ChapterPDF Available

Sexual Differentiation and Sex Differences in Neural Development

Authors:

Abstract and Figures

Sex determination occurs at the moment of conception, as a result of XX or XY chromosome pairing. From that point, the body undergoes the process of sexual differentiation, inducing the development of physical characteristics that are easily distinguishable between the sexes and are often reflected in one’s physical appearance and gender identity. Although less apparent, the brain also undergoes sexual differentiation. Sex differences in the brain are organized during a critical period of neural development and have an instrumental role in determining the physiology and behavior of an individual throughout the lifespan. Understanding the extent of sex differences in neurodevelopment also influences our understanding of the potential risk for a number of neurodevelopmental, neurological, and mental health disorders that exhibit strong sex biases. Advances made in our understanding of sexually dimorphic brain nuclei, sex differences in neural cell communication, and sex differences in the communication between the brain and peripheral organs are all research fields that have provided valuable information related to the physiological and behavioral outcomes of sex differences in brain development. More recently, investigations into the impact of epigenetic mechanisms on sexual differentiation of the brain have indicated that changes in gene expression, via epigenetic modifications, also contribute to sexual differentiation of the developing brain. Still, there are a number of important questions and ideas that have arisen from our current understanding of sex differences in neurodevelopmental processes that necessitate more time and attention in this field.
Content may be subject to copyright.
Sexual Differentiation and Sex Differences
in Neural Development
Alexandra Turano, Brittany F. Osborne, and Jaclyn M. Schwarz
Contents
1 Introduction
2 Sex Determination and Sexual Differentiation
2.1 Biological Sex is Determined at the Time of Conception
2.2 Sex Hormones Induce Sexual Differentiation of the Body
2.3 Sex Hormones Induce Sexual Differentiation of the Brain
2.4 Genetic Sex Differences Impact the Developing Brain
3 Sex Differences in the Size of Brain Nuclei
3.1 The Spinal Nucleus of the Bulbocavernosus
3.2 The Anteroventral Periventricular Nucleus of the Hypothalamus
3.3 The Bed Nucleus of the Stria Terminalis
4 Sexual Differentiation of the Brain Requires Communication Across and Involvement
of Multiple Cell Types in the Brain
4.1 An Immune Molecule and a Neurotransmitter Coordinate Masculinization
of the Preoptic Area
4.2 Sexual Differentiation Across the Synapse in the Ventromedial Nucleus
of the Hypothalamus
4.3 Development of Neuropeptides Involved in Sex-Specic Behaviors
4.4 Microglia Induce Sex-Specic Neuronal Development
4.5 Sex Differences in Cell Genesis and Fate During the Development of Cognitive
Circuits
5 A Whole Body Perspective on Sex Differences in Neural Development
5.1 The Gut Microbiome
5.2 The Liver
5.3 The Placenta
6 Epigenetics and the Sexual Differentiation of the Brain
7 Conclusions
References
Abstract Sex determination occurs at the moment of conception, as a result of XX
or XY chromosome pairing. From that point, the body undergoes the process of
A. Turano, B. F. Osborne, and J. M. Schwarz (*)
Department of Psychological and Brain Sciences, University of Delaware, Newark, DE, USA
e-mail: jschwarz@psych.udel.edu
©Springer International Publishing AG, part of Springer Nature 2018
Curr Topics Behav Neurosci
DOI 10.1007/7854_2018_56
sexual differentiation, inducing the development of physical characteristics that are
easily distinguishable between the sexes and are often reected in ones physical
appearance and gender identity. Although less apparent, the brain also undergoes
sexual differentiation. Sex differences in the brain are organized during a critical
period of neural development and have an instrumental role in determining the
physiology and behavior of an individual throughout the lifespan. Understanding
the extent of sex differences in neurodevelopment also inuences our understanding
of the potential risk for a number of neurodevelopmental, neurological, and mental
health disorders that exhibit strong sex biases. Advances made in our understanding
of sexually dimorphic brain nuclei, sex differences in neural cell communication,
and sex differences in the communication between the brain and peripheral organs
are all research elds that have provided valuable information related to the physio-
logical and behavioral outcomes of sex differences in brain development. More
recently, investigations into the impact of epigenetic mechanisms on sexual differ-
entiation of the brain have indicated that changes in gene expression, via epigenetic
modications, also contribute to sexual differentiation of the developing brain. Still,
there are a number of important questions and ideas that have arisen from our current
understanding of sex differences in neurodevelopmental processes that necessitate
more time and attention in this eld.
Keywords Hormones · Neural development · Sexual determination ·
Sexual differentiation
1 Introduction
The brain develops very early in prenatal life via a number of important processes
that produce the cells, shape the structures, and rene the connections within the
central nervous system. The goal of these neurodevelopmental processes is to create
a brain that controls our physiology and behavior for the rest of our lives; but also,
these processes produce the identifying characteristics and personality traits that we
associate with each and every one of us. One of the most notable and arguably more
important dening characteristics of ourselves is our sex. The sex of many mam-
mals, including humans, is determined at the moment of conception, via the simple
process of sex determination that almost belies its importance; however, that single
moment of sex determination subsequently shapes the physiology, brain, behavior,
and health outcomes of the individual for a lifetime.
The physical characteristics that dene our sex are often quite obvious and are
usually enhanced in our appearance by our gender identity, which is reected in our
personality, our mannerisms, and our personal style. What may be less obvious, but
equally fascinating, are the sex differences that are established in our brains during
the process of neural development. The primary goal of these sex differences in the
brain is to control our physiology and certain behaviors in a sex-specic manner, and
in humans these sex differences also dene our gender identity and inuence our
sexuality. Just as important, sex differences in the brain can inuence the risk for
certain neurodevelopmental disorders, neurological disorders, and mental health
A. Turano et al.
disorders; thus, it is important that we understand how sex differences in the
brain are established during early brain development and maintained throughout
the lifespan in order to better understand how these differences inuence our
mental health and well-being.
The goal of this chapter is to summarize our understanding of how sex differences
in the brain are established during neural development in the context of how these
sex differences regulate sex-specic behaviors later in life. We will identify histor-
ical and recent ndings that highlight the various mechanisms by which sex differ-
ences in the brain are established during development. Another goal throughout this
chapter will be to identify important future questions or areas of research in the eld
of sexual differentiation of the brain and sex differences in neural development.
2 Sex Determination and Sexual Differentiation
2.1 Biological Sex is Determined at the Time of Conception
Sex determination occurs at the moment of conception, when a females egg is
fertilized by a males sperm. The egg and sperm each contribute one half of the
genetic material to the resulting zygote. Each human zygote contains 23 chromo-
somes, including 22 pairs of autosomes and 1 pair of sex chromosomes. The egg
contributes one X chromosome to the zygote, whereas the sperm may contribute
either one X or one Y chromosome. If the sperm contributes an X chromosome, the
resulting zygote will become a female. She will possess two of the same allosomal
(sex) chromosomes, one X chromosome from the egg and one X chromosome from
the sperm. Conversely, if the sperm contributes a Y chromosome, the zygote will
become a male (Fig. 1). He will have two different allosomes, an X chromosome
from the egg and a Y chromosome from the sperm, thereby differentiating him
genetically and phenotypically from a developing female (Painter 1923). As cells
multiply and the zygote matures into an embryo, then fetus, and eventually a
newborn baby, the presence of either XX or XY chromosomes is maintained within
each and every cell, and therefore each cell sustains an autonomous chromosomal
sex throughout the lifespan that is established at the moment of conception.
Sexual differentiation is the process by which the developing embryo becomes
male or female after sex determination; and this important process occurs quite early
in embryonic development. Sexual differentiation is the process by which primary
sex characteristics develop, including the development of the gonads and sex organs,
but it is also the process by which many sex differences in the brain are established
(Fig. 1). While it is a distinct process, sexual differentiation is dependent upon the
process of sex determination. In particular, the sex-determining region of the Y
chromosome, known as the Sry gene, is critical for the process of sexual differ-
entiation of males. The Sry gene encodes for the protein known as testis-determining
factor (TDF). TDF initiates the differentiation of steroidogenic precursor cells and
primordial germ cells within the developing gonads into Sertoli cells, resulting in the
Sexual Differentiation and Sex Differences in Neural Development
formation of testes (Fig. 1). Developing females (XX) do not have Y chromosomes
and therefore, do not have the SRY gene. Consequently, devoid of TDF expression,
the precursor cells of the developing gonadal organs differentiate into granulosa
cells, resulting in the formation of ovaries (Sekido 2014; Fig. 1).
Prior to sexual differentiation and Sry gene expression, the developing gonads are
bipotential,meaning they have the ability to develop into either testes or ovaries.
Importantly, expression of the Sry gene in developing males must occur within a
sensitive window during embryogenesis in order for the bipotential gonads to
differentiate into testes as opposed to ovaries (Taketo et al. 2005). The long-held
belief was that without the presence of the Sry gene, the precursor cells of
the developing gonads passively become ovaries (Haseltine and Ohno 1981;
Fig. 1 Sex determination and sexual differentiation of the brain. Sex determination occurs at the
moment of conception, when the embryo receives either two X chromosomes (female embryo) or
an X and a Y chromosome (male embryo) from the mother and the father. If the developing embryo
has a Y chromosome, this chromosome contains the sex-determining region of the Y chromosome
(SRY gene) that is expressed early in development within specic tissues, in particular the
undifferentiated gonads. The SRY gene encodes for a protein known as the testes-determining
factor, which differentiates the gonads into a testis early in development. In the absence of the Y
chromosome, in females, the gonads differentiate into ovaries. Later, the testes begin to secrete
testosterone and dihydrotestosterone (DHT), which differentiate the body and the brain into a male
phenotype. In particular, the external genitalia are masculinized; the Wolfan duct develops into the
ducts, epididymis, vas deferens, and seminal vesicles; and, importantly for this chapter, the brain is
masculinized. In the absence of any testosterone secretion, the external genitalia develops as female,
the Müllerian duct develops as the uterus and fallopian tubes, and the brain is feminized
A. Turano et al.
Jost 1947). Contrary to this long-held belief, however, sex reversal studies investi-
gating the result of XX, loss-of-function and XY, gain-of-function mutations on
specic sex-linked genes including Wnt4,β-catenin,R-spondin1, and Foxl2 have
found that differentiation of the bipotential gonads into ovaries also requires the
active repression of testis formation (Chassot et al. 2008; Maatouk and Capel 2008;
Pailhoux et al. 2001; Sekido 2014; Vainio et al. 1999). These ndings in particular
highlight the complex process of gonadal differentiation in males and females.
Importantly, the brain is also differentiated as either male or female via active and
distinct processes; however, it has proven to be more difcult to discern the
mechanisms by which the developing brain actively differentiates as female,
which we will discuss in further detail in Sect. 6.
2.2 Sex Hormones Induce Sexual Differentiation of the Body
Once sexually differentiated, the male testes begin to secrete important molecules for
the continued sexual differentiation of the body and the brain; thus, the process of
gonadal differentiation is not concurrent with the process of brain sexual differ-
entiation, but rather one process sequentially follows the other (Fig. 1). Because the
process of sexual differentiation of the body and brain occurs at various points after
sex determination and the formation of the gonads, disruption in one or more of these
processes can result in disorders of sexual development wherein the sexof the
differentiated body or brain does not match the sexof the already differentiated
gonads. This may result in ambiguous genitalia or a dysmorphic gender identity,
both of which comprise an important, but distinct area of research from what we will
cover here.
In males, the testes produce an early surge of androgens, including testosterone
and dihydrotestosterone (DHT), that leads to the development of sex-related organs,
such as the epididymis, vas deferens, and male genitalia (Fig. 1). The testes also
secrete anti-Müllerian hormone, which is responsible for actively repressing
the Müllerian duct. The Müllerian duct is a primordial structure necessary for the
development of the female sex organs; thus its suppression is necessary for the
complete and correct development of male as opposed to female sex organs (Josso
et al. 1993). As the Müllerian duct regresses in developing males, the Wolfan duct
grows into the male accessory organs (Fig. 1). In contrast to developing males,
developing females possess ovaries. Like the testes, the ovaries also secrete sex
hormones; however, sex hormone secretion from the ovaries occurs at a much later
time at the onset of puberty (Sekido 2014). Thus, in the complete absence of
perinatal androgen exposure, the Wolfan duct regresses, and the Müllerian duct
develops, opposite to the ongoing processes that occur in developing males. Speci-
cally, in the absence of early-life androgens and anti-Müllerian hormone, female sex
organs, such as the oviduct, uterus, and female genitalia, begin to develop (Jost
1947; Fig. 1). As the appearance of these outwardly distinct male or female sex
organs begins to develop, we are able to determine the sex of a developing fetus in
Sexual Differentiation and Sex Differences in Neural Development
the womb during an ultrasound often a monumental event when anticipating the
arrival of a soon-to-be bumbling baby.
2.3 Sex Hormones Induce Sexual Differentiation of the Brain
Sexual differentiation does not end at the level of the gonads and sex organs. This
process of becoming male or female also involves the sexual differentiation of the
brain (Fig. 1). Like the gonads, the brain is initially a bipotential organ, meaning that
it has the potential to be either masculinized or feminized. The presence (or absence)
of androgens during perinatal development organizescertain structures and func-
tions of the brain in a sexually dimorphic manner. The consequences of this
organization become most apparent following activationof these sexually dimor-
phic brain regions during puberty, which is induced by the production of the sex
hormones testosterone (males) and estrogen and progesterone (females). As a result,
this theory of sexual differentiation of the brain became known as the organizational-
activational hypothesis (Phoenix et al. 1959; Fig. 2) that sex differences in the brain
are organized during early brain development and then activated by circulating
Fig. 2 The organizational-activational hypothesis of sex differences in the rodent brain. Testos-
terone levels peak around birth in the developing male rodent and naturally decrease within a few
days after birth. During that time, the testosterone permanently masculinizes the brain, organizing
the size and structure of certain brain regions necessary to perform male sex behavior. In the
absence of testosterone, the brain is feminized, permanently organizing the size and structures of the
brain regions necessary to perform female sex behavior. At puberty, the testes and the ovaries begin
to secrete sex steroid hormones. Males gradually increase production of testosterone, the levels of
which are maintained at a relatively consistent point. Females show varying levels of estradiol and
progesterone throughout the estrous cycle necessary for ovulation. The production of sex steroid
hormones at puberty activates brain regions that were organized during development, thereby
producing sex-specic sexual behaviors necessary for reproduction
A. Turano et al.
hormones after puberty. Demonstrating this theory, Phoenix et al. (1959)rst
ovariectomized female neonatal guinea pigs and then treated them with testosterone
during the organizationalperiod of brain development. Following what would
have been the onset of puberty, female guinea pigs were treated with a second
injection of testosterone, and they subsequently exhibited male-typical, as opposed
to female-typical, sex behavior (Phoenix et al. 1959). Testosterone given to the
ovariectomized neonatal female guinea pig was sufcient to masculinize the brain of
the genetically female (XX) guinea pig, causing her to display male-typical sex
behaviors following a second testosterone injection later in life. Notably, this process
of testosterone-induced sexual differentiation of the brain is conserved across many
species including other rodents, birds, and primates, including humans.
The mechanisms by which certain brain regions are organized in a sexually
dimorphic manner, resulting in sex-specic behavior, involve specic processes
that must occur at specic stages of neural development. In accordance with the
organizational-activational hypothesis and the experiments by Phoenix and col-
leagues in 1959, studies by Gorski and colleagues in 1978 indicated that ovariecto-
mized female rats also exhibited male-typical sex behavior following a pubertal
testosterone injection, if they had been exposed to testosterone early in life. Inter-
estingly, they proposed that this behavior was due to increases in the number of cells
within a brain region known as the sexually dimorphic nucleus of the preoptic area
(SDN-POA; Gorski et al. 1978). The SDN-POA is larger in males than in females,
indicative of its role in male-typical sex behavior. Concurrently, castrated males (i.e.,
devoid of perinatal testosterone exposure) exhibited no sex behavior following an
injection with testosterone in puberty and showed decreases in the number of cells
within the SDN-POA (Murakami and Arai 1989). Therefore, androgens are neces-
sary to prevent ongoing programmed cell death, or apoptosis, in the SDN-POA
during early postnatal brain development, and in doing so, androgens organize the
size and structure of the SDN-POA in a sexually dimorphic manner (Dodson and
Gorski 1993). The presence or absence of early-life androgens impacts the size of the
SDN-POA, as well as a number of other sexually dimorphic brain regions (Fig. 3);
and in doing so, it permanently affects the behavioral output of the animal in a
sex-specic manner later in life.
Notably, testosterone secreted by the testes is converted to estradiol in the
developing male rodent brain, by the p450 enzyme, aromatase. While testosterones
action on the androgen receptor contributes to masculinization of the brain, estra-
diolsaction on estrogen receptors (in particular, estrogen receptor α) is also
important for the masculinization and defeminization of the developing male rodent
brain (Arnold and Gorski 1984). During this critical period of sexual differentiation
of the brain, female rodents may be exposed to exogenous estrogens from the
mother, but, female rodents are protected from any possible estradiol-induced
changes in the developing brain by the high afnity alpha-fetoprotein (AFP) binding
protein, which binds any possible circulating estrogens during the early develop-
mental period, thus protecting females against potential masculinization and
defeminization of the brain by maternal estrogens (Bakker et al. 2006).
Sexual Differentiation and Sex Differences in Neural Development
Fig. 3 The basic processes of neural development that occur in a sex-specic manner. (a) Proliferation of cells or neurogenesis (the generation of new neurons)
occurs in a sexually dimorphic manner in brain regions including the anteroventral periventricular nucleus (AVPV), the hippocampus, and the amygdala. (b)
Differentiation of neural progenitor cells into either neurons or glia can occur in a sex-dependent manner in brain regions such as the amygdala and the
anteroventral periventricular nucleus (AVPV). (c) Testosterone or estradiol converted from testosterone promotes the outgrowth of axons seeking new synaptic
connections in the amygdala and the AVPV. (d) In the preoptic area (POA) and the ventromedial nucleus of the hypothalamus (VMN), estradiol converted from
A. Turano et al.
testosterone promotes dendritic spine synapses during postnatal brain development. (e) In the spinal nucleus of the bulbocavernosus (SNB) muscle, testosterone
promotes axon survival of neurons projecting to the muscle. In the absence of testosterone, in females, the synapses are naturally eliminated during postnatal
development. (f) Naturally occurring cell death, apoptosis occurs in a sexually dimorphic manner to produce sex differences in the size of certain brain nuclei. In
brain regions where the size is larger in males than in females (M >F), testosterone promotes the survival of the neurons and prevents apoptosis from occurring.
In brain regions where the size is larger in females than in males (F >M), testosterone promotes naturally occurring cell death during neural development
Sexual Differentiation and Sex Differences in Neural Development
The seminal study by Phoenix and colleagues in 1959, and subsequent studies by
Gorski and colleagues, provided the rst evidence of gonadal hormone-induced
sexual differentiation of the brain and behavior. This work led to an intense
investigation into the cellular mechanisms by which hormones can produce
sex-specic phenotypes of the brain and behavior during early brain development.
One of the best-known examples of sex differences in brain structure producing sex
differences in behavior is observed in birds that produce song. Male birds that
produce song as a way to attract female mates have signicantly larger nuclei
compared to females within two main neural pathways important for the learning
and the production of song (Jordan et al. 1988; Nottebohm and Arnold 1976;
Nottebohm et al. 1982). For example, the high vocal center(HVC) and the robust
nucleus of the archistriatum (RA) are two sexually dimorphic nuclei within the
efferent motor pathway responsible for the production of male bird song. Both
regions are three to six times larger in males compared to females (Nottebohm and
Arnold 1976; Nottebohm 1991). The neurons of these nuclei were later found to be
more numerous, larger in size, and more scattered in males than in females (Arnold
and Gorski 1984; Mathews et al. 1988). If females are treated with estradiol early
post-hatching and then treated with either testosterone or dihydrotestosterone (DHT)
as adults, the HVC and the RA become similar in size to that of the males, and the
treated females are then able to produce song (Gurney and Konishi 1980).
Characterization of the sexually dimorphic nuclei controlling song production in
birds provided some of the rst direct evidence linking hormone-induced changes in
brain structure to sexually dimorphic behavior but also provided converging support
for the organizational-activational hypothesis of sexual differentiation of the brain.
That said, evidence from these studies also suggested that the organizational and
activational effects of hormones cannot completely account for the sexual differen-
tiation of these brain nuclei. For example, male birds that are castrated as hatchlings,
prior to the secretion of testosterone, sing normally as adults, suggesting the possi-
bility that activation of genes on the avian sex chromosomes may also induce
masculinization of the brain in song birds (Arnold 2003; Wade et al. 2001).
2.4 Genetic Sex Differences Impact the Developing Brain
While the evidence above elucidates androgens as an essential mediator of sexual
differentiation of the brain, it is also important to keep in mind that each cells
intrinsic sex may also have the ability to contribute to the organizational-activational
hypothesis, as the presence of either XX or XY chromosomes is maintained within
every cell. In other words, organization of sexually dimorphic brain regions may not
rely on gonadal sex hormones alone but may also involve sex chromosome com-
plement, such as the direct expression of the Sry gene within male brain cells or sex
differences in gene expression that result from the process of X inactivation within
female brain cells (Arnold and Chen 2009; Carrel and Willard 2005; Dewing et al.
2006). This phenomenon has been investigated using a mouse model described as
A. Turano et al.
the four core genotypes(FCG) model, in which gonadal sex and chromosomal sex
are purposely misaligned to examine the impact of each factor either separately or
interdependently (Carruth et al. 2002; De Vries et al. 2002). This phenomenon has
also been investigated in a gynandromorphic nch, in which sex chromosome
complement was bilaterally split in a bird, but testosterone derived from the one
testis (on the genetic maleside of the body) could circulate throughout the
bloodstream and inuence the entire organism, including its brain (Agate et al.
2003). Thus, when sex differences were present across both hemispheres of the
brain in the gynandromorphic nch, despite the similar hormonal exposure, they
could be attributed to sex differences in gene expression that resulted from the avian
sex chromosomes. Notably, this nch could still produce male song, indicative of the
important role testosterone has in differentiating man regions within male brain and
associated behavior. Taken together, these models have been effective in shedding
light on how sex chromosome complement affects sex differences in gene expres-
sion and phenotypic variation in the brain beyond the scope of sex hormones alone
(Carruth et al. 2002; Xu and Disteche 2006).
It is also important to consider the evolutionary basis underlying the mechanisms
for sexual differentiation of the brain in general. Despite the fact that males and
females each have a unique set of sex chromosomes, the sex chromosomes have an
important, but relatively limited role in the process of sexual differentiation of the
body and brain in many species, including humans (McCarthy and Arnold 2011).
Instead, the sex hormones are the primary mechanism by which sexual differentia-
tion of the body and brain occurs. Sexual differentiation requires the coordinated
differentiation of a wide variety of cells and tissues throughout the body. These cells
and tissues are also mutually dependent upon each other for proper function and
coordinated behavior later in life. As a result, sex hormones may have taken onthe
important role of sexual differentiation because of their ability to target specic cells
and tissues throughout the entire body and because of their important role in
sex-specic physiology and behavior in general. Despite this, there may be many
other processes of neural development that are directly affected by the genes of the
sex chromosomes or by the sexually dimorphic expression of genes independent of
sex hormones that remain to be discovered. Thus, the complexity of sex-specic
neurodevelopmental and long-term behavioral outcomes resulting from either
chromosomal or hormonal sex, including those that do not apply directly to repro-
ductive or copulatory behavior, continues to be explored.
The organizational-activational hypothesis is supported by sex differences in the
development of a number of brain regions including the SDN-POA (described
above), the spinal nucleus of the bulbocavernosus (SNB), the bed nucleus of the
stria terminalis (BNST), and the anteroventral periventricular nucleus of the hypo-
thalamus (AVPV) (Fig. 3). These brain regions and their associated sex-specic
behaviors will be described throughout the remaining chapter. Importantly, we will
focus our discussion on the mechanisms by which neurodevelopment is inuenced
by sex hormones to induce these well-known sex differences in the brain, and in the
nal conclusions of this chapter, we will discuss important future questions and
ideas that arise from our understanding of sex differences in neurodevelopmental
processes.
Sexual Differentiation and Sex Differences in Neural Development
3 Sex Differences in the Size of Brain Nuclei
3.1 The Spinal Nucleus of the Bulbocavernosus
The spinal nucleus of the bulbocavernosus (SNB) was the rst region identied in
the rodent central nervous system (CNS) to exhibit an androgen-dependent sex
difference in cell number. Specically, males have signicantly more neurons within
the SNB than females, and treatment of neonatal females with testosterone increases
the number of neurons in the female SNB to that seen in males (Sengelaub and
Forger 2008). Breedlove and Arnold (1980) identied this collection of motor
neurons, located within the fth and sixth lumbar segments of the rat spinal cord,
which they named the SNB after the perineal striated muscles these neurons inner-
vate. They observed that these particular motor neurons were more sensitive to
treatment with testosterone than treatment with estradiol, the aromatized product
of testosterone. Interestingly, when they examined genetically male rats with a
testicular feminization mutation that severely reduced the amount of functional
androgen receptors, Breedlove and Arnold (1981) observed a reduction in, or
demasculinization of, the total cell number within the SNB of these mutated animals.
Based on these observations, they concluded that testosterone signaling at the
androgen receptor during development plays an important role in determining the
sex difference in the number of neurons within this nucleus of the CNS.
Nordeen et al. (1985) subsequently identied an epoch within the perinatal
period between embryonic day 22 (E22) and postnatal day 10 (P10) during
which the SNB cell numbers are decreased via naturally occurring apoptotic
processes, and this neurodevelopmental process is more robust in female compared
to male rats, resulting in the sex difference in cell number described above (Fig. 3).
Female rats that were treated with testosterone during the critical period of spinal
cord development lost signicantly fewer neurons in the SNB, thereby exhibiting a
masculinized SNB cell number (Nordeen et al. 1985). In concordance with these
data, Breedlove and Arnold (1983) demonstrated greater cell death of SNB motor
neuron number when treating male rats with utamide, an androgen receptor anta-
gonist, which blocks the androgen receptor. Together, these data indicate that
androgens are effective in preventing programmed cell death within the SNB during
perinatal development. Ultimately, this proposed mechanism of sexually dimorphic
cell death is involved in the establishment of the sexually dimorphic size of the SNB,
which is maintained into adulthood and allows for successful male reproductive
behavior, in particular the control of the perineal muscles.
In addition to the impact of neonatal testosterone on naturally occurring apoptosis
in the SNB, Lubischer and Arnold (1995) found that treatment of castrated juvenile
male rats with testosterone delayed naturally occurring synapse elimination in the
spinal cord by increasing the percentage of perineal muscle bers innervated by
multiple axons, thus increasing the size of the associated motor units. The observed
changes in synapse elimination persisted for at least 12 months after the completion
of testosterone treatment, demonstrating that the effects of testosterone are persistent
A. Turano et al.
and have an important role in the establishment of sex differences in the spinal cord
that are maintained long-term (Lubischer and Arnold 1995).
These data sparked an investigation into the ontogeny of androgen receptor
expression in the muscles innervated by SNB motor neurons; and, in 1997, Jordan
and colleagues denitively showed that the rat perineal muscles that are innervated
by SNB motor neurons also express androgen receptors. More importantly, the
emergence of androgen receptors on the perineal muscles actually precedes their
emergence on the motor neurons that synapse with the SNB (Jordan et al. 1997). In
fact, in the mouse SNB circuit, androgen receptors were not expressed on SNB
motor neurons until P4, but they were present on the SNB target muscles as early as
E15 (Smith et al. 2012). Ultimately, these discoveries have challenged the notion
that neonatal androgen exposure is necessary for the sex difference in cell number in
the SNB via the direct action of androgens on the motor neurons themselves,
particularly given that androgen receptors were not detected on SNB motor neurons
until P4, a time point outside of the natural surge in testosterone that occurs in
developing males. Instead, androgensaction on the perineal muscles inuences the
way by which these muscles are subsequently innervated by SNB motor neurons at a
later date, and thereby, it is androgensaction on the muscles itself that has a critical
role in establishing the male-biased sex difference in SNB cell number that is
maintained into adulthood. This is an important point to consider when understand-
ing the various mechanisms by which sex differences in the nervous system can be
established. Sex hormones, in this case testosterone, can inuence one set of target
cells (muscle tissue) to, in turn, inuence the innervation, number, and function of
motor neurons in the spinal cord, likely via the androgen-induced structural support
and growth factors that are present at the neuromuscular junction (Fig. 3). These
ndings highlight the idea that sexual differentiation of the nervous system can be
initiated early, at the time of perinatal testosterone exposure, but that the effects can
also be persistent, as testosterone is able to maintain cell survival at a later point in
development, as naturally occurring apoptosis continues postnatally.
3.2 The Anteroventral Periventricular Nucleus
of the Hypothalamus
Unlike many of the other sexually dimorphic brain regions discussed in this chapter,
the anteroventral periventricular nucleus of the hypothalamus (AVPV) is signi-
cantly larger in females than in males. This particular nucleus of the hypothalamus
plays an important role in gonadotropin release via the robust expression of
kisspeptin, found in a subset of AVPV neurons (Kauffman et al. 2007). Specically,
the Kiss1 neurons in the AVPV control a preovulatory surge in luteinizing hormone
(LH) that drives ovulation in females (Hu et al. 2015). Prior to its designation as the
AVPV, this collection of cells was originally identied by Bleier et al. (1982) as the
medial preoptic nucleus (MPN). In a comparative anatomy study, they discovered
Sexual Differentiation and Sex Differences in Neural Development
that the MPN (or AVPV) of the guinea pig, rat, and hamster contained more cells in
adult females compared to adult males (Bleier et al. 1982). The work of Bleier and
colleagues established that the sexually dimorphic nature of this specic brain region
is, indeed, conserved across a number of species, corresponding with the later
discovery of the AVPVs important role in female sex behaviors and the control of
ovulation.
In previous sections, we discussed how neonatal androgen receptor signaling
mediated the establishment of a sex difference in cell number within the SNB and the
SDN-POA via its impact on cell survival by preventing naturally occurring
apoptosis. Similar to these two brain regions, sex-specic cell death is also the
cause of differences in cell number within the AVPV. In developing rats, the
AVPV undergoes neurogenesis between E13 and E18, and treatment of females
with testosterone or estradiol does not affect the number of newly born neurons in
the AVPV (Sumida et al. 1993). In contrast, treatment with either testosterone or
estradiol increases subsequent apoptosis in the AVPV, resulting in a sex difference
in the size of the AVPV by the time it was examined at E21 (Sumida et al. 1993;
Fig. 3). Notably, this pro-apoptotic effect of testosterone in the AVPV is contrary to
the role of neonatal testosterone signaling in the SNB, which protects against
apoptosis (Arai et al. 1996; Sumida et al. 1993). Furthermore, the data indicate
that the aromatization of testosterone into estradiol and thus the actions of estradiol,
in particular, contribute to increased cell death in the developing male AVPV. These
data are corroborated by the high level of estrogen receptor alpha (ERα) expression
on the cells of the AVPV (Shughrue et al. 1997). Bax, a protein that promotes
apoptosis, is found at higher concentrations in the developing male AVPV; while
Bcl-2, a protein that promotes cell survival, is found at lower levels in the developing
male AVPV (Tsukahara 2009). Thus, estradiol aromatized from testosterone results
in the sexually dimorphic expression of these pro- and anti-apoptotic genes (see
Forger 2006 for a comprehensive review), pushing the newly born neurons in
the male AVPV toward a fate of cell death during this embryonic period of
neurodevelopment (Fig. 3).
It is also important to note that the AVPV controls the LH surge in adult females
as a result of its ability to integrate hormonal signals with circadian signals, via the
function of kisspeptin neurons localized in this brain region (Simerly 2002). The LH
surge is the result of a neuroendocrine positive feedback loop between kisspeptin
and estradiol, which is only fully functional after puberty (Clarkson et al. 2009). This
positive feedback loop of hormones acting on the AVPV results in the addition of
even more cells to the female AVPV. Thus, similar to the mechanisms and timeline
of sexual differentiation of the SNB, the sex difference in cell number in the AVPV
is initially established during the prenatal period via apoptosis; however, there
remain additional changes to cell number during puberty that are necessary for the
establishment and maintenance of sex differences in the adult AVPV necessary to
control female ovulation. Interestingly, Mohr et al. (2016) found that this increase in
cell number in the AVPV at puberty not only includes the addition of new neurons
but also includes the addition of glial cells, indicating that both neurons and glia
appear to be contributing to the establishment and maintenance of sex differences in
A. Turano et al.
the size and function of the AVPV (Mohr et al. 2016; Fig. 3). This important idea
that various cell types in the developing brain contribute to sex-dependent develop-
ment and function of the brain is seen in other examples that will be discussed in
Sect. 4.
3.3 The Bed Nucleus of the Stria Terminalis
The bed nucleus of the stria terminalis (BNST) is located at one end of the stria
terminalis, which is the bundle of axon bers that connects the BNST to the
amygdala. Because of its structure and its role in social and emotional behaviors,
the BNST has been described as the extended amygdala(Lebow and Chen 2016).
Though the BNST forms a continuum with amygdaloid structures, it is now clear
that there are at least two distinct regions, including the medial division containing
the medial amygdaloid nucleus and the medial BNST, as well as a central division
containing projections between the central amygdaloid nucleus and the lateral BNST
(reviewed in Crestani et al. 2013). The BNST contains approximately 20 distinct
nuclei that are characterized based on the various cell types and their distinct pro-
jections to other brain regions. Characterization of the various cell types and the
afferent and efferent projections of the BNST (e.g., limbic structures, hypothalamic
structures, and brain stem nuclei) suggests that it serves as a critical relay for a
variety of neuronal circuits and other systems (e.g., neuroendocrine) to coordinate
activity related to the physiology and behavior of the organism. The BNST neurons
contain a variety of neurotransmitter and neuropeptides and their respective recep-
tors including vasopressin, glutamate, GABA, noradrenaline, acetylcholine, nitric
oxide, cannabinoids, and corticotropin-releasing factor (Crestani et al. 2013).
There is a signicant sex difference in the size of the principle nucleus of the
BNST (pBNST) in addition to sexually dimorphic innervation of vasopressin neu-
rons into the BNST. Vasopressin has been studied extensively in the BNST for its
role in various physiological and behavioral functions (see Sect. 4.3). For this
reason, much of our understanding of the functional role of the BNST in males
and females throughout brain development has its foundation in the vasopressin
system. Males have signicantly more vasopressin immunoreactivity, mRNA
expression, biosynthetic capacity, and a greater ber density within the BNST
compared to females (Bales et al. 2007; Miller et al. 1989a,b). One of the main
vasopressin receptors, V1a, is more abundant in the BNST of males compared to
females in the medial posterior BNST, but not in the lateral dorsal or lateral posterior
BNST, suggesting that there is considerable heterogeneity within the subnuclei of
this region for V1a receptor distribution (Dumais and Veenema 2016). Vasopressin
neurons in the medial BNST were identied as the source of the sexually dimorphic
vasopressin ber innervation of the brain in the mid- to late 1980s with the discovery
that males have nearly twice as many vasopressin-immunoreactive bers than
females (Fig. 3). Additionally, Miller et al. (1989a,b) determined that males have
signicantly more vasopressin-expressing cells and more vasopressin grains per cell
Sexual Differentiation and Sex Differences in Neural Development
than females. These data indicate that the sex differences in vasopressin BNST ber
densities are due to males having more vasopressin-expressing neurons as well as
increased biosynthetic capacity on a per cell basis compared to females.
Progesterone is also important for establishing the sexually dimorphic expression
of vasopressin cells in the BNST. Testosterone, converted to estradiol, can induce
the synthesis of progesterone in the brain during postnatal development (Shughrue
et al. 1992), but more so, it is thought that progesterone from the pregnant mother
may induce sex differences in the developing fetal brain if sex differences are present
in the expression of progesterone receptors in the brain (Wagner et al. 1998). In fact,
there is a striking sex difference in progestin receptor (PR) expression on vasopressin
cells in the BNST (Quadros et al. 2002), and subcutaneous injection of progesterone
reduces the number of vasopressin mRNA-expressing cells in the BNST. To deter-
mine the role of PRs in the development of sex differences in the vasopressin system,
Rood et al. (2008) compared PR
lacZ
reporter mice to wild-type (WT) mice. PR
lacZ
reporter mice have a functional disruption in the PR gene that renders them unable to
respond to circulating progesterone. Surprisingly, PR
lacZ
reporter mice showed a
similar sex difference in the number of vasopressin-expressing cells in the BNST as
WT mice, suggesting that PRs do not play a major role in sexual differentiation of
vasopressin cells in the BNST; however, additional data from this study indicated
that PRs have a role in establishing the density of projections to brain regions that are
innervated by BNST originating vasopressin neurons (Rood et al. 2008). These
ndings highlight the more unconventionalmechanisms by which sex differences
in the developing brain can be established, in this case via the function of the
progesterone receptor. While it is clear that the BNST is sexually dimorphic, with
the vasopressin system playing an integral role in this sex difference, it remains to be
determined how disruptions in the development of the BNST can affect brain
development and sex-specic behavior later in life.
4 Sexual Differentiation of the Brain Requires
Communication Across and Involvement of Multiple Cell
Types in the Brain
As described in the previous sections, the sex of an organism is determined at the
moment of conception, and from that point of determination, every cell in the
organism has a sex. The sex chromosomes and, more importantly, sex hormones
differentiate the developing brain, body, and behavior of the organism early in
development. These early-life organizational and activational effects of sex hor-
mones programthe function of various cells throughout the nervous system. As a
result, it is important that we also understand the cellular mechanisms by which
basic sex differences in the brain are established, as this may provide additional
insight into the cellular mechanisms resulting in disease processes that often occur in
a sex-specic manner throughout neurodevelopment.
A. Turano et al.
The brain is arguably one of the most active endocrine organs as it produces and
responds to a diverse array of chemical messengers. Chemical communication, such
as the kind that occurs in the endocrine system, via hormones and their cognate
receptors, is integral to all levels of biological organization, from the mediation of
intracellular processes to communication between organs and even communication
between individuals. Traditionally, chemical messengers have been classied into
separate systems, mainly the nervous system (i.e., neurotransmitters), endocrine
system (i.e., hormones), and the immune system (i.e., cytokines). However, it has
become increasingly clear that there is substantial overlap and intercellular commu-
nication between these systems. In fact, many neurons express receptors for hor-
mones and cytokines; at the same time, many immune cells express receptors for
neurotransmitters and hormones. With this knowledge in mind, it is easy to imagine
how the sex hormones responsible for sexual differentiation are also capable of
affecting neural development via their actions on multiple cell types and various
chemical messenger systems to differentiate the brain and determine neural function.
4.1 An Immune Molecule and a Neurotransmitter Coordinate
Masculinization of the Preoptic Area
The medial preoptic area (POA) in the rodent brain is a sexually dimorphic cluster of
cells, located in the preoptic nucleus (Gorski 1978). Discussed earlier, the sexually
dimorphic nucleus (SDN), a small cluster of cells of the POA, is ve to seven times
larger in males compared to females and is necessary for the appropriate expression
of appetitive and consummatory male sex behavior (De Jonge et al. 1989;
Houtsmuller et al. 1994; Jarzab et al. 1990; Jeong et al. 2008; Todd et al. 2005).
In addition to this macroscopic, structural difference in the size of the SDN-POA, the
neurons throughout the POA of males also have two to three times more dendritic
spines than females. Within the rst 2 weeks of postnatal life, male rats have a higher
density of spine synapses along the dendrite on neurons in the POA compared to
female rats. Moreover, astrocytes in the male POA have longer processes with a
greater amount of branching than females (Amateau and McCarthy 2002a,b).
Masculinization of the POA is dependent on testosterone, which is converted to
estradiol in the rodent brain during the critical period of sexual differentiation. While
it is clear that naturally occurring apoptosis plays a critical role in establishing the
SDN-POA, another question remained how does estradiol induce these structural
and synaptic changes to the neurons of the male POA? The answer, which was
surprising at the time, is that these changes are induced through increases in the
synthesis of prostaglandin E2 (PGE2), a pro-inammatory signaling molecule.
PGE2 is derived from arachidonic acid following cyclinization by the cyclooxy-
genase enzymes COX-1 and COX-2, and PGE2 is best known for its role in fever
production and inammation. Amateau and McCarthy (2004) determined that acti-
vation of estrogen receptors in the POA upregulates the production of COX-1 and
Sexual Differentiation and Sex Differences in Neural Development
COX-2 leading to a nearly sevenfold increase in PGE2. Notably, PGE2 has been
shown to be synthesized by other cells, such as microglia and astrocytes, within the
POA via activation of glial adenosine receptors by ATP (Fiebich et al. 2014);
however, Amateau and McCarthy (2004) showed that the increase in PGE2 synthe-
sis in the POA was specic to neurons. Neuronal PGE2 then induces the growth of
neuronal dendritic spines via the release of glutamate and activation of AMPA
receptors (Amateau and McCarthy 2004). Astrocytes are also responsive to estra-
diol; they release glutamate following stimulation with PGE2, and their morphology
is sexually dimorphic in the POA, as described above (Amateau and McCarthy
2002a; Bezzi et al. 1998). Thus, astrocytes play a critical role in the establishment of
sexually dimorphic synaptic connectivity within the POA (Fig. 3). Treating females
with estradiol or PGE2 produces male-like patterns of dendritic spines, and blocking
COX-1 or COX-2 in males leads to female-like patterns of dendritic spines in the
POA that are maintained into adulthood (Amateau and McCarthy 2002a,2004;
Wright et al. 2008). Furthermore, PGE2 receptors 2 and 4 (EP2 and EP4) have
been shown to be the primary contributors to the masculinizing effects of PGE2 in
the POA (Wright et al. 2008). Specically, activation of EP2 or EP4 leads to
increases in protein kinase A (PKA), and if PKA signaling is disrupted during
neonatal development, the masculinization of the POA is blocked (Wright and
McCarthy, unpublished observation). Taken together, these data highlight a unique
multicellular signaling process through which sex steroid hormones can sexually
differentiate the developing brain.
4.2 Sexual Differentiation Across the Synapse
in the Ventromedial Nucleus of the Hypothalamus
A central region of the hypothalamus, the ventromedial nucleus (VMN), is a brain
region necessary for female sex behavior (Pfaff and Sakuma 1979). The VMN is
characterized by its oval shape and its sparse, thin dendrites (Millhouse 1973).
Similar to the POA, the male VMN neurons contain more than three times as
many dendritic spines and shaft synapses as those measured on female VMN
neurons (Matsumoto and Arai 1983). This sex difference in synaptic connectivity
is detected as early as postnatal day 2 and is maintained in the rodent brain until
postnatal day 100 (Matsumoto and Arai 1986; Pozzo Miller and Aoki 1991),
consistent with the long-term organizational effects of estradiol in the developing
brain. This masculinized pattern of dendritic spine morphology in the VMN can be
induced in females by treatment with either testosterone or its metabolite, estradiol,
within the rst few days of life. In adulthood, the female estrous cycle or exogenous
estradiol treatment causes changes in the dendritic spine density and patterning of the
VMN, which correlate with changes in female sex receptivity (Frankfurt et al. 1990),
indicating that changes in the synaptic connectivity of the VMN are a critical mech-
anism of inducing or constraining female sex behavior.
A. Turano et al.
Glutamate is the primary neurotransmitter in the VMN (Ziegler et al. 2002), and
estradiol mediates the changes in neuronal morphology in the developing VMN by
directly enhancing glutamate release from these developing neurons (Schwarz et al.
2008). During the critical period of sexual differentiation, estradiol exposure induces
the activation of PI3 kinase within only 3 h, which leads to a cascade of events that
can induce the defeminization of the synaptic connections in the VMN and the
defeminization of sexual behavior in adulthood. Specically, activation of PI3
kinase in the developing VMN increases the release of glutamate from presynaptic
terminals within this region. This, in turn, activates glutamatergic receptors, activat-
ing MAP kinase and masculinizing the dendritic spine number measured in the
VMN (Schwarz et al. 2008; Fig. 3). Blocking any one of these steps, including PI3
kinase activation, MAP kinase activation, or AMPA/NMDA glutamate receptor
activation, completely prevents the increase in dendritic spine proteins seen follow-
ing estradiol treatment of the female rat pup and ultimately prevents the defemini-
zation of the brain and behavior later in life. Specically, blocking NMDA receptors
during the critical period of sexual differentiation blocks estradiol-induced defemi-
nization of behavior, which results in the production of an adult rat that is capable of
expressing both male and female sex behavior (Schwarz and McCarthy 2008). These
ndings describing sexual differentiation of the VMN indicate that the effect of the
sex hormone estradiol is not limited to the cells that express the estrogen receptor but
rather that estradiol-induced masculinization in the VMN initiates the coordination
of synaptic connections and entire circuits that control male sex behavior. As stated
above, the fact that estradiol activates PI3 kinase in the developing VMN within 3 h
after estradiol exposure indicates that the effects of estradiol on modulating neuro-
transmitter function to initiate the process of sexual differentiation within the VMN
occur in a relatively rapid time course. The relatively rapid time course of estradiols
effects suggests that the developing circuitry is extremely sensitive to the organiza-
tional effects of estradiol exposure during the critical period of sexual differentiation,
and this is likely the case for other brain regions that undergo sexual differentiation
during this brief and critical developmental window.
4.3 Development of Neuropeptides Involved in Sex-Specic
Behaviors
Vasopressin and oxytocin are neuropeptides involved in the regulation of many
sexually dimorphic social behaviors including social recognition, pair bonding, and
social cognition in various species (Ross and Young 2009; Veenema and Neumann
2008). These sex differences in behavior are the result of sexually dimorphic
innervation of these two neuropeptides within specic brain regions, which are
established during brain development. The vasopressin system was rst shown to
be sexually dimorphic in 1981 by de Vries and colleagues, who discovered that
males have more vasopressin-immunoreactive bers in the lateral septum (LS) and
Sexual Differentiation and Sex Differences in Neural Development
lateral habenula than females and approximately 90% of vasopressin cells in the
BNST (see Sect. 3.3) express androgen receptors, the absolute number of which is
signicantly greater in males compared to females (Bales et al. 2007; de Vries et al.
1981). Previous work has shown that gonadectomy of neonatal male rats signi-
cantly reduces vasopressin ber density later in life, indicating that androgens are
necessary for the organization of the sexually dimorphic vasopressin system in the
BNST (De Vries et al. 1983). Interestingly, administration of testosterone to gonad-
ectomized males or intact females during the rst, second, or even third postnatal
week can induce masculinization of the vasopressin system. These ndings suggest
that not only are androgens necessary for the organization of this particular neuro-
peptide system in males but that vasopressin neurons in the female brain are also
sensitive to the effects of androgens. Gonadectomy of adult males and ovariectomy
of adult females cause vasopressin immunostaining in the BNST to disappear almost
completely within several weeks post-surgery. This nding suggests that vasopres-
sin expression in the BNST is responsive to circulating steroid hormones even into
adulthood and indicates that males and females may respond differently to testo-
sterone (or its metabolites) in order to maintain the sexually dimorphic features of the
vasopressin system in adulthood. In fact, later studies investigating this possibility
found that when adult male and female rats were gonadectomized or ovariecto-
mized, respectively, and then subsequently administered equal doses of testosterone
(i.e., via a subcutaneous implant), the sexual dimorphism in the number of vaso-
pressin cells was maintained in the BNST, supporting the hypothesis that males and
females show different sensitivities to either testosterone or the metabolites of
testosterone (De Vries et al. 1994). It is likely that this sex difference in sensitivity
to androgens in the BNST is established early in life by perinatal testosterone
exposure. Indeed, the testosterone metabolites estradiol and 5α-dihydrotestosterone
(DHT) can affect vasopressin cell number in a sex-dependent manner, such that
treatment with estradiol increases the number of vasopressin cells in the BNST in
males and females compared to controls and DHT-treated animals; however, this
effect was greater in males than females (De Vries et al. 1994). Additionally, males
treated with estradiol and DHT combined had signicantly more vasopressin cells in
the BNST than when treated with estradiol alone, while this same effect was not true
for females. These data reiterate the fact that, in addition to the circulating hormones
of the neuroendocrine system, every cell has a sex, which means that male and
female cells have the potential to respond differently to the same stimulus, thus
impacting the brain and behavior in a sexually dimorphic manner.
The nding that the vasopressin system is more prominent in males compared to
females suggests that males, more than females, may rely on this neuropeptide for
the regulation of social behavior. In support of this notion, injections of vasopressin
enhanced social recognition in both males and females, but injection of a vasopressin
receptor antagonist prevented social recognition in males only (Bluthe et al. 1994;
Dantzer and Bluthe 1993). Additionally, vasopressin mRNA expression in the
paraventricular nucleus of the hypothalamus (PVN) was determined to be positively
correlated with the amount of social play behavior in juvenile males, but not females
(Paul et al. 2014). Indeed, vasopressin has been shown to regulate some behaviors
A. Turano et al.
differently in males and females. For example, vasopressin administered to males
stimulates aggressive behavior, while this same administration of vasopressin
inhibited aggressive behavior in females (Terranova et al. 2016). Furthermore,
dominance in males (i.e., more aggressive males) is strongly associated with acti-
vation of vasopressin cell bodies in the supraoptic nucleus of the hypothalamus
compared to subordinate or control males; however, in females, activation of
vasopressin cell bodies was observed in dominant and subordinate females to the
same degree, which were both higher than that observed in control females
(Terranova et al. 2016). These ndings indicate that the vasopressin system exhibits
a fundamental sex difference in the regulation of aggressive behavior, whereby the
male vasopressin system shows a strong correlation to dominance status, but the
female vasopressin system is associated with social interaction more generally.
While the exact cellular mechanisms underlying the sexually dimorphic innervation
of the vasopressin system established by early androgen exposure are not known, it
is clear that this neuropeptide is integral to the expression of a myriad of sexually
dimorphic behaviors expressed later in life.
In contrast to vasopressin, oxytocin-synthesizing cells are greater in number in
females compared to males across a number of brain regions including the PVN,
lateral hypothalamus, lateral septum (LS), POA, and the BNST in several species.
Furthermore, the number of oxytocin-immunoreactive bers in the LS and BNST is
signicantly greater in female compared to male mice (Haussler et al. 1990) and in
the lateral hypothalamus of mandarin voles (Qiao et al. 2014); however, others have
found no sex differences in oxytocin mRNA expression (Dumais et al. 2013) nor in
the number of oxytocin neurons in humans (Wierda et al. 1991). It has been shown,
however, that administration of oxytocin or the oxytocin receptor antagonist can
increase oxytocin immunoreactivity in the PVN of female prairie voles, but not male
prairie voles (Yamamoto et al. 2004). Furthermore, when given neonatally, oxytocin
increased ERαimmunoreactivity in the ventromedial hypothalamus, and oxytocin
receptor antagonist decreased ERαimmunoreactivity in the POA of females, but not
males (Yamamoto et al. 2006). Unfortunately, there has been surprisingly little
investigation into the development of the oxytocin system in males and females,
so there is still much that remains to be elucidated regarding the neuroendocrine
regulation of oxytocin in the brain. It is also important to note that while less is
known about the oxytocin system in the brain, there is an abundance of evidence
demonstrating the sex-specic effects of oxytocin on behavior (for a review see
Caldwell 2017). For example, neonatal oxytocin exposure facilitates mate-guarding
behavior (a component of pair bonding that increases aggression and reduces social
behavior toward other voles) in female, but not male prairie voles (Bales and Carter
2003), and ICV administration of oxytocin was shown to facilitate partner preference
formation in females, but not males (Insel and Hulihan 1995; Williams et al. 1994;
Winslow et al. 1993).
Despite the evidence indicating that oxytocin may play a more prominent role in
females, Yao et al. (2017) found that oxytocin can impact certain social behaviors in
adult male, but not female, mice. Adult male mice that lack oxytocin signaling did
not prefer a female over a male social partner in a 3-chamber-social-investigation
Sexual Differentiation and Sex Differences in Neural Development
paradigm, as is typically observed in wild-type male mice. In contrast, no changes in
partner preference were observed in adult female mice with altered oxytocin signal-
ing. Also in contrast with the data described above, vasopressin did not impact
partner preference in either adult male or female mice within this specic paradigm.
Further examination indicated that social recognition in adult male mice appears to
be exclusively dependent upon oxytocin signaling in the medial amygdala, as
endogenous vasopressin was unsuccessful in deterring the negative effects of altered
oxytocin signaling on male partner preference. Notably, the authors indicated that
this male-specic effect of oxytocin on social partner preference is likely the result of
oxytocin rapidly modulating the sensory responses of aromatase neurons to social
cues; therefore, the basis for these contrasting results likely stems from the fact that
while oxytocin-synthesizing cells are greater in number in females compared to
males in the PVN, lateral hypothalamus, LS, POA, and the BNST, there are a greater
number of aromatase-expressing neurons in the medial amygdala of male mice,
compared to female mice (Wu et al. 2009; Yao et al. 2017). Importantly, these results
highlight the idea that sex differences in the development of neural circuits may
occur via neuropeptides (hormones or neurotransmitter systems) that are not directly
inuenced by the traditional testosterone-mediated mechanisms of sexual
differentiation.
4.4 Microglia Induce Sex-Specic Neuronal Development
Beyond the traditionalneural cells and sexually dimorphic circuits that have been
investigated for many years, the past decade or so has also uncovered sex differences
in the immune cells of the brain. For example, microglia, the resident immune cells
of the brain, are critical for many normal neurodevelopmental processes including
supporting neurogenesis, pruning spurious synaptic connections, and phagocytosing
apoptotic cells (Boulanger 2009; Deverman and Patterson 2009). During early
development, males have more microglia than females in certain brain regions,
including the hippocampus, the parietal cortex, and the POA (Lenz and McCarthy
2014; Schwarz et al. 2012). Interestingly, this sexually dimorphic prole of
microglia shifts dramatically throughout development, with females having more
microglia with thick processes or stout phenotype than males as adults (Schwarz
et al. 2012). Given the striking sex difference in microglia number across multiple
brain regions, and given their importance in neural development, investigators have
sought to determine the developmental prole of microglia (Hanamsagar et al. 2017)
and whether microglia have an active role in the sexual differentiation of the
developing brain. A recent study found that during hormone-induced sexual differ-
entiation of the brain, microglia in the female hippocampus have signicantly more
phagocytic cups than those in the male hippocampus (Nelson et al. 2017). When
females are treated with estradiol, the number of phagocytic cups is reduced to male
numbers (Nelson et al. 2017). In the POA, males have twice as many amoeboid
microglia as females, and estradiol treatment of female rat pups masculinizes
A. Turano et al.
microglia number and morphology in the POA (Lenz et al. 2013). Furthermore,
inhibiting microglia prevents the normal estradiol-induced masculinization of den-
dritic spines in the POA and adult copulatory behavior in males, indicating that
microglia are also an important cellular mediator of the process of brain sexual
differentiation (Lenz et al. 2013).
As the immune cells of the brain, microglia also respond to immune activation or
injury with rapid changes in morphology and the release of cytokines, processes
critical to maintaining homeostasis in the brain. Emerging literature in humans
suggests that early-life immune dysregulation leads to cognitive and behavioral
disorders that persist throughout the lifespan (Bilbo and Schwarz 2009; Frick et al.
2013; Maezawa et al. 2011; Sheridan et al. 2014) and many of these disorders show a
distinct male bias (e.g., autism spectrum disorders [ASD], ADHD, learning disabil-
ities, etc.; Schwarz and Bilbo 2012). Animal models investigating the effects of
early-life immune activation have found several sex-specicdecits in behavior. For
example, prenatal exposure to lipopolysaccharide (LPS; which activates microglia
via TLR-4 receptors) produces decits in social play behavior in juvenile males, but
not females (Taylor et al. 2012). Prenatal exposure to stress, which activates the
neuroimmune system, induces behavioral impairments including hyperactivity,
behavioral despair, and anhedonia in males, but not females (Bronson and Bale
2014; Mueller and Bale 2008). Collectively, the ndings presented here highlight the
intricate communication that must occur between many cell types in the endocrine
system, nervous system, and immune system to allow for the development of a
sexually dimorphic brain and sex-specic expression of behavior.
4.5 Sex Differences in Cell Genesis and Fate During
the Development of Cognitive Circuits
Many of the brain regions described thus far are essential for controlling a variety of
sexually dimorphic sexual and social behaviors. There are many other brain regions
whose functions are very important, and yet no overt sex differences in these
associated behaviors have been found. For example, the hippocampus and the
amygdala are both important for the formation of memory and cognition, yet there
are no overt sex differences in the basic function of these circuits. That is, both males
and females are able to perform complex cognitive tasks and remember things to a
similar degree. That said, there are sex differences in the development of the
hippocampus and amygdala that have been identied and that are worth noting in
this chapter, mainly because of their potential role in the etiology or the prevalence
of neurodevelopmental disorders that are associated with decits or alterations in
learning, cognition, or emotion and that also display strong sex biases.
During postnatal brain development, the male hippocampus is only slightly larger
than the female hippocampus, even after controlling for sex differences in body or
head size (McCarthy and Konkle 2005), and this is thought to be the result of more
Sexual Differentiation and Sex Differences in Neural Development
neurons and glia in the male compared to the female hippocampus (Conejo et al.
2003; Hilton et al. 2003; Nunez et al. 2003). Consistent with this idea, there is a
signicant sex difference in the number of newly born cells in the developing rat
hippocampus, with the neonatal male rostral CA1 and DG having more proliferating
cells than the numbers observed in females. Both androgen and estrogen treatment
increased the number of proliferating cells in females to the numbers seen in
males underscoring, once again, the sexually dimorphic role of testosterone in
neurodevelopment (Zhang et al. 2008). These data highlight the fact that, even
though the size of the hippocampus is only slightly larger in males than females,
there is a striking and signicant sex difference in the proliferation of newly born
cells during a time when cell genesis, in particular neurogenesis, in the developing
brain and hippocampus is most robust. Notably, this sex difference in cell prolifer-
ation is an effect that is driven by prenatal androgen exposure (Fig. 3), and more
importantly, it is an effect that may have long-lasting consequences for the function
of this structure throughout development.
An adjacent brain region in the temporal lobe is the amygdala, which has an
important role in the processing of emotional stimuli and learning. The amygdala
also inuences a number of sexually dimorphic behaviors including juvenile rough
and tumble play (which is greater in males and inuenced by prenatal testosterone
exposure), and it is also involved in mating, parenting, aggression, and territoriality,
all of which are sexually dimorphic. The overall size of the medial amygdala is larger
in males than in females (similar to the hippocampus; Cooke 2006); however, in
contrast to the hippocampus, the female amygdala has signicantly more cell
proliferation than males during early postnatal brain development. Interestingly,
these cells differentiate into astrocytes as opposed to neurons (Krebs-Kraft et al.
2010; Fig. 3). This sex difference is the result of differential endogenous cannabi-
noid signaling in males and females that inuences the likelihood of ongoing cell
proliferation and ultimately the fate of these cells as astrocytes.
To date, the question remains, what is the role of these sex differences in cell
number in both the hippocampus and the amygdala? It is likely that they have an
important role in generating sex differences in social and perhaps also cognitive
behaviors; however, that remains to be fully determined. Importantly, there are times
when a sex difference occurs in a brain region in order to eliminate sex differences in
behavior and in order to compensate for a sex difference in another coordinating
brain region (De Vries 2004). This may be particularly important for neural func-
tions or behaviors that are essential, like cognition (De Vries 2004), and may be one
explanation for the sex differences seen here, particularly in the hippocampus.
Perhaps more importantly, researchers should consider how these sex differences
in neural development may confer risk for decits in social or cognitive function,
particularly following a trauma or a challenge that perturbs cell genesis during early
brain development, an effect that may persist throughout the lifespan. In this case,
one sex may be more vulnerable to the trauma or challenge, resulting in a sex bias in
cognitive, social, or emotional disorders.
A. Turano et al.
5 A Whole Body Perspective on Sex Differences in Neural
Development
In accordance with the idea that every cell has a sex, there is accumulating evidence
that all organs are sexually differentiated and that even peripheral organs may affect
brain development (for a recent review, see de Vries and Forger 2015). The classic
example of peripheral organs affecting the brain is the basic process of sexual
differentiation of the brain via hormones released from the gonads, which act on
multiple cell types and receptors in the brain during development, as we described
above. The idea that sex differences in the body and brain are an important factor
affecting the health of an organism has been somewhat slow to be accepted, but
ignoring this could have detrimental effects for understanding disparate health out-
comes in men and women. In fact, there are well-known sex differences in vulner-
ability to disease, and in some cases sex is, perhaps, the most predictive factor
contributing to the prevalence of a disorder. Moreover, many of these disorders have
their origins in early development; thus an understanding of how other organs may
interact with ongoing processes that affect sexual differentiation of the brain is key to
elucidating mechanisms governing the neuroendocrine regulation of neural devel-
opment and behavior. Here we will present just a few examples of the sex differences
in peripheral organs and how they may affect brain function (Fig. 4).
Fig. 4 Examples of how peripheral organs can inuence sex-specic development of the brain. Sex
differences in various peripheral tissues and organs, including the gut microbiome, the placenta, and
the liver, can produce sex-specicinuences on the brain that consequently inuence neural
development in a sex-specic manner
Sexual Differentiation and Sex Differences in Neural Development
5.1 The Gut Microbiome
Communication between the gut microbiome and the brain is now recognized as an
integral factor that impacts brain development and behavior. Following bacterial
inoculation of the infant with the maternal vaginal microbiome at birth, communi-
cation between these ourishing gut microbes and the brain can function to integrate
various signals from the body (e.g., autonomic, neuroendocrine, etc.) that in turn
impact the brain. Importantly, the microbiome is highly dynamic, and alterations in
its composition occur throughout development, particularly during critical times of
sexual differentiation of the brain (Borre et al. 2014). Indeed, sex differences in the
gut microbiome emerge during puberty and are actively maintained into adulthood
(Yurkovetskiy et al. 2013). Specically, the female microbiome stays relatively
similar to that of the prepubertal male and female microbiome, whereas the male
microbiome begins to diverge during puberty and exhibits a distinctive phenotype by
adulthood (Yurkovetskiy et al. 2013). Interestingly, if the surge in testosterone
during puberty is prevented via gonadectomy in males, this masculinization of the
gut microbiota is eliminated in adulthood. Furthermore, transfer of the microbiome
of an adult male to a pubertal female completely masculinizes the microbiota
composition, metabolomic prole, and produces an increase in testosterone levels
that persist into adulthood suggesting that the differentiation of the microbiome
during development subsequently impacts neural and physiological function
(Markle et al. 2013). If, however, the antiandrogen utamide is administered with
the microbiome transfer, all of these changes are prevented in females demonstrating
the important, mechanistic role of testosterone in producing this microbiome-
induced masculinization (Fig. 4). In a well-established mouse model of autism
spectrum disorder (i.e., BTBR inbred mice), it was found that, while males and
females have signicantly altered microbiome composition compared to control
mice, these alterations are sex-specic (Coretti et al. 2017). These mice also display
sex differences in their inammatory prole in colon tissue, with males showing
signicantly higher levels of IL-6 and CD11c compared to females and control mice.
In another model of ASD, male mice, but not female mice, that are treated with
valproic acid on gestational day 11 show reduced social interactions and increased
gut inammation (de Theije et al. 2014a). Additionally, valproic acid signicantly
alters the composition of the gut microbiota, and the levels of social interaction
correlate with metabolites of the microbiota (de Theije et al. 2014b). These data
suggest that the gut microbiome is capable of impacting the developing brain and are
quite intriguing given the increased prevalence of ASD in males and autisms
associated risk with early-life inammation.
5.2 The Liver
Approximately 72% of genes that have been examined in the liver show sex
differences in their expression. Given the important role of the liver and many of
A. Turano et al.
these genes in drug metabolism, this striking sex difference can in turn inuence the
amount of drug that is delivered to the brain in a sex-specic manner (Yang et al.
2006). In fact, the best-known example of sex differences in liver function is drug
metabolism, which leads to differences in pharmacodynamics and pharmacokinetics
between males and females (Waxman and Holloway 2009). Sex differences in liver
metabolism can have profound effects on brain function in adults. For example, the
prescription drug zolpidem, which acts as a sleep aid by binding to GABA
A
receptors in the brain, was found to have a signicantly lower clearance rate in
women than men (Greenblatt et al. 2014), making it necessary to decrease the dosage
by 50% for women. Despite the fact that 67% of drugs that have examined sex as a
variable have shown a large difference in pharmacokinetics between men and
women, there is still very little known about how sex differences in drug metabolism
by the liver inuence how drugs affect male and female brains. The liver secretes
steroid-binding proteins and enzymes that metabolize circulating gonadal hormones,
which inuence the available levels of these hormones in the periphery of males and
females. Exactly how these liver enzymes alter the amount of gonadal hormones that
reach the brain to affect brain development in males and females is not yet known.
The maternal liver also synthesizes proteins that are critical for the development of
the fetus (Roy and Chatterjee 1983); however, it is not clear exactly how these
proteins affect long-term brain development of the fetus.
Furthermore, sexually dimorphic gene expression in the liver is regulated by
growth hormone-dependent transcription factors (Legraverend et al. 1992), which is
the result of a signicant sex difference in the expression of growth hormone itself.
Therefore, not only can sexually dimorphic genes of the liver differentially impact
brain function, but the brain (via sex differences in growth hormone expression) may
also impact the sexually dimorphic expression of genes in the liver, thereby creating
an interesting brain affects liver, liver affects brainmechanism. Still, sex differ-
ences in liver gene expression are also, in part, regulated by growth hormone-
independent transcription factors, suggesting that the sex differences in liver gene
expression extend beyond sex differences in growth hormone expression alone
(Yang et al. 2006). Clearly, more work is necessary to provide a complete under-
standing of the intricate processes by which peripheral tissues can impact sex
differences in the brain.
5.3 The Placenta
Beyond the developing fetus, the placenta is comprised of both maternal and fetal
cells. As a result, the placenta also has a sex (i.e., that of the developing fetus) that
has been shown to inuence the relative risk or resilience of the fetus to maternal
insults in a sex-specic manner (Clifton 2010; Gabory et al. 2013). For example,
when pregnant rats are exposed to a stressor during early gestation, male but not
female placentas show increases in IL-6 and IL-1βgene expression. Importantly,
following this stress, male but not female offspring display anhedonia, adverse stress
Sexual Differentiation and Sex Differences in Neural Development
responsivity, and altered responses to selective serotonin reuptake inhibitors as
adults (Mueller and Bale 2008), and blocking the action of these cytokines in the
male placenta blunts the stress phenotype in adulthood, suggesting that the changes
in immune gene expression in the placenta are the mechanistic source of these
changes in the developing brain. The O-glycotransferase enzyme, O-GlcNAc trans-
ferase (OGT), is important for regulating gene expression through chromatin
remodeling in the placenta and is naturally lower in the male placenta compared to
the female placenta. Howerton and Bale (2014) showed that key features of the male
offspring that were subjected to early prenatal stress were recapitulated in males that
had genetically reduced levels of OGT in the placenta including HPA axis
dysregulation, reduced postpubertal growth, and hypothalamic mitochondrial dys-
function. Furthermore, females in this model, instead, had a blunted HPA axis
phenotype indicating sex-specic programming of placental OGT and supporting
the possible involvement of placental X- and Y-linked genes in determining stress
axis programming in adulthood (Howerton and Bale 2014; Fig. 4).
These are just a few known examples by which peripheral tissues can inuence
the development and function of the brain in a sex-specic manner. Given that there
are many sex differences in the function of tissues and organs throughout the body,
an important area of research is to better understand how these tissues and organs
may also impact sex differences in the development of the brain, during normal brain
development, and also following early-life challenges (e.g., stress, infection, expo-
sure to drugs of abuse) or during disease states that affect the function of these
organs.
6 Epigenetics and the Sexual Differentiation of the Brain
The mechanisms described thus far that contribute to sexual differentiation of the
developing brain involve both the direct actions of sex hormones, as well as the
direct actions of the different genes located on the sex chromosomes that are inherent
in each cell (Arnold 2009; Gorski et al. 1978). Importantly, gene expression plays a
critical role in both of these interdependent mechanisms of brain organization
throughout early development. Therefore, it is perhaps not surprising that changes
in gene expression, induced via epigenetic modications, also contribute to sexual
differentiation of the developing brain.
Epigenetics refers to the mechanisms that change gene expression via the proteins
or enzymes that moderate gene expression mechanisms including modications to
DNA methylation, histones, or DNA-binding proteins. These modications take
place without causing signicant changes to the core DNA sequence (Menger et al.
2010; Nugent and McCarthy 2011), but rather these changes determine whether and
under what circumstances genes are transcribed or silenced. Epigenetic changes can
be induced as a result of the interactions between an organisms genome and changes
in the organisms internal or external environment. Notably, gonadal hormones,
which are key regulators of our internal environment, can signicantly impact
A. Turano et al.
epigenetic modications to the DNA (Auger and Auger 2011; Nugent et al. 2015).
Therefore, given the sexually dimorphic nature of gonadal hormone secretion,
epigenetic processes are likely contributors in the process of sexual differentiation,
thereby determining in which sex and under what circumstances certain genes are
expressed in the body and the brain. While we have a strong understanding of how
sex differences in the brain are established, our understanding of how these sex
differences in the brain are maintained throughout the lifespan requires further
examination (McCarthy et al. 2009). Fortunately, epigenetic modications of the
DNA, which are capable of inducing lifelong and even transgenerational changes in
gene expression (Roth et al. 2010), can and have been examined to answer this
exciting question of how sex differences may be established and maintained in the
developing brain.
We previously discussed the sexually dimorphic nature of the bed nucleus of the
stria terminalis (BNST), wherein the number of cells and the overall volume of the
primary nucleus of this brain region are signicantly greater in the male compared to
the female brain. This sex difference in the size of the BNST is a result of the
neuroprotective effects of testosterone in the developing male brain, which prevents
naturally occurring cell death in this brain region (Guillamon et al. 1988); however,
the mechanism by which testosterone protects neurons from this naturally occurring
apoptosis was previously not well understood. To address this gap in the research,
Murray et al. (2009) examined the role of epigenetic changes in the brain, speci-
cally by manipulating histone acetylation. Histone acetylation, via histone
acetyltransferases (HATs), is the process by which an acetyl group is added to the
lysine residues at the N-terminus of a histone protein, essentially looseningthe
DNA from a tightly wrapped spool. Loosening the DNA from the histone increases
the opportunity for gene transcription. Conversely, histone deacetylases (HDACs)
decrease the opportunity for gene transcription by removing that same acetyl group
from the histone protein, keeping the DNA tightly wrapped around the histone.
HDAC activity can be blocked by the administration of HDAC inhibitors, like
valproic acid (VPA), thus resulting in increased histone acetylation and increased
gene transcription. When Murray et al. (2009) treated newborn male mice with VPA,
not only was histone acetylation increased throughout the brain, but the
neuroprotective effect of testosterone on cell number in the BNST was prevented.
In other words, the VPA-treated males had reduced BNST volumes. This effect was
also observed in newborn female mice that had been treated with testosterone and
thus, prior to VPA treatment, had BNST volumes equivalent to developing males
(Murray et al. 2009). Therefore, the masculinization of the BNST, via neonatal
testosterone exposure, also involves histone deacetylation, or decreased gene expres-
sion, within this brain region that persists even after the initial testosterone exposure.
These ndings highlight the idea that androgens can disrupt an ongoing
neurodevelopmental process, in this case apoptosis, by disrupting the expression
of genes and thereby inducing a sex difference in the brain. Importantly, when these
genes are turned back on by HDAC inhibitors, like VPA, the sex difference can be
reversed (Fig. 5). Still, the question remains of whether or not this sex difference can
also be modulated outside the critical period of sexual differentiation. One might
Sexual Differentiation and Sex Differences in Neural Development
Fig. 5 Epigenetic mechanisms by which sex differences in the brain are established during neural development. (1) In males, testosterone converted to estradiol
can activate histone deacetylase (HDAC) resulting in the deacetylation of histones in the bed nucleus of the stria terminalis (BNST), which represses gene
expression necessary for programmed cell death in this brain region. (2) In females, in the absence of testosterone exposure, acetylation of histones increases
gene expression necessary for the programmed cell death in the female BNST. (3a) In males, testosterone converted to estradiol inhibits DNA methyltransferase
(Dnmt) in the preoptic area (POA) which results in the removal of methyl groups from the cytosines (C) of the DNA. (3b) In the absence of methylation in the
POA, genes necessary for masculinization of the brain and sexual behavior are expressed. (4a) In females, Dnmt is constitutively active in the POA, in the
absence of testosterone. (4b) Methylation of the DNA is maintained by Dnmt activity, which results in the repression of genes necessary for masculinization of
the POA and behavior
A. Turano et al.
assume that apoptosis in the BNST that occurs during the period of neuro-
development occurs exclusively during this critical period. However, it is also
possible that testosterone permanently alters the likelihood that the genes necessary
for programed cell death in the BNST are expressed, making it possible for gene
transcription to be turned on outside of the critical period of sexual differentiation.
If so, could cell death again be induced, thereby reversing the sex difference in the
size of this nucleus?
This fascinating question may be answered in part by a more recent study that
examined the epigenetic mechanisms by which sex differences in the structure and
function of the POA are established and maintained. As mentioned previously, the
POA exhibits a striking sex difference in the structure and function of its neurons
that is established as a result of perinatal testosterone exposure and is necessary for
the maintenance and expression of sex differences in sexual behavior later in life
(Amateau and McCarthy 2004). A study by Nugent et al. (2015) found that andro-
gens, converted to estradiol in the rodent brain, induce this sex difference in the POA
via the inhibition of DNA methyltransferase (Dnmt) activity. Dnmt is an enzyme that
methylates DNA, which is thought to silence gene transcription. Thus, androgen
exposure during the critical period of sexual differentiation induces the expression of
certain genes that are necessary to masculinize the POA by inhibiting the process of
DNA methylation. This effect may be maintained as the developing POA matures,
particularly given that others have shown that testosterone (or estradiol) can induce a
global decrease in DNA methylation that is maintained as cells multiply and mature
(Bramble et al. 2016). Furthermore, Nugent et al. (2015) found that when Dnmt3
activity was conditionally knocked out later in life (after the critical period of sexual
differentiation), in the absence of Dnmt3 and decreases in DNA methylation in the
POA, genes necessary for masculinization of the POA could be turned on, resulting
in masculinization of the brain and displays of male sex behavior. These data
highlight a few important points: (1) in the female brain, methylation represses the
genes necessary to induce masculinization of the POA; (2) during the critical period
of sexual differentiation, testosterone (via estradiol) induces the transcription of
genes that are actively epigenetically repressed by Dnmts in the female brain; and
(3) masculinization can be induced later in life if sufcient gene transcription is
re-induced in the POA, via the knockout of Dnmt3, which was actively maintaining
DNA methylation in the female brain. Furthermore, these data suggest that during
the critical period of sexual differentiation, Dnmt activity is particularly sensitive to
androgen exposure, but outside of this critical period, Dnmt activity is not as
responsive to androgen exposure and thus, estrogen exposure as well. Therefore,
one might conclude that the ability of Dnmt to be inhibited by testosterone (via
estrogen) may be one mechanism by which the critical period of sexual differenti-
ation closes,and the fate of the male and female brain is nalized.
While the relationship between neonatal testosterone and histone acetylation or
DNA methylation within the BNST and POA, respectively, has proven to be fairly
straightforward, the same cannot be said for other epigenetic changes that may
contribute to the sexual differentiation of the brain. For example, contradictory
evidence has emerged for how gonadal hormone exposure and maternal behavior
Sexual Differentiation and Sex Differences in Neural Development
may impact DNA methylation of estrogen receptor alpha (ERα) in the rodent POA.
ERαexpression in the rodent POA is critical to the organization of typical sex
behavior in both male and female rats. In combination with estradiol, which is
converted from testosterone by aromatase, ERαinitiates defeminization and mascu-
linization of the POA (Kudwa et al. 2006; McCarthy et al. 2008). Interestingly,
maternal behavior can impact the methylation of the ERαgene in the neonatal pup
(Cameron et al. 2008), thereby impacting masculinization and defeminization of the
brain and leading to male sexual behavior or female maternal behavior later in life.
The impact of maternal behavior on ERαmethylation differed depending on the
whether the behavior was simulated as opposed to naturally occurring (Kurian et al.
2010; Lenz et al. 2013); and moreover, sex differences in the methylation of the ERα
gene were largely dependent upon the developmental time point at which they were
measured (Schwarz et al. 2010). Overall, these divergent results complicate our
understanding of the ways in which DNA methylation and possibly other epigenetic
modications may be involved in the maintenance of sex differences in the brain.
There is still so much to be explored in the realm of epigenetic changes and their
potential impact on the establishment and maintenance of sex differences in the
brain. Importantly, this new eld of research has highlighted that sexual differenti-
ation of the brain requires large, complex networks of gene transcription. It has long
been thought that hormone-induced sexual differentiation of the brain involved the
transcription of genes that exclusively had a hormone response element (HRE).
Rather, we now know that hormones such as androgens (and estrogens metabolized
from testosterone) can modulate epigenetic mechanisms such as histone acetylation
and DNA methylation, which in turn produce a concert of changes in gene expres-
sion all of which are likely necessary for the masculinization of the brain. These
ndings also have the potential to increase our understanding of the mechanisms by
which the brain is feminized. While many studies have used testosterone treatment
of female rodents to induce the process of brain masculinization and examine the
expression of genes induced by testosterone during the critical period of sexual
differentiation (Armoskus et al. 2014a,b; Nakachi et al. 2015), it has been difcult to
identify the genes that are actively involved in the process of brain feminization,
given that there is no known factor that initiates the process at one specic time. Now
that research indicates that feminization of the brain requires active repression of
masculinization via DNA methylation (Nugent et al. 2015), it is possible that we
may nally be able to identify the genes that are explicitly regulated in the process of
brain feminization during this time by exploring the genes that are targeted for
methylation by androgens (Shen et al. 2015). From highly recognized mechanisms
of gene silencing due to the sexually dimorphic process of X inactivation (Deng et al.
2014) to the understudied impact of sex-specic microRNAs (miRNAs) on post-
transcriptional silencing of genes or even steroid hormone-related genes (Cochrane
et al. 2011), the exciting eld of epigenetic research still has a long way to go, but it
may be the key to unlocking a more complete understanding of the processes that
differentiate the male and female brain.
A. Turano et al.
7 Conclusions
Sexual differentiation of the brain is an important neurodevelopmental process that
results in the sex-specic development of neural structures and circuits that control
sexual behavior and associated physiology. Studying the neural mechanisms of
sexual differentiation not only informs our understanding of how these sex differ-
ences in the brain are established but also increases our understanding of how gender
identity and sexual orientation may be inuenced. In addition, our understanding of
sexual differentiation has informed our general understanding of how behavioral
circuits develop during critical periods of life. It has also informed our understanding
of how sex steroid hormones permanently inuence the function of the brain and
how various cell types communicate with each other during early brain development
to shape the sexually dimorphic brain. There are still many aspects of sexual
differentiation that have yet to be determined. In particular, these primary questions
remain: what are the mechanisms underlying brain feminization? What are the
genes, proteins, or cellular processes that actively shape the development of the
female brain? As mentioned earlier, if we could potentially identify genes that are
suppressed (e.g., methylated) or downregulated by testosterone exposure, this might
help to identify the genes that induce brain feminization.
Studies of sexual differentiation of the developing brain have revealed the
mechanisms by which hypothalamic and forebrain structures are differentiated in
order to control sexual and social behaviors. An important future area of research
will be to understand whether there are also sex differences in the development of
other circuits that are important for cognition and emotion. Alternatively, if there are
no sex differences in the development of these circuits, perhaps we can understand
how events that disrupt or perturb development can disrupt the development of
cognitive and emotional circuits in a sex-dependent manner, or disrupt the neuro-
transmitters systems that control emotion. In this way, perhaps we can understand
the sex bias in neurodevelopmental disorders such as autism, schizophrenia, and
generalized or pervasive developmental disorders that affect the development and
cognitive functioning of young boys more than girls.
Along those lines, we have yet to understand the full ontogeny of sex differences
in the brain from prenatal to postnatal stages of life and throughout pubertal and
adolescent development. Given that sexual differentiation of the brain permanently
organizes neural development in one direction (male) or the other (female), sexual
differentiation of the brain can thus determine the fate and function of cells through-
out each of these stages of life. To that end, we have yet to understand how the cells
maintain a memory of their sex, particularly as it was inuenced by sex hormones
during early brain development. This memoryof the cellshormonal sex
continuously inuences the function of the cells and the expression of genes
throughout the lifespan.
In conclusion, as we understand more about the processes underlying sexual
differentiation, we can inform our general understanding of neural development and
the formation of functional neural circuits. A secondary and important goal of
Sexual Differentiation and Sex Differences in Neural Development
understanding sex differences in neural development is to understand the possible
mechanisms by which mental health and neurodevelopmental disorders arise,
particularly those that have origins in early development and have a well-known
sex bias.
References
Agate RJ, Grisham W, Wade J, Mann S, Wingeld J, Schanen C et al (2003) Neural, not gonadal,
origin of brain sex differences in a gynandromorphic nch. Proc Natl Acad Sci U S A 100(8):
48734878. https://doi.org/10.1073/pnas.0636925100
Amateau SK, McCarthy MM (2002a) A novel mechanism of dendritic spine plasticity involving
estradiol induction of prostaglandin-E2. J Neurosci 22(19):85868596
Amateau SK, McCarthy MM (2002b) Sexual differentiation of astrocyte morphology in the
developing rat preoptic area. J Neuroendocrinol 14(11):904910
Amateau SK, McCarthy MM (2004) Induction of PGE2 by estradiol mediates developmental
masculinization of sex behavior. Nat Neurosci 7(6):643650
Arai Y, Sekine Y, Murakami S (1996) Estrogen and apoptosis in the developing sexually dimorphic
preoptic area in female rats. Neurosci Res 25(4):403407
Armoskus C, Moreira D, Bollinger K, Jimenez O, Taniguchi S, Tsai HW (2014a) Identication of
sexually dimorphic genes in the neonatal mouse cortex and hippocampus. Brain Res 1562:
2338. https://doi.org/10.1016/j.brainres.2014.03.017
Armoskus C, Mota T, Moreira D, Tsai HW (2014b) Effects of prenatal testosterone exposure on
sexually dimorphic gene expression in the neonatal mouse cortex and hippocampus. J Steroids
Horm Sci 5(3):1000139
Arnold AP (2003) The gender of the voice within: the neural origin of sex differences in the brain.
Curr Opin Neurobiol 13(6):759764
Arnold AP (2009) The organizational-activational hypothesis as the foundation for a unied theory
of sexual differentiation of all mammalian tissues. Horm Behav 55(5):570578. https://doi.org/
10.1016/j.yhbeh.2009.03.011
Arnold AP, Chen X (2009) What does the four core genotypesmouse model tell us about sex
differences in the brain and other tissues? Front Neuroendocrinol 30(1):19. https://doi.org/10.
1016/j.yfrne.2008.11.001
Arnold AP, Gorski RA (1984) Gonadal steroid induction of structural sex differences in the
central nervous system. Annu Rev Neurosci 7:413442
Auger AP, Auger CJ (2011) Epigenetic turn ons and turn offs: chromatin reorganization and
brain differentiation. Endocrinology 152(2):349353. https://doi.org/10.1210/en.2010-0793
Bakker J, DeMees C, Douhard Q, Balthazart J, Gabant P, Szpirer J (2006) Alpha-fetoprotein
protects the developing female mouse brain from masculinization and defeminization by
estrogens. Nat Neurosci 9(2):2026
Bales K, Carter CS (2003) Developmental exposure to oxytocin facilitates partner preferences in
male prairie voles (microtus ochrogaster). Behav Neurosci 117:854859
Bales KL, Plotsky PM, Young LJ, Lim MM, Grotte N, Ferrer E, Carter CS (2007) Neonatal oxy-
tocin manipulations have long-lasting, sexually dimorphic effects on vasopressin receptors.
Neuroscience 144(1):3845
Bezzi P, Carmignoto G, Pasti L, Vesce S, Rossi D, Rizzini BL et al (1998) Prostaglandins stimulate
calcium-dependent glutamate release in astrocytes. Nature 391(6664):281285
Bilbo SD, Schwarz JM (2009) Early-life programming of later-life brain and behavior: a critical role
for the immune system. Front Behav Neurosci 3:14. https://doi.org/10.3389/neuro.08.014.2009
A. Turano et al.
Bleier R, Byne W, Siggelkow I (1982) Cytoarchitectonic sexual dimorphisms of the medial preoptic
and anterior hypothalamic areas in guinea pig, rat, hamster, and mouse. J Comp Neurol 212(2):
118130
Bluthe RM, Walter V, Parnet P, Laye S, Lestage J, Verrier D et al (1994) Lipopolysaccharide
induces sickness behaviour in rats by a vagal mediated mechanism. Comptes Rendus De
L'Academie Des Sciences. Serie III Sciences De La Vie 317(6):499503
Borre YE, OKeeffe GW, Clarke G, Stanton C, Dinan TG, Cryan JF (2014) Microbiota and
neurodevelopmental windows: implications for brain disorders. Trends Mol Med 20(9):
509518. https://doi.org/10.1016/j.molmed.2014.05.002
Boulanger LM (2009) Immune proteins in brain development and synaptic plasticity. Neuron 64(1):
93109. https://doi.org/10.1016/j.neuron.2009.09.001
Bramble MS, Roach L, Lipson A, Vashist N, Eskin A, Ngun T et al (2016) Sex-specic effects of
testosterone on the sexually dimorphic transcriptome and epigenome of embryonic neural stem/
progenitor cells. Sci Rep 6:36916. https://doi.org/10.1038/srep36916
Breedlove SM, Arnold AP (1980) Hormone accumulation in a sexually dimorphic motor nucleus of
the rat spinal cord. Science 210:564566
Breedlove SM, Arnold AP (1981) Sexually dimorphic motor nucleus in the rat lumbar spinal cord:
response to adult hormone manipulation, absence in androgen-insensitive rats. Brain Res
225:297307
Breedlove SM, Arnold AP (1983) Hormonal control of a developing neuromuscular system. I.
Complete demasculinization of the male rat spinal nucleus of the bulbocavernosus using the
anti-androgen utamide. J Neurosci 3:417423
Bronson SL, Bale TL (2014) Prenatal stress-induced increases in placental inammation and
offspring hyperactivity are male-specic and ameliorated by maternal antiinammatory
treatment. Endocrinology 155(7):26352646. https://doi.org/10.1210/en.2014-1040
Caldwell HK (2017) Oxytocin and vasopressin: powerful regulators of social behavior. Neuro-
scientist 23:517528. https://doi.org/10.1177/1073858417708284
Cameron NM, Shahrokh D, Del Corpo A, Dhir SK, Szyf M, Champagne FA, Meaney MJ (2008)
Epigenetic programming of phenotypic variations in reproductive strategies in the rat through
maternal care. J Neuroendocrinol 20(6):795801. https://doi.org/10.1111/j.1365-2826.2008.
01725.x
Carrel L, Willard HF (2005) X-inactivation prole reveals extensive variability in X-linked gene
expression in females. Nature 434(7031):400404
Carruth LL, Reisert I, Arnold AP (2002) Sex chromosome genes directly affect brain sexual differ-
entiation. Nat Neurosci 5(10):933934. https://doi.org/10.1038/nn922
Chassot AA, Gregoire EP, Magliano M, Lavery R, Chaboissier MC (2008) Genetics of
ovarian differentiation: Rspo1, a major player. Sex Dev 2(45):219227. https://doi.org/10.
1159/000152038
Clarkson J, Boon WC, Simpson ER, Herbison AE (2009) Postnatal development of an estradiol-
kisspeptin positive feedback mechanism implicated in puberty onset. Endocrinology 150(7):
32143220. https://doi.org/10.1210/en.2008-1733
Clifton VL (2010) Review: Sex and the human placenta: mediating differential strategies of
fetal growth and survival. Placenta 31(Suppl):S33S39. https://doi.org/10.1016/j.placenta.
2009.11.010
Cochrane DR, Cittelly DM, Richer JK (2011) Steroid receptors and microRNAs: relationships
revealed. Steroids 76(12):110. https://doi.org/10.1016/j.steroids.2010.11.003
Conejo NM, Gonzalez-Pardo H, Pedraza C, Navarro FF, Vallejo G, Arias JL (2003) Evidence for
sexual difference in astrocytes of adult rat hippocampus. Neurosci Lett 339(2):119122
Cooke BM (2006) Steroid-dependent plasticity in the medial amygdala. Neuroscience 138(3):
9971005
Coretti L, Cristiano C, Florio E, Scala G, Lama A, Keller S et al (2017) Sex-related alterations of gut
microbiota composition in the BTBR mouse model of autism spectrum disorder. Sci Rep 7:
45356. https://doi.org/10.1038/srep45356
Crestani CC, Alves FH, Gomes FV, Resstel LB, Correa FM, Herman JP (2013) Mechanisms in
the bed nucleus of the stria terminalis involved in control of autonomic and neuroendocrine
Sexual Differentiation and Sex Differences in Neural Development
functions: a review. Curr Neuropharmacol 11(2):141159. https://doi.org/10.2174/
1570159X11311020002
Dantzer R, Bluthe RM (1993) Vasopressin and behavior: from memory to olfaction. Regul Pept
45(12):121125
De Jonge FH, Louwerse AL, Ooms MP, Evers P, Endert E, van de Poll NE (1989) Lesions of the
SDN-POA inhibit sexual behavior of male wistar rats. Brain Res Bull 23(6):483492
de Theije CG, Wopereis H, Ramadan M, van Eijndthoven T, Lambert J, Knol J et al (2014a) Altered
gut microbiota and activity in a murine model of autism spectrum disorders. Brain Behav
Immun 37:197206. https://doi.org/10.1016/j.bbi.2013.12.005
de Theije CG, Wu J, Koelink PJ, Korte-Bouws GA, Borre Y, Kas MJ et al (2014b) Autistic-like
behavioural and neurochemical changes in a mouse model of food allergy. Behav Brain Res
261:265274. https://doi.org/10.1016/j.bbr.2013.12.008
De Vries GJ (2004) Minireview: Sex differences in adult and developing brains: compensation,
compensation, compensation. Endocrinology 145(3):10631068. https://doi.org/10.1210/en.
2003-1504
de Vries GJ, Forger NG (2015) Sex differences in the brain: a whole body perspective. Biol Sex
Differ 6:15. eCollection 2015. https://doi.org/10.1186/s13293-015-0032-z
de Vries GJ, Buijs RM, Swaab DF (1981) Ontogeny of the vasopressinergic neurons of the
suprachiasmatic nucleus and their extrahypothalamic projections in the rat brain--presence of
a sex difference in the lateral septum. Brain Res 218(12):6778
De Vries GJ, Best W, Sluiter AA (1983) The inuence of androgens on the development of a
sex difference in the vasopressinergic innervation of the rat lateral septum. Brain Res
284(23):377380
De Vries GJ, Wang Z, Bullock NA, Numan S (1994) Sex differences in the effects of testosterone
and its metabolites on vasopressin messenger RNA levels in the bed nucleus of the stria
terminalis of rats. J Neurosci 14(3 Pt 2):17891794
De Vries GJ, Rissman EF, Simerly RB, Yang LY, Scordalakes EM, Auger CJ et al (2002) A model
system for study of sex chromosome effects on sexually dimorphic neural and behavioral traits.
J Neurosci 22(20):90059014
Deng X, Berletch JB, Nguyen DK, Disteche CM (2014) X chromosome regulation: diverse patterns
in development, tissues and disease. Nat Rev Genet 15(6):367378. https://doi.org/10.1038/
nrg3687
Deverman BE, Patterson PH (2009) Cytokines and CNS development. Neuron 64(1):6178. https://
doi.org/10.1016/j.neuron.2009.09.002
Dewing P, Chiang CW, Sinchak K, Sim H, Fernagut PO, Kelly S et al (2006) Direct regulation of
adult brain function by the male-specic factor SRY. Curr Biol 16(4):415420
Dodson RE, Gorski RA (1993) Testosterone propionate administration prevents the loss of neurons
within the central part of the medial preoptic nucleus. J Neurobiol 24(1):8088. https://doi.org/
10.1002/neu.480240107
Dumais KM, Veenema AH (2016) Vasopressin and oxytocin receptor systems in the brain:
sex differences and sex-specic regulation of social behavior. Front Neuroendocrinol 40:
123. https://doi.org/10.1016/j.yfrne.2015.04.003
Dumais KM, Bredewold R, Mayer TE, Veenema AH (2013) Sex differences in oxytocin receptor
binding in forebrain regions: correlations with social interest in brain region- and sex-specic
ways. Horm Behav 64(4):693701. https://doi.org/10.1016/j.yhbeh.2013.08.012
Fiebich BL, Akter S, Akundi RS (2014) The two-hit hypothesis for neuroinammation: role of
exogenous ATP in modulating inammation in the brain. Front Cell Neurosci 8:260
Forger NG (2006) Cell death and sexual differentiation of the nervous system. Neuroscience
138(3):929938
Frankfurt M, Gould E, Woolley CS, McEwen BS (1990) Gonadal steroids modify dendritic
spine density in ventromedial hypothalamic neurons: a golgi study in the adult rat.
Neuroendocrinology 51(5):530535
Frick LR, Williams K, Pittenger C (2013) Microglial dysregulation in psychiatric disease. Clin Dev
Immunol 2013:608654. https://doi.org/10.1155/2013/608654
A. Turano et al.
Gabory A, Roseboom TJ, Moore T, Moore LG, Junien C (2013) Placental contribution to the
origins of sexual dimorphism in health and diseases: sex chromosomes and epigenetics.
Biol Sex Differ 4(1):5. https://doi.org/10.1186/2042-6410-4-5
Gorski RA (1978) Sexual differentiation of the brain. Hosp Pract 13(10):5562
Gorski RA, Gordon JH, Shryne JE, Southam AM (1978) Evidence for a morphological sex
difference within the medial preoptic area of the rat brain. Brain Res 148(2):333346
Greenblatt DJ, Harmatz JS, Singh NN, Steinberg F, Roth T, Moline ML et al (2014)
Gender differences in pharmacokinetics and pharmacodynamics of zolpidem following sub-
lingual administration. J Clin Pharmacol 54(3):282290. https://doi.org/10.1002/jcph.220
Guillamon A, Segovia S, del Abril A (1988) Early effects of gonadal steroids on the neuron number
in the medial posterior region and the lateral division of the bed nucleus of the stria terminalis in
the rat. Brain Res 44(2):281290
Gurney ME, Konishi M (1980) Hormone-induced sexual differentiation of brain and behavior in
zebra nches. Science (New York, NY) 208(4450):13801383
Hanamsagar R, Alter MD, Block CS, Sullivan H, Bolton JL, Bilbo SD (2017) Generation of a
microglial developmental index in mice and in humans reveals a sex difference in maturation
and immune reactivity. Glia. https://doi.org/10.1002/glia.23176
Haseltine FP, Ohno S (1981) Mechanisms of gonadal differentiation. Science (New York, NY)
211(4488):12721278
Haussler HU, Jirikowski GF, Caldwell JD (1990) Sex differences among oxytocin-immunoreactive
neuronal systems in the mouse hypothalamus. J Chem Neuroanat 3(4):271276
Hilton GD, Nunez JL, McCarthy MM (2003) Sex differences in response to kainic acid and
estradiol in the hippocampus of newborn rats. Neuroscience 116(2):383391
Houtsmuller EJ, Brand T, de Jonge FH, Joosten RN, van de Poll NE, Slob AK (1994) SDN-POA
volume, sexual behavior, and partner preference of male rats affected by perinatal treatment with
ATD. Physiol Behav 56(3):535541
Howerton CL, Bale TL (2014) Targeted placental deletion of OGT recapitulates the prenatal stress
phenotype including hypothalamic mitochondrial dysfunction. Proc Natl Acad Sci U S A
111(26):96399644. https://doi.org/10.1073/pnas.1401203111
Hu MH, Li XF, McCausland B, Li SY, Gresham R, Kinsey-Jones JS et al (2015) Relative
importance of the arcuate and anteroventral periventricular kisspeptin neurons in control of
puberty and reproductive function in female rats. Endocrinology 156(7):26192631. https://doi.
org/10.1210/en.2014-1655
Insel TR, Hulihan TJ (1995) A gender-specic mechanism for pair bonding: oxytocin and
partner preference formation in monogamous voles. Behav Neurosci 109(4):782789
Jarzab B, Kaminski M, Gubala E, Achtelik W, Wagiel J, Dohler KD (1990) Postnatal treatment of
rats with the beta 2-adrenergic agonist salbutamol inuences the volume of the sexually di-
morphic nucleus in the preoptic area. Brain Res 516(2):257262
Jeong JK, Ryu BJ, Choi J, Kim DH, Choi EJ, Park JW et al (2008) NELL2 participates in formation
of the sexually dimorphic nucleus of the pre-optic area in rats. J Neurochem 106(4):16041613.
https://doi.org/10.1111/j.1471-4159.2008.05505.x
Jordan CL, Letinsky MS, Arnold AP (1988) Synapse elimination occurs late in the hormone-
sensitive levator ani muscle of the rat. J Neurobiol 19(4):335356. https://doi.org/10.1002/neu.
480190403
Jordan CL, Padgett B, Hershey J, Prins G, Arnold A (1997) Ontogeny of androgen receptor
immunoreactivity in lumbar motoneurons and in the sexually dimorphic levator ani muscle of
male rats. J Comp Neurol 379(1):8898. https://doi.org/10.1002/(SICI)1096-9861(19970303)
379:13.0.CO;2-E
Josso N, Lamarre I, Picard JY, Berta P, Davies N, Morichon N et al (1993) Anti-mullerian hormone
in early human development. Early Hum Dev 33(2):9199
Jost A (1947) Recherches sur la differenciation sexuelle de l'embryon de lapin. Arch Anat Microsc
Morph Exp 36:271315
Sexual Differentiation and Sex Differences in Neural Development
Kauffman AS, Gottsch ML, Roa J, Byquist AC, Crown A, Clifton DK et al (2007) Sexual differ-
entiation of Kiss1 gene expression in the brain of the rat. Endocrinology 148(4):17741783
Krebs-Kraft DL, Hill MN, Hillard CJ, McCarthy MM (2010) Sex difference in cell proliferation in
developing rat amygdala mediated by endocannabinoids has implications for social behavior.
Proc Natl Acad Sci U S A 107(47):2053520540. https://doi.org/10.1073/pnas.1005003107
Kudwa AE, Michopoulos V, Gatewood JD, Rissman EF (2006) Roles of estrogen receptors alpha
and beta in differentiation of mouse sexual behavior. Neuroscience 138(3):921928. https://doi.
org/10.1016/j.neuroscience.2005.10.018
Kurian JR, Olesen KM, Auger AP (2010) Sex differences in epigenetic regulation of the
estrogen receptor-alpha promoter within the developing preoptic area. Endocrinology 151(5):
22972305. https://doi.org/10.1210/en.2009-0649
Lebow MA, Chen A (2016) Overshadowed by the amygdala: the bed nucleus of the stria terminalis
emerges as key to psychiatric disorders. Mol Psychiatry 21(4):450463. https://doi.org/10.1038/
mp.2016.1
Legraverend C, Mode A, Westin S, Ström A, Equchi H, Zaphiropoulos PG, Gustafsson JA (1992)
Transcriptional regulation of rat P-450 2C gene subfamily members by the sexually dimorphic
pattern of growth hormone secretion. Mol Endocrinol 6(2):259266
Lenz KM, McCarthy MM (2014) A starring role for microglia in brain sex differences.
Neuroscientist 21:306321
Lenz KM, Nugent BM, Haliyur R, McCarthy MM (2013) Microglia are essential to masculinization
of brain and behavior. J Neurosci 33:27612772
Lubischer JL, Arnold AP (1995) Evidence for target regulation of the development of androgen
sensitivity in rat spinal motoneurons. Dev Neurosci 17(2):106117
Maatouk DM, Capel B (2008) Sexual development of the soma in the mouse. Curr Top Dev Biol
83:151183. https://doi.org/10.1016/S0070-2153(08)00405-5
Maezawa I, Calaore M, Wulff H, Jin LW (2011) Does microglial dysfunction play a role
in autism and Rett syndrome? Neuron Glia Biol 7(1):8597. https://doi.org/10.1017/
S1740925X1200004X
Markle JG, Frank DN, Mortin-Toth S, Robertson CE, Feazel LM, Rolle-Kampczyk U et al (2013)
Sex differences in the gut microbiome drive hormone-dependent regulation of autoimmunity.
Science (New York, NY) 339(6123):10841088. https://doi.org/10.1126/science.1233521
Mathews GA, Brenowitz EA, Arnold AP (1988) Paradoxical hypermasculinization of the
zebra nch song system by an antiestrogen. Horm Behav 22(4):540551
Matsumoto A, Arai Y (1983) Sex difference in volume of the ventromedial nucleus of the
hypothalamus in the rat. Endocrinol Jpn 30(3):277280
Matsumoto A, Arai Y (1986) Development of sexual dimorphism in synaptic organization in the
ventromedial nucleus of the hypothalamus in rats. Neurosci Lett 68(2):165168
McCarthy MM, Arnold AP (2011) Reframing sexual differentiation of the brain. Nat Neurosci
14(6):677683. https://doi.org/10.1038/nn.2834
McCarthy MM, Konkle AT (2005) When is a sex difference not a sex difference? Front Neuro-
endocrinol 26(2):85102
McCarthy MM, Schwarz JM, Wright CL, Dean SL (2008) Mechanisms mediating oestradiol
modulation of the developing brain. J Neuroendocrinol 20(6):777783. https://doi.org/10.
1111/j.1365-2826.2008.01723.x
McCarthy MM, Auger AP, Bale TL, De Vries GJ, Dunn GA, Forger NG et al (2009) The
epigenetics of sex differences in the brain. J Neurosci 29(41):1281512823. https://doi.org/
10.1523/JNEUROSCI.3331-09.2009
Menger Y, Bettscheider M, Murgatroyd C, Spengler D (2010) Sex differences in brain epigenetics.
Epigenomics 2(6):807821. https://doi.org/10.2217/epi.10.60
Miller MA, Urban JH, Dorsa DM (1989a) Steroid dependency of vasopressin neurons in the
bed nucleus of the stria terminalis by in situ hybridization. Endocrinology 125(5):23352340.
https://doi.org/10.1210/endo-125-5-2335
A. Turano et al.
Miller MA, Vician L, Clifton DK, Dorsa DM (1989b) Sex differences in vasopressin neurons in the
bed nucleus of the stria terminalis by in situ hybridization. Peptides 10(3):615619
Millhouse OE (1973) The organization of the ventromedial hypothalamic nucleus. Brain Res 55(1):
7187
Mohr MA, Garcia FL, DonCarlos LL, Sisk CL (2016) Neurons and glial cells are added to
the female rat anteroventral periventricular nucleus during puberty. Endocrinology 157(6):
23932402. https://doi.org/10.1210/en.2015-2012
Mueller BR, Bale TL (2008) Sex-specic programming of offspring emotionality after stress
early in pregnancy. J Neurosci 28(36):90559065. https://doi.org/10.1523/JNEUROSCI.1424-
08.2008
Murakami S, Arai Y (1989) Neuronal death in the developing sexually dimorphic periventricular
nucleus of the preoptic area in the female rat: effect of neonatal androgen treatment.
Neurosci Lett 102(23):185190
Murray EK, Hien A, de Vries GJ, Forger NG (2009) Epigenetic control of sexual differentiation of
the bed nucleus of the stria terminalis. Endocrinology 150(9):42414247. https://doi.org/10.
1210/en.2009-0458
Nakachi Y, Iseki M, Yokoo T, Mizuno Y, Okazaki Y (2015) Gene expression prole of the neonatal
female mouse brain after administration of testosterone propionate. J Sex Med 12(4):887896.
https://doi.org/10.1111/jsm.12802
Nelson LH, Warden S, Lenz KM (2017) Sex differences in microglial phagocytosis in the
neonatal hippocampus. Brain Behav Immun 64:1122
Nordeen EJ, Nordeen KW, Sengelaub DR, Arnold AP (1985) Androgens prevent normally occur-
ring cell death in a sexually dimorphic spinal nucleus. Science (New York, NY) 229(4714):
671673
Nottebohm F (1991) Reassessing the mechanisms and origins of vocal learning in birds.
Trends Neurosci 14(5):206211
Nottebohm F, Arnold AP (1976) Sexual dimorphism in vocal control areas of the songbird brain.
Science (New York, NY) 194(4261):211213
Nottebohm F, Kelley DB, Paton JA (1982) Connections of vocal control nuclei in the canary
telencephalon. J Comp Neurol 207(4):344357. https://doi.org/10.1002/cne.902070406
Nugent BM, McCarthy MM (2011) Epigenetic underpinnings of developmental sex differences in
the brain. Neuroendocrinology 93(3):150158. https://doi.org/10.1159/000325264
Nugent BM, Wright CL, Shetty AC, Hodes GE, Lenz KM, Mahurkar A et al (2015) Brain femin-
ization requires active repression of masculinization via DNA methylation. Nat Neurosci 18(5):
690697. https://doi.org/10.1038/nn.3988
Nunez JL, Alt JJ, McCarthy MM (2003) A new model for prenatal brain damage. I. GABAA
receptor activation induces cell death in developing rat hippocampus. Exp Neurol 181(2):
258269
Pailhoux E, Parma P, Sundstrom J, Vigier B, Servel N, Kuopio T et al (2001) Time course of
female-to-male sex reversal in 38,XX fetal and postnatal pigs. Dev Dyn 222(3):328340. https://
doi.org/10.1002/dvdy.1194
Painter TS (1923) Studies in mammalian spermatogenesis. II. The spermatogenesis of man. J Exp
Zool 37:291336. https://doi.org/10.1002/jez.1400370303
Paul MJ, Terranova JI, Probst CK, Murray EK, Ismail NI, de Vries GJ (2014) Sexually dimorphic
role for vasopressin in the development of social play. Front Behav Neurosci 8:58. https://doi.
org/10.3389/fnbeh.2014.00058
Pfaff DW, Sakuma Y (1979) Decit in the lordosis reex of female rats caused by lesions in the
ventromedial nucleus of the hypothalamus. J Physiol 288:203210
Phoenix CH, Goy RW, Gerall AA, Young WC (1959) Organizing action of prenatally administered
testosterone propionate on the tissues mediating mating behavior in the female guinea pig.
Endocrinology 65:369382
Sexual Differentiation and Sex Differences in Neural Development
Pozzo Miller LD, Aoki A (1991) Stereological analysis of the hypothalamic ventromedial nucleus.
II. Hormone-induced changes in the synaptogenic pattern. Brain Res Dev Brain Res 61(2):
189196
Qiao X, Yan Y, Wu R, Tai F, Hao P, Cao Y, Wang J (2014) Sociality and oxytocin and vasopressin
in the brain of male and female dominant and subordinate mandarin voles. J Comp Physiol A
Neuroethol Sens Neural Behav Physiol 200(2):149159. https://doi.org/10.1007/s00359-013-
0870-2
Quadros PD, Pfau JL, Goldstein AY, de Vries GJ, Wagner CK (2002) Sex differences in proges-
terone receptor expression: a potential mechanism for estradiol-mediated sexual differentiation.
Endocrinology 143(10):37273739
Rood BD, Murray EK, Laroche J, Yang MK, Blaustein JD, De Vries GJ (2008) Absence of
progestin receptors alters distribution of vasopressin bers but not sexual differentiation of
vasopressin system in mice. Neuroscience 154(3):911921. https://doi.org/10.1016/j.neurosci
ence.2008.03.087
Ross HE, Young LJ (2009) Oxytocin and the neural mechanisms regulating social cognition and
afliative behavior. Front Neuroendocrinol 30(4):534547. https://doi.org/10.1016/j.yfrne.
2009.05.004
Roth TL, Roth ED, Sweatt JD (2010) Epigenetic regulation of genes in learning and memory.
Essays Biochem 48(1):263274. https://doi.org/10.1042/bse0480263
Roy AK, Chatterjee B (1983) Sexual dimorphism in the liver. Annu Rev Physiol 45:3750. https://
doi.org/10.1146/annurev.ph.45.030183.000345
Schwarz JM, Bilbo SD (2012) Sex, glia, and development: interactions in health and disease. Horm
Behav 62(3):243253. https://doi.org/10.1016/j.yhbeh.2012.02.018
Schwarz JM, McCarthy MM (2008) The role of neonatal NMDA receptor activation in defemin-
ization and masculinization of sex behavior in the rat. Horm Behav 54(5):662668. https://doi.
org/10.1016/j.yhbeh.2008.07.004
Schwarz JM, Liang SL, Thompson SM, McCarthy MM (2008) Estradiol induces hypothalamic
dendritic spines by enhancing glutamate release: a mechanism for organizational sex differ-
ences. Neuron 58(4):584598. https://doi.org/10.1016/j.neuron.2008.03.008
Schwarz JM, Nugent BM, McCarthy MM (2010) Developmental and hormone-induced epigenetic
changes to estrogen and progesterone receptor genes in brain are dynamic across the life span.
Endocrinology 151(10):48714881. https://doi.org/10.1210/en.2010-0142
Schwarz JM, Sholar PW, Bilbo SD (2012) Sex differences in microglial colonization of the devel-
oping rat brain. J Neurochem 120(6):948963. https://doi.org/10.1111/j.1471-4159.2011.
07630.x
Sekido R (2014) The potential role of SRY in epigenetic gene regulation during brain sexual differ-
entiation in mammals. Adv Genet 86:135165. https://doi.org/10.1016/B978-0-12-800222-3.
00007-3
Sengelaub DR, Forger NG (2008) The spinal nucleus of the bulbocavernosus: rsts in androgen-
dependent neural sex differences. Horm Behav 53(5):596612. https://doi.org/10.1016/j.yhbeh.
2007.11.008
Shen EY, Ahern TH, Cheung I, Straubhaar J, Dincer A, Houston I et al (2015) Epigenetics and sex
differences in the brain: a genome-wide comparison of histone-3 lysine-4 trimethylation
(H3K4me3) in male and female mice. Exp Neurol 268:2129. https://doi.org/10.1016/j.
expneurol.2014.08.006
Sheridan GK, Wdowicz A, Pickering M, Watters O, Halley P, O'Sullivan NC et al (2014) CX3CL1
is up-regulated in the rat hippocampus during memory-associated synaptic plasticity. Front Cell
Neurosci 8:233. https://doi.org/10.3389/fncel.2014.00233
Shughrue PJ, Sar M, Stumpf WE (1992) Progestin target cell distribution in forebrain and midbrain
regions of the 8-day postnatal mouse brain. Endocrinology 130(6):36503659
Shughrue PJ, Lane MV, Merchenthaler I (1997) Comparative distribution of estrogen receptor-
alpha and -beta mRNA in the rat central nervous system. J Comp Neurol 388(4):507525
Simerly RB (2002) Wired for reproduction: organization and development of sexually dimorphic
circuits in the mammalian forebrain. Annu Rev Neurosci 25(1):507536. https://doi.org/10.
1146/annurev.neuro.25.112701.142745
A. Turano et al.
Smith MR, Hamson DK, Poort JE, Jordan CL, Breedlove SM (2012) Ontogeny of androgen
receptor expression in spinal nucleus of the bulbocavernosus motoneurons and their target
muscles in male mice. Neurosci Lett 513(2):119123. https://doi.org/10.1016/j.neulet.2012.
01.067
Sumida H, Nishizuka M, Kano Y, Arai Y (1993) Sex differences in the anteroventral periventricular
nucleus of the preoptic area and in the related effects of androgen in prenatal rats. Neurosci Lett
151(1):4144
Taketo T, Lee CH, Zhang J, Li Y, Lee CY, Lau YF (2005) Expression of SRY proteins in both
normal and sex-reversed XY fetal mouse gonads. Dev Dyn 233(2):612622. https://doi.org/10.
1002/dvdy.20352
Taylor PV, Veenema AH, Paul MJ, Bredewold R, Isaacs S, de Vries GJ (2012) Sexually dimorphic
effects of a prenatal immune challenge on social play and vasopressin expression in
juvenile rats. Biol Sex Differ 3(1):15. https://doi.org/10.1186/2042-6410-3-15
Terranova JI, Song Z, Larkin TE 2nd, Hardcastle N, Norvelle A, Riaz A, Albers HE (2016)
Serotonin and arginine-vasopressin mediate sex differences in the regulation of dominance
and aggression by the social brain. Proc Natl Acad Sci U S A 113(46):1323313238
Todd BJ, Schwarz JM, McCarthy MM (2005) Prostaglandin-E2: a point of divergence in estradiol-
mediated sexual differentiation. Horm Behav 48(5):512521. https://doi.org/10.1016/j.yhbeh.
2005.07.011
Tsukahara S (2009) Sex differences and the roles of sex steroids in apoptosis of sexually dimorphic
nuclei of the preoptic area in postnatal rats. J Neuroendocrinol 21(4):370376. https://doi.org/
10.1111/j.1365-2826.2009.01855.x
Vainio SJ, Itaranta PV, Perasaari JP, Uusitalo MS (1999) Wnts as kidney tubule inducing factors.
Int J Dev Biol 43(5):419423
Veenema AH, Neumann ID (2008) Central vasopressin and oxytocin release: regulation of complex
social behaviours. Prog Brain Res 170:261276. https://doi.org/10.1016/S0079-6123(08)
00422-6
Wade CB, Robinson S, Shapiro RA, Dorsa DM (2001) Estrogen receptor (ER)alpha and ERbeta
exhibit unique pharmacologic properties when coupled to activation of the mitogen-activated
protein kinase pathway. Endocrinology 142(6):23362342
Wagner CK, Nakayama AY, de Vries GJ (1998) Potential role of maternal progesterone in the
sexual differentiation of the brain. Endocrinology 139(8):36583661
Waxman DJ, Holloway MG (2009) Sex differences in the expression of hepatic drug metabolizing
enzymes. Mol Pharmacol 76(2):215228. https://doi.org/10.1124/mol.109.056705
Wierda M, Goudsmit E, Van der Woude PF, Purba JS, Hofman MA, Bogte H, Swaab DF (1991)
Oxytocin cell number in the human paraventricular nucleus remains constant with aging and in
Alzheimers disease. Neurobiol Aging 12(5):511516
Williams JR, Insel TR, Harbaugh CR, Carter CS (1994) Oxytocin administered centrally
facilitates formation of a partner preference in female prairie voles (microtus ochrogaster).
J Neuroendocrinol 6(3):247250
Winslow JT, Hastings N, Carter CS, Harbaugh CR, Insel TR (1993) A role for central vasopressin
in pair bonding in monogamous prairie voles. Nature 365(6446):545548. https://doi.org/10.
1038/365545a0
Wright CL, Burks SR, McCarthy MM (2008) Identication of prostaglandin E2 receptors
mediating perinatal masculinization of adult sex behavior and neuroanatomical correlates.
Dev Neurobiol 68(12):14061419. https://doi.org/10.1002/dneu.20665
Wu MV, Manoli DS, Fraser EJ, Coats JK, Tollkuhn J, Honda S, Harada N, Shah NM (2009)
Estrogen masculinizes neural pathways and sex-specic behaviors. Cell 139(7):6172
Xu J, Disteche CM (2006) Sex differences in brain expression of X- and Y-linked genes. Brain Res
1126(1):5055
Yamamoto Y, Cushing BS, Kramer KM, Epperson PD, Hoffman GE, Carter CS (2004) Neonatal
manipulations of oxytocin alter expression of oxytocin and vasopressin immunoreactive cells in
Sexual Differentiation and Sex Differences in Neural Development
the paraventricular nucleus of the hypothalamus in a gender-specic manner. Neuroscience
125(4):947955. https://doi.org/10.1016/j.neuroscience.2004.02.028
Yamamoto Y, Carter CS, Cushing BS (2006) Neonatal manipulation of oxytocin affects expression
of estrogen receptor alpha. Neuroscience 137(1):157164
Yang X, Schadt EE, Wang S, Wang H, Arnold AP, Ingram-Drake L et al (2006) Tissue-specic
expression and regulation of sexually dimorphic genes in mice. Genome Res 16(8):9951004
Yao S, Bergan J, Lanjuin A, Dulac C (2017) Oxytocin signaling in the medial amygdala is required
for sex discrimination of social cues. elife 6:pii: e31373. https://doi.org/10.7554/eLife.31373
Yurkovetskiy L, Burrows M, Khan AA, Graham L, Volchkov P, Becker L et al (2013) Gender bias
in autoimmunity is inuenced by microbiota. Immunity 39(2):400412. https://doi.org/10.1016/
j.immuni.2013.08.013
Zhang JM, Konkle AT, Zup SL, McCarthy MM (2008) Impact of sex and hormones on new cells in
the developing rat hippocampus: a novel source of sex dimorphism? Eur J Neurosci 27(4):
791800. https://doi.org/10.1111/j.1460-9568.2008.06073.x
Ziegler DR, Cullinan WE, Herman JP (2002) Distribution of vesicular glutamate transporter mRNA
in rat hypothalamus. J Comp Neurol 448(3):217229
A. Turano et al.
... For example, males have a Y chromosome which possess male-specific gene (SRY gene), and in Sex Differences: Neurodevelopment, Fig. 2 Gonadal hormones in neurodevelopment: The MIS (Müllerian Inhibiting Substance) and testosterone released from testes forms the Fallopian Tubes and male reproductive system in males whereas Müllerian duct and the female reproductive system is developed in females with no MIS females a particular X chromosome gets deactivated to balance out the X chromosome expression in males. Hence, along with gonadal hormones sex chromosomes are accountable for sexual dimorphism in the brain (Fig. 3) (Turano et al., 2018). ...
... This variation occurs due to the testosterone signaling in AVPV during neurodevelopment; it leads to apoptosis of neurons within this area in males, which is diametrically opposite to the function it performs in SNB. Along with testosterone, estradiol also promotes cell death in the developing stage of AVPV (Turano et al., 2018). ...
... The sex fate of different individuals is determined at the time of conception, and these very sex chromosomes, or to be more precise sex hormones, determine the difference in development and behavior via enhancing the cell functions of the nervous system. This makes the understanding of the underlying cellular mechanisms responsible for these sex differences, more important (Turano et al., 2018). ...
... Early testosterone exposure affects outcomes, including gender identity, sexual orientation, and children's play behavior [26]. Epigenetic alterations also influence gene expression, which aids in the development of the brain's sexual differentiation [27]. The two recognized epigenetic processes causing SG differences are genomic imprinting and the inactivation of one of the two X chromosomes in females. ...
... Three main areas are being explored related to the physiological and behavioral outcomes of SG differences in brain development: brain nuclei, neural cell communication, and the communication between the brain hemispheres [27]. Sex differences in brain tissue micro-structure suggest a lesser vulnerability to age-related changes in women [30]. ...
Article
Full-text available
The societal implication of sex and gender (SG) differences in brain are profound, as they influence brain development, behavior, and importantly, the presentation, prevalence, and therapeutic response to diseases. Technological advances have enabled speed up identification and characterization of SG differences during development and in psychopathologies. The main aim of this review is to elaborate on new technological advancements, such as genomics, imaging, and emerging biobanks, coupled with bioinformatics analyses of data generated from these technologies have facilitated the identification and characterization of SG differences in the human brain through development and psychopathologies. First, a brief explanation of SG concepts is provided, along with a developmental and evolutionary context. We then describe physiological SG differences in brain activity and function, and in psychopathologies identified through imaging techniques. We further provide an overview of insights into SG differences using genomics, specifically taking advantage of large cohorts and biobanks. We finally emphasize how bioinformatics analyses of big data generated by emerging technologies provides new opportunities to reduce SG disparities in health outcomes, including major challenges.
... According to the study by Whitehouse et al., a high level of testosterone in the male umbilical cord was a risk factor for speech developmental delay at the age of 1, 2 and 3 years old [27]. Research also suggests the effect of epigenetic mechanisms on sex differences in the human brain [28,29]. In the study of Martínez-Nadal et al., cesarean delivery was associated with the risk of developmental delay [30], especially in the gross motor area, which is consistent with our study. ...
Article
Full-text available
Background: Studies on perinatal risk factors and the developmental delay of children have been inconclusive and few studies have assessed the association between infants and toddlers' body mass index (BMI) and developmental outcomes. Methods: We conducted a cross-sectional study of children aged 1-36 months who had a routine physical examination in the child health departments of hospitals from March 2018 to November 2021 in 16 provinces, 4 autonomous regions and 2 municipalities directly under the central government by using the Infant Toddler Growth Development Screening Test (ITGDST). Normal children were defined as those with scores ≥ mean - 2 standard deviations (SD), while children with developmental delay were those with scores < mean-2SD in terms of overall development, gross motor, fine motor and language development. Binary logistic regression was used to analyze the risk factors of gross motor, fine motor, language and overall neurodevelopment. Results: After removing some provinces with a small sample size and children with incomplete data, 178,235 children with 12 complete variables were included in the final analysis. The rate of overall developmental delay was 4.5%, while 12.5% of children had at least one developmental delay aspect. Boys, parity, advanced maternal age, multiple birth, cesarean section, neonatal injury, family heredity history, microcephaly, abnormal BMI at birth and at physical examination after controlling the confounding of other factors had a significant effect on development delay (overall neurodevelopment, gross motor, fine motor or language development). Per capita gross domestic product was a protective factor for the children's neuropsychological development. Conclusions: This study reveals significant associations of perinatal factors and BMI with developmental delay in the Chinese children aged 1-36 months, which may be crucial for early intervention.
... To detect sex differences in human early neural development when sex chromosome complement is the only determinant of sex differences (Turano et al., 2018), the XX and XY hiPSC lines were directed to differentiate towards the neural lineage and subsequently compared. After 14 days of differentiation, the majority of cells acquired neural progenitor phenotype, as demonstrated by the expression of PSA-NCAM (92% ± 3.5%; FACS analysis) and NESTIN (>90%; immunostaining) ( Figures 4A and 4B). ...
Article
Full-text available
Biological sex is a fundamental trait influencing development, reproduction, pathogenesis, and medical treatment outcomes. Modeling sex differences is challenging because of the masking effect of genetic variability and the hurdle of differentiating chromosomal versus hormonal effects. In this work we developed a cellular model to study sex differences in humans. Somatic cells from a mosaic Klinefelter syndrome patient were reprogrammed to generate isogenic induced pluripotent stem cell (iPSC) lines with different sex chromosome complements: 47,XXY/46,XX/46,XY/45,X0. Transcriptional analysis of the hiPSCs revealed novel and known genes and pathways that are sexually dimorphic in the pluripotent state and during early neural development. Female hiPSCs more closely resembled the naive pluripotent state than their male counterparts. Moreover, the system enabled differentiation between the contributions of X versus Y chromosome to these differences. Taken together, isogenic hiPSCs present a novel platform for studying sex differences in humans and bear potential to promote gender-specific medicine in the future.
... Nevertheless, both dimensions mutually interact. Morphological, physiological, and plastic behavioral sex differences have been described consistently in many animal groups across taxa (e.g., Iwaniuk, 2017;Vallortigara and Versace, 2017;Cummings, 2018;Darda et al., 2018;Turano et al., 2018;Luders and Kurth, 2020;Kurth et al., 2020). These could derive from the different roles both sexes perform in their social environment (e.g., brood/nest care, social training by mothers, defense and food provisioning by fathers), which are frequently instrumental in shaping and being shaped by personality traits. ...
Article
Full-text available
Across taxa, mate choice is a highly selective process involving both intra- and intersexual selection processes aiming to pass on one’s genes, making mate choice a pivotal tool of sexual selection. Individuals adapt mate choice behavior dynamically in response to environmental and social changes. These changes are perceived sensorily and integrated on a neuronal level, which ultimately leads to an adequate behavioral response. Along with perception and prior to an appropriate behavioral response, the choosing sex has (1) to recognize and discriminate between the prospective mates and (2) to be able to assess and compare their performance in order to make an informed decision. To do so, cognitive processes allow for the simultaneous processing of multiple information from the (in-) animate environment as well as from a variety of both sexual and social (but non-sexual) conspecific cues. Although many behavioral aspects of cognition on one side and of mate choice displays on the other are well understood, the interplay of neuronal mechanisms governing both determinants, i.e., governing cognitive mate choice have been described only vaguely. This review aimed to throw a spotlight on neuronal prerequisites, networks and processes supporting the interaction between mate choice, sex roles and sexual cognition, hence, supporting cognitive mate choice. How does neuronal activity differ between males and females regarding social cognition? Does sex or the respective sex role within the prevailing mating system mirror at a neuronal level? How does cognitive competence affect mate choice? Conversely, how does mate choice affect the cognitive abilities of both sexes? Benefitting from studies using different neuroanatomical techniques such as neuronal activity markers, differential coexpression or candidate gene analyses, modulatory effects of neurotransmitters and hormones, or imaging techniques such as fMRI, there is ample evidence pointing to a reflection of sex and the respective sex role at the neuronal level, at least in individual brain regions. Moreover, this review aims to summarize evidence for cognitive abilities influencing mate choice and vice versa. At the same time, new questions arise centering the complex relationship between neurobiology, cognition and mate choice, which we will perhaps be able to answer with new experimental techniques.
Article
Maternal vitamin B12 deficiency has been associated with disturbed cognitive functioning in offspring at different ages during childhood. However, this association has not been explored in pre‐school‐age children. The objective of this study was to examine the association between maternal vitamin B12 levels at the beginning and end of pregnancy and cognitive functioning in their children at 4 years of age. This longitudinal prospective study included a subsample of pregnant women and their children aged 4 years ( n = 249) who participated in the ECLIPSES Study conducted in the province of Tarragona, Spain, from 2013 to 2017. Maternal vitamin B12 concentrations were determined in the first and third trimesters, and sociodemographic, nutritional and psychological data were collected. The children's cognitive functioning was assessed using the Wechsler Preschool and Primary Scale of Intelligence (WPPSI‐IV) and subtests of the Neuropsychological Assessment of Development (NEPSY‐II). The multivariable models showed a significant relationship between vitamin B12 and the working memory index in the first trimester of the pregnancy but not in the third trimester. Children of mothers in the second vitamin B12 level tertile (314–413 pg/mL) ( β = 6.468, 95% confidence interval [CI]: = 2.054, 10.882, p = 0.004) and third vitamin B12 level tertile (≥414 pg/mL) ( β = 4.703, 95% CI: = 0.292, 9.114, p = 0.037) scored higher in the working memory index of the WPPSI‐IV than the children of mothers with vitamin B12 levels in the first tertile (<314 pg/mL). Maintaining an adequate level of maternal vitamin B12 during early pregnancy contributes to improved performance in childhood working memory at 4 years of age.
Article
Sex/gender differences in cognitive sciences are riddled by conflicting perspectives. At the center of debates are clinical, social, and political perspectives. Front and center, evolutionary and biological perspectives have often focused on 'nature' arguments, while feminist and constructivist views have often focused on 'nurture arguments regarding cognitive sex differences. In the current narrative review, we provide a comprehensive overview regarding the origins and historical advancement of these debates while providing a summary of the results in the field of sexually polymorphic cognition. In so doing, we attempt to highlight the importance of using transdisciplinary perspectives which help bridge disciplines together to provide a refined understanding the specific factors that drive sex differences a gender diversity in cognitive abilities. To summarize, biological sex (e.g., birth-assigned sex, sex hormones), socio-cultural gender (gender identity, gender roles), and sexual orientation each uniquely shape the cognitive abilities reviewed. To date, however, few studies integrate these sex and gender factors together to better understand individual differences in cognitive functioning. This has potential benefits if a broader understanding of sex and gender factors are systematically measured when researching and treating numerous conditions where cognition is altered.
Article
Full-text available
Biological sex contributes to phenotypic sex effects through genetic (sex chromosomal) and hormonal (gonadal) mechanisms. There are profound sex differences in the prevalence and progression of age-related brain diseases, including neurodegenerative diseases. Inflammation of neural tissue is one of the most consistent age-related phenotypes seen with healthy aging and disease. The pro-inflammatory environment of the aging brain has primarily been attributed to microglial reactivity and adoption of heterogeneous reactive states dependent upon intrinsic (i.e., sex) and extrinsic (i.e., age, disease state) factors. Here, we review sex effects in microglia across the lifespan, explore potential genetic and hormonal molecular mechanisms of microglial sex effects, and discuss currently available models and methods to study sex effects in the aging brain. Despite recent attention to this area, significant further research is needed to mechanistically understand the regulation of microglial sex effects across the lifespan, which may open new avenues for sex informed prevention and treatment strategies.
Article
Exposure to selective serotonin reuptake inhibitors can affect hormone‐dependent processes, such as the brain sexual differentiation. Since the use of these antidepressants cause concern during lactation, we evaluated the possible effects of venlafaxine on lactational exposure and its late repercussions on reproductive parameters in male rats. Lactating rats were exposed to venlafaxine (3.85, 7.7, or 15.4 mg/kg/body weight; gavage), from lactational day 1 to 20. Venlafaxine and O‐desmethylvenlafaxine residues were found in all milk samples of dams treated, demonstrating the lactational transfer of this antidepressant to the offspring. Although the maternal behavior was normal, the dams presented an increase in urea and uric acid levels in the groups treated with 7.7 and 15.4, respectively, as well as a spleen weight increased in the 3.85, and 15.4 groups. The male offspring showed a decrease in play behavior parameters in the intermediate dose group. Sperm analysis indicated a reduction in sperm motility in all treated groups. The androgen receptor expression in the hypothalamus was decreased in the highest dose group, although the sexual behavior had not been affected. In conclusion, venlafaxine was transferred through breast milk and promoted changes in play behavior, sperm quality, and hypothalamic AR content, which may indicate an incomplete masculinization of the brain of male offspring. Since the use of SNRIs during lactation cause concern, we evaluated the possible effects of the antidepressant venlafaxine on reproductive parameters in male rats. Lactating rats were exposed to venlafaxine (3.85, 7.7, or 15.4 mg/kg/body weight; gavage), from lactational day 1 to 20. Venlafaxine treatment caused mild signs of maternal toxicity and altered important parameters evaluated in the male offspring, such as sperm parameters and AR expression in the hypothalamus. Thus, venlafaxine exposure was detrimental in these conditions.
Article
Full-text available
The neural control of social behaviors in rodents requires the encoding of pheromonal cues by the vomeronasal system. Here we show that the typical preference of male mice for females is eliminated in mutants lacking oxytocin, a neuropeptide modulating social behaviors in many species. Ablation of the oxytocin receptor in aromatase-expressing neurons of the medial amygdala (MeA) fully recapitulates the elimination of female preference in males. Further, single- unit recording in the MeA uncovered significant changes in the sensory representation of conspecific cues in the absence of oxytocin signaling. Finally, acute manipulation of oxytocin signaling in adults is sufficient to alter social interaction preferences in males as well as responses of MeA neurons to chemosensory cues. These results uncover the critical role of oxytocin signaling in a molecularly defined neuronal population in order to modulate the behavioral and physiological responses of male mice to females on a moment-to-moment basis.
Article
Full-text available
Alterations of microbiota-gut-brain axis have been invoked in the pathogenesis of autism spectrum disorders (ASD). Mouse models could represent an excellent tool to understand how gut dysbiosis and related alterations may contribute to autistic phenotype. In this study we paralleled gut microbiota (GM) profiles, behavioral characteristics, intestinal integrity and immunological features of colon tissues in BTBR T + tf/J (BTBR) inbred mice, a well established animal model of ASD. Sex differences, up to date poorly investigated in animal models, were specifically addressed. Results showed that BTBR mice of both sexes presented a marked intestinal dysbiosis, alterations of behavior, gut permeability and immunological state with respect to prosocial C57BL/6j (C57) strain. Noticeably, sex-related differences were clearly detected. We identified Bacteroides, Parabacteroides, Sutterella, Dehalobacterium and Oscillospira genera as key drivers of sex-specific gut microbiota profiles associated with selected pathological traits. Taken together, our findings indicate that alteration of GM in BTBR mice shows relevant sex-associated differences and supports the use of BTBR mouse model to dissect autism associated microbiota-gut-brain axis alteration.
Article
Full-text available
The mechanisms by which sex differences in the mammalian brain arise are poorly understood, but are influenced by a combination of underlying genetic differences and gonadal hormone exposure. Using a mouse embryonic neural stem cell (eNSC) model to understand early events contributing to sexually dimorphic brain development, we identified novel interactions between chromosomal sex and hormonal exposure that are instrumental to early brain sex differences. RNA-sequencing identified 103 transcripts that were differentially expressed between XX and XY eNSCs at baseline (FDR = 0.10). Treatment with testosterone-propionate (TP) reveals sex-specific gene expression changes, causing 2854 and 792 transcripts to become differentially expressed on XX and XY genetic backgrounds respectively. Within the TP responsive transcripts, there was enrichment for genes which function as epigenetic regulators that affect both histone modifications and DNA methylation patterning. We observed that TP caused a global decrease in 5-methylcytosine abundance in both sexes, a transmissible effect that was maintained in cellular progeny. Additionally, we determined that TP was associated with residue-specific alterations in acetylation of histone tails. These findings highlight an unknown component of androgen action on cells within the developmental CNS, and contribute to a novel mechanism of action by which early hormonal organization is initiated and maintained.
Article
Full-text available
Significance There are profound sex differences in the expression of social behavior and in the incidence of many psychiatric disorders, and yet little is known about how the brain mechanisms underlying these phenomena differ in females and males. Here, we report that serotonin (5-HT) and arginine–vasopressin (AVP) act in opposite ways within the hypothalamus to regulate dominance and aggression in females and males. Dominance and aggression are promoted by 5-HT in females and by AVP in males. Because dominance and aggressiveness have been linked to the resistance to stress-related psychiatric disorders, these disorders may be more effectively treated with 5-HT–targeted drugs in females and AVP-targeted drugs in males.
Article
Evidence suggests many neurological disorders emerge when normal neurodevelopmental trajectories are disrupted, i.e., when circuits or cells do not reach their fully mature state. Microglia play a critical role in normal neurodevelopment and are hypothesized to contribute to brain disease. We used whole transcriptome profiling with Next Generation sequencing of purified developing microglia to identify a microglial developmental gene expression program involving thousands of genes whose expression levels change monotonically (up or down) across development. Importantly, the gene expression program was delayed in males relative to females and exposure of adult male mice to LPS, a potent immune activator, accelerated microglial development in males. Next, a microglial developmental index (MDI) generated from gene expression patterns obtained from purified mouse microglia, was applied to human brain transcriptome datasets to test the hypothesis that variability in microglial development is associated with human diseases such as Alzheimer's and autism where microglia have been suggested to play a role. MDI was significantly increased in both Alzheimer's Disease and in autism, suggesting that accelerated microglial development may contribute to neuropathology. In conclusion, we identified a microglia-specific gene expression program in mice that was used to create a microglia developmental index, which was applied to human datasets containing heterogeneous cell types to reveal differences between healthy and diseased brain samples, and between males and females. This powerful tool has wide ranging applicability to examine microglial development within the context of disease and in response to other variables such as stress and pharmacological treatments.
Article
For many, the terms oxytocin and vasopressin immediately evoke images of animals interacting with one another, as both of these neuropeptides have been implicated as being part of the neurochemical "glue" that socially binds animals. However, social environments and social interactions are complex and include behaviors that bring animals together as well as behaviors that keep animals apart. It is at the intersection of social context, social experience, and an individual's sex that oxytocin and vasopressin act to modulate social behavior and social cognition. In this review, this complexity will be explored across mammalian species, with a focus on social memory, cooperative behaviors, and competitive behaviors. Implications for humans as well as future directions will also be considered.
Article
Microglia regulate brain development through many processes, such as promoting neurogenesis, supporting cell survival, and phagocytizing progenitor, newly-born, and dying cells. Many of these same developmental processes show robust sex differences, yet very few studies have assessed sex differences in microglia function during development. Hormonally-induced sexual differentiation of the brain occurs during the perinatal period, thus we examined sex differences in microglial morphology, phagocytosis, and proliferation in the hippocampus during the early postnatal period. We found that the neonatal female hippocampus had significantly more microglia with phagocytic cups than the male hippocampus. We subsequently found that female microglia phagocytized more neural progenitor cells and healthy cells compared to males, but there were no sex differences in the number of newly-born or dying cells targeted by microglial phagocytosis. We found that the number of phagocytic microglia in females was reduced to male-typical levels by treatment with estradiol, the hormone responsible for masculinizing the rodent brain. Females also had higher expression of several phagocytic pathway genes in the hippocampus compared to males. In contrast to robust sex differences in phagocytic microglia, we found no sex differences in the number of microglia with amoeboid, transitioning, or ramified morphologies or differences in three-dimensional reconstructions of microglial morphology. While we did not find a baseline sex difference in microglial proliferation during or following the prenatal gonadal hormone surge in males, we found that estradiol treatment increased microglia proliferation in females. Overall, these data show that there are important sex differences in microglia function in the hippocampus during the early neonatal period.
Article
The onset of the sexually dimorphic pattern of GH secretion and increased hepatic GH-binding capacity in rats at puberty is temporally correlated with the developmental induction of three hepatic cytochrome P-450s with steroid hydroxylase activity, P-450 IIC11, P-450 IIC12, and P-450 IIC13, and one cytochrome P-450 with vitamin A hydroxylase activity, P-450 IIC7. In this study we demonstrate that expression of the 2C11, 2C12, and 2C13 genes is modulated by GH at the level of transcriptional initiation both in vivo and in primary cultures of adult hepatocytes. In an effort to define the minimum sequence responsible for the inductive effects of GH, we have analyzed the ability of a 0.7-kilobase fragment isolated from the 5'-flank of the 2C12 gene, including the natural promoter, to drive transcription of a 320-basepair G-less cassette in vitro. We were unable to detect any substantial difference in RNA polymerase-II-dependent transcriptional efficiency toward the 2C12 promoter between liver nuclear extracts...
Article
The anteroventral periventricular nucleus (AVPV) orchestrates the neuroendocrine positive feedback response that triggers ovulation in female rodents. The AVPV is larger and more cell-dense in females than in males, and during puberty, only females develop the capacity to show a positive feedback response. We previously reported a potential new mechanism to explain this female-specific gain of function during puberty, namely a female-biased sex difference in the pubertal addition of new cells to the rat AVPV. Here we first asked whether this sex difference is due to greater cell proliferation and/or survival in females. Female and male rats received the cell birthdate marker bromodeoxyuridine (BrdU; 200 mg/kg, ip) on postnatal day (P)30; brains were collected at short and long intervals after BrdU administration to assess cell proliferation and survival, respectively. Overall, females had more BrdU-ir cells in the AVPV than did males, with no sex differences in the rate of cell attrition over time. Thus, the sex difference in pubertal addition of AVPV cells appears to be due to greater cell proliferation in females. Next, to determine the phenotype of pubertally born AVPV cells, daily BrdU injections were given to female rats on P28-56 and tissue was collected on P77 to assess colocalization of BrdU and markers for mature neurons or glia. Of the pubertally born AVPV cells, ∼15% differentiated into neurons, ∼19% into astrocytes, and ∼23% into microglia. Thus, both neuro- and gliogenesis occur in the pubertal female rat AVPV and potentially contribute to maturation of female reproductive function.