Sei sulla pagina 1di 30

Author's personal copy

5.01 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms,


and Meaning
Margaret M McCarthy, University of Maryland School of Medicine, Baltimore, MD, USA
Geert J De Vries and Nancy G Forger, Neuroscience Institute, Georgia State University, Atlanta, GA, USA
Ó 2017 Elsevier Inc. All rights reserved.

5.01.1 Introduction 4
5.01.2 Ten Things We Know 4
5.01.2.1 Hormones Cause Sex Differences by Acting during Development as well as in Adulthood 4
5.01.2.2 There Are Sex Differences in Behavior 5
5.01.2.3 There Are Sex Differences in Physiology 6
5.01.2.4 There Are Sex Differences in Disease Susceptibility 6
5.01.2.5 There Are Sex Differences in Neural Structure, Glial Structure, and Connectivity 7
5.01.2.5.1 Neural Structure 7
5.01.2.5.2 Glial Cells 7
5.01.2.5.3 Synaptic Connectivity 8
5.01.2.6 There Are Sex Differences in Neurochemistry 8
5.01.2.6.1 Sex Differences in g-Aminobutyric Acid Signaling during Development 8
5.01.2.6.2 Sex Differences in Neurotransmitters in Adulthood 8
5.01.2.6.3 Sex Differences in Steroid Receptor Expression 9
5.01.2.7 Androgens as well as Estrogens Play a Role in Sexual Differentiation of the Brain 9
5.01.2.8 Sex Chromosome Complement Contributes to Sexual Differentiation 10
5.01.2.9 Sex Differences Are Context Dependent 11
5.01.2.10 Sexual Differentiation Depends on Four Key Processes 12
5.01.3 Recent Progress in Understanding the Four Key Processes 12
5.01.3.1 Neurogenesis 12
5.01.3.2 Cell Migration 14
5.01.3.3 Cell Death 14
5.01.3.4 Differentiation of Circuits 15
5.01.3.4.1 Axonal Growth 16
5.01.3.4.2 Dendritic Growth and Branching 16
5.01.3.4.3 Synaptogenesis 16
5.01.3.4.4 Differentiation of Neurochemical Phenotype 18
5.01.3.5 The Role of Epigenetic Mechanisms in the Four Key Processes 19
5.01.3.5.1 How Can Effects of Sexual Differentiation Last a Lifetime? 19
5.01.3.5.2 Effects of Manipulating Epigenetic Mechanisms on Sexual Differentiation 20
5.01.3.5.3 Sexual Differentiation of Epigenetic Marks across the Genome 20
5.01.4 Six Unanswered Questions 21
5.01.4.1 What Are the Actual Steroid Levels in the Brain during the Sensitive Period of Sexual Differentiation? 21
5.01.4.2 Do Noncoding RNAs Contribute to Sex Differences in Brain and Behavior? 22
5.01.4.3 Are Sex Differences Necessary? 22
5.01.4.4 Do Sex Differences in Brain Structure Beget Sex Differences in Brain Function? 23
5.01.4.4.1 ‘Canalization’ May Explain the Size and Variability of Sex Differences in Brain Structure and Behavior 24
5.01.4.5 Is Partner Preference Sexually Dimorphic? 24
5.01.4.6 Is the Human Brain Sexual Differentiated? 25
References 26

Abbreviations
AH Anterior hypothalamus GABA g-Aminobutyric acid
AVPV Anteroventral periventricular nucleus GH Growth hormone
BSTp Principal nucleus of the bed nucleus of the stria GnRH Gonadotropin-releasing hormone
terminalis INAH3 Third interstitial nucleus of the anterior
COX-2 Cyclooxygenase-2 hypothalamus
FAK Focal adhesion kinase LH Luteinizing hormone

Hormones, Brain, and Behavior, 3rd edition, Volume 5 http://dx.doi.org/10.1016/B978-0-12-803592-4.00091-2 3

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
4 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

LTP Long-term potentiation POA Preoptic area


MAP kinase Mitogen-activated protein kinase SDN Sexually dimorphic nucleus
MeA Medial amygdala SDN-POA Sexually dimorphic nucleus of the preoptic area
mPOA Medial preoptic area SNB Spinal nucleus of the bulbocavernosus
PGE2 Prostaglandin E2 VMN Ventromedial nucleus of the hypothalamus
PI3 kinase Phosphoinositide 3-kinase

5.01.1 Introduction and is irreversible (hence, differentiation). Phoenix and


colleagues called developmental effects of hormones organiza-
It is now over 55 years since publication of the seminal Phoenix, tional (Figure 1). In mammals, the focus of this chapter, sexual
Goy, Gerall, and Young manuscript (Phoenix et al., 1959), an differentiation occurs secondarily to sex determination, which
influential paper in which four scientists articulated the organi- is the process whereby the chromosomal or genetic sex of an
zational hypothesis of hormonally controlled sexual differentia- individual determines the fate of the bipotential gonad as
tion of the brain. For many years the quest to understand how either a testis or an ovary. The Sry gene on the Y chromosome
the brain differed in males and females and led to sex differences initiates testicular formation (Koopman et al., 1991), and the
in physiology and behavior was the exclusive domain of scien- process of sexual differentiation is then directed predominantly
tists self-identifying as ‘neuroendocrinologists’ or ‘behavioral by steroids produced by the gonads. However, the impact of
endocrinologists.’ This is beginning to change as major grant steroids is largely one sided in that it is the production of testos-
funding agencies in the United States, Canada, and Europe are terone that directs the brain (and body) toward a male pheno-
increasingly requiring the inclusion of both male and female type and the lack of testosterone production by the ovary that
subjects in preclinical or basic science research. The result is directs the brain (and body) toward a female phenotype. Many
a growing awareness combined with a burgeoning knowledge of the morphological and behavioral sex differences described
base of the pervasive impact of sex on the brain and its myriad below are due to organizational effects of hormones directing
functions. The goal of this chapter is to explore what we have the development of neural circuitry that will generate male-
learned in the years hence. We first list 10 concepts that have or female-typical functions and behaviors in adulthood.
stood the test of time and can safely be considered ‘known.’ Not all sex differences in brain function depend on organi-
The last of these is our view of what we consider the four core zational effects, however. As steroid hormones freely cross the
cellular processes that may be influenced by steroids during blood–brain barrier, sex differences in levels of circulating
sexual differentiation of the brain: (1) cell birth, (2) cell death, gonadal hormones can be expected to cause sex differences in
(3) cell migration, and (4) the differentiation of circuits. The neural function in steroid-sensitive areas at the time of
current challenge in the field is to identify the mechanisms by hormone exposure. For example, to engage in male sexual
which steroids alter these four core processes, and a substantial behavior, animals have to be exposed to testosterone not
portion of this chapter is devoted to recent advances in that only during development, but in adulthood as well. This adult
domain. Progress is being made, but much remains to be effect is transient, and Phoenix and colleagues therefore called
done. Moreover, there are fundamental questions remaining, such effects activational (Figure 1). There are also activational
and we conclude this chapter by highlighting several of these. effects of steroids that are independent of developmental
effects. For example, male baboons show more yawning than
do females. This sex difference appears to depend entirely on
5.01.2 Ten Things We Know sex differences in circulating testosterone as castrated male
and ovariectomized female baboons yawn at equally high
Making a list is both appealing and daunting. It is appealing levels after testosterone treatment in adulthood (Phoenix and
because it highlights what we have accomplished and gives Chambers, 1982). Similarly, some sex differences in brain
us confidence in the foundations on which further studies are morphology are due to activational effects of steroids (see
built. It is daunting because one might miss something and below), and there is emerging evidence that sex differences in
inadvertently leave a critical finding off the list. It is also daunt- cognition may be largely a function of steroids circulating in
ing because it requires that we examine the strength of a partic- adulthood (reviewed in McCarthy and Konkle, 2005). In this
ular finding in the harsh light of new data. Do the initial chapter, we will refer to ‘organizational’ and ‘activational’
conclusions still hold? Were the techniques used appropriate effects as ‘programming’ and ‘acute,’ respectively, to bring
for the question? Are the findings robust and reliable? Both them in line with terminology used in other areas of develop-
the appealing and daunting aspects serve a useful purpose, mental biology (cf. Forger et al., 2015a).
and with this in mind, we venture forward. It is important to distinguish ‘sexual differentiation’ from
‘sex differences.’ There are many differences between adult
males and females that are a consequence of transient changes
5.01.2.1 Hormones Cause Sex Differences by Acting during
in hormones, reproductive status, or age; these are sex differences,
Development as well as in Adulthood
but are not by our definition sexually differentiated traits. Although
For the purposes of this chapter, we define sexual differentia- one could argue with our somewhat restricted view of hormon-
tion as a process that occurs during a restrictive developmental period ally mediated sexual differentiation of the brain, it serves as

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 5

Figure 1 Sexual differentiation of behavior occurs during a perinatal sensitive period. In rodents the parameters of the sensitive period for sexual
differentiation are operationally defined with the opening of the window being the onset of androgen production by the fetal testis of males late in gesta-
tion and a second surge around birth. The closing of the sensitive period is the developmental time point at which females can no longer be masculin-
ized by treating them with exogenous testosterone. The developmental hormone exposure is akin to early-life programming and determines adult
hormonal effects on sexual and parental behaviors. During the juvenile period, there are not gonadal steroids present but there is a sex difference in
play intensity and frequency which was established by the early-life programming effects of steroids. In rodents, testosterone is aromatized to estradiol
within the brain, and many effects of masculinization are due to the actions of estradiol, while others are directly the result of androgen action.

a useful construct on which to base a discussion of underlying Males and females are frequently viewed as differing along
mechanisms. a continuum, with maleness at one end and femaleness on
the other. Individual phenotypes could exist anywhere along
the continuum but on average, males would predominate at
5.01.2.2 There Are Sex Differences in Behavior
one end and females at the other. For instance, maternal
Sex differences in behavior are ubiquitous among sexually behavior, as demonstrated by pup retrieval, is much more prev-
reproducing species and presumably contribute in some way alent in female rats (that have recently given birth) than in
to the reproductive success of the organism. Males and females males, which almost never spontaneously retrieve pups. Alter-
have evolved different reproductive strategies, necessitating natively, the sexes may differ on a dependent measure, but only
different morphologies and behaviors. In mammals, sex differ- slightly. In standardized tests of language, for example, girls
ences are commonly seen in sexual behavior, aggression, play score significantly higher than boys, although in this case the
behavior, parental behavior, and performance on cognitive mean sex difference is relatively small, and there is substantial
tasks, to name just a few examples. Since behavior is controlled overlap between the sexes (see Spelke, 2005 for review). None-
by the nervous system, it would follow that the brain must theless, both of these behaviors might be said to vary along
differ in males and females. However, sex differences in a single masculine–feminine axis.
behavior were recognized long before we considered the brain In other cases, multiple axes have been proposed. The best
to be a sexually dimorphic organ. A scientific approach to the example of this is the case of sex behavior in the laboratory
topic began with observations of effects of hormones on rat where sexual differentiation is often viewed as involving
behavior in domesticated animals such as roosters and two axes and at least three distinct processes: (1) feminization,
extended to controlled laboratory studies of inbred rodents the default developmental program required for a female
in the early 1900s. We now consider it self-evident that behavioral response in adulthood (i.e., the display of lordosis
hormones act on the brain to cause sex differences in behavior, under the proper hormonal milieu), (2) masculinization,
but this was not the initial presumption. A school of thought a hormonally induced process allowing for male sexual
prevailed for some time that held that hormonal effects on behavior in adulthood (e.g., the display of mounting,
peripheral structures, such as the penis, determined behavior thrusting, and ejaculation), and (3) defeminization, a hormon-
(Beach et al., 1969). In fact, the iconic Phoenix et al. (1959) ally induced process to eliminate, mask, or override the femini-
paper laying out the organizational/activational hypothesis of zation process. Emerging evidence suggests that the processes
hormone action makes little reference to the brain, but instead of masculinization and defeminization occur relatively inde-
refers to the ‘tissues’ controlling behavior. There is now ample pendently and may target distinct neural circuitry. Although
evidence, discussed below, that these ‘tissues’ include the brain. males are normally both defeminized and masculinized, it is
Relevant to the understanding of sex differences in behavior possible to separate these. In mice, for example, estrogen
is defining the axis or axes along which the behavior can vary. receptor alpha activation appears required for masculinization,

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
6 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

while estrogen receptor beta has a major role in defeminiza- caused a permanent absence of cyclicity in female rats
tion, so that mice with a mutated estrogen receptor beta gene (Barraclough and Leathem, 1954; Barraclough, 1961). After
are masculinized, but not defeminized (Kudwa et al., 2006). identifying the anterior preoptic area (POA) as the site respon-
Recent evidence suggests that the cellular mechanisms medi- sible for cyclicity in gonadotropic hormone release, they
ating masculinization and defeminization are also unique concluded: “the anterior preoptic area is undifferentiated at
(see below). birth with regard to its subsequent control of gonadotropin
One could make the argument that other examples of sexu- secretion.” They also concluded that in the absence of testos-
ally dimorphic behavior, such as juvenile play, also vary along terone the anterior POA differentiates to sustain cyclicity,
multiple axes. For instance, rough and tumble play by juveniles whereas in the presence of testosterone it “becomes refractory
is more intense and more frequent in males than females, to both intrinsic and extrinsic activation, and the more tonic
across many species (Meaney, 1989; Hines, 2006). Although type of male gonadotropin secretion is observed” (Barraclough
males show more rough play, females are not idle while the and Gorski, 1961). We now know that in response to gonado-
males are playing; they are doing something else (in monkeys, tropin-releasing hormone (GnRH) from the anterior POA, the
grooming-type ‘play’; in humans, this might be doll play). If pituitary synthesizes and releases the gonadotropins, luteiniz-
‘play’ is looked at comprehensively, this could suggest two ing hormone (LH), and follicle-stimulating hormone. Both
axes (e.g., rough play and grooming play), which may, like act at a distant target, the gonads, and are essential links in
male and female sexual behavior, be separable. Are males the hypothalamic–pituitary–gonadal axis.
masculinized on rough play and defeminized on doll play? If As with sex behavior, the female pattern, defined as the
so, this would make play analogous to the multiple axes view ability to respond to an increase in estradiol level with a surge
of sex behavior. If not, this raises the question of whether the in LH release (positive feedback), is the default, and the male
differentiation of sexual behavior is really qualitatively pattern (defined as lack of positive feedback to estradiol) is
different from the differentiation of other behaviors, such as induced by the actions of testicular steroids during develop-
play. For example, if female sex behavior and male sex behavior ment. When systematically compared, however, the sensitive
are considered as two different behaviors entirely, then we are period for steroid-mediated behavioral defeminization ends
back to single axes with each individual more feminine or earlier than the sensitive period for defeminization of gonado-
masculine on each of these two behaviors; the term ‘defemini- tropin secretion (Diaz et al., 1995).
zation’ would not be necessary. There are also interesting sex differences in the secretion of
A reader of the sex differences literature will often find other pituitary hormones, including prolactin and growth
defeminization used interchangeably with masculinization, or hormone (GH). For example, males have more frequent and
demasculinization used synonymously with feminization. The higher amplitude pulses of GH than females (Norstedt and
term demasculinization is generally not used when discussing Palmiter, 1984), and these differences are determined at least
the differentiation of sex behavior, because the masculinization in part by gonadal secretions during perinatal life (Jansson
of behavior has historically been viewed as an active process, et al., 1984). The sex difference in GH secretion has important
and you cannot deprogram a process that has to be induced implications for liver function and body weight: about
in the first place. However, work on the cellular underpinnings 75–85% of the genes in the liver exhibit a sex difference in
of sexually dimorphic behaviors suggests otherwise. Neurons expression (Rinn and Snyder, 2005), and these differences are
underlying at least some aspects of male sexual behavior indirectly driven by the sex difference in GH secretion pattern.
initially form in both sexes and then die in females, analogous Thus sexual differentiation of the secretion pattern of the ante-
to differentiation of the Wolffian ducts. This suggests that a case rior pituitary has lifelong consequences for the entire body.
for demasculinization can be made. Thus, while few would argue
that there are lasting and important differences in behavior
5.01.2.4 There Are Sex Differences in Disease Susceptibility
between the sexes, there is less agreement about how terms
are applied. One can argue whether this is ‘just semantics,’ or The literature on human neurological and psychiatric disorders
a significant impediment to clear thinking about the processes demonstrates some of the most compelling evidence for sex
involved. differences in the brain. Many, if not most, neurological
diseases exhibit sex differences in incidence, severity, or disease
course (see Table 1). For example, men are much more likely
5.01.2.3 There Are Sex Differences in Physiology
than women to be diagnosed with Parkinson’s disease
Even before the work of Phoenix et al. (1959), there was (Strickland and Bertoni, 2004; Shulman, 2007), while females
evidence that the brain was sexually differentiated. Female outnumber males in the incidence of Alzheimer’s disease and
mammals exhibit profound cyclical variations in the release multiple sclerosis (Fratiglioni et al., 1997; Barrett, 1999; Swaab
of gonadal and pituitary gonadotropic hormones, whereas et al., 2003; Schwendimann and Alekseeva, 2007). Among
males do not. Harris and Jacobsohn (1952) demonstrated psychiatric disorders, women exhibit higher rates of anorexia,
that male pituitary transplants placed directly under the hypo- bulimia, depression, and essentially all anxiety disorders
thalamus sustained ovarian cycles in female rats. This dis- (Kessler et al., 1994; Kornstein et al., 2002; Nestler, 2002;
proved earlier claims that the pituitary was responsible for Altemus, 2006; Södersten et al., 2006), whereas men show
sex differences in gonadotropic hormone release and instead a greater incidence of gender identity disorder, early onset
pointed to the brain as the site of sexual differentiation. Barra- schizophrenia, antisocial personality disorders, and conduct
clough and his collaborators then showed that a single injec- disorders (Häfner et al., 1993; Kessler et al., 1994; American
tion of testosterone propionate given shortly after birth Psychiatric Association DSM-IV-TR, 2000).

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 7

Table 1 Many neurological and psychiatric disorders connectivity (e.g., axon projection patterns or synapse
exhibit a sex difference in incidence number), or neurochemistry (e.g., the expression of neuropep-
tides, neurotransmitters, or receptors).
Disorder Relative incidence

Developmental disorders 5.01.2.5.1 Neural Structure


Attention-deficit hyperactivity M>F Several of the earliest recognized and most intensively studied
Asperger’s syndrome M>F neural sex differences were differences in the volume of cell
Autism M>F groups. For example, as first described by Nottebohm and
Dyslexia M>F Arnold (1976) over 40 years ago, there are striking sex differ-
Gilles de la Tourette M>F ences in the volume of several nuclei controlling singing in
Mental retardation M>F
canaries and zebra finches. These volume differences result
Stuttering M>F
from sex differences in cell number, soma size, and dendritic
Adult-onset disorders
Alzheimer’s disease F > Ma complexity (reviewed in Arnold, 1992). Similarly, the sexually
Amyotrophic lateral sclerosis M>F dimorphic nucleus of the medial preoptic area (SDN-POA) of
Anorexia nervosa F>M rats is several times larger in volume in males than in
Antisocial personality M>F females, and this sex difference is also reflected by sex
Bulimia F>M differences in the number and size of neurons in this region
Conduct disorder M>F (Gorski et al., 1978, 1980). Homologous sex differences in
Depression F>M cell groups of the POA are seen in many mammals, including
Gender identity disorder M>F gerbils, guinea pigs, ferrets, sheep, hyenas, monkeys, and
Multiple sclerosis F>M
humans (reviewed in Forger, 2001).
Parkinson’s disease M>F
One of the best-understood sex differences in neural struc-
Schizophrenia (early onset) M>F
Sleep apnea M>F ture involves a sexually dimorphic neuromuscular system.
The spinal nucleus of the bulbocavernosus (SNB) is a cluster
a
Estimates of gender-specific incidence for Alzheimer’s disease of motoneurons in the lumbar spinal cord that innervates stri-
vary. After accounting for sex differences in life span, the risk ated muscles involved in copulation. Although first described
appears to be between 1.5 and 3.0 times higher in women.
in rats (Breedlove and Arnold, 1980), males of many species
have more SNB motoneurons (or the homologous motoneu-
rons in Onuf’s nucleus) than do females. Perhaps owing to
Perhaps most striking, essentially all neurodevelopmental disor-
the relative simplicity of the circuit and the accessibility of
ders are more prevalent in boys than in girls (Gissler et al., 1999;
motoneurons for experimental manipulation, more may be
Zup and Forger, 2002; Swaab et al., 2003; Rutter et al.,
known about cellular mechanisms involved in sexual differen-
2003). This is true for cerebral palsy, mental retardation,
tiation of the SNB than for any other neural sex difference
dyslexia, stuttering, phonological disorder, sleep terrors, atten-
(reviewed in Sengelaub and Forger, 2008). Well-established
tion-deficient disorder, Asperger’s syndrome, autism, and tic
sex differences in volume and/or cell number also exist in the
disorders. In fact, of all the disorders of infancy or childhood
medial amygdala (MeA; reviewed in Cooke, 2006), anteroven-
listed in the Diagnostic and Statistical Manual of Mental Disorders
tral periventricular nucleus of the hypothalamus (AVPV), and
(DSM-IV-TR), published by the American Psychiatric
principal nucleus of the bed nucleus of the stria terminalis
Association, the only exception to the pattern is Rett
(BSTp; reviewed in Forger, 2006). In the AVPV and BSTp, the
syndrome, which affects girls almost exclusively. However,
differences are sexually differentiated, as the sex differences in
this is due to the fact that the disorder is X chromosome-
volume and cell number can be reversed by appropriate
linked; male fetuses inheriting the Rett syndrome mutation
gonadal steroid hormone treatments during development;
are more severely affected than are female fetuses, and rarely
MeA volume, on the other hand, may be an example of a sex
survive to term (Schanen, 1999).
difference due principally to acute effects of hormones (Cooke
The male bias in neurodevelopmental disorders is especially
et al., 1999).
notable since sex differences in childhood disorders are less
likely than are adult-onset disorders to be due to differences
5.01.2.5.2 Glial Cells
in circulating hormones or to experience, and are more likely
Structural sex differences in the nervous system are not limited
to reflect pre- or postnatal brain development. The bias is not
to neurons. For example, astrocytes in the arcuate nucleus and
well understood and strikes us as one of the most understudied
POA of rats are much more complex in males than in females,
areas in psychiatry or medicine.
and this sex difference, too, is dependent on differential expo-
sure to gonadal steroid hormones in males and females
(reviewed in McCarthy, 2008). Another cell type, the microglia,
5.01.2.5 There Are Sex Differences in Neural Structure, Glial
which is actually an immune cell, is also found in different
Structure, and Connectivity
forms in males and females in some brain regions, and this
Many of the sex differences in behavior, physiology, and neuro- difference is determined by early gonadal steroid exposure. In
logical disorders described above can presumably be traced to the POA, microglia are critical contributors to the establish-
sex differences in the nervous system. These neural sex differ- ment of a higher density of dendritic spine synapses on
ences may take many forms, but can be broadly conceptualized neurons of males (Lenz et al., 2013), highlighting the impor-
as differences in structure (e.g., the size or number of neurons), tance of cell-to-cell communication.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
8 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

5.01.2.5.3 Synaptic Connectivity permeable ionophore. Estrogens enhance the depolarizing


A persistent problem in the study of sex differences in the brain action of GABA by increasing intracellular chloride. As a result,
is the impact of technical limitations directing the nature of the when the channel opens, the shift in the membrane potential is
research, and this is particularly true in the study of synaptic greater than it would be without the action of estradiol, and
connectivity. There are several ways to quantify synapses. The there is a stronger and more enduring opening of voltage-
gold standard is electron microscopy as this is the only way gated calcium channels (Nunez and McCarthy, 2008). Since
to see synaptic vesicles and thereby say definitively that males have higher estradiol content in the brain during the
a synapse is truly present. This technique was used very effec- developmental period when GABA action is depolarizing,
tively by Arai and Matsumoto in the early 1980s in a series of they also have greater intracellular calcium influx. This can be
reports revealing profoundly sexually dimorphic patterning in both trophic and promote the growth of axons, dendrites,
the arcuate nucleus, ventromedial nucleus of the hypothalamus and synapses, but can also be excitotoxic if in excess (Auger
(VMN), and amygdala (for review, see Matsumoto, 2000). The et al., 2001; Nunez et al., 2003). Moreover, early postnatally,
down side of quantitative electron microscopy is the small testosterone and its metabolite estradiol elevate the levels of
region sampled and the labor involved, making it difficult to GABA and its synthesizing enzymes, GAD 65 and GAD 67, in
examine entire brain regions or to compare large numbers of several hypothalamic nuclei and the hippocampus of male
animals. Golgi impregnation of neurons offers a way to more rats (Davis et al., 1996c, 1999). So males both make more
quickly quantify the number or density of dendritic GABA and respond to it more strongly, a scenario which
spines and to completely reconstruct dendritic trees. Dynamic increases excitation broadly in the developing male brain and
changes in spine density across the life span and a sex difference may contribute to the greater vulnerability of the male to dele-
in branching were revealed in the VMN using this technique terious effects of insult or injury.
(Pozzo-Miller and Aoki, 1991; Mong et al., 1999; Todd et al.,
2007). Disadvantages of Golgi impregnation, however, are 5.01.2.6.2 Sex Differences in Neurotransmitters in Adulthood
that only dendritic spines, and not synapses, can be measured, Numerous sex differences in neurotransmitter and neurotrans-
and there is also the risk that what is presumed to be a random mitter receptor expression have been reported throughout the
selection process (only a subset of neurons take up the Golgi brain. These include modulatory systems not traditionally
material), may actually reflect a bias that we do not yet under- linked with reproductive behavior, such as histamine neurons
stand. Quantitative Western blot for synaptic proteins can serve in the hypothalamus, noradrenaline neurons in the locus
as a proxy measure of synapses, and when used to assess the coeruleus, serotonin neurons in the dorsal raphe, and dopa-
protein spinophilin, which is heavily localized to dendritic mine in the substantia nigra and ventral tegmental area in
spines, has proved useful in cross-group comparisons animals as well as humans (De Vries, 1990; De Vries and
(Amateau and McCarthy, 2002a). Again, however, only spine Simerly, 2002; Cosgrove et al., 2007). The ways in which
synapses are detected, leaving completely unexplored the neurotransmitter systems can vary are continuing to be discov-
hormonal regulation of axosomatic synapses. In summary, ered. In a manner similar to that seen for depolarizing actions
no single technique is perfect, and our ability to quantify of GABA during development, a component of the stress axis is
specific types of synapses across groups is limited. Nonetheless, profoundly different in males and females as a result of
we actually do know a fair amount about sex differences in a change in sensitivity to corticotropin-releasing factor (CRF),
synaptic profiles, which are proving to be robust and may be which acts as a modulatory neuropeptide within the brain.
of profound significance (see Section 5.01.3.4.3). Adrenergic neurons of the locus coeruleus express receptors
for CRF, and when an animal is stressed, the receptors are inter-
nalized away from the membrane in males but trafficked to the
5.01.2.6 There Are Sex Differences in Neurochemistry
membrane in females (Bangasser et al., 2010). The impact of
There is a wealth of information on sex differences in neuro- this differential distribution of CRF receptor is further ampli-
chemistry. In fact, some of the earliest reports on sex differences fied by ramped-up noradrenergic transmission in females
in the brain concerned higher serotonin levels in female versus (Bangasser, 2013; see also Forger et al., 2015b for further
male brains (Kato, 1960). Perusal of PubMed suggests that at discussion). This example is probably one of many as
least a quarter of the reports on sex differences in the brain researchers become increasingly aware of the importance of
concern sex differences in neurochemistry (e.g., levels and sex as a biological variable in modulating the nervous system.
distribution of neurotransmitter and their receptors, steroid Because of the widespread influences of these modulatory
receptors, enzymes involved in steroid and neurotransmitter systems, the impact of the sex differences could be profound
synthesis and catabolism, growth factors, and cytokines). and may be related to the development of behavioral disorders
Several concrete examples are offered below. that show sex differences in incidence, as described in Section
5.01.2.4 (see also Cosgrove et al., 2007; Becker and Hu, 2008).
5.01.2.6.1 Sex Differences in g-Aminobutyric Acid Signaling Sex differences in neurochemistry may also offer clues as to
during Development the significance of sex differences in brain structure. An
Contrary to its ubiquitous inhibitory effect in the mature brain, example of this concerns the AVPV. Female mice have more
g-aminobutyric acid (GABA) has emerged as the dominant neurons in the AVPV than males (Forger et al., 2004). A subset
source of excitation in the developing brain. This is due to of these neurons express Kiss1 mRNA and its gene product,
a reversal of the transmembrane chloride concentration kisspeptin, and females mice have 10 times as many kisspeptin
gradient, resulting in depolarization of the membrane poten- neurons in the AVPV as do males (Clarkson and Herbison,
tial upon opening of the GABAA receptor which is a chloride 2006). In rats, the sex difference in kisspeptin expression is

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 9

close to absolute, with males expressing almost no kisspeptin the almost absolute sex difference in the expression of proges-
(Kauffman et al., 2007). In both species, this sex difference terone receptors in the medial POA (mPOA) around birth.
depends on early exposure to gonadal steroids (Kauffman Perinatally, male rats express a much higher level of proges-
et al., 2007; Gonzalez-Martinez et al., 2008). Kisspeptin- terone receptors than females (Wagner et al., 1998). This
containing projections are found in close apposition to suggests that exposure of pups to maternal progesterone during
GnRH neurons (Clarkson and Herbison, 2006), and kisspeptin gestation may have different consequences in males than in
triggers an LH surge in both rats and mice by stimulating GnRH females. The role of this sex difference in sexual differentiation
neurons (Gottsch et al., 2004; Irwig et al., 2004). Thus, the sex of the mPOA is unclear, but treating males postnatally with the
difference in kisspeptin cell number may contribute to the sex progesterone receptor blocker, RU486, reduced male sexual
difference in LH release described in Section 5.01.2.3. Interest- behavior in adulthood (Lonstein et al., 2001). Although other
ingly, estradiol treatment increases Kiss1 mRNA in the AVPV interpretations are possible, these data are consistent with the
but decreases it in the arcuate nucleus, which contains a nondi- idea that activation of the differently expressed progesterone
morphic group of kisspeptin neurons; it has been suggested receptor is involved in sexual differentiation of the brain.
that the former group contributes to the surge and the latter
to the negative feedback of gonadotropin secretion (Dungan
5.01.2.7 Androgens as well as Estrogens Play a Role
et al., 2006). It is tempting to speculate that the higher number
in Sexual Differentiation of the Brain
of kisspeptin neurons in the AVPV gives females the ability to
respond to estradiol treatment with an LH surge, an ability In early work on sexual differentiation, testosterone was shown
that males lack. Kisspeptin is probably not the only factor, to masculinize and defeminize the behavior of female rats or
however. For example, female rats also have twice as many guinea pigs (Phoenix et al., 1959; Beach et al., 1969). Testos-
neurons in AVPV that coexpress markers of both glutamatergic terone is an androgen, which acts primarily by binding to intra-
and GABAergic signaling (Ottem et al., 2004). These dual cellular androgen receptors. Thus, the early literature
phenotype neurons may synapse on GnRH neurons to control understandably often discussed effects of ‘androgenization,’
the switch from negative to positive feedback of estradiol that or the ‘androgenized’ female. However, it soon became evident
occurs around ovulation (see also Section 5.01.3.4.1). that estrogens, in particular 17b estradiol (hereafter, ‘estra-
diol’), mimicked many effects of testosterone, and usually at
5.01.2.6.3 Sex Differences in Steroid Receptor Expression lower doses. This was somewhat puzzling because all mamma-
Sex differences in gonadal steroid receptor expression have lian fetuses are exposed to elevated estrogens, produced by the
been reported for several forebrain nuclei of adult mice and maternal ovaries; how are female fetuses protected from the
rats. For example, androgen receptor expression is greater in effects of estrogen exposure? The apparent paradox was
the BSTp of males than of females (Simerly et al., 1990a; resolved by two suggestions: first, that a-fetoprotein, an
Roselli, 1991; Shah et al., 2004). In this case, the mechanism estrogen-binding protein produced at high levels in developing
may be related to the same mechanism that causes the differ- animals, binds circulating estrogens and prevents them from
ence in overall cell number in the BSTp (i.e., cell death) because entering the brain. Second, the aromatization hypothesis
mice lacking the prodeath gene Bax do not show a sex differ- proposed that testosterone, released by the perinatal testes,
ence in overall cell number (Forger et al., 2004; see Section crosses the blood–brain barrier and is converted to estradiol
5.01.3.3.), or in the number of androgen receptor expressing in target cells by the aromatase enzyme; estradiol then acts
cells in the BSTp (Holmes et al., 2009a). For other brain areas, via estrogen receptors to masculinize the brain and behavior
different mechanisms seem to apply. For example, estrogen (for reviews, see Goy and McEwen, 1980; Lephart, 1996).
receptor immunoreactivity is about 20 times greater in the The importance of a-fetoprotein has recently been
BSTp of adult, gonadally intact female mice than of males, confirmed by examining the brain and behavior of a-fetopro-
and in this case the difference is due to the acute suppression tein knockout mice. Females that develop in the absence of
of receptor expression by testicular hormones (Kelly et al., functional a-fetoprotein show very low, male-like levels of
2013). Estrogen and progesterone receptor binding and/or lordosis as adults and high levels of male sexual behavior, as
mRNA expression in the VMN, preoptic periventricular, and well as male-like neuroendocrine responses (Bakker et al.,
medial preoptic nuclei also appear to be higher in females 2006; González-Martínez et al., 2008; Taziaux and Bakker,
(Simerly et al., 1990; De Vries and Simerly, 2002), but in 2015). In support of the aromatization hypothesis, brain
many of these cases, it is unclear whether the sex difference is masculinization can be blocked by treating neonatal male
caused by programming or acute effects of gonadal hormones. rats with an aromatase inhibitor or estrogen antagonist (e.g.,
Regardless of the underlying mechanisms, if expression of Döhler et al., 1986); the SDN-POA is feminized in male rats
hormone receptors causes differences in hormone sensitivity, treated with antisense oligonucleotides to the estrogen
they are likely to contribute to sex differences in neural receptor (McCarthy et al., 1993); and the behavior of mice is
function. demasculinized or feminized by a knockout of the estrogen
This may also be the case for sex differences in steroid receptor a or b gene, respectively (Kudwa et al., 2006). These
receptor expression during development. Gonadal hormones findings are bolstered by reports that nonaromatizable
cause sex differences in the expression of estrogen receptors, androgens, such as dihydrotestosterone, in many cases do not
androgen receptors, and aromatase in the developing hypothal- mimic effects of testosterone on the brain or behavior.
amus (MacLusky et al., 1985; DonCarlos and Handa, 1994; The SNB has long been known as an exception to the rule
McAbee and DonCarlos, 1998), suggesting that the hormones that estrogenic metabolites of testosterone are responsible for
influence their own subsequent effectiveness. A special case is masculinization and defeminization of the rodent brain and

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
10 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

behavior. SNB cell number in rats was the first neural sex differ- gonads into testes or ovaries; the resulting differences in
ence shown to differentiate under the control of androgens, gonadal secretions cause all other differences.”
acting via classical intracellular androgen receptors (Breedlove The Jost doctrine, however, may not be as generally appli-
and Arnold, 1981; Breedlove et al., 1982). It is now clear that cable as first thought. Several sex differences, for example, in
androgens act at virtually every level of the SNB system, in bird plumage, birdsong, and the development of pouch versus
development and throughout adult life (reviewed in Sengelaub scrotum in wallabies, are likely to be caused directly by sex
and Forger, 2008). Although for some time the SNB was chromosomal genes and not by differences in gonadal
considered the only morphological sex difference that depends hormones (Arnold, 1996). To test whether this may also apply
on androgens and androgen receptors, more recently new roles to sex differences in the brain, Arnold and colleagues intro-
of androgens have been identified. For example, male rats lack- duced a model system (known as the Four Core Genotypes)
ing functional androgen receptors exhibit partial demasculini- that can distinguish between direct actions of sex chromosomal
zation of the volume of the BSTp, VMN, and posterodorsal complement (XX vs XY) and gonadal hormones on sexually
MeA (Morris et al., 2005; Durazzo et al., 2007; Dugger et al., dimorphic traits (De Vries et al., 2002). Female mice with the
2007). Androgens may also maintain normal hippocampal familiar XX genotype were crossed with males with an XY
structure in male rats and monkeys (MacLusky et al., Sry genotype. These mice lacked the Sry gene on the Y chromo-
2006) and play a role in the sexual differentiation of vaso- some, which normally directs the growth of the testis, but
pressin innervation in the brain (Han and de Vries, 2003). developed a male phenotype anyway because they had an Sry
This should not come as a surprise, since it has been known transgene inserted on an autosomal chromosome. The
for many years that androgens are required for complete offspring included XX and XY mice of either gonadal sex
masculinization of behavior, including juvenile play (Meaney, depending on whether they inherited the autosomal Sry trans-
1989) and male sexual behavior in some strains of rats (e.g., gene or not. Comparing XX and XY mice within sex (defined
van der Schoot, 1980; the full masculine phenotype can be on basis of gonad) revealed a number of differences that appear
achieved with neonatal estradiol treatment in other strains, to be caused by differences in sex chromosomal complement
however, see Amateau and McCarthy, 2004; Todd et al., independent of gonad type. For example, vasopressin expres-
2005). More recently, androgen receptors have been implicated sion is more masculinized in XY females than in XX females,
in the sexual differentiation of social and olfactory preferences and in XY males than in XX males (De Vries et al., 2002). Addi-
in mice (Bodo and Rissman, 2007, 2008) and the neuroendo- tional effects on aggression, habit formation, body weight, and
crine response to a mild stressor in rats (Zuloaga et al., 2011). response to brain injury have since been detected (for reviews,
Importantly, although estrogens are critical for differentia- see Arnold and Chen, 2009; Forger et al., 2015a). Thus we now
tion of the rat and mouse brain, androgens may play a more know that while hormones act broadly throughout the brain,
central role in sexual differentiation of the primate brain (see every cell in the brain has a genetic sex as a result of chromo-
Wallen, 2005; Thornton et al., 2009). Human a-fetoprotein some complement (Figure 2).
does not have a high affinity for estrogens (Swartz and Soloff, The introduction of the Four Core Genotypes model
1974), suggesting that the brains of human fetuses of both provides an important new tool for distinguishing roles of
sexes are exposed to estrogens. Experiments of nature also hormones and genes in establishing or maintaining sex differ-
cast doubt on the importance of estrogens for sexual differenti- ences. While first viewed as a means for identifying specific
ation in primates. About half a dozen men with inactivating genes on the X or Y chromosome that direct sex-specific devel-
mutations of the aromatase gene have been reported, and all opment, the model has led to far more nuanced interpretations
appear to be heterosexual with normal sexual function of the ways that the sex chromosomes contribute to sex differ-
(reviewed in Jones et al., 2006), although overall libido may ences. One example concerns the X chromosome. In females,
be reduced (Carani et al., 2005). A role for androgen receptors one X chromosome in each cell is largely silenced in a process
in sexual differentiation of humans is supported by the obser- referred to as random X inactivation, but this process is not
vation that genetic males (XY) with an inactivating mutation of complete, with as many as 15% of X genes in humans escaping
the androgen receptor gene are not only phenotypically female inactivation (Berletch et al., 2010). This means that female cells
with respect to body type, but also female-typical in terms of have a ‘double dose’ of some X genes. But there are also
gender identity, sexual orientation, and gender role (Hines multiple other means by which the X can impact an organism.
et al., 2003; Hines, 2008). These include parent of origin allelic variation (also known as
imprinting), cellular mosaicism (a result of random inactiva-
tion of the maternal versus paternal X), and the creation of
5.01.2.8 Sex Chromosome Complement Contributes
an epigenetic sink whereby the inactive X sequesters a dispro-
to Sexual Differentiation
portionate amount of the proteins that regulate gene expres-
Alfred Jost’s proposal in the 1940s (Jost, 1947) that testes are sion (see for review Arnold et al., 2016).
crucial for the development of the male phenotype, and that The importance of the number of X chromosomes
without them the body develops in a female direction, has becomes apparent in consideration of XO women, referred
been driving research in the field of sexual differentiation for to as Turners Syndrome, and XXY men, referred to as Kleinfel-
more than half a century. The more recent discoveries of genes ter’s Syndrome in humans. It is difficult to determine,
that direct the differentiation of the primordial gonad into the however, whether the phenotypic differences in these
testis (Koopman et al., 1991) only require reformulating the syndromes are due to an altered sex chromosome comple-
theory on the mechanisms of sexual differentiation into: “sex ment or due to changes in gonadal hormones, because in
chromosomal genes determine the differentiation of the both cases there is a retardation or regression of gonadal

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 11

Figure 2 Both hormones and chromosome complement contribute to sex differences. The mode and means of hormone action have greatly
expanded in the past decade as we now know that steroids can act both via nuclear transcription factor receptors and membrane receptors. Steroids
can also be locally synthesized within the brain, which challenges us to understand what truly local levels of steroid concentrations are. We now also
appreciate that every cell in the brain has a sex due to the chromosome complement being either XX or XY in mammals. Many effects of chromo-
some complement have been identified but precisely how they exert these effects is still being determined. Interactions between steroid hormones
and chromosome complement further add to the complexity.

development. Arnold and colleagues have addressed this of males (Wersinger et al., 1997; Jyotika et al., 2007), although
confound with a variant on the Four Core Genotypes model other groups find clear a clear sex difference of male sexual
called the Sex Chromosome Trisomy model which produces behavior in mice (Vale et al., 1973; Bakker et al., 2006). Strain
XX, XY, XXY, and XYY mice. By focusing on genes that escape differences may account for this variability, but testing condi-
X inactivation and would therefore be elevated in XXY versus tions are also likely to play a role. For example, sex differences
XYY or XY males they hope to identify which genes explain the in response to sleep deprivation depend on whether animals
multiple phenotypic effects evident in individuals with sex are stressed or not, with stressed animals showing bigger sex
chromosome aneuploidies (Arnold et al., 2016). differences in sleep recovery than unstressed animals (Koehl
A different approach is to examine naturally occurring sex et al., 2006). Developmental history may also contribute to
chromosome aneuploidies in humans and ask how they variability. Male prairie voles are spontaneously parental,
impact brain parameters compared to normative controls. whereas virgin females avoid or attack pups (Lonstein and
Cortical thickness asymmetries are a well-documented pheno- De Vries, 1999a). This sex difference depends on rearing condi-
typic variable in the frontal and occipital cortex of humans tions, however, because if females are raised to adulthood in
(Lin et al., 2015) and in some regions are found only in males. the presence of their parents, they too are parental (Lonstein
The number of sex chromosomes impacted many of the and De Vries, 2001). Even what appear to be very subtle
cortical thickness asymmetries observed in a highly region- changes in developmental history can make a big difference.
specific and divergent manner for X versus Y, but quite inter- In one study, prairie voles showed sex differences in partner
estingly did not alter those regions in which there was a sex preference formation, or not, depending on whether they
difference. Moreover, the same aneuploidy (i.e., an extra were transported by hand or in a plastic cup during routine
X chromosome) had different effects in males versus females, cage changing (Bales et al., 2007).
highlighting the complex interplay of genetic and gonadal sex Context also plays a role in sex differences in human
(Raznahan et al., 2015). behavior. For example, one of the most consistent cognitive
sex differences is found in the Mental Rotations Test. In this
famous test, subjects have to mentally rotate a block figure to
5.01.2.9 Sex Differences Are Context Dependent
match it with a congruent object in a line-up of similarly
If the pioneers of our field had chosen to work with mice rather shaped but not identical objects. Males outperform females
than with rats and guinea pigs, things might have been very on this task in a wide range of studies. Interestingly, a much
different. Mice do not show consistent sex differences in smaller male advantage is seen if instead of interconnected
male sexual behavior. For example, several reports suggest cubes, the figures take on a human shape (Alexander and
that female mice treated with testosterone or estradiol in adult- Evardone, 2008). Similarly, the male advantage in certain
hood show mounting and thrusting in response to a receptive math tests is eliminated or reduced if female subjects are told
female at levels that are similar to, or even higher than, those in advance that females do as well or better than males on these

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
12 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

tests, but sex differences are exacerbated if females are told the programmed, by gonadal steroids. Such a difference could arise
reverse (for review, see Spelke, 2005). In all of these cases because steroid hormones cause more neurons destined for the
context, whether defined as the conditions during the test or brain region of interest to be born in males (i.e., a hormonal
those leading up to the test, determines the outcome. To effect on neurogenesis), or because steroids affect the move-
what extent these different behavioral outcomes are reflected ment and aggregation of cells, such that more cells overall
in context-dependent brain structure or neurochemistry is not come to be associated with the brain region of interest in males.
known. But when studying sex differences, it is important to The difference could also arise because fewer cells die in males
frame any findings as to whether they are persistent or tran- (an effect on cell death), or because gonadal steroids cause
sient, context dependent or independent, or the result of direct existing cells to express neurotransmitter ‘x’ (i.e., a hormonal
effects of hormones and chromosome complement or indirect effect on neurochemical differentiation).
as a function of somatic sex differences or in the case of Much of the recent progress in understanding sexual differ-
humans, gender differences (Figure 3). entiation comes from determining which of these basic devel-
opmental processes is controlled by gonadal steroids. At this
point, there is firm evidence that cell death and the morpholog-
5.01.2.10 Sexual Differentiation Depends on Four Key
ical differentiation of circuits contribute to permanent neural
Processes
sex differences (see Section 5.01.3.3), and in at least one system
As is evident from the preceding sections, a large number of sex it now seems clear that hormones direct neurochemical differ-
differences of every stripe have been identified in the mamma- entiation (see Section 5.01.3.4.4). There are also now estab-
lian brain. The sheer number and variety of the reported differ- lished incidences of sex differences in neuro- and glial-genesis
ences can be bewildering. It may be useful to recognize that in the developing hippocampus and amygdala. Some of this
sexual differentiation of any neural trait must, in principle, be evidence is presented below.
due to effects on one (or a combination) of four basic cellular
processes: cell birth, cell migration, cell death, and the differen-
tiation of neural circuits (Figure 4). This last category includes 5.01.3 Recent Progress in Understanding the Four
the morphological differentiation of cells (i.e., dendritic extent, Key Processes
axon outgrowth, spine, and synapse formation) as well as
5.01.3.1 Neurogenesis
neurochemical differentiation (i.e., which neurotransmitters,
neuropeptides, or hormone receptors are expressed). For some neural sex differences, the birth of the cells in ques-
To understand at a mechanistic level how hormones differ- tion is complete before the time the gonads differentiate. In
entiate male and female brains, it is important to identify these cases, it is very unlikely that gonadal hormones could
which of the four key processes are involved. Disentangling cause the sex difference by controlling neurogenesis, and so
the various possibilities can be surprisingly difficult, however. this mechanism can effectively be eliminated. Motoneurons
For example, assume that a sex difference is found in which that populate the SNB, for example, undergo their final mitosis
males have more neurons than do females that express neuro- by the 14th day of gestation in rats (Breedlove et al., 1983), yet
transmitter ‘x’ in a brain region of interest. If this sex difference testosterone is not produced by the differentiating testes until
persists even after adult hormone levels are made equivalent, several days later. Thus, it is very unlikely that in this system
but can be reversed by developmental exposure to steroids, gonadal steroid hormones could create a sex difference in cell
we would say that the trait is sexually differentiated, or number via an alteration of neurogenesis. It is even more

Figure 3 Not all sex differences are made equal. Sex differences come in many varieties. They may be dimorphic, i.e., in distinct forms in each sex,
or they may just vary along a continuum. Some sex differences only emerge under certain conditions and are therefore context dependent, while
others will disappear under certain conditions as the two sexes converge on the same endpoint. Early-life programming of sex differences usually
results in persistent changes, while others appear only transiently in response to the changing hormonal milieu.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 13

Figure 4 Four key principles by which sex differences can be established. There are four major pathways by which neuroanatomical sex differences
can be established. The first is cell genesis, the birth of new cells that will eventually become neurons or glia. To date, there is relatively little
evidence for neurogenesis being an important contributor to the establishment of sex differences in the numbers of neurons or glia in a particular
brain region, but this is beginning to change with new evidence emerging in the telencephalon and in the adult brain. A second potential source for
sex differences in the numbers of neurons in a particular brain region is differential migration into that region. Stuart Tobet and colleagues have
observed hormonally modulation of migratory paths of hypothalamic neurons, but whether this ultimately contributes to sex differences remains to
be determined. The third and one of the best documented sources of sex differences in the number of neurons in a brain region is differential cell
death, with either more cells dying in females, as observed for the sexually dimorphic nucleus, spinal nucleus of the bulbocavernosus, and bed
nucleus of the stria terminalis, or more cells dying in males, as seen in the anteroventral periventricular nucleus. These sex differences in cell death
are attributed to the actions of the gonadal steroids, testosterone, and estradiol, but the precise cellular mediators remain unknown. The fourth key
principle is broadly defined as cellular differentiation and refers to both the neurochemical phenotype of particular neurons and the nature and extent
of synaptic connections made by particular neurons. Sex differences in both neurochemical phenotype and synaptic connectivity are profound, and
advances are being made in identifying the particular cellular mechanisms by which these differences are achieved.

unlikely that postnatal exogenous steroid treatments, which within in the CA1 region, many of the newly generated cells
can masculinize SNB cell number, could affect the birth of become neurons (Zhang et al., 2008; Bowers et al., 2010).
SNB cells. This sex difference in neurogenesis is due primarily to endoge-
In other cases, however, gonadal steroid production and the nously produced estrogens because the administration of an
birth of neurons destined for the area in question may overlap. aromatase inhibitor or estrogen receptor antagonist at birth
This is true for the SDN-POA, for example (Jacobson and decreases the generation of new cells in males while having
Gorski, 1981; Davis et al., 1996a). Using tritiated thymidine no effect in females (Bowers et al., 2010). Since sex differences
or bromodeoxyuridine to label dividing cells, one can test for in cell number in the hippocampus of adult rats are small, the
sex differences in neurogenesis. There are pitfalls, however, greater cell genesis in males presumably is offset by increased
and a sex difference in the number of BrdUþ or thymidine- cell elimination at some point in life.
labeled cells at sacrifice does not necessarily mean that A sex difference in cell genesis has also been observed in the
neurogenesis was different. For example, if the same number developing amygdale; only in this instance the number of
of cells are born in males and females, but these cells migrate newborn cells is greater in females than males (Krebs-Kraft
to different brain areas or die at different rates in males than et al., 2010). Some of these newborn cells are destined to be
in females, a sex difference in labeled cells will result, but astrocytes while others will become neurons. Evidence suggests
does not reflect a sex difference in neurogenesis. Investigators that a sex difference in endocannabinoid tone in the devel-
have addressed this problem by examining the number of oping amygdala is the determinant of the greater number of
cells labeled after very short survival times (i.e., presumably newborn cells in the females. Males have a higher endocanna-
before cell death or migration could occur). Using this type binoid tone than females in this brain region, and treating
of analysis, perinatal neurogenesis has been eliminated as females with mimics of endocannabinoids reduces the number
important for sexual differentiation of the SDN-POA or AVPV of newborn cells to that seen in males. Determining if gonadal
(for review, see Forger, 2006). steroids modulate endocannabinoid levels is an important next
The hippocampus is one brain region where a sex difference step in identifying the cellular mechanisms by which this sex
in early-life neurogenesis has been clearly demonstrated. After difference occurs.
an injection of BrdU at birth, male rats have about twice as The genesis of new cells, some of which are neurons, has also
many labeled cells in the hippocampus as do females and, been observed in the AVPV and MeA of adolescent rats; this ‘late’

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
14 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

neurogenesis may contribute to the development or mainte- during a perinatal period of naturally occurring cell death
nance of neuroanatomical sex differences (Ahmed et al., 2008). (Oppenheim, 1991; Mosley et al., 2016). Gonadal steroids
can increase or decrease the rate of this cell death in some
neural regions, and the hormonal control of cell death is
5.01.3.2 Cell Migration
currently the best-established mechanism for creating neural
Many cells in sexually dimorphic neural regions of mice and sex differences in rodents (reviewed in Forger, 2006; Forger
rats (e.g., the POA, BST, AVPV, and SNB) have migrated to their et al., 2015a,b). It is not clear whether cell death actually is
respective location prior to testis differentiation and the time of the most important mechanism underlying neural sex differ-
the perinatal sex difference in testosterone synthesis (Bayer, ences, or whether we know a fair amount about the contribu-
1987; Bayer and Altman, 1987; Han and De Vries, 1999); the tion of cell death to sexual differentiation because it is
gross position of these cells is therefore unlikely to be affected relatively easy to find evidence for differential death in males
by gonadal hormones. However, cell addition in some brain and females. For example, dying cells can be quantified
regions continues throughout perinatal development (see based on morphology (pyknosis), or using well-established
discussion in Gotsiridze et al., 2007), and even for early- techniques such as TUNEL (terminal deoxynucleotidyl nick
arriving cells, gonadal steroids could cause subtle changes in end labeling) and immunohistochemistry for activated
cell position within a general brain region. For example, adult caspase-3, which take advantage of biochemical markers of
male and female rats have similar numbers of estrogen receptor apoptosis.
b-expressing cells in the AVPV, but these cells are more medially Although cell death is a well-established mechanism for
located in females (Orikasa et al., 2002). This sex difference in generating neural sex differences, if the evidence is viewed
the position of cells, however, is not necessarily due to sex with a critical eye, the case is clear in only a handful of cases.
differences in migration. For example, if testosterone increases These include some of the most ‘famous’ sex differences,
the survival of cells in the lateral AVPV, then males will wind up including the SNB, SDN-POA, BSTp, and the AVPV in rats
with more laterally positioned cells. To conclude that cell (Nordeen et al., 1985; Davis et al., 1996a; Chung et al.,
migration is sexually dimorphic and/or regulated by gonadal 2000; Sumida et al., 1993). There is also some evidence for
steroids, therefore, more direct evidence is necessary. a role of cell death in sexual differentiation of rat visual
Stuart Tobet and colleagues have addressed this issue using cortex (Nuñez et al., 2001), the gerbil POA (Holman et al.,
video microscopy to follow the movements of neurons in orga- 1996), at least one birdsong nucleus (Kirn and DeVoogd,
notypic brain slices. In an analysis of the POA/anterior hypo- 1989), and frog laryngeal motoneurons (Kay et al., 1999).
thalamus (AH) of embryonic mice, sex differences in the In each of these cases, a sex difference in cell number in
migratory pathway of labeled cells were found, such that signif- adulthood is correlated with a sex difference in the number
icantly more medial-to-lateral migration was seen in embry- of dying cells at some point in development. For example,
onic males than in females (Henderson et al., 1999), which adult males have more SNB motoneurons than do females,
correlates with a transient sex difference in the location of and females have more pyknotic cells in the SNB region
phenotypically identified cells in the POA/AH 2 days later around the time of birth (Nordeen et al., 1985). Females
(Wolfe et al., 2005). In addition, the movement of neurons also have more dying cells than do males in the early
in slice cultures of the POA/AH is affected by estradiol, but postnatal BSTp (Chung et al., 2000; Gotsiridze et al., 2007;
not by DHT (Knoll et al., 2007), and antagonists to GABA Ahern et al., 2013), and this correlates with greater BSTp cell
receptors may also influence cell migration in the hypothal- number in adult males. In both the SNB and BSTp,
amus (McClellan et al., 2008). This latter observation is testosterone treatments that eliminate the sex difference in
intriguing because it suggests that sex differences in neural cell number in adulthood also eliminate the sex difference
activity could influence cell positioning. Although the func- in dying cells during early development (Nordeen et al.,
tional significance of sex differences in cell migration is not 1985; Chung et al., 2000), adding weight to the conclusion
clear, one could imagine that the positioning of cells might that differential cell death is the mechanism underlying
affect afferent input, cell–cell communication on a local scale, sexual differentiation.
or exposure to diffusible signals. In the case of the estrogen There are limitations to this type of analysis, however. For
receptor b-expressing cells in AVPV described above, for example, counts of the number of pyknotic, TUNEL-positive,
example, the closer proximity of these neurons to the ventricle or caspase-3 positive cells are necessarily ‘snapshots’ of the
in females could increase access to chemosignals in the cerebro- number of cells dying (or, more specifically, at a particular
spinal fluid. stage of apoptosis) at a given moment. In order to determine
The technical obstacles to obtaining direct evidence for sex whether a sex difference in dying cells can account
differences in neuronal migration have discouraged all but quantitatively for a sex difference in neuron number in
a few hearty souls from pursuing this line of investigation. adulthood, one would have to know how long a cell appears
Therefore, whether this key process is in fact a critical contrib- apoptotic in the material under study. This variable has not
utor to programmed sex differences in the brain remains to been determined for any sexually dimorphic region
be determined. (although see Nuñez et al., 2000) and estimates of how long
a cell takes to undergo apoptosis vary greatly. Thus, it is
difficult to rule out other mechanisms. For example, in the
5.01.3.3 Cell Death
rat locus coeruleus, adult females have more neurons than do
About half of the neurons originally produced during devel- males, and cell death is greater in males on the day of birth
opment of the vertebrate nervous system are eliminated (Guillamon et al., 1988). One might be tempted to stop

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 15

there and declare that the mechanism underlying the sex


difference has been demonstrated. However, a more detailed
study of developmental changes in cell number suggests that
the adult sex difference may actually result from
a postpubertal addition of cells in the locus coeruleus of
females (Pinos et al., 2001). Similarly, a greater number of
cells in visual cortex of males correlate with greater cell death
in this region in females postnatally, but a more complete
developmental profile suggests that the mechanism
accounting for the adult sex difference is complex (Nuñez
et al., 2002).
Genetically modified mice have been used to get around
some of these problems and to test more directly the contribu-
tion of cell death to neural sex differences. For example, Bax,
a prodeath gene of the Bcl-2 family is singularly important
for apoptosis in neural development, and deletion of this single
gene nearly eliminates perinatal cell death in many brain
regions (White et al., 1998; Ahern et al., 2013). By examining
Bax knockout mice, it is therefore possible to ask which neural
sex differences depend on differential cell death in males and
females. In mice lacking Bax, the number of cells in the SNB,
AVPV, and BSTp is significantly increased (Forger et al., 2004;
Jacob et al., 2005) and, more important, sex differences in total
cell number in each of these regions are eliminated (Figure 5).
This demonstrates that Bax protein is required for sexually
dimorphic cell death in the mouse forebrain and spinal cord.
Because the animals in these studies were gonadally intact,
Bax gene status overrides endogenous testosterone levels in
determining cell number. Thus, one obligatory gene in the
sexually dimorphic cell death pathway – Bax – has been
identified.
The study of mutant mice has also identified sex differences
Figure 5 The proapoptotic gene, bax, is required for sex differences
not dependent on Bax and demonstrates that the control of cell
in neuron number. (a) Cell death throughout the developing mouse
number may vary not only from region to region, but also
brain is dependent on the prodeath gene, Bax. Dying cells (black dots),
among subtypes of cells within a single region (see below, labeled by immunohistochemistry for activated caspase-3, are much
and Zup et al., 2003; Forger et al., 2004; Semaan et al., 2010; more numerous in newborn wild-type mice (Baxþ/þ) than in Bax
Gilmore et al., 2012). knockout mice (Bax/). (b) In adulthood, cell number in the BNSTp
The next question is how testosterone regulates cell death. was greater in wild-type males than in females. Deletion of the Bax
Estrogenic metabolites play the predominant role in sexual gene increased cell number overall and eliminated the sex difference,
differentiation of the AVPV, SDN-POA, and BSTp of rodents suggesting this sex difference depends on developmental cell death.
(McCarthy et al., 1993; Bodo et al., 2006; Hisasue et al., The number of animals per group is indicated at the base of each bar.
2010; Orikasa and Sakuma, 2010; Tsukahara et al., 2011), so n.s., Not significant. (a) Reprinted with permission of Wiley from Ahern,
T.H., Krug, S., Carr, A.V., Murray, E.K., Fitzpatrick, E., Bengston, L.,
the aromatization of testosterone and the activation of
McCutcheon, J., De Vries, G.J., Forger, N.G., 2013. Cell death atlas of
estrogen receptors is a first step. In some cases, effects of
the postnatal mouse ventral forebrain and hypothalamus: effects of age
estrogen receptor activation on Bcl-2 family proteins may be and sex. J. Comp. Neurol. 521, 2551–2569; (b) Reprinted with the
direct, since putative estrogen response elements have been permission of the National Academy of Sciences of the USA from
described in the Bcl-2 and Bcl-xL genes (Pike, 1999; Lin Forger, N.G., Rosen, G.J., Waters, E.M., Jacob, D., Simerly, R.B.,
et al., 2006). Moreover, estradiol increases Bcl-2 and de Vries, G.J., 2004. Deletion of Bax eliminates sex differences in the
decreases Bax expression in the SDN-POA of newborn rats mouse forebrain. Proc. Natl. Acad. Sci. U.S.A. 101, 13666–13671.
(Tsukahara, 2009). Alternatively, the estrogen regulation of
Bcl-2 family members may be parallel to, or downstream of,
5.01.3.4 Differentiation of Circuits
other signaling pathways. In the neonatal rat AVPV, for
example, the use of targeted apoptosis microarrays identified The building of circuits involves at least three morphological
a sex difference in the activity of the tumor necrosis factor components: axonal growth, dendritic growth, and synapto-
(TNF) alpha–TNF receptor 2–NF-kappa B cell survival genesis. In addition to the morphological differentiation of
pathway, as well as a difference in Bax expression (Krishnan circuits, the expression of neurotransmitters and neuropeptides
et al., 2009). Using transcriptome analysis, other novel will determine circuit functionality. There is evidence for
proapoptotic proteins have also been implicated in male- steroids regulating all three morphological processes, although
specific cell death in the neonatal AVPV (Del Pino Sans the preponderance of the data pertains to synaptogenesis and
et al., 2015). particularly to synapses on dendritic spines. This is largely

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
16 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

due to technical limitations. As noted above, it is simply easier To explore the basis of this striking sex difference further, Sim-
to quantify dendritic spines than to measure changes in axonal erly and colleagues cocultured BST and AVPV explants from
and dendritic growth. Nonetheless, there is good evidence for neonatal male and female rats. They found a robust outgrowth
steroid modulation of axonal growth, with one of the best from BST to AVPV only when the AVPV came from a male or
examples being the building of the functional circuit relevant a testosterone-treated female (Ibanez et al., 2001); the sex/
to control of the LH surge. There is also good evidence for hormone status of the BST had no effect on outgrowth. Thus,
effects of steroids on the production of neurotransmitter/ a diffusible factor from the AVPV is responsible for controlling
neuropeptides, as described below. the growth of BST neurons, but the precise identity of this factor
remains to be determined.
5.01.3.4.1 Axonal Growth
The sexual differentiation of neural circuits was presaged by 5.01.3.4.2 Dendritic Growth and Branching
landmark studies of Dominique Toran-Allerand on the The first indication that gonadal steroids could influence
profound induction of neurite outgrowth by estradiol from dendritic growth came from studies on motoneurons of the
organotypic explant cultures of the POA, hypothalamus, and SNB. For example, the castration of adult male rats led to
cerebral cortex (for review, see Toran-Allerand, 2005). Neurite a 50% reduction in the overall extent of SNB dendritic trees,
is the appropriate term in these cases, as processes emerging which could be prevented by treating castrates with testos-
from neurons grown in culture are not readily distinguished terone (Kurz et al., 1986). A similar reduction in dendritic
as dendrites or axons. Explant cultures do, however, allow for extent was seen in white-footed mice exposed to short day
direct application of steroids to the medium and precise quan- lengths characteristic of winter (Forger and Breedlove, 1987),
tification of neurite length. In cells cultured from the VMN, suggesting that hormonally mediated growth and retraction
estradiol interacts with neurotropic factor signaling to control of SNB dendrites may be a normal feature in seasonally
neurite outgrowth, and effects of estradiol appear to be via breeding rodents. Gonadal hormones also control the initial
a membrane receptor (Carrer et al., 2005). Estradiol can also outgrowth of SNB dendrites in development; in this case,
have inhibitory effects on neurite growth, for example, on sero- however, both estrogens and androgens support growth
tonergic neurons derived from the embryonic mesencephalon (Goldstein and Sengelaub, 1994). Interestingly, although
(Lu et al., 2004). Unfortunately, studying neurite development SNB motoneurons abundantly express the androgen receptor,
in the actual brain turns out to be quite challenging, and prog- they do not express estrogen receptors. The effect of estrogens
ress has been relatively slow. on SNB dendrites appears to be mediated instead by hormone
We discussed above one of the most pronounced sex differ- action at the target muscles. Estrogen administration at the
ences in physiology, the control of gonadotropin secretion muscle supports dendritic growth of SNB motoneurons, and
from the anterior pituitary. Male rodents have continuously the local blockade of estrogen receptors at SNB target muscles
pulsatile LH release, whereas females exhibit an LH surge in results in severely reduced dendritic trees (Nowacek and
response to estradiol. Although, as mentioned above, it is clear Sengelaub, 2006).
that this sex difference is controlled by the brain, finding out
just where in the brain proved far more complicated than origi- 5.01.3.4.3 Synaptogenesis
nally supposed, and may not yet be entirely solved. The obvious Sexual differentiation of the brain occurs during a perinatal
candidates, the GnRH neurons themselves, are largely indistin- sensitive period. This is analogous to other developmental
guishable in males and females, suggesting that differences in processes that are restricted to sensitive developmental periods,
afferent input to these neurons is the key variable. One brain such as organization of the barrel cortex in response to whisker
region that may play this role is the AVPV (Wiegand and activation, or organization of the visual system by light and
Terasawa, 1982; Ronnekleiv and Kelly, 1986; Petersen and color. In both cases, sensory stimuli act to refine an overly
Barraclough, 1989). The results of anterograde labeling experi- exuberant innervation of a target region by pruning excessive
ments indicate that neurons in the AVPV provide direct inputs or superfluous inputs. Similarly, SNB motoneuron dendrites
to GnRH-containing neurons in the preoptic region (Gu and are pruned from early exuberant growth, and this pruning is
Simerly, 1997). A subset of these neurons are probably the steroid sensitive (Goldstein et al., 1990). However, steroids
sexually dimorphic kisspeptin and/or glutamatergic/GABAergic also build sexually dimorphic circuits by forming new synapses
neurons discussed in Section 5.01.2.6.2. In addition, ‘on demand,’ as will be discussed in this section. Depending on
descending projections from the AVPV to dopaminergic the brain region, estradiol can either increase or decrease the
neurons in the arcuate nucleus are about eight times more density and/or number of dendritic spines and the attendant
abundant in females than in males (Simerly, 1998). Thus, synapses. In addition, strikingly different mechanisms are
axonal output from the AVPV is highly sexually dimorphic, utilized in different brain regions. We highlight this point by
and this may account for the sex difference in control of reviewing effects of estradiol on dendritic spine formation in
reproductive function. three regions: the arcuate nucleus, POA, and VMN.
In addition, AVPV itself receives sexually dimorphic input.
For example, a projection from the BSTp to AVPV is about 5.01.3.4.3.1 Arcuate Nucleus
10-fold greater in male rats than in females (Gu and Simerly, The arcuate nucleus of the hypothalamus represents an impor-
1997; Hutton et al., 1998). This is entirely due to a hormonally tant site for neuroendocrine integration and contains a variety
mediated sex difference in axon outgrowth during develop- of neurons that control hormone secretion from the anterior
ment; the projection apparently never forms in females, but pituitary (Everitt et al., 1986; Swanson, 1986). The nucleus is
is induced in females treated with testosterone at birth. comprised of numerous cell types, including dopaminergic,

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 17

enkephalinergic, and GABAergic neurons, and many of these surprising that the POA is a site of major sex differences in
coexpress releasing peptides such as corticotropin-releasing morphometry. For example, the sex difference in SDN-POA
hormone, GH-releasing hormone, and thyrotropin-releasing volume, discussed above, is perhaps the most celebrated sex
hormone. The arcuate sends its strongest projections to other difference in the mammalian brain. Close scrutiny of the
parts of the periventricular zone of the hypothalamus (Simerly, POA also reveals additional sex differences at the synaptic level.
1995; Li et al., 1999) and plays a central role in reproduction, As in the arcuate nucleus, the POA of males contains more
feeding, and the response to stress. There is a marked and complex, stellate astrocytes than that of females (Amateau and
permanent sexual dimorphism in the arcuate nucleus, in which McCarthy, 2002a,b). But the relationship between astrocyte
females have a two- to threefold higher density of axodendritic morphology and dendritic spine patterning is the opposite of
spine synapses than males, while males have two- to threefold that seen in the arcuate: male rats have a two- to threefold
more axosomatic synapses than do females. The pattern is greater density of dendritic spines in the POA than do females
reversed in males castrated as neonates or females treated (Amateau and McCarthy, 2002b). The induction of spines in
with testosterone at birth (Matsumoto and Arai, 1980). Given the POA of males is permanent, with the pattern established
that axosomatic synapses are generally inhibitory while axo- in the first few days of postnatal life, and persisting until at least
dendritic spine synapses are excitatory, sex differences in the 90 days of age (Amateau and McCarthy, 2004).
relative number of somatic versus spine synapses are likely to The sexual differentiation of POA neurons is a complex
have a profound impact on neuronal excitability, as well as process that begins with estradiol upregulation of the enzyme
on the source of afferent input to arcuate neurons. One such cyclooxygenase-2 (COX-2), a nodal point in the production
sexually dimorphic projection, from AVPV to the arcuate of prostaglandins and the thromboxanes (Hoffman, 2000).
nucleus, was discussed above. Induction of COX-2 is strongly yoked with an inducible form
The testosterone-induced decrease in arcuate spine synapses of prostaglandin E2 synthase, leading to the preferential
correlates with a marked sex difference in the morphology of production of prostaglandin E2 (PGE2) over other prosta-
astrocytes in the same brain region (Mong et al., 1996, noids. Estradiol treatment of neonatal female rats increases
1999): males have more complex astrocytes with a stellate PGE2 levels in the POA, and this appears to be a direct result
morphology, compared to the relatively simple, bipolar shape of estradiol induction of COX-2 gene transcription (Amateau
of astrocytes in the arcuate of females. This sex difference is also and McCarthy, 2004). There are two effects of PGE2 on astro-
determined by estradiol acting within the first few days of life. cytes: increased stellation (Amateau and McCarthy, 2002a)
In the adult arcuate, this same population of astrocytes is and release of glutamate (Bezzi et al., 1998; Sangrizi et al.,
capable of physically stripping synapses and allowing for rees- 1999). The glutamate released by the astrocytes in response
tablishment later. This feature is unique to females and is to PGE2 is speculated to then activate AMPA receptors on the
a component of the remodeling that occurs across the estrous neighboring (or originating) neuron to induce the formation
cycle (Garcia-Segura et al., 1994). Whether there is an analo- of dendritic spines (Figure 6). The source of the PGE2 produc-
gous but permanent process in which astrocytes are differenti- tion is hard to determine precisely but a critical contribution of
ated by estradiol to suppress the formation of dendritic spine the brain’s innate immune cells, the microglia, has been estab-
formation during development in males, remains unknown. lished (Lenz et al., 2013). Microglia had previously been
Paramount for establishing a causal link between astrocyte considered only in the context of injury but have recently
morphology and the number of dendritic spines on neurons received renewed attention as important sculptors of neural
is determining the primary site of estradiol action. Interestingly, circuits (Schafer et al., 2012). In the POA, males have more
the estradiol-induced astrocyte stellation was found to require microglia and they both respond to and produce more PGE2
activation of the GABAA receptor (Mong et al., 2002). And than in females, thereby providing a positive feedback mecha-
while astrocytes express GABAA receptors, they do not make nism for maintaining elevated prostaglandin during the sensi-
GABA. The rate-limiting enzyme in GABA synthesis, GAD, is tive period for sexual differentiation. The magnitude of the
found only in neurons, and estradiol increases the amount induction of these dendritic spines by prostaglandins is tightly
and activity of GAD (see Section 5.01.2.6.1). Therefore, estro- correlated with the expression of male sexual behavior in adult-
gens apparently alter the morphology of arcuate astrocytes by hood. Compared to animals with a low density of POA
stimulating the synthesis of GABA, which is released from dendritic spines, animals with high spine densities exhibit
neurons to act on neighboring astrocytes to induce stellation. shorter latencies and higher frequencies of mounting and
It has yet to be determined whether astrocyte stellation then thrusting in tests with receptive females (Wright et al., 2008).
results in the permanent suppression of dendritic spine synapse
formation. 5.01.3.4.3.3 Ventromedial Nucleus of the Hypothalamus
As the POA is critical to the control of male reproductive
5.01.3.4.3.2 The Preoptic Area of the Hypothalamus behavior, the VMN is central to female reproductive behavior
The POA has long been established as crucial for the control of (reviewed in Pfaff et al., 1994). The neural outputs of the
male copulatory behavior (Davidson, 1966; Hansen et al., VMN have been examined in detail, and they provide an
1982). This region also influences other sexually dimorphic anatomical substrate for relaying the acute effects of ovarian
functions, including the control of gonadotropin release, steroid hormones onto neural systems mediating lordosis, as
maternal behavior, and female copulatory behavior, by well as a means of coordinating female copulatory receptivity
virtue of its projections to other sexually dimorphic nuclei with gonadotropin secretion. Despite this, and the intimate
involved in the regulation of these sex-specific functions connections between the VMN and other sexually dimorphic
(Swanson, 1986; Simerly and Swanson, 1988). Thus it is not nuclei, the VMN is not dramatically dimorphic in terms of

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
18 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

in the phosphorylated form of its substrate, Akt, within 1 h of


estradiol treatment. PI3 kinase has previously been implicated
in neurotransmission, but precisely how it is promoting the
release of glutamate in this instance remains poorly under-
stood. A final piece of evidence makes this series of findings
even more intriguing: there is no requirement for activation
of estrogen receptors in the postsynaptic neuron in which
dendritic spines are being induced by estradiol. The toxin,
latrotoxin, acts presynaptically to stimulate glutamate release
and mimics the effects of estradiol on spine formation in the
VMN. The increase in dendritic spines in latrotoxin-treated
animals was not altered by blocking estrogen action, indi-
cating that there was no contribution of a postsynaptic effect
of estradiol (Schwarz et al., 2008).
From a behavioral perspective, these observations are
equally intriguing. Given that we had previously determined
that masculinization of sex behavior could be achieved inde-
Figure 6 Prostaglandins induce masculinization of the preoptic area
neurons and sex behavior. Estradiol in the neonatal preoptic area (POA) pendently of gonadal steroids by manipulating prostaglandins,
of males binds to the estrogen receptor and induces transcription of the and that males that were treated neonatally with a prosta-
gene for cyclooxygenase-2 (COX-2), resulting in increased production glandin inhibitor were not masculinized but were defeminized
of prostaglandin E2 (PGE2). The astrocytes in the male POA are more (Todd et al., 2005), we concluded that steroid action in the
complex, with longer and more frequently branching processes than in VMN must be relevant to the process of defeminization. Thus
females, and this sex difference is mediated in part by PGE2, and in there seem to be separate, albeit related, cellular processes
part by glutamate. More importantly, PGE2 induces the formation of that control masculinization versus defeminization of sexual
dendritic spine synapses on POA neurons, and this is also mediated in behavior.
part by glutamate. The organization of a greater density of dendritic
spines in the male POA results in masculinization of sex behavior in the
5.01.3.4.4 Differentiation of Neurochemical Phenotype
adult animal.
In addition to the examples of morphological differentiation
discussed above, there are even more reports on sex differences
in neurotransmitters or neurotransmitter receptor expression
cell number or nuclear volume. It does, however, exhibit a sex (De Vries, 1990; Fink et al., 1998; Rhodes and Rubin, 1999).
difference in dendritic spines: neurons of the VMN of males However, two fundamentally different sets of processes could
have more excitatory dendritic spine synapses than females, cause any of these differences: processes that determine the
and VMN dendrites are longer and branch more frequently in absolute number of cells capable of expressing a specific attri-
males (Schwarz et al., 2008). bute (i.e., birth, death, or migration of cells), or processes
An obvious question is whether prostaglandins mediate that act on preexisting cells to alter their phenotype (Tobet
the sex difference in dendritic spines in the VMN, as in the and Hanna, 1997; De Vries and Simerly, 2002; Forger, 2006).
POA. The answer here is no; nor does GABA appear to play The former set of processes enriches or depletes areas with cells
any significant role in synaptogenesis during development of that show specific features, e.g., cells that have relatively large
this brain region (Todd et al., 2007). There is again, however, dendritic trees, whereas the second set of processes changes
a critical role for glutamate, but it is a very distinct role from the features of existing cells, e.g., extend or prune dendritic
that in the POA. Recent studies indicate that estradiol trees. As described above, it has been surprisingly difficult to
promotes the release of glutamate from VMN nerve terminals. disentangle these two possibilities, in part because gonadal
The enhanced release is independent of protein synthesis but hormones often trigger sexual differentiation before the
requires activation of phosphoinositide 3-kinase (PI3 kinase). neurons of interest assume their final phenotype.
The impact of the increased glutamate release is activation of A case in point is the sexually dimorphic vasopressin inner-
MAP kinase in the postsynaptic neuron, followed by the vation of the brain. This innervation shows one of the most
induction of spinophilin synthesis and the construction and consistently found neural sex differences among vertebrates
maturation of dendritic spine synapses (Schwarz et al., (Figure 7; De Vries and Panzica, 2006), with males having
2008). There are many novel aspects of this mechanism for more vasopressin neurons in the BST and MeA and denser
establishing a sexually dimorphic pattern of synaptic connec- projections from these areas than do females across many
tivity in the VMN, one of the most notable being the apparent mammalian species (De Vries and Panzica, 2006). Nonmam-
lack of any involvement of astrocytes. Equally notable is the malian vertebrates show similar sex differences in homologous
rapidity of the effects, with estradiol significantly enhancing vasotocin projections (Moore et al., 2000; Goodson and Bass,
glutamate release within 3 h, and the observation that protein 2001; De Vries and Panzica, 2006). This sex difference has
synthesis is not required for the effect of estradiol. Both of been well studied in rats, where exposure to gonadal steroids
these findings suggest a nongenomic mechanism of action, during perinatal life determines the number of vasopressin cells
which is somewhat surprising given the permanent nature of found in adults (Wang et al., 1993; Han and De Vries, 2003).
the effect. The estrogen receptor is required and appears to acti- Differential cell birth and migration are very unlikely to
vate PI3 kinase directly, as demonstrated by a marked increase contribute to the sex differences, as vasopressin cells are born

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 19

birth (Wang et al., 1993; Han and De Vries, 2003), yet the vast
majority of presumptive vasopressin neurons do not begin
expressing vasopressin until several days (in males) to weeks
(in females) later (Szot and Dorsa, 1993). Thus, one cannot
identify the cells of interest during the time the sex difference
in their number is determined. It has also been difficult to
rule out a role for cell death. Mice are especially useful for
differentiating between these two possibilities, as there are
several genetically engineered strains in which cell death is
altered (Lindsten et al., 2005; see Section 5.01.3.3). One can
therefore use such mice as a litmus test for cell death. Mice over-
expressing cell death-reducing factors would probably show
no, or reduced, sexual differentiation if differentiation
depended on cell death. The same would be true for mice lack-
ing proteins required for neuronal cell death, such as the Bax
protein described above. Vasopressin expression was studied
by comparing wild-type mice with two genetically altered
strains: those that overexpress the antiapoptotic factor Bcl-2
specifically in neurons, or mice with a null mutation in the
gene encoding Bax. Neuronal cell death is markedly reduced
throughout the brain, and sex differences in cell number are
reduced or abolished in several brain areas in both mutants
(see Section 5.01.3.3). These mutations do indeed increase
the total number of cells that produce vasopressin. Critically,
however, the sex difference in cell number remains intact
(Figure 7; De Vries et al., 2008). This leaves sexual differentia-
tion of cellular phenotype as the only remaining plausible
mechanism for sexual differentiation of vasopressin expression.
Numerous other neurotransmitters and neuropeptides (e.g.,
Figure 7 Sex difference in the vasopressin (AVP) innervation of the dopamine, neurotensin, substance P, and enkephalin) show
lateral septum (LS). The top panels show the AVP innervation in the LS
sex differences in cell number (reviewed in De Vries, 1990;
of a female (left) and male rat (right). Note the higher density of vaso-
Forger et al., 2015a), and for none of these has the cellular
pressin fibers in male brains. The graph shows the number of AVP cells
in the bed nucleus of the stria terminalis of male and female wild type mechanism of sexual differentiation been established. The
and Bax/ knockout mice. Note that while the overall number of AVP search for these mechanisms may be amenable to the same
cells is higher in bax/ mice, the sex difference in cell number strategy described here. If, however, one successfully eliminates
remains. cell birth, migration, and death as possible mechanisms, that is
only a modest first step. One then faces the task of identifying
which of a host of possible molecular mechanisms contributes
on embryonic days 12 and 13 (al-Shamma and De Vries, 1996; to the phenotype of differentiating neurons. Understanding the
Han and De Vries, 1999), at least a week before gonadal process underlying phenotypic decisions constitutes one of the
hormones trigger their sexual differentiation (Wang et al., major challenges of neuroscience today, and hormone-
1993; Han and De Vries, 2003). This leaves differential cell sensitive systems are poised to make a major contribution
death or phenotypic differentiation as the two most likely because this process can often be manipulated (hormonally)
causes. in sexually dimorphic systems with relative ease. One thing
Circumstantial evidence favors the latter. Essentially all seems certain: to the extent that sexual differentiation of cell
vasopressin cells in the BST coexpress the neuropeptide gala- phenotype involves long-term changes in gene expression,
nin, but not all galanin cells coexpress vasopressin (Planas epigenetic modifications will be involved, as discussed in the
et al., 1995a). Because the total number of galanin cells does next section.
not differ between males and females, it was hypothesized
that, during development, higher levels of testosterone act on
5.01.3.5 The Role of Epigenetic Mechanisms in the Four Key
existing galaninergic cells to increase the percentage that will
Processes
coexpress vasopressin (Planas et al., 1995b). In support, vaso-
pressin and galanin neurons in the BST and MeA of rats show 5.01.3.5.1 How Can Effects of Sexual Differentiation Last
the same unusual birth profile, with both types of neurons a Lifetime?
born days earlier than most surrounding cells (Han and How can we explain the enduring effects of gonadal hormones
De Vries, 1999), consistent with the idea that these neurons on cell morphology (e.g., changes in dendritic spines) or
belong to a single cohort of pluripotent cells. neurochemical phenotype (e.g., as seen in the vasopressin
The hypothesis that testosterone instructs these cells to system) described above? In some systems, hormones may
become vasopressinergic, however, is difficult to test directly. induce permanent changes in the phenotype of neurons by
Gonadal steroids determine vasopressin cell number soon after altering the structures that they innervate. Target-dependent

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
20 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

differentiation of neurotransmitter phenotype has been found brain regions that are not sexually dimorphic, excluding
in the sympathetic nervous system, for example (Ernsberger a nonspecific effect on cell survival.
and Rohrer, 1999; Landis, 1990). Gonadal hormones could Matsuda et al. (2011) took things further by administering
also alter neuronal morphology or phenotype by acting directly a different HDAC inhibitor, trichostatin A, to newborn rats.
on those cells. Either way, epigenetic mechanisms (modifica- When compared to animals receiving control treatments, tri-
tions of chromatin that do not involve changes to the under- chostatin A-treated rats showed marked impairments in male
lying DNA code) are implicated. sexual behavior in adulthood (Matsuda et al., 2011).
The DNA that is common to every cell in the body is epige- Importantly, the intracerebroventricular administration of
netically modified during early development to generate the antisense oligonucleotides to block the endogenous
vast array of different cell types in the mature organism, and production of HDACs also had the same effect, eliminating
many of the cell fate ‘decisions’ are essentially irreversible concerns that the inhibition of masculinization was related to
(i.e., once a liver cell, always a liver cell). Similarly, to the extent side effects of trichostatin A. These authors also showed that
that gonadal steroid exposure during neonatal life leads to histones associated with the estrogen receptor alpha and
permanent changes in neuronal cell morphology or neuro- aromatase genes are differentially acetylated in the mPOA of
chemistry, epigenetic mechanisms are likely to be involved. perinatal males and females, although whether this is related
Even effects of gonadal steroids on cell death may involve to the effects of HDAC blockade on sexual function is not
epigenetic mechanisms. For example, testosterone prevents known. Thus, both the Matsuda et al. (2011) and the Murray
neurons from dying during the cell death period in the et al. (2009) studies suggest that HDAC activity (and, hence,
neonatal SDN-POA and BSTp of mice and rats. However, the the repression of some gene or genes) is required for
hormone signal that is present on the day of birth does not masculinization of brain morphology and behavior.
cause a sex difference in cell death until 5–7 days later A more recent study by Nugent et al. (2015), however,
(Gotsiridze et al., 2007; Davis et al., 1996a) indicating that found greater DNA methylation in the mPOA of neonatal
a cellular ‘memory’ for the hormone exposure is required. female rats. To test the functional significance of this sex differ-
Indeed, several recent studies confirm the requirement for ence, neonatal rats were treated with a DNA methyltransferase
epigenetic modifications in sexual differentiation of the brain (DNMT) inhibitor during the first 2 days of life, and dendritic
(for reviews, see McCarthy et al., 2009; Forger, 2016). spine density (which is normally higher in the POA of males,
see Section 5.01.3.4.3.2) was examined. DNMT inhibition
5.01.3.5.2 Effects of Manipulating Epigenetic Mechanisms on masculinized spine density in females and eliminated the sex
Sexual Differentiation difference (Nugent et al., 2015). DNMT-treated females also
The fundamental structural unit of chromatin is the nucleo- displayed male sexual behavior in adulthood (Nugent et al.,
some, comprising a length of about 146 base pairs of DNA 2015). This strongly suggests that active repression of some
wrapped around an octamer of histone proteins. Covalent gene(s) by DNA methylation is required for normal female
modifications of the DNA, or the histone protein tails that development.
protrude from the nucleosome, influence accessibility to tran- The alert reader may have noticed that the repression of
scriptional machinery and gene expression (for reviews, see gene transcription is implicated in normal masculine develop-
Felsenfeld and Groudine, 2003; Jiang et al., 2008). ment by the HDAC inhibitor studies, but in normal feminine
The best understood of the histone tail modifications is development by the DNMT study (Murray et al., 2009;
acetylation. Histone acetyltransferases add acetyl groups to Matsuda et al., 2011; Nugent et al., 2015). The effect of a given
histone tails, which generally enhances transcriptional activity, epigenetic perturbation may depend on species, age, brain
whereas histone deacetylases (HDACs) remove acetyl groups region, and even on the dependent variable measured within
and generally reduce transcription (Cosgrove and Wolberger, a given brain region. Indeed, the same HDAC inhibitor that
2005). Interestingly, steroid hormone receptors, including prevents masculinization of total cell number in the mouse
androgen and estrogen receptors, recruit coactivators or core- BSTp promotes masculinization of the vasopressin projections
pressors to target genes that alter histone acetylation emanating from the BSTp (Murray et al., 2009, 2011). Prelim-
(Kishimoto et al., 2006; Tetel, 2009), and blocking these cofac- inary findings also suggest that neonatal DNMT inhibition in
tors can prevent effects of testosterone on morphological and mice masculinizes or feminizes neurochemical differentiation
behavioral sexual differentiation (Auger et al., 2000, 2002). in the mPOA, depending on which neurochemical is examined
Given the role of histone acetylation in steroid receptor (M. Mosley and N. Forger, unpublished; Forger, 2016).
action, we hypothesized that this epigenetic mark might be
important for hormonally mediated sexual differentiation. 5.01.3.5.3 Sexual Differentiation of Epigenetic Marks across
As described in Section 5.01.3.3, the size of the BSTp the Genome
depends on testosterone-regulated cell death in early devel- In a complementary approach, a handful of studies have
opment. To test whether histone acetylation plays a role, recently reported sex differences and effects of early hormone
mice were treated with the HDAC inhibitor, valproic acid, exposure on specific epigenetic marks across the genome. For
during the first 2 days of life, and the BSTp was examined example, Gharahmani et al. (2014) examined the effects of
at weaning. The valproic acid treatment transiently increased testosterone treatment on the day of birth on the DNA methyl-
histone acetylation and prevented masculinization of BSTp ome in the BST/mPOA and striatum of mice. They found
volume and cell number in both males and testosterone- a small number of genes (<70) that were differentially methyl-
treated females (Murray et al., 2009). Valproic acid had no ated between control females and testosterone-treated females
effect on the BSTp of control females, or on cell number in on postnatal day 4, and roughly 1000 differentially methylated

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 21

genes in adulthood. A similar pattern – few differences in DNA Rhoda et al., 1984). These papers found higher serum testos-
methylation on day 4 but tenfold more in adulthood – was terone and hypothalamic estradiol in males and have been
seen when comparing males and females (Gharamani et al., taken by many as proof that male brains are organized by
2014). Thus, effects of neonatal testosterone exposure were steroids derived from the testes. These findings have stood
late-emerging, in what might be termed an ‘epigenetic echo’ the test of time and proven enormously useful. The risk of
of the hormone exposure. Parallel effects were seen in a detailed iconic papers, however, is that findings that are true for one
study of epigenetic changes to the promoter regions of the set of circumstances are often inappropriately generalized to
estrogen receptors, alpha and beta, and the progesterone other circumstances. For 30 years, it has been generally
receptor from birth to adulthood. Sex differences were present assumed that the entire male brain is exposed to significantly
at some times and not others, with new ones appearing long higher levels of both androgens and estrogens than the female
after the perinatal hormone exposure (Schwarz et al., 2010). brain during the perinatal sensitive period. Moreover, it has
Another interesting finding of the Gharahmani et al. (2014) been a foregone conclusion that the source of ‘gonadal’ steroids
study is that those genes exhibiting a sex difference (or an effect found in the brain during the developmental window is the
of neonatal testosterone treatment in females) on DNA meth- circulation, which in turn reflects steroidogenesis by the testis.
ylation were unevenly distributed: 85–90% of them were hypo- There are two pieces of information that suggest this
methylated in control females. This was true at both ages simplistic view should be reevaluated. One is the evidence
(postnatal day 4 and adulthood) and in both brain regions that neurons can synthesize steroids de novo from cholesterol
(striatum and BST/mPOA). Similarly, a genome-wide analysis all the way to estradiol (Rune et al., 2002; Prange-Kiel et al.,
across fetal development in the human brain identified a subset 2003; Hojo et al., 2004). Thus the brain can serve as its own
of genes with a significant sex-by-age interaction on DNA meth- gonad. This is best exemplified in the songbird brain where local
ylation: among these, the pattern noted was progressive hypo- estradiol synthesis is critical to masculinization of one of the
methylation in females (Spiers et al., 2015). song control nuclei (Holloway and Clayton, 2001). This may
Based on a transcriptome analysis, differences in DNA also be true in the mammalian brain, but remains to be more
methylation between groups in the mouse study (Gharahmani definitively established. The second piece of evidence is that
et al., 2014) were not a good predictor of differences in gene when steroid levels in the brain are measured in multiple
expression. This implies that, for the set of differentially meth- regions and multiple time points, the concept of uniformly
ylated genes, females use less DNA methylation to achieve the higher levels in perinatal males is not supported. In fact, in
same rate of transcription. This resonates with findings in some brain regions outside the diencephalon, levels of estradiol
a recent genome-wide study of the distribution of the histone can be high in both sexes and higher than in the POA or hypo-
mark H3K4me3 (three methyl groups on lysine 4 of histone thalamus (Amateau et al., 2004). However, there is a problem
3): genes with a sex difference in H3K4me3 in adult mice with the methods used to obtain both sets of evidence.
showed a strong sex bias (75% of them had more of the histone Radioimmunoassay as a way to measure small peptides,
mark in females), and sex differences in the epigenetic mark did and eventually steroids, was a methodological advance so
not predict differences in gene expression (Shen et al., 2015). major that it garnered a Noble Prize for Roslyn Yallow in
Taken together, males and females may use different epigenetic 1977. Still in common use today, this technique allows for
mechanisms to control gene expression, even when that expres- reasonably precise quantification of steroids in the circulation
sion does not vary (see Forger, 2016). where only the presence of binding globulins is a potential
and easily remedied confound. It is far more difficult, however,
to precisely measure steroids in brain tissue where the dense
5.01.4 Six Unanswered Questions concentration of lipids can create false positives, or the strin-
gent extraction protocols required can cause loss of substantial
Despite the obvious progress in understanding mechanisms of signal in the small samples generated from individual brain
sexual differentiation over the past several years, a number of regions. Quantification techniques that are based on the phys-
big questions remain unanswered and, often, unasked. We ical properties of the steroids themselves, such as liquid or gas
discuss several of these below. With the goal of holding up chromatography followed by mass spectrometry, are the most
a magnifying glass to cherished or unexamined beliefs, the precise techniques available, but remain prohibitive for most
tone here is both more speculative and more provocative investigators due to a combination of expense, time, and acces-
than above. sibility. Moreover, the sensitivity of mass spectrometry is at
least 10-fold less than that of radioimmunoassay. As a result,
there is currently a large gap in our knowledge of what are
5.01.4.1 What Are the Actual Steroid Levels in the Brain
the precise levels of steroids in particular brain regions during
during the Sensitive Period of Sexual Differentiation?
the dynamic period of sexual differentiation. Not knowing
Every discipline has its iconic papers, findings so fundamental this information severely handicaps our ability to determine
that the same manuscript is cited over and over again without whether specific parts of the brain are either making steroids
question. As opposed to dogma, these studies do not them- de novo or regulating the rate of metabolism, uptake or binding
selves postulate any theory or guiding principle but instead to increase or decrease exposure in a meaningful way. More
provide data that can be used to support such. In the field of importantly, not knowing when there are or are not sex differ-
sexual differentiation, a small number of studies measuring ences in tissue steroid levels at key developmental time
serum testosterone and hypothalamic estradiol levels in the points precludes us from discovering potentially novel mecha-
neonatal rat have served such a role (Weisz and Ward, 1980; nisms of sexual differentiation of the brain.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
22 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

5.01.4.2 Do Noncoding RNAs Contribute to Sex Differences in equivalent gene expression outcome in males and females.
Brain and Behavior? In the mPOA and VMH of the rat brain, for example, sex differ-
ences in microRNA expression did not correlate with differ-
In addition to the iconic epigenetic modifications discussed
ences in the expression of mRNAs predicted to be targets of
above (DNA methylation and the covalent modification of
those microRNAs (Topper et al., 2015).
histone tails), other epigenetic mechanisms play important
roles in the control of gene expression. The best-known
5.01.4.3 Are Sex Differences Necessary?
example relates to the most pervasive sex difference in mamma-
lian biology: the silencing of one randomly selected X chromo- Sex differences are seen in essentially all sexually reproduc-
some in every cell of females and in no cells of males (Lyon, ing species. Given the near universality of differences
1999). The process of random X inactivation is initiated by between males and females, they seem likely to play a crucial
the transcription from one X chromosome of a long, noncod- role. What this role is, however, can be surprisingly unclear.
ing RNA known as XIST (X inactivation-specific transcript; This is particularly true for neural sex differences, and in
reviewed in Jeon et al., 2012; Gendrel and Heard, 2014). Section 5.01.4.4, we describe some of the difficulties in
XIST coats its chromosome of origin and recruits further epige- ascribing function to known sex differences in the nervous
netic marks that cause profound condensation of the inacti- system.
vated X (Jeon et al., 2012; Gendrel and Heard, 2014). The A related issue is the fact that the degree of sexual dimor-
widely accepted purpose of random X inactivation is to make phism varies enormously. At one extreme are orb-weaving
male and female cells more similar by equalizing the func- spiders in which females may be >12 times the size of males
tional dosage of X chromosome genes. Thus, somewhat ironi- (Hormiga et al., 2000) or, as described above, brain regions
cally, an essentially absolute sex difference in an epigenetic in rats that are several times larger in volume in one sex than
processes (the coating of one X chromosome with XIST and the other. At the other end of the spectrum are species that
the countless epigenetic modifications that follow) functions appear to exhibit few sex differences or, within a species, sex
to make the sexes more alike. differences that are extremely subtle, yet reliable. These obser-
Recently, important roles for other noncoding RNAs in the vations raise the question of whether, or when, sex differences
control of gene expression have been discovered. For example, are necessary. Much of what we know about sexual differentia-
microRNAs are short (22–23 nucleotides), noncoding RNA tion of the brain relies on observations of rats, mice, and guinea
molecules that are transcribed from DNA, bind to complemen- pigs. These are polygynous species in which, under the favor-
tary sites in the 30 untranslated region of specific messenger able conditions of the laboratory, essentially all individuals
RNAs (mRNAs), and target them for destruction (Bartel, that achieve adult body size become reproductive and attempt
2009). A single microRNA can silence up to 100 different to mate. In species exhibiting cooperative breeding, in which
mRNAs and thus has the potential for a profound impact on some members of a social group suppress reproduction to
the posttranscriptional control of gene expression. Whether contribute to the rearing of the offspring of others, other rules
microRNAs should be considered an epigenetic mechanism is may apply. Cooperative breeding with reproductive suppres-
up for debate. If your definition of ‘epigenetics’ requires modi- sion is seen in a wide variety of mammalian taxa, including
fications to chromatin, not all microRNAs qualify. However, the rodents, canids, veverrids, and many examples among the
some microRNAs also induce DNA or histone changes leading primates (Jennions and MacDonald, 1994; Solomon and
to chromatin condensation, which clearly fits the bill (Sato French, 1997). The most extreme example is seen in naked
et al., 2011). mole-rats (Heterocephalus glaber), which are uniquely social
Sex differences have recently been described in the expres- mammals that exhibit the closest mammalian equivalent to
sion of specific microRNAs in whole brain (Morgan and Bale, eusociality. Naked mole-rats live in large subterranean colonies
2011) or cortex (Murphy et al., 2014) of developing and adult averaging 70–90 individuals, including a single breeding
rats. At least some of the sex differences may be due to female (the queen) and 1–3 breeding males (Jarvis, 1981).
programming effects of gonadal steroids, because the adminis- All other members of the colony are reproductively suppressed,
tration of an aromatase inhibitor markedly alters the brain but contribute to overall maintenance and survival of the
microRNA patterns of neonatal male rats (Morgan and Bale, colony, including caring for pups (Brett, 1991; Lacey and
2011). Sex differences have also been reported for microRNA Sherman, 1991). Subordinates are not sterile, however, and
transcription in response to challenges such as endocrine- can become breeders if removed from the colony and placed
disrupting chemicals (Topper et al., 2015) or ischemia (Lusardi with an opposite-sex partner.
et al., 2014). The profile of circulating microRNAs also differs Naked mole-rats are remarkably sexually monomorphic.
in male and female rats after stroke and may be a biomarker Among subordinates, there are no sex differences in overall
of the severity of stroke outcome (Selvamani et al., 2014). body size, anogenital distance, or the expression of a large
As far as we are aware, no study has yet shown an effect of variety of behaviors (Lacey et al., 1991; Peroulakis et al.,
manipulating the expression of a particular microRNA on 2002). In addition, naked mole-rats lack many of the sexual
a behavioral or brain sex difference. In other words, we are still dimorphisms in the brain and spinal cord that are seen in other
awaiting the demonstration that microRNAs contribute to mammals. For example, vasopressin innervation of the septum,
sexual differentiation of the brain. Some sex differences in and morphology of the perineal muscles and motoneurons do
microRNA levels may not cause differences in gene expression not differ between males and females (Peroulakis et al., 2002;
but, as proposed above for sex differences DNA and histone Seney et al., 2006; Rosen et al., 2007; Holmes et al., 2009b).
methylation, may reflect different epigenetic pathways to an In addition, in a stereological analysis of several brain regions

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 23

known to be sexually dimorphic in other rodents, male and that sex differences in vasopressin innervation beget sex differ-
female naked mole-rats did not differ on any measure (Holmes ences in function. They also invoke the idea that the size and
et al., 2007). Instead, brain morphology varied by social status direction of sex differences in vasopressin innervation correlate
(i.e,. differences between breeding and nonbreeding animals, with the size and direction of sex differences in behavior.
regardless of sex). Even the number of kisspeptin neurons, Research in prairie voles, however, suggests that sex differences
which is so dimorphic in other rodents (Section 5.01.2.6.2), in vasopressin innervation cannot be that easily interpreted.
does not differ between subordinate male and female naked Prairie voles resemble spotted hyenas in that they are highly
mole-rats, but depends on breeding status (Zhou et al., social (Getz et al., 1981; Carter et al., 1995). They differ in that
2013). These findings suggest that status, rather than sex, plays social behaviors, including aggressive and parental behavior,
the predominant role in determining neural structure in this are remarkably similar between male and female voles
uniquely social mammal. (Villalba et al., 1997; Lonstein and De Vries, 1999a). Paradox-
The reduction in sex differences in naked mole-rats may be ically, however, the sex difference in vasopressin expression in
related to their unique reproductive strategy, as in nature, it is prairie voles is larger than what has been reported for any other
estimated that fewer than 5% of all naked mole-rats ever mammal to date (Bamshad et al., 1993, 1994; Wang et al.,
achieve reproductive status (Jarvis et al., 1994). Perhaps in 1994b). Evidence suggests that this sex difference in vaso-
this case, it makes ‘sense’ to rely more on acute than on pressin innervation contributes to the absence of sex differences
programming effects of steroids to control brain morphology in parental behavior typically seen in other rodents (Lonstein
and function. Although naked mole-rats present an extreme and De Vries, 2000). Mating appears to release vasopressin in
case, cooperative breeding is widespread. It would be inter- the septum of male prairie voles (Bamshad et al., 1994;
esting to know to what extent programming versus acute effects Wang et al., 1994b). This may stimulate parental behavior,
of steroids shape the brain in other species with social organi- because injecting exogenous vasopressin into the septum
zation that differs from that characteristic of common lab increases paternal responsiveness and V1a receptor antagonists
animals. reduce it (Wang et al., 1994a). This indicates that the vaso-
pressin innervation of the septum is required for parental
behavior in males. As in other female rodents, hormonal
5.01.4.4 Do Sex Differences in Brain Structure Beget Sex
changes associated with pregnancy and parturition appear
Differences in Brain Function?
necessary to trigger parental behavior in female prairie voles
The landmark studies by Phoenix et al. (1959) suggested that (Lonstein and De Vries, 1999b; Hayes and De Vries, 2007).
sex differences in neural structure could explain the permanent Because male prairie voles will never get pregnant let alone
effects of hormones on behavior. When such differences were give birth, they may need a different strategy to boost their
found, they were typically linked to sex differences in function. parental responsiveness. If the sex difference in vasopressin
For example, sex differences in the medial POA were related to expression plays this role, then vasopressin can cause as well
sex differences in male sexual behavior and the regulation of as prevent sex differences in social behavior.
gonadotropic hormone release (Yahr, 1988; Kelley, 1988; This hypothesis is perfectly testable. One would predict in
De Vries, 1990; Breedlove, 1992). This made sense at the the former case (vasopressin causes sex differences) that block-
time, as the medial POA had indeed been implicated in these ing vasopressin neurotransmission would blunt or eliminate
functions. However, there is good evidence that sex differences sex differences and in the latter case (vasopressin prevents sex
in the brain do not only serve to generate sex differences in differences) that blocking vasopressin would cause sex differ-
physiology and behavior. The opposite may be true as well; ences in vasopressin-dependent functions that were not there
that is, structural sex differences may enable the brain to before. In fact, such tests have already been done. For example,
generate similar behaviors in males and females in spite of vasopressin antagonists block social recognition memory in
them being exposed to different hormonal and physiological male but not in female rats, thereby causing a sex difference
conditions (De Vries and Boyle, 1998; De Vries, 2004). that was previously absent (Bluthe and Dantzer, 1990).
We will illustrate this with the sexually dimorphic vasopressin Similarly, a knockout in the V1a receptor gene reduces anxiety
innervation. in male but not in female mice (Bielsky et al., 2005), exactly
As mentioned above, most vertebrate species exhibit a sex what one would predict if a system is more important for
difference in vasopressin innervation, such that males have a function in one sex than in the other. Therefore, the possi-
a greater density of vasopressin fibers projecting from the BST bility that sex differences can cause as well as prevent sex differ-
and MeA to the septum than do females (De Vries and Panzica, ences in function should be considered as two obligate
2006). This has been linked to the higher aggression seen in alternative hypotheses when regarding the function of vaso-
males (Koolhaas et al., 1990, 1991). Interestingly, in Syrian pressin innervation in the brain. It is unlikely that the vaso-
hamsters, which lack vasopressin projections of the BST and pressin innervation is the only neural system for which this is
MeA altogether (Albers et al., 1991), females are as aggressive true. Of the hundreds of sex differences that have been found
as males and tend to dominate males of similar body weight in the central nervous system, only a handful can be clearly
(Payne and Swanson, 1970; Huhman et al., 2003). Similarly, linked to sex differences in behavior. Perhaps the clearest
female spotted hyenas are socially dominant and much more example is the cluster of motoneurons in the spinal cord that
aggressive than males (Hamilton et al., 1986; Glickman et al., innervates muscle groups with a clearly sexually dimorphic
2006), and the traditionally found sex difference in septal function (i.e., the SNB). There is also a strong positive
innervation is absent or reversed in this species (Rosen et al., correlation between the density of dendritic spine synapses
2006). Each of these examples is compatible with the idea on POA neurons and measures of male sexual behavior

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
24 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

(Wright et al., 2008). But of almost all others, we do not know sexual partners and the majority of females prefer males, then
exactly what advantage the dimorphism gives males and this sounds like a profoundly sexually dimorphic and presum-
females in terms of function. Considering the possibility that ably differentiated response. Antecedent to this view would be
these dimorphisms can cause as well as prevent sex differences the assumption that distinct biological processes drive the
in function may help answer this question. neural substrate of partner preference to either a male bias or
a female bias. The existence of distinct processes for a male pref-
5.01.4.4.1 ‘Canalization’ May Explain the Size and Variability erence versus a female preference provides a ready explanation
of Sex Differences in Brain Structure and Behavior for why some females prefer other females, some males prefer
Most neuroanatomical sex differences, be they in synaptic males, and why some individuals have no preference.
profile, neurochemical phenotype, astrocyte morphology, or However, if we take the view that the majority of animals prefer
neurotransmitter levels, are of a relatively modest magnitude. the opposite sex as partners, then there is no sex difference as
Typical are one- to twofold differences in an endpoint mean the same drive exists in males and females but it is manifest
between males and females. But more importantly, the variance differently as a function of one’s own sex. This means that
associated with the mean is often very low, suggesting there is a component of the neural response is computation of one’s
not a continuum of responses with a great deal of overlap own sex, which then determines the response to others’ sex.
between males and females but instead a clustering of males Given the intensity and early onset of both internal and
and females at opposite ends of the spectrum. This is suggestive external influences of sex on brain development, this is not
of a process termed canalization that has been used by evolu- outside the realm of possibility. In humans, we are unlikely
tionary biologists to explain the robustness of species pheno- to ever be able to definitively separate the impact of nature
typic traits in the face of constant challenges from a changing from nurture, and our best alternative is the study of naturally
environment (Gursky et al., 2012). Species phenotypic traits occurring or experimentally manipulated variation in sexual
are maintained in a narrow ‘canal’ by various factors that range preference in animals.
from chaperone proteins to miRNAs to even transposons (see The current state of the art of partner preference research is
for review McCarthy et al., 2015). It is possible that phenotypic found on several fronts. These include studies of the program-
traits of the brains of males and females are also constrained to ming effects of gonadal steroids and early experience on partner
specific ‘canals’ but also not allowed to stray too far apart. preference, the neuroanatomical loci controlling partner prefer-
Behavior, on the other hand, is not so tightly constrained as it ence, and the study of naturally occurring variation in partner
is influenced by a multitude of physiological and contextual preference in animal models. Consistent evidence supports
factors that must be integrated by the nervous system before the view that partner preference is organized by gonadal
a response is made. Thus it would be predicted that there is steroids, such that perinatal androgens, with aromatization to
much more overlap in the behavioral phenotypes of males estrogens in rodents, direct the formation of preference for
and females, especially outside the context of reproduction, a female sexual partner (Brand et al., 1991; Vega Matuszczyk
and this does indeed seem to be the case. et al., 1988). In many mammals, odors are the primary signal
indicating sex. Preference can be assessed by determining the
amount of time a test subject prefers to spend with male versus
5.01.4.5 Is Partner Preference Sexually Dimorphic?
female stimulus animals or by the amount of time spent inves-
Sexual orientation, also referred to as sexual partner preference, tigating odors generated by stimulus animals. Male- versus
is defined by the sex of the individuals that are arousing or attrac- female-specific odors can induce a differential brain response
tive to the reference individual, whether it be an individual of the in the same animal, and likewise, animals of opposite sex
opposite sex (heterosexual), the same sex (homosexual), or both will respond to the same odor differently (Bakker et al.,
sexes (bisexual). The estimated frequency of homosexuality in 1996; Woodley and Baum, 2004). The latter speaks to the
humans ranges from 2% to 10%, suggesting that the large sexual differentiation of partner preference and suggests that
majority of males are sexually oriented toward females and the the olfactory system may be the initiation point for subsequent
majority of females are sexually oriented toward males. The over- behavioral responses. In many species, if olfaction is blocked,
whelming prevalence of one sex preferring the other is a constant there is no partner preference to measure.
across all vertebrate species, as would quite naturally be expected Olfaction is important to humans as well, but visual stimuli
from the point of view of reproductive success. Nonetheless, are far more potent and the arousal potential of same-sex versus
what draws the majority of attention is the much less frequent opposite-sex images depends on the partner preference of the
phenotype of same-sex preference. Notably, the biological basis observer (see for review Baum, 2006). Zebra finches are also
of sexual orientation is a matter of impassioned debate only heavily dependent upon vision for expressing partner preference,
when it involves discussion of the etiology of homosexuality. and steroids influence partner preference in this species as well
Few seem to question whether opposite-sex orientation is bio- (Adkins-Regan and Leung, 2006). The effect is context depen-
logical. But actually we understand little about opposite-sex dent, however, because early experience, i.e., being raised in an
attraction, and it can be argued that understanding the biological environment with a skewed sex ratio, can also strongly influence
basis of same-sex orientation would be greatly advanced by adult partner preference in zebra finches.
understanding opposite-sex orientation. Thus, a fundamental The neuroanatomical substrate of partner preference begins
question is whether sexual orientation per se is sexually with that portion of the brain detecting and decoding the
differentiated. sex-specific sensory signals originating from the stimulus
The answer to this may depend on how you pose the ques- animal, be they olfactory, visual, or auditory. But from there,
tion. If we state that the majority of males prefer females as all signals appear to converge on the POA, and in particular

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 25

an SDN within the POA (see Baum, 2006). An SDN is present 5.01.4.6 Is the Human Brain Sexual Differentiated?
in the POA not only in rats, but also in sheep, gerbils, ferrets,
To some an affirmative answer to this question is self-evident:
hamsters, and humans. Lesions of the SDN and its surround
how could the human brain not be subject to the same process
in rats and ferrets either eliminate or reverse sexual preference
that occurs in the majority if not all other mammals as well as
(for review, see Baum, 2006). In humans, the third interstitial
many birds, reptiles, and amphibians? Even invertebrates have
nucleus of the hypothalamus (INAH3) may be homologous
brains that differ in males and females. But others argue, ‘not so
to the SDN-POA of rats and is larger in men than women
fast, humans are exceptional in many ways.’ We are the only
(Allen et al., 1989). Levay (1991) found that INAH3 is
species with complex computational abilities and a sophisti-
smaller in homosexual men than in heterosexual men, and
cated language that includes generation of an historical record.
a second study found a mean difference in the same direction
We also have rich cultural and societal rituals and expectations
that did not, however, reach statistical significance (Byne
that include prescriptions of the appropriate behavior for boys
et al., 2001). Thus, the size of INAH3 may be a marker of
and girls, men and women. These rules and expectations are
partner preference in men, although this conclusion is not
imposed on children even before they are born with the choice
without its detractors.
of a gender-typical name and continue with modes of dress and
Another approach is the use of biomarkers to determine if
even the manner in which adults interact with an infant of one
an individual was exposed to an endocrine environment
sex versus the other. Thus as we have mentioned many times in
in utero that varies from the norm for that sex. These biomarkers
this chapter, it is impossible to parse out the influence of envi-
include long bone length, hand digit ratio, and the detection of
ronment and experience from biology when considering the
small noises made by the inner ear. In general, these studies
brain and behavior of humans. Nonetheless, it is worth a try.
support the conclusion that the prenatal hormonal milieu
One powerful approach is the study of children at a young
contributes to the propensity to show same-sex orientation
age. While the influence of environment and experience cannot
(Balthazart, 2016). But in many human and animal studies,
be eliminated, it is at least lessened compared to that of a full-
a major and unavoidable confound is either the use of surgical
grown adult. Toy choice reflects an interest in different types of
manipulations, such as lesions, or the health status of the
objects, and this varies on average between boys and girls. The
affected individuals, such as the number of HIV-infected
data are messy, with many boys willing to play with girls’ toys
subjects in the homosexual group in human studies. Neither
on occasion, and vice versa. One of the strongest influences is
of these criticisms applies to the study of a naturally occurring
whether a child has a sibling of the opposite sex, demonstrating
variant of homosexuality, the male-preferring domestic ram.
the importance both of exposure and of modeling the behavior
In at least two different study populations, approximately 8%
of other children. Nonetheless, on average, boys spend more
of rams prefer to mount other male rams. The frequency of
time with certain types of toys and girls with others. Melissa
the phenotype is similar to that observed in humans, and there
Hines has spent most of her career studying this phenomenon
are no clear external markers of male-preferring rams. Analyses
and whether prenatal exposure to androgens in girls can influ-
of the brain reveal that the SDN of male-preferring rams is
ence toy choice. She and others have consistently found that
smaller than that of rams that prefer ewes, and it contains fewer
androgen-exposed girls shift their toy preference to that of
aromatase-expressing neurons. This suggests reduced neuronal
boys (reviewed in Hines, 2006). So does this mean there is
exposure to estradiol developmentally and in adulthood may
a ‘toy’ nucleus in the brain that directs boys to like trucks and
be a critical variable in the establishment of same-sex prefer-
girls to like dolls? That seems unlikely. Hines and colleagues
ence in this species (Roselli et al., 2004).
recently added a new dimension to the sexual differentiation
Thus, on balance, we can conclude that partner preference
of play with the observation that girls prefer to mimic the
is sexually differentiated and that there is an important role for
behavior of other girls or women and that it is this aspect of
gonadal steroid exposure in the organization of partner prefer-
their brain development that is influenced by hormones, not
ence, but early experience may also be important. The primary
the desire to play with a particular object or in a particular
detection of the sensory stimulus emanating from a potential
way (Hines et al., 2016). In girls exposed to androgens in utero,
partner is a critical initiating step but the integration and
the desire to mimic other females is lost, and for reasons not
response to the stimulus appear to be encoded within the
well understood, their preference shifts to toys normally
POA. While these are important advances, there remains
preferred by boys.
much to be learned. Work on the genetics of partner prefer-
The work of Hines and others speaks to the behavior of
ence generated a great deal of interest in the early 1990s
humans, but, as noted above, brain and behavior are not
(Hammer et al., 1993), but there has been little progress on
always closely aligned. Many neuroanatomical sex differences
that front. There is a continuing interest in the role of birth
have been reported in the human brain, but the majority of
order and correlations with handedness, particularly for
these relied on postmortem tissue. Because of this, the majority
male homosexuality, and the proposal of the maternal
also involved the brains of adults although one remarkable
immune hypothesis (Blanchard et al., 2006). But again, we
study looked at a nucleus in the hypothalamus from many
know far less than we should. Moreover, the preponderance
different individuals ranging from birth to old age and found
of information is weighted toward understanding male, as
a sex difference that did not appear until late adolescence
opposed to female partner preferences, although this may be
and then waned again in older adults (Swaab et al., 2003).
defensible given the health implications for male versus
More recently, researchers have taken advantage of noninvasive
female homosexuality. Regardless, progress on both is likely
imaging techniques that allow for longitudinal analyses of the
to remain slow given the paucity of researchers and resources
same individual as they mature. The magnitude and direction
currently dedicated to this topic.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
26 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

of sex differences varies with the mode of data acquisition and Ahmed, E.I., Zehr, J.L., Schulz, K.M., Lorenz, B.H., DonCarlos, L.L., Sisk, C.L., 2008.
analyses (i.e., correcting for total brain volume or not), but Pubertal hormones modulate the addition of new cells to sexually dimorphic brain
regions. Nat. Neurosci. 11, 995–997.
differences in the developmental trajectory of the peak of
Albers, H.E., Rowland, C.M., Ferris, C.F., 1991. Arginine-vasopressin immunoreactivity
cortical gray matter are reliably found and modulated by is not altered by photoperiod or gonadal hormones in the Syrian hamster (Meso-
androgen receptor allelic variation as well as androgen levels. cricetus auratus). Brain Res. 539, 137–142.
White matter increases more rapidly in male brains as develop- Alexander, G.M., Evardone, M., 2008. Blocks and bodies: sex differences in a novel
ment proceeds and combined with differences in gray matter version of the Mental Rotations Test. Horm. Behav. 53, 177–184.
Allen, L.S., Hines, M., Shryne, J.E., Gorski, R.A., 1989. Two sexually dimorphic cell
amplify the magnitude of sex differences across the life span groups in the human brain. J. Neurosci. 9, 497–506.
(reviewed in Giedd et al., 2012). Altemus, M., 2006. Sex differences in depression and anxiety disorders: potential
More recently advances in imaging have allowed for measure- biological determinants. Horm. Behav. 50, 534–538.
ment of functional connectivity. Images from almost 1000 Amateau, S.K., McCarthy, M.M., 2002a. A novel mechanism of dendritic spine
plasticity involving estradiol induction of prostaglandin-E2. J. Neurosci. 22,
humans revealed a profound sex difference in the ‘connectome,’
8586–8596.
with females showing strong interhemispheric connectivity and Amateau, S.K., McCarthy, M.M., 2002b. Sexual differentiation of astrocyte morphology
males the opposite, strong intrahemispheric connections in the developing rat preoptic area. J. Neuroendocrinol. 14, 904–910.
(Ingalhalikar et al., 2014). The authors interpreted their findings Amateau, S.K., McCarthy, M.M., 2004. Induction of PGE2 by estradiol mediates
as supportive of the view that females engage multiple tasks at developmental masculinization of sex behavior. Nat. Neurosci. 7, 643–650.
Amateau, S.K., Alt, J.J., Stamps, C.L., McCarthy, M.M., 2004. Brain estradiol content
once and are highly social while males are focused systematizers. in newborn rats: sex differences, regional heterogeneity, and possible de novo
This stereotypical view generated a firestorm of criticism. In synthesis by the female telencephalon. Endocrinology 145, 2906–2917.
response the authors went on to use a computerized battery of American Psychiatric Association, 2000. Diagnostic and Statistical Manual of
neurocognitive tasks in combination with imaging and largely Mental Disorders, fourth ed., text revision. American Psychiatric Association,
Washington, DC.
supported their original conclusions (Tunc et al., 2016), but the
Arnold, A.P., Chen, X., 2009. What does the “four core genotypes” mouse model tell
controversy here is certainly not resolved. us about sex differences in the brain and other tissues? Front. Neuroendocrinol.
On the opposite end of the spectrum and equally controver- 30, 1–9.
sial was a recent report that reexamined several studies Arnold, A.P., et al., 2016. The importance of having two X chromosomes. Philos.
involving imaging and gender-typical activities. Here the Trans. R. Soc. Lond. B Biol. Sci. 371. Article ID 20150113.
Arnold, A.P., 1992. Developmental plasticity in neural circuits controlling birdsong:
authors concluded that there was no clear predictability of sexual differentiation and the neural basis of learning. J. Neurobiol. 23,
sex based on the mean responses (Joel et al., 2015). Instead, 1506–1528.
they concluded, every brain is a mosaic of male-like, female- Arnold, A.P., 1996. Genetically triggered sexual differentiation of brain and behavior.
like, and neutral features, and therefore there is no such thing Horm. Behav. 30, 495–505.
Auger, A.P., Tetel, M.J., McCarthy, M.M., 2000. Steroid receptor coactivator-1
as a ‘male brain’ or a ‘female brain.’ In some ways this conclu-
(SRC-1) mediates the development of sex-specific brain morphology and
sion is intuitively obvious and consistent with the high degree behavior. Proc. Natl. Acad. Sci. U.S.A. 97, 7551–7555.
of regionally specific mechanisms establishing sex differences Auger, A.P., Perrot-Sinal, T.S., McCarthy, M.M., 2001. Excitatory versus inhibitory
in the brain as determined in rodent models. But the finding GABA as a divergence point in steroid-mediated sexual differentiation of the brain.
was largely misinterpreted by the lay media as demonstrating Proc. Natl. Acad. Sci. U.S.A. 98, 8059–8064.
Auger, A.P., Perrot-Sinal, T.S., Auger, C.J., Ekas, L.A., Tetel, M.J., McCarthy, M.M.,
there are no sex differences in the brain, which was not the 2002. Expression of the nuclear receptor coactivator, cAMP response element-
case even in the Joel et al. study. binding protein, is sexually dimorphic and modulates sexual differentiation of
This serves as a fitting conclusion to our long treatise on the neonatal rat brain. Endocrinology 143, 3009–3016.
modes, mechanisms, and meanings of sex differences in the Bakker, J., Baum, M.J., Slob, A.K., 1996. Neonatal inhibition of brain estrogen
synthesis alters adult neural Fos responses to mating and pheromonal stimulation
brain as it so aptly demonstrates how much we still have to
in the male rat. Neuroscience 74, 251–260.
learn. In some ways, the topic of sex differences in the brain Bakker, J., De Mees, C., Douhard, Q., Balthazart, J., Gabant, P., Szpirer, J.,
remains as controversial today as when the first reports were Szpirer, C., 2006. Alpha-fetoprotein protects the developing female mouse brain
made in the late 1960s early 1970s. With the changing policies from masculinization and defeminization by estrogens. Nat. Neurosci. 9,
at major granting agencies, there is likely to be more, not fewer 220–226.
Bales, K.L., Lewis-Reese, A.D., Pfeifer, L.A., Kramer, K.M., Carter, C.S., 2007. Early
reports of brain and behavior sex differences. This makes even experience affects the traits of monogamy in a sexually dimorphic manner. Dev.
more salient the admonishment that scientists bear the burden Psychobiol. 49, 335–342.
of assuring their work is not used or interpreted inappropriately Balthazart, J., 2016. Sex differences in partner preference in humans and animals.
(Maney, 2016). It is essential that we ‘get it right’ as the impli- Philos. Trans. R. Soc. Lond. B Biol. Sci. 371. Article ID 20150118.
Bamshad, M., Novak, M.A., De Vries, G.J., 1993. Sex and species differences in the
cations of sex differences research reach far beyond the labora-
vasopressin innervation of sexually naive and parental prairie voles, Microtus
tory to medical, educational, and public health policies that ochrogaster, and meadow voles, Microtus pennsylvanicus. J. Neuroendocrinol. 5,
impact the daily lives of all members of society. 247–255.
Bamshad, M., Novak, M.A., De Vries, G.J., 1994. Cohabitation alters vasopressin
innervation and paternal behavior in Prairie voles, Microtus ochrogaster. Physiol.
Behav. 56, 751–758.
References Bangasser, D.A., et al., 2010. Sex differences in corticotropin-releasing factor receptor
signaling and trafficking: potential role in female vulnerability to stress-related
Adkins-Regan, E., Leung, C.H., 2006. Sex steroids modulate changes in social and psychopathology. Mol. Psychiatry 15, 877.
sexual preference during juvenile development in zebra finches. Horm. Behav. 50, Bangasser, D.A., 2013. Sex differences in stress-related receptors: ‘micro’ differences
772–778. with ‘macro’ implications for mood and anxiety disorders. Biol. Sex. Differ. 4, 2.
Ahern, T.H., Krug, S., Carr, A.V., Murray, E.K., Fitzpatrick, E., Bengston, L., Barraclough, C.A., Gorski, R.A., 1961. Evidence that the hypothalamus is responsible
McCutcheon, J., De Vries, G.J., Forger, N.G., 2013. Cell death atlas of the for androgen-induced sterility in the female rat. Endocrinology 68, 68–79.
postnatal mouse ventral forebrain and hypothalamus: effects of age and sex. Barraclough, C.A., Leathem, J.H., 1954. Infertility induced in mice by a single injection
J. Comp. Neurol. 521, 2551–2569. of testosterone propionate. Proc. Soc. Exp. Biol. Med. 85, 673–674.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 27

Barraclough, C.A., 1961. Production of anovulatory, sterile rats by single injections of Chung, W.C., Swaab, D.F., De Vries, G.J., 2000. Apoptosis during sexual differenti-
testosterone propionate. Endocrinology 68, 62–67. ation of the bed nucleus of the stria terminalis in the rat brain. J. Neurobiol. 43,
Barrett, A.M., 1999. Probable Alzheimer’s disease: gender-related issues. J. Gend. 234–243.
Specif. Med. 2, 55–60. Clarkson, J., Herbison, A.E., 2006. Postnatal development of kisspeptin neurons in
Bartel, D.P., 2009. MicroRNAs: target recognition and regulatory functions. Cell 136, mouse hypothalamus; sexual dimorphism and projections to gonadotropin-
215–233. releasing hormone neurons. Endocrinology 147, 5817–5825.
Baum, M.J., 2006. Mammalian animal models of psychosexual differentiation: when is Cooke, B.M., Tabibnia, G., Breedlove, S.M., 1999. A brain sexual dimorphism
‘translation’ to the human situation possible? Horm. Behav. 50, 579–588. controlled by adult circulating androgens. Proc. Natl. Acad. Sci. U.S.A. 96,
Bayer, S.A., Altman, J., 1987. Development of the preoptic area: time and site of 7538–7540.
origin, migratory routes, and settling patterns of its neurons. J. Comp. Neurol. 265, Cooke, B.M., 2006. Steroid-dependent plasticity in the medial amygdala. Neuroscience
65–95. 138, 997–1005.
Bayer, S.A., 1987. Neurogenetic and morphogenetic heterogeneity in the bed nucleus Cosgrove, M.S., Wolberger, C., 2005. How does the histone code work? Biochem. Cell
of the stria terminalis. J. Comp. Neurol. 265, 47–64. Biol. 83, 468–476.
Beach, F.A., Noble, R.G., Orndoff, R.K., 1969. Effects of perinatal androgen treatment Cosgrove, K.P., Mazure, C.M., Staley, J.K., 2007. Evolving knowledge of sex
on responses of male rats to gonadal hormones in adulthood. J. Comp. Physiol. differences in brain structure, function, and chemistry. Biol. Psychiatry 62,
Psychol. 68, 490–497. 847–855.
Becker, J.B., Hu, M., 2008. Sex differences in drug abuse. Front. Neuroendocrinol. 29, Davidson, J.M., 1966. Activation of the male rat’s sexual behavior by intracerebral
36–47. implantation of androgen. Endocrinology 79, 783–794.
Berletch, J.B., Yang, F., Disteche, C.M., 2010. Escape from X inactivation in mice and Davis, E.C., Popper, P., Gorski, R.A., 1996a. The role of apoptosis in sexual differ-
humans. Genome Biol. 11, 213. entiation of the rat sexually dimorphic nucleus of the preoptic area. Brain Res. 734,
Bezzi, P., Carmignoto, G., Pasti, L., Vesce, S., Rossi, D., Rizzini, B.L., Pozzan, T., 10–18.
Volterra, A., 1998. Prostaglandins stimulate calcium-dependent glutamate release Davis, E.C., Shryne, J.E., Gorski, R.A., 1996b. Structural sexual dimorphisms in the
in astrocytes. Nature 391, 281–285. anteroventral periventricular nucleus of the rat hypothalamus are sensitive to
Bielsky, I.F., Hu, S.B., Young, L.J., 2005. Sexual dimorphism in the vasopressin gonadal steroids perinatally, but develop peripubertally. Neuroendocrinology 63,
system: lack of an altered behavioral phenotype in female V1a receptor knockout 142–148.
mice. Behav. Brain Res. 164, 132–136. Davis, A.M., Grattan, D.R., Selmanoff, M., McCarthy, M.M., 1996c. Sex differences in
Blanchard, R., Cantor, J.M., Bogaert, A.F., Breedlove, S.M., Ellis, 2006. Interaction of glutamic acid decarboxylase mRNA in neonatal rat brain: implications for sexual
fraternal birth order and handedness in the development of male homosexuality. differentiation. Horm. Behav. 30, 538–552.
Horm. Behav. 49, 405–414. Davis, A.M., Ward, S.C., Selmanoff, M., Herbison, A.E., McCarthy, M.M., 1999.
Bluthe, R.-M., Dantzer, R., 1990. Social recognition does not involve vasopressinergic Developmental sex differences in amino acid neurotransmitter levels in hypotha-
neurotransmission in female rats. Brain Res. 535, 301–304. lamic and limbic areas of rat brain. Neuroscience 90, 1471–1482.
Bodo, C., Rissman, E.F., 2007. Androgen receptor is essential for sexual differ- De Vries, G.J., Boyle, P.A., 1998. Double duty for sex differences in the brain. Behav.
entiation of responses to olfactory cues in mice. Eur. J. Neurosci. 25, Brain Res. 92, 205–213.
2182–2190. De Vries, G.J., Panzica, G.C., 2006. Sexual differentiation of central vasopressin and
Bodo, C., Rissman, E.F., 2008. The androgen receptor is selectively involved in vasotocin systems in vertebrates: different mechanisms, similar endpoints.
organization of sexually dimorphic social behaviors in mice. Endocrinology 149, Neuroscience 138, 947–955.
4142–4150. De Vries, G.J., Simerly, R.B., 2002. Anatomy, development, and function of sexually
Bodo, C., Kudwa, A.E., Rissman, E.F., 2006. Both estrogen receptor-alpha and -beta dimorphic neural circuits in the mammalian brain. In: Pfaff, D.W., Arnold, A.P.,
are required for sexual differentiation of the anteroventral periventricular area in Etgen, A.M., Fahrbach, S.E., Moss, R.L., Rubin, R.T. (Eds.), Hormones, Brain, and
mice. Endocrinology 147, 415–420. Behavior, Development of Hormone-Dependent Neuronal Systems, vol. 4.
Bowers, J.M., Waddell, J., McCarthy, M.M., 2010. A developmental sex difference in Academic Press, San Diego, pp. 137–191.
hippocampal neurogenesis is mediated by endogenous oestradiol. Biol. Sex. Differ. De Vries, G.J., Rissman, E.F., Simerly, R.B., Yang, L.Y., Scordalakes, E.M.,
1, 8. Auger, C.J., Swain, A., Lovell-Badge, R., Burgoyne, P.S., Arnold, A.P., 2002.
Brand, T., Kroonen, J., Mos, J., Slob, A.K., 1991. Adult partner preference and sexual A model system for study of sex chromosome effects on sexually dimorphic neural
behavior of male rats affected by perinatal endocrine manipulations. Horm. Behav. and behavioral traits. J. Neurosci. 22, 9005–9014.
25, 323–341. De Vries, G.J., Jardon, M., Reza, M., Rosen, G.J., Immerman, E., Forger, N.G., 2008.
Breedlove, S.M., Arnold, A.P., 1980. Hormone accumulation in a sexually dimorphic Sexual differentiation of vasopressin innervation of the brain: cell death versus
motor nucleus of the rat spinal cord. Science 210, 564–566. phenotypic differentiation. Endocrinology 149 (9), 4632–4637.
Breedlove, S.M., Arnold, A.P., 1981. Sexually dimorphic motor nucleus in the rat De Vries, G.J., 1990. Sex differences in neurotransmitters in the brain.
lumbar spinal cord: response to adult hormone manipulation, absence in androgen- J. Neuroendocrinol. 2, 1–13.
insensitive rats. Brain Res. 225, 297–307. De Vries, G.J., 2004. Sex differences in adult and developing brains; compensation,
Breedlove, S.M., Jacobson, C.D., Gorski, R.A., Arnold, A.P., 1982. Masculinization of compensation, compensation. Endocrinology 145, 1063–1068.
the female rat spinal cord following a single neonatal injection of testosterone Del Pino Sans, J., Krishnan, S., Aggison, L.K., Adams, H.L., Shrikant, M.M., López-
propionate but not estradiol benzoate. Brain Res. 237, 173–181. Giráldez, F., Petersen, S.L., 2015. Microarray analysis of neonatal rat ante-
Breedlove, S.M., Jordan, C.L., Arnold, A.P., 1983. Neurogenesis of motoneurons in the roventral periventricular transcriptomes identifies the proapoptotic Cugbp2 gene as
sexually dimorphic spinal nucleus of the bulbocavernosus in rats. Brain Res. 285, sex-specific and regulated by estradiol. Neuroscience 303, 312–322.
39–43. Diaz, D.R., Fleming, D.E., Rhees, R.W., 1995. The hormone-sensitive early postnatal
Breedlove, S.M., 1992. Sexual dimorphism in the vertebrate nervous system. periods for sexual differentiation of feminine behavior and luteinizing hormone
J. Neurosci. 12, 4133–4142. secretion in male and female rats. Brain Res. Dev. Brain Res. 86, 227–232.
Brett, R.A., 1991. The population structure of naked mole-rat colonies. In: Döhler, K.D., Coquelin, A., Davis, F., Hines, M., Shryne, J.E., Sickmöller, P.M.,
Sherman, P.W., Jarvis, J.U.M., Alexander, R.D. (Eds.), The Biology of the Naked Jarzab, B., Gorski, R.A., 1986. Pre- and post-natal influence of an estrogen
Mole-Rat. Princeton University Press, New Jersey, pp. 97–136. antagonist and an androgen antagonist on differentiation of the sexually dimorphic
Byne, W., Tobet, S., Mattiace, L.A., Lasco, M.S., Kemether, E., Edgar, M.A., nucleus of the preoptic area in male and female rats. Neuroendocrinology 42,
Morgello, S., Buchsbaum, M.S., Jones, L.B., 2001. The interstitial nuclei of the 443–448.
human anterior hypothalamus: an investigation of variation with sex, sexual DonCarlos, L.L., Handa, R.J., 1994. Developmental profile of estrogen receptor mRNA
orientation, and HIV status. Horm. Behav. 40, 86–92. in the preoptic area of male and female neonatal rats. Brain Res. Dev. Brain Res.
Carani, C., Granata, A.R., Rochira, V., Caffagni, G., Aranda, C., Antunez, P., 79, 83–289.
Maffei, L.E., 2005. Sex steroids and sexual desire in a man with a novel mutation Dugger, B.N., Morris, J.A., Jordan, C.L., Breedlove, S.M., 2007. Androgen receptors
of aromatase gene and hypogonadism. Psychoneuroendocrinology 30, 413–417. are required for full masculinization of the ventromedial hypothalamus (VMH) in
Carrer, H.F., Cambiasso, M.J., Gorosito, S., 2005. Effects of estrogen on neuronal rats. Horm. Behav. 51, 195–201.
growth and differentiation. J. Steroid Biochem. Mol. Biol. 93, 319–323. Dungan, H.M., Clifton, D.K., Steiner, R.A., 2006. Minireview: kisspeptin neurons as
Carter, C.S., DeVries, A.C., Getz, L.L., 1995. Physiological substrates of mammalian central processors in the regulation of gonadotropin-releasing hormone secretion.
monogamy: the prairie vole model. Neurosci. Biobehav. Rev. 19, 303–314. Endocrinology 147, 1154–1158.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
28 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

Durazzo, A., Morris, J.A., Breedlove, S.M., Jordan, C.L., 2007. Effects of the testicular Gorski, R.A., Gordon, J.H., Shryne, J.E., Southam, A.M., 1978. Evidence for
feminization mutation (tfm) of the androgen receptor gene on BSTMPM volume and a morphological sex difference within the medial preoptic area of the rat brain.
morphology in rats. Neurosci. Lett. 419, 168–171. Brain Res. 148, 333–346.
Ernsberger, U., Rohrer, H., 1999. Development of the cholinergic neurotransmitter Gorski, R.A., Harlan, R.E., Jacobson, C.D., Shryne, J.E., Southam, A.M., 1980.
phenotype in postganglionic sympathetic neurons. Cell Tissue Res. 297, 339–361. Evidence for the existence of a sexually dimorphic nucleus in the preoptic area of
Everitt, B.J., Meister, B., Hökfelt, T., Melander, T., Terenius, L., Rökaeus, A., the rat. J. Comp. Neurol. 193, 529–539.
Theodorsson-Norheim, E., Dockray, G., Edwardson, J., Cuello, C., 1986. The Gotsiridze, T., Kang, N., Jacob, D., Forger, N.G., 2007. Development of sex differences
hypothalamic arcuate nucleus-median eminence complex: immunohistochemistry in the principal nucleus of the bed nucleus of the stria terminalis of mice: role of
of transmitters, peptides and DARPP-32 with special reference to coexistence in Bax-dependent cell death. Dev. Neurobiol. 67, 355–362.
dopamine neurons. Brain Res. 396, 97–155. Gottsch, M.L., Cunningham, M.J., Smith, J.T., Popa, S.M., Acohido, B.V.,
Felsenfeld, G., Groudine, M., 2003. Controlling the double helix. Nature 421, Crowley, W.F., Seminara, S., Clifton, D.K., Steiner, R.A., 2004. A role for kiss-
448–453. peptins in the regulation of gonadotropin secretion in the mouse. Endocrinology
Fink, G., Sumner, B.E., McQueen, J.K., Wilson, H., Rosie, R., 1998. Sex steroid control 145, 4073–4077.
of mood, mental state and memory. Clin. Exp. Pharmacol. Physiol. 25, 764–775. Goy, R.W., McEwen, B.S., 1980. Sexual Differentiation of the Brain. The MIT Press,
Forger, N.G., Breedlove, S.M., 1987. Seasonal variation in mammalian striated muscle Cambridge MA.
mass and motoneuron morphology. J. Neurobiol. 18, 155–165. Gu, G.B., Simerly, R.B., 1997. Projections of the sexually dimorphic anteroventral
Forger, N.G., Rosen, G.J., Waters, E.M., Jacob, D., Simerly, R.B., de Vries, G.J., 2004. periventricular nucleus in the female rat. J. Comp. Neurol. 384, 142–164.
Deletion of Bax eliminates sex differences in the mouse forebrain. Proc. Natl. Acad. Guillamón, A., de Blas, M.R., Segovia, S., 1988. Effects of sex steroids on the
Sci. U.S.A. 101, 13666–13671. development of the locus coeruleus in the rat. Brain Res. 468, 306–310.
Forger, N.G., De Vries, G.J., Breedlove, S.M., 2015a. Sexual Differentiation of Brain Gursky, V.V., Surkova, S.Y., Samsonova, M.G., 2012. Mechanisms of developmental
and Behavior. Knobil and Neill’s Physiology of Reproduction, fourth ed. Elsevier, robustness. Bio Syst. 109, 329–335.
Amsterdam, pp. 2109–2155. Häfner, H., Riecher-Rössler, A., an der Hieden, W., Maurer, K., Fätkenheuer, B.,
Forger, N.G., Strahan, J.A., Castillo-Ruiz, A., 2015b. Cellular and molecular mecha- Löffler, W., 1993. Generating and testing a causal explanation of the gender
nisms of sexual differentiation in the mammalian nervous system. Front. Neuro- difference in age at first onset of schizophrenia. Psychol. Med. 23, 925–940.
endocrinol. 40, 67–86. Hamer, D.H., Hu, S., Magnuson, V.L., Hu, N., Pattatucci, A.M., 1993. A linkage
Forger, N.G., 2001. Development of sex differences in the nervous system. In: between DNA markers on the X chromosome and male sexual orientation. Science
Blass, E. (Ed.), Handbook of Behavioral Neurobiology, vol. 13. Kluwer Academic/ 261, 321–327.
Plenum Publishers, New York, pp. 143–198. Hamilton, W.J., Tilson, R.L., Frank, L.G., 1986. Sexual monomorphism in spotted
Forger, N.G., 2006. Cell death and sexual differentiation of the nervous system. hyenas, Crocuta crocuta. Ethology 71, 63–73.
Neuroscience 138, 929–938. Han, T.M., De Vries, G.J., 1999. Neurogenesis of galanin cells in the bed nucleus of
Forger, N.G., 2016. Epigenetic mechanisms in sexual differentiation of the brain and the stria terminalis and centromedial amygdala in rats: a model for sexual differ-
behaviour. Philos. Trans. R. Soc. Lond. B Biol. Sci. 371 (1688). entiation of phenotype. J. Neurobiol. 38, 491–498.
Fratiglioni, L., Viitanen, M., von Strauss, E., Tontodonati, V., Herlitz, A., Winblad, B., Han, T.M., De Vries, G.J., 2003. Organizational effects of testosterone, estradiol, and
1997. Very old women at highest risk of dementia and Alzheimer’s disease: dihydrotestosterone on vasopressin mRNA expression in the bed nucleus of the
incidence data from the Kungsholmen Project, Stockholm. Neurology 48, stria terminalis. J. Neurobiol. 54, 502–510.
132–138. Hansen, S., Köhler, C., Goldstein, M., Steinbusch, H.V.M., 1982. Effects of ibotenic
Garcia-Segura, L.M., Chowen, J.A., Dueñas, M., Torres-Aleman, I., Naftolin, F., 1994. acid-induced neuronal degeneration in the medial preoptic area and the lateral
Gonadal steroids as promoters of neuro-glial plasticity. Psychoneuroendocrinology hypothalamic area on sexual behavior in the male rat. Brain Res. 239,
19, 445–453. 213–232.
Gendrel, A.V., Heard, E., 2014. Noncoding RNAs and epigenetic mechanisms during Harris, G.W., Jacobsohn, D., 1952. Functional grafts of the anterior pituitary gland.
X-chromosome inactivation. Annu. Rev. Cell Dev. Biol. 30, 561–580. Proc. R. Soc. Lond. B. Biol. Sci. 139, 263–276.
Getz, L.L., Carter, C.S., Gavish, L., 1981. The mating system of prairie vole, Microtus Hayes, U.L., De Vries, G.J., 2007. Role of pregnancy and parturition in induction of
ochrogaster: field and laboratory evidence for pair bonding. Behav. Ecol. Sociobiol. maternal behavior in prairie voles (Microtus ochrogaster). Horm. Behav. 51,
8, 189–194. 265–272.
Ghahramani, N.M., Ngun, T.C., Chen, P.Y., Tian, Y., Krishnan, S., Muir, S., Rubbi, L., Henderson, R.G., Brown, A.E., Tobet, S.A., 1999. Sex differences in cell migration in
Arnold, A.P., de Vries, G.J., Forger, N.G., Pellegrini, M., Vilain, E., 2014. the preoptic area/anterior hypothalamus of mice. J. Neurobiol. 41, 252–266.
The effects of perinatal testosterone exposure on the DNA methylome of the mouse Hines, M., Ahmed, S.F., Hughes, I.A., 2003. Psychological outcomes and gender-
brain are late-emerging. Biol. Sex. Differ. 5, 8. related development in complete androgen insensitivity syndrome. Arch. Sex.
Giedd, J.N., Raznahan, A., Mills, K.L., Lenroot, R.K., 2012. Review: magnetic reso- Behav. 32, 93–101.
nance imaging of male/female differences in human adolescent brain anatomy. Hines, M., Pasterski, V., Spencer, D., Neufeld, S., Patalay, P., Hindmarsh, P.C.,
Biol. Sex. Differ. 3, 19. Hughes, I.A., Acerini, C.L., 2016. Prenatal androgen exposure alters girls’
Gilmore, R.F., Varnum, M.M., Forger, N.G., February 2012. Effects of blocking responses to information indicating gender-appropriate behaviour. Philos. Trans. R.
developmental cell death on sexually dimorphic calbindin cell groups in the preoptic Soc. Lond. B Biol. Sci. 371. Article ID 20150125.
area and bed nucleus of the stria terminalis. Biol. Sex. Differ. 15 (3), 5. Hines, M., 2006. Prenatal testosterone and gender-related behaviour. Eur. J. Endo-
Gissler, M., Järvelin, M.R., Louhiala, P., Hemminiki, E., 1999. Boys have more health crinol. 155 (Suppl. 1), S115–S121.
problems in childhood than girls: follow-up of the 1987 Finnish birth cohort. Acta Hines, M., 2008. Early androgen influences on human neural and behavioural
Pediatr. 88, 310–314. development. Early Hum. Dev. 84, 805–807.
Glickman, S.E., Cunha, G.R., Drea, C.M., Conley, A.J., Place, N.J., 2006. Mammalian Hisasue, S., Seney, M.L., Immerman, E., Forger, N.G., 2010. Control of cell number in
sexual differentiation: lessons from the spotted hyena. Trends Endocrinol. Metab. the bed nucleus of the stria terminalis of mice: role of testosterone metabolites and
17, 349–356. estrogen receptor subtypes. J. Sex. Med. 7 (4 Pt 1), 1401–1409.
Goldstein, L.A., Sengelaub, D.R., 1994. Differential effects of dihydrotestosterone and Hoffmann, C., 2000. COX-2 in brain and spinal cord implications for therapeutic use.
estrogen on the development of motoneuron morphology in a sexually dimorphic rat Curr. Med. Chem. 7, 1113–1120.
spinal nucleus. J. Neurobiol. 25, 878–892. Hojo, Y., Hattori, T.A., Enami, T., Furukawa, A., Suzuki, K., Ishii, H.T., Mukai, H.,
Goldstein, L.A., Kurz, E.M., Sengelaub, D.R., 1990. Androgen regulation of dendritic Morrison, J.H., Janssen, W.G., Kominami, S., Harada, N., Kimoto, T., Kawato, S.,
growth and retraction in the development of a sexually dimorphic spinal nucleus. 2004. Adult male rat hippocampus synthesizes estradiol from pregnenolone by
J. Neurosci. 10, 935–946. cytochromes P45017alpha and P450 aromatase localized in neurons. Proc. Natl.
Gonzalez-Martinez, D., De Mees, C., Douhard, Q., Szpirer, C., Bakker, J., 2008. Acad. Sci. U.S.A. 101, 865–870.
Absence of gonadotropin-releasing Hormone 1 and Kiss1 activation in alpha- Holloway, C.C., Clayton, D.F., 2001. Estrogen synthesis in the male brain triggers
fetoprotein knockout mice: prenatal estrogens defeminize the potential to show development of the avian song control pathway in vitro. Nat. Neurosci. 4,
preovulatory luteinizing hormone surges. Endocrinology 149, 2333–2340. 170–175.
Goodson, J.L., Bass, A.H., 2001. Social behavior functions and related anatomical Holman, S.D., Collado, P., Skepper, J.N., Rice, A., 1996. Postnatal development of
characteristics of vasotocin/vasopressin systems in vertebrates. Brain Res. Rev. a sexually dimorphic, hypothalamic nucleus in gerbils: a stereological study of
35, 246–265. neuronal number and apoptosis. J. Comp. Neurol. 376, 315–325.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 29

Holmes, M.M., Rosen, G.J., Jordan, C.L., De Vries, G.J., Goldman, B.D., Forger, N.G., Kauffman, A.S., Gottsch, M.L., Roa, J., Byquist, A.C., Crown, A., Clifton, D.K.,
2007. Social control of brain morphology in a eusocial mammal. Proc. Natl. Acad. Hoffman, G.E., Steiner, R.A., Tena-Sempere, M., 2007. Sexual differentiation of
Sci. U.S.A. 104, 10548–10552. Kiss1 gene expression in the brain of the rat. Endocrinology 148, 1774–1783.
Holmes, M.M., Goldman, B.D., Forger, N.G., 2008. Social status and sex indepen- Kay, J.N., Hannigan, P., Kelley, D.B., 1999. Trophic effects of androgen: development
dently influence androgen receptor expression in the eusocial naked mole-rat brain. and hormonal regulation of neuron number in a sexually dimorphic vocal motor
Horm. Behav. 54 [March 28:epub ahead of print]. nucleus. J. Neurobiol. 40, 375–385.
Holmes, M.M., McCutcheon, J., Forger, N.G., 2009a. Sex differences in NeuN- and Kelley, D.B., 1988. Sexually dimorphic behavior. Annu. Rev. Neurosci. 11, 225–251.
androgen receptor-positive cells in the bed nucleus of the stria terminalis are due Kelly, D.A., Varnum, M.M., Krentzel, A.A., Krug, S., Forger, N.G., 2013. Differential
to Bax-dependent cell death. Neuroscience 158, 1251–1256. control of sex differences in estrogen receptor a in the bed nucleus of the stria
Holmes, M.M., Goldman, B.D., Goldman, S.L., Seney, M.L., Forger, N.G., 2009b. terminalis and anteroventral periventricular nucleus. Endocrinology 154,
Neuroendocrinology and sexual differentiation in eusocial mammals. Front. Neu- 3836–3846.
roendocrinol. 305, 19–33. Kessler, R.C., McGonagle, K.A., Zhao, S., Nelson, C.B., Hughes, M., Eshleman, S.,
Holmes, M.M., Van Mil, S., Bulkowski, C., Goldman, S.L., Goldman, B.D., Forger, N.G., Wittchen, H.U., Kendler, K.S., 1994. Lifetime and 12-month prevalence of DSM-III-
2013. Androgen receptor distribution in the social decision-making network of R psychiatric disorders in the United States. Results from the National Comorbidity
eusocial naked mole-rats. Behav. Brain Res. 256, 214–218. Survey. Arch. Gen. Psychiatry 51, 8–19.
Hormiga, G., Scharff, N., Coddington, J.A., 2000. The phylogenetic basis of sexual size Keverne, E.B., Curley, J.P., 2008. Epigenetics, brain evolution and behaviour. Front.
dimorphism in orb-weaving spiders (Araneae, Orbiculariae). Syst. Biol. 49, 435–462. Neuroendocrinol. 29, 398–412.
Hsu, H.-K., Shao, P.-L., Tsai, K.-L., Shih, H.-C., Lee, T.-Y., Hsu, C., 2005. Gene Kim, S.J., Linden, D.J., 2007. Ubiquitous plasticity and memory storage. Neuron 56,
regulation by NMDA receptor activation in the SDN-POA neurons of male rats 582–592.
during sexual development. J. Mol. Endocrinol. 34, 433–445. Kimchi, T., Xu, J., Dulac, C., 2007. A functional circuit underlying male sexual behavior
Huhman, K.L., Solomon, M.B., Janicki, M., Harmon, A.C., Lin, S.M., Israel, J.E., in the female mouse brain. Nature 448, 1009–1015.
Jasnow, A.M., 2003. Conditioned defeat in male and female Syrian hamsters. Kirn, J.R., DeVoogd, T.J., 1989. Genesis and death of vocal control neurons during
Horm. Behav. 44, 293–299. sexual differentiation in the zebra finch. J. Neurosci. 9, 3176–3187.
Hutton, L.A., Gu, G.B., Simerly, R.B., 1998. Development of a sexually dimorphic Kishimoto, M., Fujiki, R., Takezawa, S., Sasaki, Y., Nakamura, T., Yamaoka, K.,
projection from the bed nuclei of the stria terminalis to the anteroventral peri- Kitagawa, H., Kato, S., 2006. Nuclear receptor mediated gene regulation through
ventricular nucleus in the rat. J. Neurosci. 18, 3003–3013. chromatin remodeling and histone modifications. Endocr. J. 53, 157–172.
Ibanez, M.A., Gu, G., Simerly, R.B., 2001. Target dependent sexual differentiation of Knoll, J.G., Wolfe, C.A., Tobet, S.A., 2007. Estrogen modulates neuronal movements
a limbic-hypothalamic pathway. J. Neurosci. 21, 5652–5659. within the developing preoptic area-anterior hypothalamus. Eur. J. Neurosci. 26,
Ingalhalikar, M., Smith, A., Parker, D., Satterthwaite, T.D., Elliott, M.A., Ruparel, K., 1091–1099.
Hakonarson, H., Gur, R.E., Gur, R.C., Verma, R., 2014. Sex differences in the structural Koehl, M., Battle, S., Meerlo, P., 2006. Sex differences in sleep: the response to sleep
connectome of the human brain. Proc. Natl. Acad. Sci. U.S.A. 14, 823–828. deprivation and restraint stress in mice. Sleep 29, 1224–1231.
Irwig, M.S., Fraley, G.S., Smith, J.T., Acohido, B.V., Popa, S.M., Cunningham, M.J., Koolhaas, J.M., Van den Brink, T.H.C., Roozendaal, B., Boorsma, F., 1990. Medial
Gottsch, M.L., Clifton, D.K., Steiner, R.A., 2004. Kisspeptin activation of gonad- amygdala and aggressive behavior: interaction between testosterone and vaso-
otropin releasing hormone neurons and regulation of KiSS-1 mRNA in the male rat. pressin. Aggress. Behav. 16, 223–229.
Neuroendocrinology 80, 264–272. Koolhaas, J.M., Moor, E., Hiemstra, Y., Bohus, B., 1991. The testosterone dependent
Jacob, D.A., Bengston, C.L., Forger, N.G., 2005. Effects of Bax gene deletion on vasopressinergic neurons in the medial amygdala and lateral septum: involvement
muscle and motoneuron degeneration in a sexually dimorphic neuromuscular in social behaviour of male rats. In: Jard, S., Jamison, R. (Eds.), Vasopressin,
system. J. Neurosci. 25, 5638–5644. Colloque INSERM, vol. 208. John Libbey Eurotext Ltd, London, pp. 213–219.
Jacobson, C.D., Gorski, R.A., 1981. Neurogenesis of the sexually dimorphic nucleus of Koopman, P., Gubbay, J., Vivian, N., Goodfellow, P., Lovell-Badge, R., 1991. Male devel-
the preoptic area in the rat. J. Comp. Neurol. 196, 519–529. opment of chromosomally female mice transgenic for Sry. Nature 351, 117–121.
Jansson, J.O., Ekberg, S., Isaksson, O.G., Edén, S., 1984. Influence of gonadal Kornstein, S.G., Sloan, D.M., Thase, M.E., 2002. Gender-specific differences in
steroids on age- and sex-related secretory patterns of growth hormone in the rat. depression and treatment response. Psychoparmacol. Bull. 36, 99–112.
Endocrinology 114, 1287–1294. Krebs-Kraft, D.L., Hill, M.N., Hillard, C.J., McCarthy, M.M., 2010. Sex difference in cell
Jarvis, J.U.M., O’Riain, M.J., Bennett, N.C., Sherman, P.W., 1994. Mammalian proliferation in developing rat amygdala mediated by endocannabinoids has
eusociality: a family affair. Trends Ecol. Evol. 9, 47–51. implications for social behavior. Proc. Natl. Acad. Sci. U.S.A. 107, 20535–20540.
Jarvis, J.U.M., 1981. Eusociality in a mammal: cooperative breeding in naked mole-rat Krishnan, S., Intlekofer, K.A., Aggison, L.K., Petersen, S.L., 2009. Central role of
colonies. Science 212, 571–573. TRAF-interacting protein in a new model of brain sexual differentiation. Proc. Natl.
Jeffrey, K.J., 2007. Integration of the sensory inputs to place cells: what, where, why, Acad. Sci. U.S.A. 106, 16692–16697.
and how? Hippocampus 19, 775–785. Kudwa, A.E., Michopoulos, V., Gatewood, J.D., Rissman, E.F., 2006. Roles of estrogen
Jennions, M.D., Macdonald, D.W., 1994. Cooperative breeding in mammals. Trends receptors alpha and beta in differentiation of mouse sexual behavior. Neuroscience
Ecol. Evol. 9, 89–93. 138, 921–928.
Jenuwein, T., Allis, C.D., 2001. Translating the histone code. Science 293, Kurz, E.M., Sengelaub, D.R., Arnold, A.P., 1986. Androgens regulate the dendritic
1074–1080. length of mammalian motoneurons in adulthood. Science 232, 395–398.
Jeon, Y., Sarma, K., Lee, J.T., 2012. New and Xisting regulatory mechanisms of X Lacey, E.A., Sherman, P.W., 1991. Social organization of naked mole-rat colonies:
chromosome inactivation. Curr. Opin. Genet. Dev. 22, 62–71. evidence for divisions of labor. In: Sherman, P.W., Jarvis, J.U.M., Alexander, R.D.
Jiang, Y., Langley, B., Lubin, F.D., Renthal, W., Wood, M.A., Yasui, D.H., Kumar, A., (Eds.), The Biology of the Naked Mole-Rat. Princeton University Press, New Jersey,
Nestler EJ,Akbarian, S., Beckel-Mitchener, A.C., 2008. Epigenetics in the nervous pp. 275–336.
system. J. Neurosci. 28, 11753–11759. Lacey, E.A., Alexander, R.D., Braude, S.H., Sherman, P.W., Jarvis, J.U.M., 1991. An
Joel, D., Berman, Z., Tavor, I., Wexler, N., Gaber, O., Stein, Y., Shefi, N., Pool, J., ethogram for the naked mole-rat: nonvocal behaviors. In: Sherman, P.W.,
Urchs, S., Margulies, D.S., Liem, F., Hänggi, J., Jäncke, L., Assaf, Y., 2015. Sex Jarvis, J.U.M., Alexander, R.D. (Eds.), The Biology of the Naked Mole-Rat.
beyond the genitalia: the human brain mosaic. Proc. Natl. Acad. Sci. U.S.A. 112, Princeton University Press, New Jersey, pp. 209–242.
15468–15473. Landis, S.C., 1990. Target regulation of neurotransmitter phenotype. Trends Neurosci.
Jones, M.E., Boon, W.C., Proietto, J., Simpson, E.R., 2006. Of mice and men: the 13, 344–350.
evolving phenotype of aromatase deficiency. Trends Endocrinol. Metab. 17, Lenz, K.M., Nugent, B.M., Haliyur, R., McCarthy, M.M., 2013. Microglia are essential
55–64. to masculinization of brain and behavior. J. Neurosci. 33, 2761–2772.
Jost, A., 1947. Recherches sur la differenciation sexuelle de lembryon de lapin .1.intro- Lephart, E.D., 1996. A review of brain aromatase cytochrome P450. Brain Res. Rev.
duction et embryologie genitale normale. Arch. Anat. Microsc. Morphol. Exp. 36, 22, 1–26.
151–200. LeVay, S., 1991. A difference in hypothalamic structure between heterosexual and
Jyotika, J., McCutcheon, J., Laroche, J., Blaustein, J.D., Forger, N.G., 2007. Deletion homosexual men. Science 253, 1034–1037.
of the Bax gene disrupts sexual behavior and modestly impairs motor function in Li, C., Chen, P., Smith, M.S., 1999. Morphological evidence for direct interaction
mice. Dev. Neurobiol. 67, 1511–1519. between arcuate nucleus neuropeptide Y (NPY) neurons and gonadotropin-
Kato, R., 1960. Serotonin content of rat brain in relation to sex and age. J. Neurochem. releasing hormone neurons and the possible involvement of NPY Y1 receptors.
5, 202. Endocrinology 140, 5382–5390.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
30 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

Lin, C.C., Tsai, Y.L., Huang, M.T., Lu, Y.P., Ho, C.T., Tseng, S.F., Teng, S.C., 2006. Morgan, C.P., Bale, T.L., 2011. Early prenatal stress epigenetically programs dys-
Inhibition of estradiol-induced mammary proliferation by dibenzoylmethane through masculinization in second-generation offspring via the paternal lineage.
the E2-ER-ERE-dependent pathway. Carcinogenesis 27, 131–136. J. Neurosci. 31, 11748–11755.
Lin, A., et al., 2015. Mapping the stability of human brain asymmetry across five sex- Morris, J.A., Jordan, C.L., Dugger, B.N., Breedlove, S.M., 2005. Partial demasculi-
chromosome aneuploidies. J. Neurosci. 35, 140–145. nization of several brain regions in adult male (XY) rats with a dysfunctional
Lindsten, T., Zong, W.X., Thompson, C.B., 2005. Defining the role of the Bcl-2 family androgen receptor gene. J. Comp. Neurol. 487, 217–226.
of proteins in the nervous system. Neuroscientist 11, 10–15. Mosley, M., Shah, C., Morse, K.A., Miloro, S.A., Holmes, M.M., Ahern, T.H.,
Lonstein, J.S., De Vries, G.J., 1999a. Comparison of the parental behavior of pair- Forger, N.G., 2016. Patterns of cell death in the perinatal mouse forebrain.
bonded female and male prairie voles (Microtus ochrogaster). Physiol. Behav. J. Comp. Neurol. May 19 [epub ahead of print].
66, 33–40. Murphy, S.J., Lusardi, T.A., Phillips, J.I., Saugstad, J.A., 2014. Sex differences in
Lonstein, J.S., De Vries, G.J., 1999b. Sex differences in the parental behaviour of adult microRNA expression during development in rat cortex. Neurochem. Int. 77, 24–32.
virgin prairie voles: independence from gonadal hormones and vasopressin. Murray, E.K., Hien, A., de Vries, G.J., Forger, N.G., 2009. Epigenetic control of sexual
J. Neuroendocrinol. 11, 441–449. differentiation of the bed nucleus of the stria terminalis. Endocrinology 150,
Lonstein, J.S., De Vries, G.J., 2000. Sex differences in the parental behavior of 4241–4247.
rodents. Neurosci. Biobehav. Rev. 24, ]669–686. Murray, E.K., Varnum, M.M., Fernandez, J.L., de Vries, G.J., Forger, N.G., 2011.
Lonstein, J.S., De Vries, G.J., 2001. Social influences on parental and nonparental Effects of neonatal treatment with valproic acid on vasopressin immunoreactivity
responses toward pups in virgin female prairie voles (Microtus ochrogaster). and olfactory behaviour in mice. J. Neuroendocrinol. 23, 906–914.
J. Comp. Psychol. 115, 53–61. Nestler, E.F., Barrot, M., DiLeone, R.J., Eisch, A.J., Gold, S.J., Monteggia, L.M., 2002.
Lonstein, J.S., Quadros, P.S., Wagner, C.K., 2001. Effects of neonatal RU486 on adult Neurobiology of depression. Neuron 34, 13–25.
parental, sexual and fearful behavior in rats. Behav. Neurosci. 115, 58–70. Nordeen, E.J., Nordeen, K.W., Sengelaub, D.R., Arnold, A.P., 1985. Androgens
Lu, H., Nishi, M., Matsuda, K., Kawata, M., 2004. Estrogen reduces the neurite growth prevent normally occurring cell death in a sexually dimorphic spinal nucleus.
of serotonergic cells expressing estrogen receptors. Neurosci. Res. 50, 23–28. Science 229, 671–673.
Lusardi, T.A., Murphy, S.J., Phillips, J.I., Chen, Y., Davis, C.M., Young, J.M., Norstedt, G., Palmiter, R., 1984. Secretory rhythm of growth hormone regulates sexual
Thompson, S.J., Saugstad, J.A., 2014. MicroRNA responses to focal cerebral differentiation of mouse liver. Cell 36, 805–812.
ischemia in male and female mouse brain. Front. Mol. Neurosci. 7, 11. Nottebohm, F., Arnold, A.P., 1976. Sexual dimorphism in vocal control areas of the
Lyon, M.F., 1999. X-chromosome inactivation. Curr. Biol. 9, R235–R237. songbird brain. Science 194, 211–213.
MacLusky, N.J., Philip, A., Hurlburt, C., Naftolin, F., 1985. Estrogen formation in the Nowacek, A.S., Sengelaub, D.R., 2006. Estrogenic support of motoneuron dendritic
developing rat brain: sex differences in aromatase activity during early post-natal growth via the neuromuscular periphery in a sexually dimorphic motor system.
life. Psychoneuroendocrinology 10, 355–361. J. Neurobiol. 66, 962–976.
MacLusky, N.J., Hajszan, T., Prange-Kiel, J., Leranth, C., 2006. Androgen modulation Nugent, B.M., Wright, C.L., Shetty, A.C., Hodes, G.E., Lenz, K.M., Mahurkar, A.,
of hippocampal synaptic plasticity. Neuroscience 138, 957–965. Russo, S.J., Devine, S.E., McCarthy, M.M., 2015. Brain feminization requires
Maney, D.L., 2016. Perils and pitfalls of reporting sex differences. Philos. Trans. R. active repression of masculinization via DNA methylation. Nat. Neurosci. 18,
Soc. Lond. B Biol. Sci. 371. Article ID 20150119. 690–697.
Matsuda, K.I., Mori, H., Nugent, B.M., Pfaff, D.W., McCarthy, M.M., Kawata, M., Nunez, J.L., McCarthy, M.M., 2008. Androgens predispose males to GABAA-mediated
2011. Histone deacetylation during brain development is essential for permanent excitotoxicity in the developing hippocampus. Exp. Neurol. 210, 699–708.
masculinization of sexual behavior. Endocrinology 152, 2760–2767. Nuñez, J.L., Jurgens, H.A., Juraska, J.M., 2000. Androgens reduce cell death in the
Matsumoto, A., Arai, Y., 1980. Sexual dimorphism in ‘wiring patter’ in the hypotha- developing rat visual cortex. Brain Res. Dev. Brain Res. 125, 83–88.
lamic arcuate nucleus and its modification by neonatal hormonal environment. Nuñez, J.L., Lauschke, D.M., Juraska, J.M., 2001. Cell death in the development of
Brain Res. 190, 238–242. the posterior cortex in male and female rats. J. Comp. Neurol. 436, 32–41.
Matsumoto, A., 2000. Sexual Differentiation of the Brain. CRC Press, Boca Raton, FL. Nuñez, J.L., Sodhi, J., Juraska, J.M., 2002. Ovarian hormones after postnatal day 20
McAbee, M.D., DonCarlos, L.L., 1998. Ontogeny of region-specific sex differences in reduce neuron number in the rat primary visual cortex. J. Neurobiol. 52, 312–321.
androgen receptor messenger ribonucleic acid expression in the rat forebrain. Nunez, J.L., Alt, J., McCarthy, M.M., 2003. A new model for prenatal brain damage: I.
Endocrinology 139, 1738–1745. GABAA receptor activation induces cell death in developing rat hippocampus. Exp.
McCarthy, M.M., Konkle, A.T., 2005. When is a sex difference not a sex difference? Neurol. 181, 258–269.
Front. Neuroendocrinol. 26, 85–102. Oppenheim, R.W., 1991. Cell death during development of the nervous system. Annu.
McCarthy, M.M., Schlenker, E.H., Pfaff, D.W., 1993. Enduring consequences of Rev. Neurosci. 14, 453–501.
neonatal treatment with antisense oligonucleotides to estrogen receptor messenger Orikasa, C., Sakuma, Y., 2010. Estrogen configures sexual dimorphism in the preoptic
ribonucleic acid on sexual differentiation of rat brain. Endocrinology 133, 433–439. area of C57BL/6J and ddN strains of mice. J. Comp. Neurol. 518, 3618–3629.
McCarthy, M.M., Auger, A.P., Bale, T.L., De Vries, G.J., Dunn, G.A., Forger, N.G., Orikasa, C., Kondo, Y., Hayashi, S., McEwen, B.S., Sakuma, Y., 2002. Sexually
Murray, E.K., Nugent, B.M., Schwarz, J.M., Wilson, M.E., 2009. The epigenetics of dimorphic expression of estrogen receptor beta in the anteroventral periventricular
sex differences in the brain. J. Neurosci. 29, 12815–12823. nucleus of the rat preoptic area: implication in luteinizing hormone surge. Proc.
McCarthy, M.M., Pickett, L.A., VanRyzin, J.W., Kight, K.E., 2015. Surprising origins of Natl. Acad. Sci. U.S.A. 99, 3306–3311.
sex differences in the brain. Horm. Behav. 76, 3–10. Ottem, E.N., Godwin, J.G., Krishnan, S., Petersen, S.L., 2004. Dual-phenotype GABA/
McCarthy, M.M., 2008. Estradiol and the developing brain. Physiol. Rev. 88, 91–124. glutamate neurons in adult preoptic area: sexual dimorphism and function.
McClellan, K.M., Calver, A.R., Tobet, S.A., 2008. GABA(B) receptors role in cell J. Neurosci. 24, 8097–8105.
migration and positioning within the ventromedial nucleus of the hypothalamus. Payne, A.P., Swanson, H.H., 1970. Agonistic behaviour between pairs of hamsters of
Neuroscience 151, 1119–1131. the same and opposite sex in a neutral observation area. Behaviour 36, 260–269.
Meaney, M.J., 1989. The sexual differentiation of social play. Psychiatr. Dev. 7, Peroulakis, M.E., Goldman, B., Forger, N.G., 2002. Perineal muscles and motoneurons
247–261. are sexually monomorphic in the naked mole-rat (Heterocephalus glaber).
Mong, J.A., McCarthy, M.M., 1999. Steroid-induced developmental plasticity in J. Neurobiol. 51, 33–42.
hypothalamic astrocytes: implications for synaptic patterning. J. Neurobiol. 40, Petersen, S.L., Barraclough, C.A., 1989. Suppression of spontaneous LH surges in
602–619. estrogen-treated ovariectomized rats by microimplants of antiestrogens into the
Mong, J.A., Kurzweil, R.L., Davis, A.M., Rocca, M.S., McCarthy, M.M., 1996. Evidence preoptic brain. Brain Res. 484, 279–289.
for sexual differentiation of glia in rat brain. Horm. Behav. 30, 553–562. Pfaff, D.W., Schwartz-Giblin, S., McCarthy, M.M., Kow, L.M., 1994. Cellular and
Mong, J.A., Glaser, E., McCarthy, M.M., 1999. Gonadal steroids promote glial molecular mechanisms of female reproductive behaviors. In: Knobil, E., Neill, J.D.
differentiation and alter neuronal morphology in the developing hypothalamus in (Eds.), Physiology of Reproduction. Raven Press, New York, pp. 107–220.
a regionally specific manner. J. Neurosci. 19, 1464–1472. Phoenix, C.H., Chambers, K.C., 1982. Sexual behavior in adult gonadectomized female
Mong, J.A., Nuñez, J.L., McCarthy, M.M., 2002. GABA mediates steroid-induced pseudohermaphrodite, female, and male rhesus macaques (Macaca mulatta)
astrocyte differentiation in the neonatal rat hypothalamus. J. Neuroendocrinol. treated with estradiol benzoate and testosterone propionate. J. Comp. Physiol.
14, 45–55. Psychol. 96, 823–833.
Moore, F.L., Richardson, C., Lowry, C.A., 2000. Sexual dimorphism in numbers of Phoenix, C.H., Goy, R.W., Gerall, A.A., Young, W.C., 1959. Organizing action of
vasotocin-immunoreactive neurons in brain areas associated with reproductive prenatally administered testosterone propionate on the tissues mediating mating
behaviors in the rough skin newt. Gen. Comp. Endocrinol. 117, 281–298. behavior in the female guinea pig. Endocrinology 65, 369–382.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning 31

Pike, C.J., 1999. Estrogen modulates neuronal Bcl-xL expression and beta-amyloid- Selvamani, A., Williams, M.H., Miranda, R.C., Sohrabji, F., 2014. Circulating miRNA
induced apoptosis: relevance to Alzheimer’s disease. J. Neurochem. 72, profiles provide a biomarker for severity of stroke outcomes associated with age
1552–1563. and sex in a rat model. Clin. Sci. (Lond.) 127, 77–89.
Pinos, H., Collado, P., Rodríguez-Zafra, M., Rodríguez, C., Segovia, S., Guillamón, A., Semaan, S.J., Murray, E.K., Poling, M.C., Dhamija, S., Forger, N.G., Kauffman, A.S.,
2001. The development of sex differences in the locus coeruleus of the rat. Brain 2010. BAX-dependent and BAX-independent regulation of Kiss1 neuron
Res. Bull. 56, 73–78. development in mice. Endocrinology 151, 5807–5817.
Planas, B., Kolb, P.E., Raskind, M.A., Miller, M.A., 1995a. Vasopressin and galanin Seney, M.L., Goldman, B.D., Forger, N.G., 2006. Breeding status affects motoneuron
mRNAs coexist in the nucleus of the horizontal diagonal band: a novel site of number and muscle size in naked mole-rats: recruitment of perineal motoneurons?
vasopressin gene expression. J. Comp. Neurol. 361, 48–56. J. Neurobiol. 66, 1354–1364.
Planas, B., Kolb, P.E., Raskind, M.A., Miller, M.A., 1995b. Sex difference in coex- Sengelaub, D.R., Forger, N.G., 2008. The spinal nucleus of the bulbocavernosus:
pression by galanin neurons accounts for sexual dimorphism of vasopressin in the firsts in androgen-dependent neural sex differences. Horm. Behav. 53,
bed nucleus of the stria terminalis. Endocrinology 136, 727–733. 596–612.
Pozzo-Miller, L.D., Aoki, A., 1991. Stereological analysis of the hypothalamic Shah, N.M., Pisapia, D.J., Maniatis, S., Mendelsohn, M.M., Nemes, A., Axel, R., 2004.
ventromedial nucleus II. Hormone induced changes in the synaptogenic pattern. Visualizing sexual dimorphism in the brain. Neuron 43, 313–319.
Brain Res. Dev. Brain Res. 61, 189–196. Shen, E.Y., Ahern, T.H., Cheung, I., Straubhaar, J., Dincer, A., Houston, I., de
Prange-Kiel, J., Wehrenberg, U., Jarry, H., Rune, G.M., 2003. Para/autocrine regu- Vries, G.J., Akbarian, S., Forger, N.G., 2015. Epigenetics and sex differences in
lation of estrogen receptors in hippocampal neurons. Hippocampus 13, 226–234. the brain: a genome-wide comparison of histone-3 lysine-4 trimethylation
Raznahan, A., Lue, Y., Probst, F., Greenstein, D., Giedd, J., Wang, C., Lerch, J., (H3K4me3) in male and female mice. Exp. Neurol. 268, 21–29.
Swerdloff, R., 2015. Triangulating the sexually dimorphic brain through high- Shulman, L.M., 2007. Gender differences in Parkinson’s disease. Gend. Med. 4, 8–18.
resolution neuroimaging of murine sex chromosome aneuploidies. Brain Struct. Simerly, R.B., Swanson, L.W., 1988. Projections of the medial preoptic nucleus:
Funct. 220, 3581–3593. a Phaseolus vulgaris leucoagglutinin anterograde tract-tracing study in the rat.
Ren, K., Dubner, R., 2007. Pain facilitation and activity-dependent plasticity in pain J. Comp. Neurol. 270, 209–242.
modulatory circuitry: role of BDNF-TrkB signaling and NMDA receptors. Mol. Simerly, R.B., Chang, C., Muramatsu, M., Swanson, L.W., 1990. Distribution of
Neurobiol. 35, 224–235. androgen and estrogen receptor mRNA-containing cells in the rat brain: an in situ
Rhoda, J., Corbier, P., Roffi, J., 1984. Gonadal steroid concentrations in serum and hybridization study. J. Comp. Neurol. 294, 76–95.
hypothalamus of the rat at birth: aromatization of testosterone to 17b-estradiol. Simerly, R.B., 1990a. Organization and regulation of sexually dimorphic neuroendo-
Endocrinology 114, 1754–1760. crine pathways. Behav. Brain Res. 92, 195–203.
Rhodes, M.E., Rubin, R.T., 1999. Functional sex differences (‘sexual diergism’) of Simerly, R.B., 1990b. Hormonal control of neuropeptide gene expression in sexually
central nervous system cholinergic systems, vasopressin, and hypothalamic- dimorphic olfactory pathways. Trends Neurosci. 13, 104–110.
pituitary- adrenal axis activity in mammals: a selective review. Brain Res. Brain Simerly, R.B., 1995. Anatomical substrates of hypothalamic integration. In:
Res. Rev. 30, 135–152. Paxinos, G. (Ed.), The Rat Nervous System. Academic Press, San Francisco,
Rinn, J.L., Snyder, M., 2005. Sexual dimorphism in mammalian gene expression. pp. 353–376.
Trends Genet. 21, 298–305. Simerly, R.B., 1998. Organization and regulation of sexually dimorphic neuroendocrine
Ronnekleiv, O.K., Kelly, M.J., 1986. Luteinizing hormone-releasing hormone neuronal pathways. Behav. Brain Res. 92, 195–203.
system during the estrous cycle of the female rat: effects of surgically induced Södersten, P., Bergh, C., Zandian, M., 2006. Understanding eating disorders. Horm.
persistent estrus. Neuroendocrinology 43, 564–576. Behav. 50, 572–578.
Roselli, C.E., Larkin, K., Resko, J.A., Stellflug, J.N., Stormshak, F., 2004. The volume Solomon, N.G., French, J.A., 1997. The study of mammalian cooperative breeding. In:
of a sexually dimorphic nucleus in the ovine medial preoptic area/anterior hypo- Solomon, N.G., French, J.A. (Eds.), Cooperative Breeding in Mammals. Cambridge
thalamus varies with sexual partner preference. Endocrinology 145, 478–483. University Press, Cambridge, pp. 1–10.
Roselli, C.E., 1991. Sex differences in androgen receptors and aromatase activity in Speert, D.B., Konkle, A.T., Zup, S.L., Schwarz, J.M., Shiroor, C., Taylor, M.E.,
microdissected regions of the rat brain. Endocrinology 128, 1310–1316. McCarthy, M.M., 2007. Focal adhesion kinase and paxillin: novel regulators of
Rosen, G.J., De Vries, G.J., Villalba, C., Weldele, M.L., Place, N.J., Coscia, E.M., brain sexual differentiation? Endocrinology 148, 3391–3401.
Glickman, S.E., Forger, N.G., 2006. The distribution of vasopressin in the forebrain Spelke, E.S., 2005. Sex differences in intrinsic aptitude for mathematics and science?:
of spotted hyenas. J. Comp. Neurol. 498, 80–92. a critical review. Am. Psychol. 60, 950–958.
Rosen, G.J., DeVries, G.J., Goldman, S.L., Goldman, B.D., Forger, N.G., 2007. Spiers, H., Hannon, E., Schalkwyk, L.C., Smith, R., Wong, C.C., O’Donovan, M.C.,
Distribution of vasopressin in the brain of the eusocial naked mole-rat. J. Comp. Bray, N.J., Mill, J., 2015. Methylomic trajectories across human fetal brain
Neurol. 500, 1093–1105. development. Genome Res. 25, 338–352.
Rune, G.M., Wehrenberg, U., Prange-Kiel, J., Zhou, L., Adelmann, G., Frotscher, M., Strickland, D., Bertoni, J.M., 2004. Parkinson’s prevalence estimated by a state
2002. Estrogen up-regulates estrogen receptor alpha and synaptophysin in slice registry. Mov. Disord. 19, 318–323.
cultures of rat hippocampus. Neuroscience 113, 167–175. Sumida, H., Nishizuka, M., Kano, Y., Arai, Y., 1993. Sex differences in the ante-
Rutter, M., Caspi, A., Moffitt, T.E., 2003. Using sex differences in psychopathology to roventral periventricular nucleus of the preoptic area and in the related effects of
study causal mechanisms: unifying issues and research strategies. J. Child. androgen in prenatal rats. Neurosci. Lett. 151, 41–44.
Psychol. Psychiatry 44, 1092–1115. van der Schoot, P., 1980. Effects of dihydrotestosterone and oestradiol on sexual
al-Shamma, H.A., De Vries, G.J., 1996. Neurogenesis of the sexually dimorphic differentiation in male rats. J. Endocrinol. 84, 397–407.
vasopressin cells of the bed nucleus of the stria terminalis and amygdala of rats. Swaab, D.F., Chung, W.C.J., Kruijver, F.P.M., Hofman, M.A., Hestiantoro, A., 2003.
J. Neurobiol. 29, 91–98. Sex differences in the hypothalamus in the different stages of human life. Neu-
Sanzgiri, R.P., Araque, A., Haydon, P.G., 1999. Prostaglandin E(2) stimulates gluta- robiol. Aging 24, S1–S16.
mate receptor-dependent astrocyte neuromodulation in cultured hippocampal cells. Swanson, L.W., 1986. Organization of mammalian neuroendocrine system. In:
J. Neurobiol. 41, 221–229. Björklund, A., Hökfelt, T., Swanson, L.W. (Eds.), Handbook of Physiology – The
Sato, F., Tsuchiya, S., Meltzer, S.J., Shimizu, K., 2011. MicroRNAs and epigenetics. Nervous System, vol. 4. Waverly Press, Baltimore, pp. 317–363.
FEBS J. 278, 1598–1609. Swartz, S.K., Soloff, M.S., 1974. The lack of estrogen binding by human alpha-
Schafer, D.P., et al., 2012. Microglia sculpt postnatal neural circuits in an activity and fetoprotein. J. Clin. Endocrinol. Metab. 39, 589–591.
complement-dependent manner. Neuron 74, 691–705. Szot, P., Dorsa, D.M., 1993. Differential timing and sexual dimorphism in the
Schanen, N.C., 1999. Molecular approaches to the Rett syndrome gene. J. Child. expression of the vasopressin gene in the developing rat brain. Brain. Res. Dev.
Neurol. 14, 806–814. Brain Res. 73, 177–183.
Schwarz, J.M., Liang, S.L., Thompson, S.M., McCarthy, M.M., 2008. Estradiol induces Tanapat, P., Hastings, N.B., Reeves, A.J., Gould, E., 1999. Estrogen stimulates
hypothalamic dendritic spines by enhancing glutamate release: a mechanism for a transient increase in the number of new neurons in the dentate gyrus of the adult
organizational sex differences. Neuron 58, 1–14. female rat. J. Neurosci. 19, 5792–5801.
Schwarz, J.M., Nugent, B.M., McCarthy, M.M., 2010. Developmental and hormone- Taziaux, M., Bakker, J., 2015. Absence of female-typical pheromone-induced hypo-
induced epigenetic changes to estrogen and progesterone receptor genes in thalamic neural responses and kisspeptin neuronal activity in a-fetoprotein
brain are dynamic across the life span. Endocrinology 151, 4871–4881. knockout female mice. Endocrinology 156, 2595–2607.
Schwendimann, R.N., Alekseeva, N., 2007. Gender issues in multiple sclerosis. Int. Tetel, M.J., 2009. Nuclear receptor coactivators: essential players for steroid hormone
Rev. Neurobiol. 79, 377–392. action in the brain and in behaviour. J. Neuroendocrinol. 21, 229–237.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


Author's personal copy
32 Sexual Differentiation of the Brain: A Fresh Look at Mode, Mechanisms, and Meaning

Thornton, J., Zehr, J.L., Loose, M.D., 2009. Effects of prenatal androgens on rhesus Wang, Z.X., Smith, W., Major, D.E., De Vries, G.J., 1994b. Sex and species differ-
monkeys: a model system to explore the organizational hypothesis in primates. ences in the effects of cohabitation on vasopressin messenger RNA expression in
Horm. Behav. 5, 633–645. the bed nucleus of the stria terminalis in prairie voles (Microtus ochrogaster) and
Tobet, S.A., Hanna, I.K., 1997. Ontogeny of sex differences in the mammalian meadow voles (Microtus pennsylvanicus). Brain Res. 650, 212–218.
hypothalamus and preoptic area. Cell Mol. Neurobiol. 17, 565–601. Weisz, J., Ward, I.L., 1980. Plasma testosterone and progesterone titers of pregnant
Tobet, S.A., Chickering, T.W., King, J.C., Stopa, E.G., Kim, K., Kuo-Leblank, V., rats, their male and female fetuses and neonatal offspring. Endocrinology 106,
Schwarting, G.A., 1996. Expression of gamma-aminobutyric acid and 306–313.
gonadotropin-releasing hormone during neuronal migration through the olfactory Wersinger, S.R., Sannen, K., Villalba, C., Lubahn, D.B., Rissman, E.F., De Vries, G.J.,
system. Endocrinology 137, 5415–5420. 1997. Masculine sexual behavior is disrupted in male and female mice lacking
Todd, B.J., Schwarz, J.M., McCarthy, M.M., 2005. Prostaglandin-E2: a point of a functional estrogen receptor alpha gene. Horm. Behav. 32, 176–183.
divergence in estradiol-mediated sexual differentiation. Horm. Behav. 48, White, F.A., Keller-Peck, C.R., Knudson, C.M., Korsmeyer, S.J., Snider, W.D., 1998.
512–521. Widespread elimination of naturally occurring neuronal death in Bax-deficient mice.
Todd, B.J., Schwarz, J.M., Mong, J.A., McCarthy, M.M., 2007. Glutamate AMPA/ J. Neurosci. 18, 1428–1439.
kainate receptors, not GABAA receptors, mediate estradiol-induced sex differences Wiegand, S.J., Terasawa, E., 1982. Discrete lesions reveal functional heterogeneity of
in the hypothalamus. Dev. Neurobiol. 67, 304–315. suprachiasmatic structures in regulation of gonadotropin secretion in the female
Topper, V.Y., Walker, D.M., Gore, A.C., 2015. Sexually dimorphic effects of gestational rat. Neuroendocrinology 34, 395–404.
endocrine-disrupting chemicals on microRNA expression in the developing rat Wolfe, C.A., Van Doren, M., Walker, H.J., Seney, M.L., McClellan, K.M., Tobet, S.A.,
hypothalamus. Mol. Cell. Endocrinol. 414, 42–52. 2005. Sex differences in the location of immunochemically defined cell populations
Toran-Allerand, C.D., Singh, M., Sétáló Jr., G., 1999. Novel mechanisms of estrogen in the mouse preoptic area/anterior hypothalamus. Dev. Brain Res. 157, 34–41.
action in the brain: new players in an old story. Front. Neuroendocrinol. 20, Woodley, S.K., Baum, M.J., 2004. Differential activation of glomeruli in the ferret’s
97–121. main olfactory bulb by anal scent gland odours from males and females: an early
Toran-Allerand, C.D., 2005. Estrogen and the brain: beyond ER-alpha, ER-beta, and step in mate identification. Eur. J. Neurosci. 20, 1025–1032.
17beta-estradiol. Ann. N.Y. Acad. Sci. 1052, 136–144. Wray, S., Gainer, H., 1987. Effect of neonatal gonadectomy on the postnatal devel-
Tsukahara, S., Tsuda, M.C., Kurihara, R., Kato, Y., Kuroda, Y., Nakata, M., Xiao, K., opment of LHRH cell subtypes in male and female rats. Neuroendocrinology 45,
Nagata, K., Toda, K., Ogawa, S., 2011. Effects of aromatase or estrogen receptor 13–419.
gene deletion on masculinization of the principal nucleus of the bed nucleus of the Wray, S., Hoffman, G., 1986. A developmental study of the quantitative distribution of
stria terminalis of mice. Neuroendocrinology 94, 137–147. LHRH neurons within the central nervous system of postnatal male and female rats.
Tsukahara, S., 2009. Sex differences and the roles of sex steroids in apoptosis of J. Comp. Neurol. 252, 522–531.
sexually dimorphic nuclei of the preoptic area in postnatal rats. J. Neuroendocrinol. Wright, C.L., Burks, S.R., McCarthy, M.M., 2008. Identification of prostaglandin E2
21, 370–376. receptors mediating perinatal masculinization of adult sex behavior and neuroan-
Tunç, B., Solmaz, B., Parker, D., Satterthwaite, T.D., Elliott, M.A., Calkins, M.E., atomical correlates. Dev. Neurobiol. 68.
Ruparel, K., Gur, R.E., Gur, R.C., Verma, R., 2016. Establishing a link between sex Yahr, P., 1988. Sexual differentiation of behavior in the context of developmental
differences in the structural connectome and behavior. Philos. Trans. R. Soc. Lond. psychobiology. In: Blass, E. (Ed.), Handbook of Behavioral Neurobiology, vol. 9.
B Biol. Sci. 37. Article ID 20150111. Plenum, New York, pp. 197–243.
Vale, J.R., Ray, D., Vale, C.A., 1973. The interaction of genotype and exogenous Zhang, J.M., Konkle, A.T., Zup, S.L., McCarthy, M.M., 2008. Impact of sex and
neonatal androgen and estrogen: sex behavior in female mice. Dev. Psychobiol. 6, hormones on new cells in the developing rat hippocampus: a novel source of sex
319–327. dimorphism? Eur. J. Neurosci. 27, 791–800.
Vega Matuszczyk, J., Fernandez-Guasti, A., Larsson, K., 1988. Sexual orientation, Zhou, S., Holmes, M.M., Forger, N.G., Goldman, B.D., Lovern, M.B., Caraty, A.,
proceptivity, and receptivity in the male rat as a function of neonatal hormonal Kalló, I., Faulkes, C.G., Coen, C.W., 2013. Socially regulated reproductive devel-
manipulation. Horm. Behav. 22, 362–378. opment: analysis ofGnRH-1 and kisspeptin neuronal systems in cooperatively
Villalba, C., Boyle, P.A., De Vries, G.J., 1997. Effects of the selective serotonin breeding naked mole-rats (Heterocephalus glaber). J. Comp. Neurol. 521,
reuptake inhibitor, fluoxetine, on social behaviors in male and female prairie voles 3003–3029.
(Microtus ochrogaster). Horm. Behav. 32, 184–191. Zuloaga, D.G., Poort, J.E., Jordan, C.L., Breedlove, S.M., 2011. Male rats with the
Wagner, C.K., Nakayama, A.Y., De Vries, G.J., 1998. Potential role of maternal testicular feminization mutation of the androgen receptor display elevated anxiety-
progesterone in the sexual differentiation of the brain. Endocrinology 139, related behavior and corticosterone response to mild stress. Horm. Behav. 60,
3658–3661. 380–388.
Wallen, K., 2005. Hormonal influences on sexually differentiated behavior in nonhuman Zup, S.L., Forger, N.G., 2002. Hormones and sexual differentiation. In:
primates. Front. Neuroendocrinol. 26, 7–26. Ramachandran, V.S. (Ed.), Encyclopedia of the Human Brain. Academic Press,
Wang, Z., Bullock, N.A., De Vries, G.J., 1993. Sexual differentiation of vasopressin pp. 323–341.
projections of the bed nucleus of the stria terminals and medial amygdaloid nucleus Zup, S.L., Carrier, H., Waters, E.M., Tabor, A., Bengston, L., Rosen, G.J.,
in rats. Endocrinology 132, 2299–2306. Simerly, R.B., Forger, N.G., 2003. Overexpression of bcl-2 reduces sex differences
Wang, Z.X., Ferris, C.F., De Vries, G.J., 1994a. The role of septal vasopressin in neuron number in the brain and spinal cord. J. Neurosci. 23, 2357–2362.
innervation in paternal behavior in prairie voles (Microtus ochrogaster). Proc. Natl.
Acad. Sci. U.S.A. 91, 400–404.

Hormones, Brain and Behavior, Third Edition, 2017, 3–32


View publication stats
Powered by TCPDF (www.tcpdf.org)

Potrebbero piacerti anche