TH expression in the
AVPV of male mice |
[70] |
Placenta-specific
Akr1c14 KO |
Male mouse-specific abnormalities in
cerebellar white matter |
[75] |
Escape from X chromosome
inactivation |
Prenatal stress model |
Placental OGT expression levels are
twice as high for female fetuses as for male fetuses; this
results in sexually distinct gene expression in trophoblasts
through epigenetic modulation by histone methylation |
[79,80] |
Integrated stress response
pathway |
MIA (Poly[I:C]) |
Hyperactivation of the ISR pathway in
male MIA offspring, resulting in reduced nascent protein
synthesis in the brain |
[85] |
Immune pathways |
Prenatal GBS infection |
Heightened levels of pro-inflammatory
cytokines and chemokines such as IL-1β and CINC-1/CXCL1
in male fetuses |
[98] |
MIA (LPS) |
Male MIA offspring exhibit heightened
cortical hypoxia, reduced mitosis of radial glial cells,
disrupted E/I balance within the brain, severe placental
necrosis, elevated inflammation, and reduced placental
growth |
[99] |
MIA (two-hit model) |
The anti-inflammatory cytokines IL-10
and TGF-β1 are decreased in male offspring but increased
in female mice |
[100] |
Objectives
This review aims to elucidate the potential contributing mechanisms underlying
the sex-biased prevalence of ASD as demonstrated in preclinical rodent models.
These include synaptic dysfunction, microglial abnormalities, the influence of
sex hormones, escape from X chromosome inactivation, the integrated stress
response (ISR) pathway, and immune pathways.
Methods
Ethics statement
The present study is based on a review of the literature; consequently, neither
approval from an institutional review board nor the acquisition of informed
consent was necessary.
Study design
This study is a narrative review.
Literature search: information sources and search strategies
We searched the PubMed database for articles published from 1990 up to April
2024. Only articles published in English were included.
Results
This review encompasses a total of 104 articles. A list of articles pertaining to
each topic is available in the references.
Synaptic dysfunction
Previous studies have reported that those with ASD exhibit different brain
connectivity patterns compared to typically developing individuals [
12]. Patterns of widespread cortical
underconnectivity, local overconnectivity, or a combination of these suggest
that disrupted brain connectivity may represent a potential neural signature of
ASD [
13]. Brain connectivity is largely
determined by the characteristics of neurons and synapses, with synapses being
highly specialized, asymmetric cell-to-cell junctions that constitute the
fundamental units of brain communication [
14].
According to the Simons Foundation Autism Research Initiative (SFARI) gene
database, hundreds of genes have been identified as being associated with ASD
[
15–
18]. Among these, genes such as those of the SH3 and
multiple ankyrin repeat domains (
SHANK) family, fragile X
mental retardation 1 (
FMR1), and chromodomain helicase
DNA-binding protein 8 (
CHD8) are linked to common cellular
pathways that converge at synapses [
19–
21]. This
convergence suggests that synaptic dysfunction may contribute to the development
of ASD, potentially leading to functional and cognitive impairments [
14].
SHANK, also known as ProSAP, is a family of postsynaptic proteins found at
glutamatergic synapses and includes three major isoforms: SHANK1, SHANK2, and
SHANK3. These proteins act as master scaffolding proteins at excitatory synapses
[
22,
23]. They interact with over 30 synaptic proteins across multiple
domains and are critical for synaptic formation, glutamate receptor trafficking,
and neuronal signaling [
24]. Genetic
screenings have identified mutations, rare variants, or disruptions of the
SHANK3 gene in patients with ASD [
22]. Mice with a genetic disruption of
Shank3 display compulsive/repetitive behaviors and social
interaction deficits, which reflect clinical features of ASD [
25]. Studies using
Shank3
knockout (KO) mouse models have reported sexually dimorphic phenotypes [
26,
27]. Matas et al. found that male
Shank3 KO mice
with a mutation in the C-terminal regions (exons 21–22) [
28] exhibit more pronounced gait deficits
than their female siblings [
27]. Further
research into cerebellar glutamate levels and postsynaptic receptors showed that
metabotropic glutamate receptor 5 levels were reduced only in male
Shank3 KO mice, suggesting a potential cause for the varied
behavioral outcomes [
27].
CHD8, a chromatin remodeling factor, is essential for regulating the
transcription of a wide variety of genes [
29,
30], including
approximately 1,000 ASD risk genes identified in the SFARI gene database [
30]. Mice with homozygous deletions of
Chd8 die early in embryonic development [
31], while those with heterozygous
mutations or gene knockdown display a range of ASD-like phenotypes [
32–
36]. These include impaired social interaction, repetitive
behaviors, and cognitive impairments, resembling characteristics of individuals
with
CHD8 mutations [
20,
30,
37–
39]. Jung
et al. found that a heterozygous mutation in
Chd8, specifically
the substitution of asparagine with lysine at position 2373—the first
mutation identified as an ASD risk factor in human
CHD8—results in sexually dimorphic effects that range
from transcriptional to behavioral changes in mice [
40]. Male
Chd8+/N2373K mice
exhibited various abnormal behaviors at the pup, juvenile, and adult stages,
such as increased ultrasonic vocalizations when seeking their mother, heightened
attachment upon reunion with their mother, and increased self-grooming when
isolated. In contrast, their female counterparts did not exhibit these
behaviors. This behavioral disparity is thought to be associated with sexual
dimorphism in neuronal activity, synaptic transmission, and transcriptomic
profiles.
The
FMR1 gene encodes the fragile X mental retardation protein
(FMRP), which acts as a messenger RNA-binding translational suppressor. It also
modulates activity-dependent calcium signaling during critical developmental
periods [
41]. In mice, FMRP is most
abundantly expressed in the hippocampus and cerebral cortex, with peak levels
occurring between 2 to 4 weeks postnatally—a crucial time frame for
synaptic development and maturation [
42].
A deficiency in FMRP results in abnormal synaptic plasticity and structural
remodeling [
43]. Notably, male
Fmr1 KO mice exhibit more severe anxiety, deficiencies in
social preference, and repetitive behaviors than their female counterparts
[
44]. Differential gene expression
analysis of the hippocampus in wild-type (WT) versus
Fmr1 KO
mice revealed that in male
Fmr1 KO mice,
Ctnnb1 and
Grin1—genes considered
high-confidence risk factors for ASD—are highly upregulated. In contrast,
female
Fmr1 KO mice exhibited upregulation of genes such as
Homer1,
Ptgs2, and
Drd1,
which are strong ASD risk gene candidates, as well as
Pik3ca
and
Csnk1g1, which provide suggestive evidence of risk for ASD.
These findings suggest that the loss of FMRP leads to sexually dimorphic
phenotypes, potentially due to different patterns of gene expression regulation
resulting from the absence of FMR1.
The collective evidence from these reports suggests that synaptic dysfunction and
disrupted connectivity could be responsible for sex-specific functional and
cognitive impairments observed in ASD [
45].
Microglial abnormalities
The balance between excitation and inhibition (E/I) in neural circuits is
critical for maintaining brain homeostasis [
46]. Disruption of this E/I balance has been implicated as a
potential cause of behavioral phenotypes associated with ASD [
47]. Microglia, the phagocytic cells that
reside in the brain from the developmental period, engulf the synaptic
materials, thus pruning synapses and supporting synaptic maturation. [
48]. When microglial function is
compromised, improper synaptic pruning can disrupt the E/I balance and
potentially contribute to the pathogenesis of ASD [
49]. Notably, microglia exhibit sexually dimorphic
transcriptional and translational profiles [
50]. Furthermore, the morphology and number of microglia in the
developing rat brain differ between male and female rats [
51]. During the early postnatal period, male rats have
significantly higher numbers of microglia compared to female rats. These
sex-based differences in microglial numbers appear to be functionally related to
sex-specific behaviors [
52].
The contactin-associated protein 2 (
CNTNAP2) gene encodes the
CASPR2 protein, which is a neurexin-related synaptic cell adhesion molecule. A
study utilizing high-density single nucleotide polymorphisms identified
CNTNAP2 as a strong candidate gene implicated in the
etiology of ASD [
53]. Subsequent
loss-of-function studies in
Cntnap2 KO mice demonstrated that
the absence of
Cntnap2 leads to a decrease in dendritic spine
density [
54], disruptions in synaptic
function [
55], imbalances in E/I
signaling, and impaired neural oscillations [
56]. These
Cntnap2 KO mice also display core
ASD-like behavioral phenotypes, including impairments in sociability and
repetitive behaviors [
57]. Dawson et al.
found that male
Cntnap2 KO mice exhibited pronounced social
deficits, whereas their female counterparts did not. Further investigation into
the anterior cingulate cortex—a region critical for social behavior
regulation through its connections with other intracortical and subcortical
areas—revealed a more activated morphology and increased phagocytosis of
synaptic structures in male KO mice compared to WT mice, a distinction not
observed in female KO versus WT mice [
58].
In addition to genetic models, differences in microglial morphology and function
have been observed in preclinical models that incorporate environmental risk
factors. High levels of air pollution, particularly during development [
59,
60], and maternal stress (MS) during gestation [
61,
62] have been linked to an increased risk of ASD. Smith et al.
investigated the combined effects of these two risk factors and found that
prenatal exposure to air pollution—specifically diesel exhaust particles
(DEP)—along with MS in mice led to sociability deficits exclusively in
male offspring [
63]. These behavioral
impairments were paralleled by alterations in microglial morphology and gene
expression, with DEP/MS exposure resulting in a hyper-ramified microglial
phenotype in male but not female animals.
The collective evidence from these reports suggests that sexually dimorphic
microglial activity could play a role in the etiology of ASD. This activity may
disrupt the development of neural circuits responsible for social behavior by
excessively pruning synapses during a critical period of development [
49].
Hormones
Sex steroid hormones are known to contribute to sex differences in neural
activity and behaviors in mammals through their interactions with specific
nuclear hormone receptors [
64].
Testosterone plays a crucial role during prenatal development in shaping sex
differences, influencing brain structure, neurotransmitter and receptor levels,
neurogenesis, immune responses, neuropeptide signaling, and cellular processes
such as apoptosis, migration, and differentiation [
65]. Clinical reports have correlated high levels of
testosterone with autistic behavior [
66,
67], and this association
is supported by Erdogan et al., who showed that prenatal testosterone exposure
led to ASD-like behaviors in the offspring of Wistar rats [
68]. Both male and female rats exposed to testosterone
exhibited reduced interaction times with a stranger rat during the three-chamber
sociability and social novelty test, indicating a decrease in social interaction
and a phenotype with characteristics resembling ASD. In line with these
findings, studies on the valproic acid (VPA)-induced ASD mouse model, which is
based on a medication known to increase the risk of ASD in humans [
69], have shown that elevated plasma
testosterone levels resulting from VPA treatment led to significantly lower
levels of tyrosine hydroxylase (TH) expression in the anteroventral
periventricular nucleus of male mice. In contrast, TH levels in female mice were
unaffected [
70].
Allopregnanolone (ALLO), a 3ɑ, 5ɑ progesterone metabolite [
71], is a key GABAergic neurosteroid [
72,
73]. Reduced ALLO levels are correlated with a greater severity of
restricted and repetitive behaviors [
74].
Penn and colleagues have shown that ALLO plays a vital role as a placental
hormone in shaping the fetal brain, leading to sexually dimorphic behavioral
outcomes [
75]. Specifically, a deficiency
of placental ALLO in mice resulted in male-specific abnormalities in cerebellar
white matter and core ASD symptoms, such as diminished social preference and
increased repetitive behaviors. Notably, this study observed sex-linked
dysregulation of myelin proteins in the cerebellar vermis of preterm infants,
which aligns with human data.
These results highlight the influence of hormones in molding the early brain
environment, potentially leading to sexually dimorphic behavioral outcomes.
Escape from X chromosome inactivation
In a mouse model of early prenatal stress, male offspring exposed to MS during
gestation exhibited certain NDD phenotypes [
76,
77]. The placenta plays a
critical role during pregnancy, acting as a mediator in response to disturbances
within the intrauterine environment [
78].
MS leads to sexually dimorphic changes in the placental expression of
O-linked-N-acetylglucosamine transferase (OGT), an X-linked gene essential for
the regulation of proteins involved in chromatin remodeling [
79]. Notably, OGT escapes X chromosome
inactivation in the placenta, resulting in placental levels that are
approximately twice as high in female animals than in male animals. Crucially,
this finding also translates to humans: levels of both OGT and its biochemical
marker, O-GlcNAcylation, have been found to be considerably lower for male
fetuses and are further reduced by prenatal stress [
79]. Nugent et al. demonstrated that OGT levels establish a
sex-specific gene expression pattern in trophoblasts through regulation of a
canonical histone repressive mark, H3K27me3 [
80]. Higher placental levels of H3K27me3 for female offspring
provided a protective effect against the altered hypothalamic programming
associated with prenatal stress exposure. Consequently, lower levels of OGT may
predispose male offspring to a higher risk of ASD. Future studies should explore
the molecular mechanisms underlying this increased male susceptibility.
Integrated stress response pathway
Maternal immune activation (MIA) during pregnancy is linked to a heightened risk
of ASD in offspring [
81]. This phenomenon
has been extensively investigated using a rodent MIA model. In this model,
pregnant mice received intraperitoneal injections of polyinosinic:polycytidylic
acid (poly[I:C]), a synthetic analog of double-stranded RNA that simulates viral
infection. The offspring exhibited significant neurodevelopmental impairments,
including diminished social interaction and increased repetitive behaviors
[
82–
84]. Kalish et al. found that MIA exerts a sexually
dimorphic effect in utero, leading to different behavioral outcomes. Male
offspring exhibited MIA-induced behavioral abnormalities, whereas female
offspring did not [
85]. Notably, when
gene expression was examined at the single-cell level, changes in the fetal
cortex were observed to be sexually dimorphic. In male fetuses, these changes
were predominantly characterized by reduced gene expression related to protein
translation, followed by an overactive ISR pathway. In eukaryotic cells, the ISR
signaling pathway regulates protein synthesis in response to various stresses,
both physiological and pathological, to restore cellular homeostasis [
86]. Dysregulation of protein synthesis has
been implicated in ASD-related traits and other neurological disorders [
87–
89]. Male-specific activation of the ISR is dependent on the
maternal induction of interleukin (IL)-17A following MIA, which has been shown
to be necessary for the development of MIA-induced ASD-like behaviors in mouse
offspring [
83]. The genetic and
pharmacological inhibition of ISR pathway hyperactivation was sufficient to
protect male offspring from MIA-induced behavioral abnormalities. This study
offers valuable insights into potential preventative strategies for ASD-like
phenotypes that may result from prenatal immune activation.
Immune pathways
Preterm delivery is associated with a higher likelihood of ASD in children
compared to those born at full term [
90–
93].
Chorioamnionitis, caused by Group B
Streptococcus (GBS;
Streptococcus agalactiae), is one of the most common
maternal infections and accounts for 40% to 70% of preterm births [
94–
96]. This condition typically involves an inflammatory intrauterine
environment, even in the absence of bacterial translocation from mother to fetus
[
97]. Allard et al. demonstrated that
prenatal infection with live GBS in rats resulted in social impairments in male
but not in female offspring [
98]. A
prominent inflammatory state was noted in male animals, with higher levels of
the pro-inflammatory cytokine IL-1β and the cytokine-induced neutrophil
chemoattractant-1 (CINC-1/CXCL1), compared to female rats. These findings
suggest that sex-specific inflammatory profiles may contribute to the observed
sexually dimorphic behavioral outcomes [
98].
Consistent with this notion, in a model of MIA induced by lipopolysaccharide
(LPS), a toll-like receptor 4 agonist, Braun et al. investigated sex-specific
pro-inflammatory responses in both the placenta and fetus, as well as their
effects on behavioral outcomes. Male offspring of mothers exposed to LPS
exhibited behavioral abnormalities in social interaction and learning, as well
as increased repetitive behavior, whereas female offspring were unaffected
[
99]. Male MIA offspring showed
increased cortical hypoxia, decreased mitosis of radial glial cells, and
disrupted E/I balance in the brain. Additionally, severe placental necrosis,
heightened inflammation, and reduced placental growth were specifically observed
in male mice affected by MIA, suggesting that unique sex-specific placental
characteristics may make male offspring more susceptible to intrauterine
disturbances.
Carlezon Jr et al. demonstrated sex-specific behavioral effects and immune
responses in the brain using a combined rodent “two-hit” immune
activation model. This model involved treatment with poly (I:C) to induce MIA
and the administration of LPS to produce postnatal immune activation [
100]. Exposure to early-life immune
activation (EIA) was shown to lead to reduced social interaction and increased
repetitive behaviors in male animals, while female rodents displayed no
significant changes. Molecular studies indicated that EIA resulted in pronounced
sex-specific alterations in the expression of inflammation-related genes in the
brain. Both male and female rodents exposed to EIA exhibited elevated levels of
pro-inflammatory factors in the brain, such as tumor necrosis factor alpha,
inducible nitric oxide synthase, IL-6, and IL-1β. Conversely, the
expression of anti-inflammatory factors like IL-10 and transforming growth
factor beta 1 was reduced in male mice but elevated in female animals [
100].
The collective findings of these studies suggest that sexually dimorphic
inflammatory responses could potentially contribute to the sex-specific effects
of MS on the neurobehavioral outcomes of offspring.
Discussion
Implication and suggestion
In this review, we have comprehensively examined the potential mechanisms by
which genetic variants and environmental factors contribute to sex differences,
as demonstrated by preclinical models. Our analysis encompassed both intrinsic
differences in the brain, such as synaptic connectivity and microglial activity,
and the potential influence of extrinsic factors, including sex hormones and the
placenta. These elements may either increase male susceptibility or bolster
female resilience. Notably, beyond the intrinsic factors of the brain, the
hormonal profile, epigenetic landscape, and immune pathways associated with the
placenta have been implicated in contributing to sexually dimorphic outcomes in
mouse models exhibiting ASD-like behaviors. This indicates that a deeper
understanding of the placenta as a temporary but dynamic interface during
prenatal development could provide valuable insights into the sex biases
observed in NDDs.
Although preclinical mouse models of ASD have important limitations, such as
their inability to engage neural circuitry comparable to that observed in humans
or to recapitulate all human ASD phenotypes [
8], they remain valuable tools for gaining mechanistic insights into
ASD pathogenesis and for developing potential therapeutic approaches [
10,
11]. Further investigation is imperative to unravel the mechanistic
basis of sexual dimorphism in ASD, as well as in other NDDs. For example,
utilizing large-scale transcriptomics from postmortem brain studies [
101], generating a single-cell atlas
[
102], and conducting multi-omic
profiling of somatic mutations [
103] in
brains exhibiting ASD could help uncover sex-specific changes in genes or
pathways during early neurodevelopment. Given that ASD is a developmental
disorder, it is crucial to conduct further longitudinal investigations to
understand how risk factors evolve across the developmental trajectory. For
instance, a longitudinal cohort study of children with developmental
disabilities suggested that
de novo protein-truncating variants
were correlated with clinical characteristics [
104].
Conclusion
Overall, understanding sex-specific mechanisms is pivotal for comprehending the
fundamental causes of ASD and may illuminate the pathologies of other diseases
characterized by prominent sex biases.
Authors' contributions
-
Project administration: not applicable
Conceptualization: Lee T, Kim E
Methodology & data curation: not applicable
Funding acquisition: Kim E
Writing – original draft: Lee T
Writing – review & editing: Lee T, Kim E
Conflict of interest
-
Eunha Kim serves as a consultant for Interon Laboratories.
Funding
-
Eunha Kim received support through a Korea University grant (K2225821).
Data availability
-
Not applicable.
Acknowledgments
We would like to express our gratitude to all members of the Kim laboratory for their
insightful comments. We also give special thanks to Hyun Je for providing technical
assistance in structuring the manuscript.
Supplementary materials
-
Not applicable.
References
- 1. Morris-Rosendahl DJ, Crocq MA. Neurodevelopmental disorders—the history and future of a
diagnostic concept. Dialogues Clin Neurosci 2020;22(1):65-72.
- 2. American Psychiatric Association. Diagnostic and statistical manual of mental disorders
(DSM-5®); 5th ed. Washington: American Psychiatric Association; 2013
- 3. Fombonne E. Epidemiology of pervasive developmental disorders. Pediatr Res 2009;65(6):591-598.
- 4. Werling DM, Geschwind DH. Sex differences in autism spectrum disorders. Curr Opin Neurol 2013;26(2):146-153.
- 5. Loomes R, Hull L, Mandy WPL. What is the male-to-female ratio in autism spectrum disorder? A
systematic review and meta-analysis. J Am Acad Child Adolesc Psychiatry 2017;56(6):466-474.
- 6. Hull L, Petrides KV, Mandy W. The female autism phenotype and camouflaging: a narrative
review. Rev J Autism Dev Disord 2020;7(4):306-317.
- 7. Zwaigenbaum L, Bryson SE, Szatmari P, Brian J, Smith IM, Roberts W, et al. Sex differences in children with autism spectrum disorder
identified within a high-risk infant cohort. J Autism Dev Disord 2012;42(12):2585-2596.
- 8. Silverman JL, Thurm A, Ethridge SB, Soller MM, Petkova SP, Abel T, et al. Reconsidering animal models used to study autism spectrum
disorder: current state and optimizing future. Genes Brain Behav 2022;21(5):e12803.
- 9. Pensado-López A, Veiga-Rúa S, Carracedo Á, Allegue C, Sánchez L. Experimental models to study autism spectrum disorders: hiPSCs,
rodents and zebrafish. Genes 2020;11(11):1376
- 10. Kazdoba TM, Leach PT, Yang M, Silverman JL, Solomon M, Crawley JN. Translational mouse models of autism: advancing toward
pharmacological therapeutics. Curr Top Behav Neurosci 2016;28:1-52.
- 11. Jabarin R, Netser S, Wagner S. Beyond the three-chamber test: toward a multimodal and objective
assessment of social behavior in rodents. Mol Autism 2022;13(1):41
- 12. Mohammad-Rezazadeh I, Frohlich J, Loo SK, Jeste SS. Brain connectivity in autism spectrum disorder. Curr Opin Neurol 2016;29(2):137-147.
- 13. Maximo JO, Cadena EJ, Kana RK. The implications of brain connectivity in the neuropsychology of
autism. Neuropsychol Rev 2014;24(1):16-31.
- 14. Guang S, Pang N, Deng X, Yang L, He F, Wu L, et al. Synaptopathology involved in autism spectrum
disorder. Front Cell Neurosci 2018;12:470
- 15. Banerjee-Basu S, Packer A. SFARI Gene: an evolving database for the autism research
community. Dis Model Mech 2010;3(3-4):133-135.
- 16. Abrahams BS, Arking DE, Campbell DB, Mefford HC, Morrow EM, Weiss LA, et al. SFARI Gene 2.0: a community-driven knowledgebase for the autism
spectrum disorders (ASDs). Mol Autism 2013;4(1):36
- 17. Iossifov I, O'Roak BJ, Sanders SJ, Ronemus M, Krumm N, Levy D, et al. The contribution of de novo coding mutations to
autism spectrum disorder. Nature 2014;515(7526):216-221.
- 18. Choi L, An JY. Genetic architecture of autism spectrum disorder: lessons from
large-scale genomic studies. Neurosci Biobehav Rev 2021;128:244-257.
- 19. Wang T, Guo H, Xiong B, Stessman HAF, Wu H, Coe BP, et al. De novo genic mutations among a Chinese autism
spectrum disorder cohort. Nat Commun 2016;7:13316
- 20. Stessman HAF, Xiong B, Coe BP, Wang T, Hoekzema K, Fenckova M, et al. Targeted sequencing identifies 91 neurodevelopmental-disorder
risk genes with autism and developmental-disability biases. Nat Genet 2017;49(4):515-526.
- 21. Ellingford RA, Panasiuk MJ, de Meritens ER, Shaunak R, Naybour L, Browne L, et al. Cell-type-specific synaptic imbalance and disrupted homeostatic
plasticity in cortical circuits of ASD-associated Chd8
haploinsufficient mice. Mol Psychiatry 2021;26(7):3614-3624.
- 22. Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, Fauchereau F, et al. Mutations in the gene encoding the synaptic scaffolding protein
SHANK3 are associated with autism spectrum disorders. Nat Genet 2007;39(1):25-27.
- 23. Betancur C, Buxbaum JD. SHANK3 haploinsufficiency: a
"common" but underdiagnosed highly penetrant monogenic cause
of autism spectrum disorders. Mol Autism 2013;4(1):17
- 24. Monteiro P, Feng G. SHANK proteins: roles at the synapse and in autism spectrum
disorder. Nat Rev Neurosci 2017;18(3):147-157.
- 25. Peca J, Feliciano C, Ting JT, Wang W, Wells MF, Venkatraman TN, et al. Shank3 mutant mice display autistic-like
behaviours and striatal dysfunction. Nature 2011;472(7344):437-442.
- 26. Maloney SE, Sarafinovska S, Weichselbaum C, McCullough KB, Swift RG, Liu Y, et al. A comprehensive assay of social motivation reveals sex-specific
roles of autism-associated genes and oxytocin. Cell Rep Methods 2023;3(6):100504
- 27. Matas E, Maisterrena A, Thabault M, Balado E, Francheteau M, Balbous A, et al. Major motor and gait deficits with sexual dimorphism in a
Shank3 mutant mouse model. Mol Autism 2021;12(1):2
- 28. Moretto E, Murru L, Martano G, Sassone J, Passafaro M. Glutamatergic synapses in neurodevelopmental
disorders. Prog Neuropsychopharmacol Biol Psychiatry 2018;84(Pt B):328-342.
- 29. Bourgeron T. From the genetic architecture to synaptic plasticity in autism
spectrum disorder. Nat Rev Neurosci 2015;16(9):551-563.
- 30. Barnard RA, Pomaville MB, O'Roak BJ. Mutations and modeling of the chromatin remodeler CHD8 define an
emerging autism etiology. Front Neurosci 2015;9:477
- 31. Nishiyama M, Nakayama K, Tsunematsu R, Tsukiyama T, Kikuchi A, Nakayama KI. Early embryonic death in mice lacking the
β-catenin-binding protein Duplin. Mol Cell Biol 2004;24(19):8386-8394.
- 32. Katayama Y, Nishiyama M, Shoji H, Ohkawa Y, Kawamura A, Sato T, et al. CHD8 haploinsufficiency results in autistic-like phenotypes in
mice. Nature 2016;537(7622):675-679.
- 33. Durak O, Gao F, Kaeser-Woo YJ, Rueda R, Martorell AJ, Nott A, et al. Chd8 mediates cortical neurogenesis via transcriptional
regulation of cell cycle and Wnt signaling. Nat Neurosci 2016;19(11):1477-1488.
- 34. Gompers AL, Su-Feher L, Ellegood J, Copping NA, Asrafuzzaman Riyadh M, Stradleigh TW, et al. Germline Chd8 haploinsufficiency alters brain
development in mouse. Nat Neurosci 2017;20(8):1062-1073.
- 35. Platt RJ, Zhou Y, Slaymaker IM, Shetty AS, Weisbach NR, Kim JA, et al. Chd8 mutation leads to autistic-like behaviors
and impaired striatal circuits. Cell Rep 2017;19(2):335-350.
- 36. Suetterlin P, Hurley S, Mohan C, Riegman KLH, Pagani M, Caruso A, et al. Altered neocortical gene expression, brain overgrowth and
functional over-connectivity in Chd8 haploinsufficient
mice. Cereb Cortex 2018;28(6):2192-2206.
- 37. Bernier R, Golzio C, Xiong B, Stessman HA, Coe BP, Penn O, et al. Disruptive CHD8 mutations define a subtype of
autism early in development. Cell 2014;158(2):263-276.
- 38. O'Roak BJ, Vives L, Fu W, Egertson JD, Stanaway IB, Phelps IG, et al. Multiplex targeted sequencing identifies recurrently mutated
genes in autism spectrum disorders. Science 2012;338(6114):1619-1622.
- 39. Merner N, Forgeot d'Arc B, Bell SC, Maussion G, Peng H, Gauthier J, et al. A de novo frameshift mutation in chromodomain
helicase DNA-binding domain 8 (CHD8): a case report and literature
review. Am J Med Genet A 2016;170(5):1225-1235.
- 40. Jung H, Park H, Choi Y, Kang H, Lee E, Kweon H, et al. Sexually dimorphic behavior, neuronal activity, and gene
expression in Chd8-mutant mice. Nat Neurosci 2018;21(9):1218-1228.
- 41. Davis JK, Broadie K. Multifarious functions of the fragile X mental retardation
protein. Trends Genet 2017;33(10):703-714.
- 42. Pacey LKK, Xuan ICY, Guan S, Sussman D, Mark Henkelman R, Chen Y, et al. Delayed myelination in a mouse model of fragile X
syndrome. Hum Mol Genet 2013;22(19):3920-3930.
- 43. Qin M, Entezam A, Usdin K, Huang T, Liu ZH, Hoffman GE, et al. A mouse model of the fragile X premutation: effects on behavior,
dendrite morphology, and regional rates of cerebral protein
synthesis. Neurobiol Dis 2011;42(1):85-98.
- 44. Wang Z, Qiao D, Chen H, Zhang S, Zhang B, Zhang J, et al. Effects of Fmr1 gene mutations on sex
differences in autism-like behavior and dendritic spine development in mice
and transcriptomic studies. Neuroscience 2023;534:16-28.
- 45. Luo J, Norris RH, Gordon SL, Nithianantharajah J. Neurodevelopmental synaptopathies: insights from behaviour in
rodent models of synapse gene mutations. Prog Neuropsychopharmacol Biol Psychiatry 2018;84(Pt B):424-439.
- 46. Tatti R, Haley MS, Swanson OK, Tselha T, Maffei A. Neurophysiology and regulation of the balance between excitation
and inhibition in neocortical circuits. Biol Psychiatry 2017;81(10):821-831.
- 47. Hollestein V, Poelmans G, Forde NJ, Beckmann CF, Ecker C, Mann C, et al. Excitatory/inhibitory imbalance in autism: the role of glutamate
and GABA gene-sets in symptoms and cortical brain structure. Transl Psychiatry 2023;13(1):18
- 48. Paolicelli RC, Bolasco G, Pagani F, Maggi L, Scianni M, Panzanelli P, et al. Synaptic pruning by microglia is necessary for normal brain
development. Science 2011;333(6048):1456-1458.
- 49. Koyama R, Ikegaya Y. Microglia in the pathogenesis of autism spectrum
disorders. Neurosci Res 2015;100:1-5.
- 50. Guneykaya D, Ivanov A, Hernandez DP, Haage V, Wojtas B, Meyer N, et al. Transcriptional and translational differences of microglia from
male and female brains. Cell Rep 2018;24(10):2773-2783.E6.
- 51. Schwarz JM, Sholar PW, Bilbo SD. Sex differences in microglial colonization of the developing rat
brain. J Neurochem 2012;120(6):948-963.
- 52. McCarthy MM, Nugent BM, Lenz KM. Neuroimmunology and neuroepigenetics in the establishment of sex
differences in the brain. Nat Rev Neurosci 2017;18(8):471-484.
- 53. Alarcón M, Abrahams BS, Stone JL, Duvall JA, Perederiy JV, Bomar JM, et al. Linkage, association, and gene-expression analyses identify
CNTNAP2 as an autism-susceptibility
gene. Am J Hum Genet 2008;82(1):150-159.
- 54. Varea O, Martin-de-Saavedra MD, Kopeikina KJ, Schürmann B, Fleming HJ, Fawcett-Patel JM, et al. Synaptic abnormalities and cytoplasmic glutamate receptor
aggregates in contactin associated protein-like 2/Caspr2
knockout neurons. Proc Natl Acad Sci USA 2015;112(19):6176-6181.
- 55. Anderson GR, Galfin T, Xu W, Aoto J, Malenka RC, Südhof TC. Candidate autism gene screen identifies critical role for
cell-adhesion molecule CASPR2 in dendritic arborization and spine
development. Proc Natl Acad Sci USA 2012;109(44):18120-18125.
- 56. Lazaro MT, Taxidis J, Shuman T, Bachmutsky I, Ikrar T, Santos R, et al. Reduced prefrontal synaptic connectivity and disturbed
oscillatory population dynamics in the CNTNAP2 model of
autism. Cell Rep 2019;27(9):2567-2578.E6.
- 57. Peñagarikano O, Geschwind DH. What does CNTNAP2 reveal about autism spectrum
disorder? Trends Mol Med 2012;18(3):156-163.
- 58. Dawson MS, Gordon-Fleet K, Yan L, Tardos V, He H, Mui K, et al. Sexual dimorphism in the social behaviour of
Cntnap2-null mice correlates with disrupted synaptic
connectivity and increased microglial activity in the anterior cingulate
cortex. Commun Biol 2023;6(1):846
- 59. Rahman MM, Shu YH, Chow T, Lurmann FW, Yu X, Martinez MP, et al. Prenatal exposure to air pollution and autism spectrum disorder:
sensitive windows of exposure and sex differences. Environ Health Perspect 2022;130(1):017008-1-017008-9.
- 60. Volk HE, Lurmann F, Penfold B, Hertz-Picciotto I, McConnell R. Traffic-related air pollution, particulate matter, and
autism. JAMA Psychiatry 2013;70(1):71-77.
- 61. Roberts AL, Koenen KC, Lyall K, Ascherio A, Weisskopf MG. Women's posttraumatic stress symptoms and autism spectrum
disorder in their children. Res Autism Spectr Disord 2014;8(6):608-616.
- 62. Kinney DK, Munir KM, Crowley DJ, Miller AM. Prenatal stress and risk for autism. Neurosci Biobehav Rev 2008;32(8):1519-1532.
- 63. Smith CJ, Rendina DN, Kingsbury MA, Malacon KE, Nguyen DM, Tran JJ, et al. Microbial modulation via cross-fostering prevents the effects of
pervasive environmental stressors on microglia and social behavior, but not
the dopamine system. Mol Psychiatry 2023;28(6):2549-2562.
- 64. Gegenhuber B, Wu MV, Bronstein R, Tollkuhn J. Gene regulation by gonadal hormone receptors underlies brain sex
differences. Nature 2022;606(7912):153-159.
- 65. Ferri SL, Abel T, Brodkin ES. Sex differences in autism spectrum disorder: a
review. Curr Psychiatry Rep 2018;20(2):9
- 66. Baron-Cohen S, Auyeung B, Nørgaard-Pedersen B, Hougaard DM, Abdallah MW, Melgaard L, et al. Elevated fetal steroidogenic activity in autism. Mol Psychiatry 2015;20(3):369-376.
- 67. Majewska MD, Hill M, Urbanowicz E, Rok-Bujko P, Bieńkowski P, Namyslowska I, et al. Marked elevation of adrenal steroids, especially androgens, in
saliva of prepubertal autistic children. Eur Child Adolesc Psychiatry 2014;23(6):485-498.
- 68. Erdogan MA, Bozkurt MF, Erbas O. Effects of prenatal testosterone exposure on the development of
autism-like behaviours in offspring of Wistar rats. Int J Dev Neurosci 2022;83(2):201-215.
- 69. Christensen J, Grønborg TK, Sørensen MJ, Schendel D, Parner ET, Pedersen LH, et al. Prenatal valproate exposure and risk of autism spectrum disorders
and childhood autism. JAMA 2013;309(16):1696-1703.
- 70. Grgurevic N. Testing the extreme male hypothesis in the valproate mouse model;
sex-specific effects on plasma testosterone levels and tyrosine hydroxylase
expression in the anteroventral periventricular nucleus, but not on parental
behavior. Front Behav Neurosci 2023;17:1107226
- 71. Diviccaro S, Cioffi L, Falvo E, Giatti S, Melcangi RC. Allopregnanolone: an overview on its synthesis and
effects. J Neuroendocrinol 2021;34(2):e12996.
- 72. Lambert JJ, Belelli D, Hill-Venning C, Peters JA. Neurosteroids and GABAA receptor function. Trends Pharmacol Sci 1995;16(9):295-303.
- 73. Pinna G, Uzunova V, Matsumoto K, Puia G, Mienville JM, Costa E, et al. Brain allopregnanolone regulates the potency of the GABAA
receptor agonist muscimol. Neuropharmacology 2000;39(3):440-448.
- 74. Chew L, Sun KL, Sun W, Wang Z, Rajadas J, Flores RE, et al. Association of serum allopregnanolone with restricted and
repetitive behaviors in adult males with autism. Psychoneuroendocrinology 2021;123:105039
- 75. Vacher CM, Lacaille H, O'Reilly JJ, Salzbank J, Bakalar D, Sebaoui S, et al. Placental endocrine function shapes cerebellar development and
social behavior. Nat Neurosci 2021;24(10):1392-1401.
- 76. Mueller BR, Bale TL. Early prenatal stress impact on coping strategies and learning
performance is sex dependent. Physiol Behav 2007;91(1):55-65.
- 77. Mueller BR, Bale TL. Sex-specific programming of offspring emotionality after stress
early in pregnancy. J Neurosci 2008;28(36):9055-9065.
- 78. Marsit CJ, Maccani MA, Padbury JF, Lester BM. Placental 11-beta hydroxysteroid dehydrogenase methylation is
associated with newborn growth and a measure of neurobehavioral
outcome. PLoS ONE 2012;7(3):e33794.
- 79. Howerton CL, Morgan CP, Fischer DB, Bale TL. O-GlcNAc transferase (OGT) as a placental biomarker of maternal
stress and reprogramming of CNS gene transcription in
development. Proc Natl Acad Sci USA 2013;110(13):5169-5174.
- 80. Nugent BM, O'Donnell CM, Neill Epperson C, Bale TL. Placental H3K27me3 establishes female resilience to prenatal
insults. Nat Commun 2018;9(1):2555
- 81. Estes ML, Kimberley McAllister A. Maternal immune activation: implications for neuropsychiatric
disorders. Science 2016;353(6301):772-777.
- 82. Smith SEP, Li J, Garbett K, Mirnics K, Patterson PH. Maternal immune activation alters fetal brain development through
interleukin-6. J Neurosci 2007;27(40):10695-10702.
- 83. Choi GB, Yim YS, Wong H, Kim S, Kim H, Kim SV, et al. The maternal interleukin-17a pathway in mice promotes autism-like
phenotypes in offspring. Science 2016;351(6276):933-939.
- 84. Shin Yim Y, Park A, Berrios J, Lafourcade M, Pascual LM, Soares N, et al. Reversing behavioural abnormalities in mice exposed to maternal
inflammation. Nature 2017;549(7673):482-487.
- 85. Kalish BT, Kim E, Finander B, Duffy EE, Kim H, Gilman CK, et al. Maternal immune activation in mice disrupts proteostasis in the
fetal brain. Nat Neurosci 2021;24(2):204-213.
- 86. Pakos‐Zebrucka K, Koryga I, Mnich K, Ljujic M, Samali A, Gorman AM. The integrated stress response. EMBO Rep 2016;17(10):1374-1395.
- 87. Kelleher RJ 3rd, Bear MF. The autistic neuron: troubled translation? Cell 2008;135(3):401-406.
- 88. Torossian A, Saré RM, Loutaev I, Smith CB. Increased rates of cerebral protein synthesis in
Shank3 knockout mice: implications for a link between
synaptic protein deficit and dysregulated protein synthesis in autism
spectrum disorder/intellectual disability. Neurobiol Dis 2021;148:105213
- 89. Wood H. Integrated stress response mediates cognitive decline in Down
syndrome. Nat Rev Neurol 2020;16(1):3
- 90. Dudova I, Kasparova M, Markova D, Zemankova J, Beranova S, Urbanek T, et al. Screening for autism in preterm children with extremely low and
very low birth weight. Neuropsychiatr Dis Treat 2014;10:277-282.
- 91. Guy A, Seaton SE, Boyle EM, Draper ES, Field DJ, Manktelow BN, et al. Infants born late/moderately preterm are at increased risk for a
positive autism screen at 2 years of age. J Pediatr 2015;166(2):269-275.E3.
- 92. Kuzniewicz MW, Wi S, Qian Y, Walsh EM, Armstrong MA, Croen LA. Prevalence and neonatal factors associated with autism spectrum
disorders in preterm infants. J Pediatr 2014;164(1):20-25.
- 93. Limperopoulos C, Bassan H, Sullivan NR, Soul JS, Robertson RL Jr, Moore M, et al. Positive screening for autism in ex-preterm infants: prevalence
and risk factors. Pediatrics 2008;121(4):758-765.
- 94. Jain VG, Willis KA, Jobe A, Ambalavanan N. Chorioamnionitis and neonatal outcomes. Pediatr Res 2022;91(2):289-296.
- 95. Horvath B, Lakatos F, Tóth C, Bödecs T, Bódis J. Silent chorioamnionitis and associated pregnancy outcomes: a
review of clinical data gathered over a 16-year period. J Perinat Med 2014;42(4):441-447.
- 96. Larsen JW, Sever JL. Group B Streptococcus and pregnancy: a
review. Am J Obstet Gynecol 2008;198(4):440-450.
- 97. Nasef N, Shabaan AE, Schurr P, Iaboni D, Choudhury J, Church P, et al. Effect of clinical and histological chorioamnionitis on the
outcome of preterm infants. Am J Perinatol 2013;30(01):059-068.
- 98. Allard MJ, Bergeron JD, Baharnoori M, Srivastava LK, Fortier LC, Poyart C, et al. A sexually dichotomous, autistic-like phenotype is induced by
group B Streptococcus maternofetal immune
activation. Autism Res 2016;10(2):233-245.
- 99. Braun AE, Carpentier PA, Babineau BA, Narayan AR, Kielhold ML, Moon HM, et al. "Females are not just 'Protected'
males": sex-specific vulnerabilities in placenta and brain after
prenatal immune disruption. eNeuro 2019;6(6):ENEURO.0358-19.2019
- 100. Carlezon WA Jr, Kim W, Missig G, Finger BC, Landino SM, Alexander AJ, et al. Maternal and early postnatal immune activation produce
sex-specific effects on autism-like behaviors and neuroimmune function in
mice. Sci Rep 2019;9(1):16928
- 101. Werling DM, Pochareddy S, Choi J, An JY, Sheppard B, Peng M, et al. Whole-genome and RNA sequencing reveal variation and
transcriptomic coordination in the developing human prefrontal
cortex. Cell Rep 2020;31(1):107489
- 102. Kim S, Lee J, Koh IG, Ji J, Kim HJ, Kim E, et al. An integrative single-cell atlas to explore the cellular and temporal
specificity of neurological disorder genes during human brain development
[Internet]; Cold Spring Harbor (NY): Cold Spring Harbor Laboratory; c2024 [cited 2024 Apr 11]. Available from: https://www.biorxiv.org/content/10.1101/2024.04.09.588220v1.
- 103. Chung C, Yang X, Bae T, Vong KI, Mittal S, Donkels C, et al. Comprehensive multi-omic profiling of somatic mutations in
malformations of cortical development. Nat Genet 2023;55(2):209-220.
- 104. Lee T, Lee H, Kim S, Park KJ, An JY, Kim HW. Brief report: risk variants could inform early neurodevelopmental
outcome in children with developmental disabilities. J Autism Dev Disord 2022;Sep 7;[Epub].