Vous êtes sur la page 1sur 20

YFRNE 451 No.

of Pages 20, Model 5G


21 October 2010

Frontiers in Neuroendocrinology xxx (2010) xxx–xxx


1

Contents lists available at ScienceDirect

Frontiers in Neuroendocrinology
journal homepage: www.elsevier.com/locate/yfrne

2 Review

3 The genetics of sex differences in brain and behavior


4 Tuck C. Ngun 1, Negar Ghahramani 1, Francisco J. Sánchez, Sven Bocklandt, Eric Vilain ⇑
5 David Geffen School of Medicine at UCLA, Gonda Center, Room 5506, 695 Charles Young Drive South, Los Angeles, CA 90095-7088, United States

6
a r t i c l e i n f o a b s t r a c t
8
2 2
9 Article history: Biological differences between men and women contribute to many sex-specific illnesses and disorders. 23
10 Available online xxxx Historically, it was argued that such differences were largely, if not exclusively, due to gonadal hormone 24
secretions. However, emerging research has shown that some differences are mediated by mechanisms 25
11 Keywords: other than the action of these hormone secretions and in particular by products of genes located on the X 26
12 Sexual differentiation and Y chromosomes, which we refer to as direct genetic effects. This paper reviews the evidence for direct 27
13 Brain anatomy genetic effects in behavioral and brain sex differences. We highlight the ‘four core genotypes’ model and sex 28
14 Sex differences
differences in the midbrain dopaminergic system, specifically focusing on the role of Sry. We also discuss 29
15 Sexual orientation
16 Gender identity
novel research being done on unique populations including people attracted to the same sex and people 30
17 Sex chromosomes with a cross-gender identity. As science continues to advance our understanding of biological sex differ- 31
18 SRY ences, a new field is emerging that is aimed at better addressing the needs of both sexes: gender-based biol- 32
19 Dopamine ogy and medicine. Ultimately, the study of the biological basis for sex differences will improve healthcare for 33
20 Behavior both men and women. 34
21
! 2010 Elsevier Inc. All rights reserved. 35

36
37
38 1. Introduction non-gonadal cells and result in sex differences in the functions of 62
those cells. First, we will highlight some sex differences at the biolog- 63
39 Men and women are different in many ways. These differences ical level and at the psychological level. Then, we will review the ‘clas- 64
40 include both biological phenotypes [e.g. 191] and psychological sic’ view that dominated the field of sex differences—that most sex 65
41 traits [e.g. 200]. Some of these differences are influenced by envi- differences, especially those concerned with reproductive physiology 66
42 ronmental factors [1,340]. Yet, there are fundamental differences and behavior, were due to the action of hormones produced by the go- 67
43 between the sexes that are rooted in biology. nads. Next, we will present the emerging view that ‘direct genetic ef- 68
44 Of particular interest are sex differences that have been identified fects’ play an important role as well. Finally, we will discuss novel 69
45 in the brain. Although the brains of men and women are highly sim- approaches to studying sex differences by focusing on unique groups 70
46 ilar, they show consistent differences that have important implica- of individuals: people with sex-chromosome variations (e.g., Klinefel- 71
47 tions for each sex. That is, brain sex differences uniquely affect ter’s Syndrome and Turner Syndrome), people with genetic mutations 72
48 biochemical processes, may contribute to the susceptibility to spe- in the sexual development pathway, people with an atypical sexual- 73
49 cific diseases, and may influence specific behaviors. Such biological orientation, and people who experience a cross-gender identity. 74
50 differences should never be used to justify discrimination or sexism.
51 However, we believe that a thorough understanding of these differ- 2. Biological sex differences 75
52 ences can inform researchers and clinicians so that they can better
53 address important issues. Two examples include how genetic sex There are many biological differences between males and fe- 76
54 can lead to differences between the sexes in the etiology and the pro- males that are beyond the obvious differences at a gross, macro le- 77
55 gression of disease and how differences in neural development may vel (e.g., height, weight, and external genitalia). Specifically, there 78
56 result in differences in cognition and behavior. are several important physiological differences that have critical 79
57 In this paper, we will review sex differences in brain and behavior implications including the susceptibility to different diseases and 80
58 that are not due to the action of hormones secreted by the gonads— the ability to metabolize different medications. In this section we 81
59 which has been the dominant mechanism associated with such will highlight some sex differences in neuroanatomy and 82
60 differences—but from what we term ‘direct genetic effects.’ These neurochemistry. 83
61 are effects that arise from the expression of X and Y genes within

2.1. Neuroanatomy 84
Q1 ⇑ Corresponding author. Fax: +1 (310) 794 5446.
E-mail address: evilain@ucla.edu (E. Vilain). The two sexes have similar but not identical brains. Most brain 85
1
These authors contributed equally to this work. studies have focused on gross manifestations of these differences— 86

0091-3022/$ - see front matter ! 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.yfrne.2010.10.001

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

2 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

Table 1
Selected neuroanatomical sex differences in the rat.

Structure/region Known roles Sex difference Basis of difference


Sexually dimorphic The POA is implicated in the regulation of male copulatory 2.6 times larger in males [118] Perinatal aromatized androgen decreases
nucleus of the behavior [225]. Lesions of the SDN alone slow acquisition neuronal apoptotic rates in males [317]
preoptic area of this behavior. Potential human equivalent is INAH-3 [4]
(SDN-POA)
Anteroventral Involved in regulating the luteinizing hormone surge in 2.2 times larger in females with a Degeneration of cells in this region is greater
periventricular females [317] and male copulatory behavior [262] higher cell density [45] in males [308] due to prenatal action of
nucleus (AVPV) androgen
Bed nucleus of stria Plays a role in the control of male sexual behavior [100], The principal nucleus (BNSTp) is The larger volume in males is due to
terminalis release of gonadotropin [32], and modulation of stress larger in volume in males [85] sexually different apoptotic rates caused by
(BNST) [329,134] testosterone [109]
Corpus callosum Conducts information between the two halves of the Larger in neonatal males [351] Organizational effects of testosterone lead
cortex [304] to masculinization while feminization
appears to be dependent on estrogens
[106,105]
Arcuate nucleus Helps regulate the estrus cycle [203], appetite and body Neurokin-B neurons innervate Dihydrotestosterone is responsible for the
(ARC) weight [217] capillary vessels in the masculine projection pattern [67]
ventromedial ARC in post-pubertal
males only [66]
Amygdala Strongly associated with emotion, decision-making and Adult males have a larger medial Treatment of females with estradiol
Pavlovian conditioning [288] nucleus than adult females [221] masculinizes this nucleus [221]
The posterodorsal aspect of the Activational effects of circulating androgens
medial amygdala is 65% larger in accounts for the larger region in males [73]
males [148]
Cerebral cortex Connected to a wide range of processes from memory [20] Right posterior cortex is thicker Gonadal hormones play a role in
to language [33] to emotional processing [237] than left but only in males [90] maintaining the sex difference (ovariectomy
masculinizes the cortex of females) [90]
Ventromedial Involved in the control of lordosis, mounting, and Females have less synapses in the Organizational effects of aromatized
hypothalamic norepinephrine release [102]. High concentrations of ventrolateral VMN compared to testosterone appear to be crucial in
nucleus (VMN) steroid receptor mRNA have been observed in the males [211] establishing the masculine trait [253]
ventrolateral VMN [297]
Substantia nigra Made up almost entirely of dopaminergic neurons. Females have 20% fewer A genetic component has been
pars compacta Dopamine is involved in control of motor activity [123] dopaminergic neurons [86] demonstrated in mice [60]
*Note: This table highlights some prominent sex differences in the rat brain but it is by no means exhaustive. Conflicting evidence concerning the examples reported here
(particularly in the SDN-POA) exist, and the interpretation of the data is often more complicated than this summary implies.

87 namely the size of specific regions or nuclei. Yet, there is mounting nology continue to advance, we will eventually know how to make 119
88 evidence of sex differences at a finer level including differences in sense of the mounting evidence of sex differences in the brain. For 120
89 synaptic patterns [120,66] and neuronal density [117,211,338]. It now, it is reasonable to suspect that such differences may help ac- 121
90 is beyond the scope of this article to provide a comprehensive re- count for observed sex differences in behavior, neurological dis- 122
91 view of all known neuoranatomical differences. We have provided eases, and cognitive abilities. 123
92 notable sex differences in the rat brain in Table 1. There are also
93 excellent resources for those who are interested in delving deeper 2.2. Neurochemistry 124
94 into this topic [146,98,28].
95 We have chosen to focus on neuroanatomical differences in the Males and females exhibit different patterns of transmitting, 125
96 rat because the biological significance and origins of these differ- regulating, and processing biomolecules. Table 2 presents some 126
97 ences are much clearer than in humans. Neuroanatomical differ- of the neurochemical sex differences that have been identified. 127
98 ences in humans are also well-studied although ethical reasons As a specific example, we focus below on the monoaminergic 128
99 preclude the experimental manipulations that have led to the find- system, which has been implicated in several neurological 129
100 ings detailed in Table 1. This significantly limits the conclusions diseases and mental disorders that differentially affect men and 130
101 that can be drawn from any observations made in humans. women. 131
102 Although these neuroanatomical differences are intriguing, Monoamines are a class of small-molecule neurotransmitters 132
103 most are limited because the practical or functional significance that are involved in the control of a variety of processes including 133
104 of these findings are unknown. Discovering the significance of reproduction and sexual behavior [183,170], respiration [112], and 134
105 these differences is often difficult, even in rodents. de Vries and stress responses [163]. Monoamines have also been implicated in 135
106 Sodersten have eloquently outlined the challenges facing research- numerous mental disorders, including ones that differentially af- 136
107 ers who want to understand the link between sex differences in fect men and women [283,303]. Likewise, sex differences in the 137
108 structure and behavior [82]. A highly relevant case study high- monoaminergic systems in the rat are well-documented. Reisert 138
109 lighted in their review concerns the sexually dimorphic nucleus and Pilgrim provided a comprehensive review of arguments for 139
110 of the preoptic area (SDN-POA). The preoptic area (POA) has been the genetic bases of these differences [259]. 140
111 implicated in the regulation of male copulatory behavior [225], Monoamines are subdivided into two groups—catecholamines 141
112 but the link (if any) between the sex difference in SDN-POA size and indolamines—based on their molecular structure. The main 142
113 and behavior remains elusive. Masculinizing the size of the SDN- catecholamines are dopamine (DA), norepinephrine (NE) and epi- 143
114 POA in female rats does not result in a corresponding masculiniza- nephrine, which are synthesized from the amino acid tyrosine. 144
115 tion and defeminization of behavior [159]. Instead, the SDN-POA Fig. 1 highlights some of the known sex differences of the dopami- 145
116 may be related to inhibition of female sexual behaviors nergic system. Regulation of dopamine can potentially control the 146
117 [252,141], which might not have been an obvious hypothesis given levels of the other two catecholamines as they are derived from 147
118 what was known about the POA previously. As science and tech- dopamine. 148

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 3

Table 2
Selected neurochemical sex differences in the brain.

Neurochemical Known roles Species Selected sex differences


system/
pathway
Catecholamines Involved in the control of a variety of processes including Rat Male have higher norepinephrine (NE) levels in the amygdala and
(also see reproduction and sexual behavior [183,170], respiration [112], and hypothalamus at day 25. Direction of this sex difference is reversed
Fig. 1) stress responses [163] at day 300 [296]
In response to chronic physical stress, dopamine (DA) activity is
upregulated only in males whereas NE activity is increased only in
females [201]
Human Women appear to be more dependent than men on NE for long-
term emotional memory formation [323]
Serotonin Modulates a wide variety of processes including mood, aggression, Rat Sex differences in the serotonergic system are found at multiple
perception, reward, and attention [39] and levels [234,333,348,305,220]. See Fig. 2 for an illustration of some of
human these differences
Aromatase Plays a key role in sexual differentiation of the brain by converting Rat Aromatase activity is higher in males than females in many regions
testosterone to 17b-estradiol [231] including the anterior hypothalamus, BNST and POA [269]
Only males experience spikes in the expression of brain-specific and
total aromatase during embryonic development and shortly after
[71]
Vasopressin VP in the central nervous system (CNS) has been linked to learning, Rat The number of vasopressin-positive cells is two to three times
(VP) memory and motor behavior [263]. It has also been connected to higher in males than in females [81]
the control of social behaviors such as pair-bonding, parenting and Vasopressin-positive projections are also two to three times denser
aggression [151] in males [81]
Intrahypothalamic release of VP due to an increase of plasma
osmolality is higher in females [239]
Human Some studies have found that plasma VP concentrations are higher
in men than in women [263]
Cholinergic The cholinergic system helps regulate the sleep-wake cycle and Rat Levels of acetylcholine (ACh) are higher in females, regardless of
system modulates synaptic plasticity implicated in memory, learning, and estrous cycle, than in males [153]. The maximal level of Ach in
development [77,165]. Sex differences are found at many points in females was found at proestrus
the cholinergic system (reviewed in Rhodes and Rubin [263]) The binding affinity of muscarinic Ach receptors is lower in females
than in males [18]. Estrogens appear to modulate the binding
activity of these receptors [96]
Human Men are more sensitive to cholinergic stimulation than women
[273]
Opioid system Opioids are a class of chemical for which receptors are found Rat Generally, l and j class opioids seem more effective in males than
throughout the CNS [346,206]. Opioids exert an analgesic effect and and females although in some cases the effectiveness is equal [76]. In a
also play a role in stress response and reproduction [315] mouse minority of cases, they are more effective in females
Human l-opioids appear more effective in women than in men [76]
l-opioids show significantly higher binding potential in women in
the amygdala, thalamus and the cerebellum [352]. The sex
difference in the first two regions is reversed after menopause

149 Catecholamines are released by the adrenal glands usually in re- form better on specific visuospatial aspects (e.g., mental rotation) 173
150 sponse to stress, which affects males and females differently. For compared to women; and women perform better on specific verbal 174
151 instance, chronic physical stress impairs memory in male rats only tasks (e.g., verbal fluency) compared to men [155]. Furthermore, 175
152 [201]. The sexes also show differing neurochemical responses: there is a large sex difference in sexual interests and behaviors, 176
153 Dopamine activity is upregulated in males only whereas norepi- such as interest in casual sex, interest in multiple sex partners, 177
154 nephrine is upregulated in females only (Fig. 1A). Sex differences and interest in visual-sexual stimuli (e.g., pornography) 178
155 have also been found in the regulation and modification of dopa- [198,281]. Other examples are summarized in Table 3. 179
156 mine (see Fig. 1B and C). Specifically, the enzyme tyrosine hydrox- Some contend that these differences are due to social systems 180
157 ylase (TH), which is involved in dopamine synthesis [193], is and gender socialization [cf. 310,53,238]. Nevertheless, biological 181
158 regulated by Sry—the male sex determination gene—which is not traits likely contribute to many sex differences. Thus, a thorough 182
159 present in females. Additionally, levels of norepinehrine in the understanding of the main determinants involved in expression 183
160 amygdala differ between the sexes as a result of age. Thus, it is of such sex differences can help us better explain the relationship 184
161 likely that brain catecholaminergic responses to stress might also between brain, behavior, and environment. In addition, it allows us 185
162 differ between the sexes. to determine how one’s sex potentially influences the risk of devel- 186
163 Another monoamine is serotonin, which is an indolamine. Un- oping disorders that manifest and progress differently in men and 187
164 like catecholamines, serotonin is derived from the amino acid tryp- women. Such knowledge can better inform the treatment of these 188
165 tophan. The serotonergic system shows sex differences (Fig. 2), diseases. Tables 3 and 4 illustrate several factors (e.g. hormones 189
166 though many of these differences remain unlinked to behavioral and genes) that may be causally linked to expression of sex differ- 190
167 differences between men and women. Nevertheless, differences ences in behavior and disease, respectively. 191
168 in this system likely have consequences given the link between
169 serotonin and numerous mental disorders [48,275]. 4. The classical view on sex differences 192

170 3. Psychological and behavioral sex differences Researchers have examined what contributes to the differences 193
we see between males and females. Certainly for humans, social 194
171 In addition to biological differences, men and women differ in environments influence some of these differences. For instance, so- 195
172 many psychological and behavioral aspects. For instance, men per- cial stratifications (e.g., social class and the distribution of social 196

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

4 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

Fig. 1. The catecholaminergic pathway is sexually differentiated. TH: Tyrosine hydroxylase, L-DOPA: L-dihydroxyphenylalanine, NE: norepinephrine. (A) Chronic physical
stress results in sexually dimorphic responses. Dopamine (DA) activity is upregulated exclusively in males (light blue arrow) while norepinephrine activity is upregulated
exclusively in females (yellow arrow) [201]. Only males experience impaired memory. (B) Control of TH expression differs between the sexes. SRY, the testis determining
gene, which is not found in females, directly regulates TH expression in males [86,219]. 17b-estradiol increases TH expression only in males (light blue arrows) [257].
Aromatase activity is more responsive to dihydrotestosterone (DHT) in males than in females (dark blue arrow) [270]. (C) Male rats have higher NE levels than female ones in
the amygdala (A) and hypothalamus (HT) early in life [296]. When the rats reach day 300, the direction of this difference is reversed.

Fig. 2. Serotonin (5-HT) is sexually differentiated on multiple levels. In addition to the differences illustrated above, some of the loci that influence 5-HT levels in the blood
are also sexually dimorphic [333]. Refs.: 1 – [348], 2 – [305], 3 – [234], 4 – [220].

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
j.yfrne.2010.10.001
Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/

21 October 2010
YFRNE 451
Table 3
Sex differences in behavioral traits in humans.

Trait Sex bias Evidence for the role of hormones Evidence for the role of genetic factors Other factors affecting sex differences in
behavior
Cognition Men do better at spatial tasks [326] and Prenatal hormone effects shown from No reliable evidence for the effect of sex-chromosome genes proven Greater brain asymmetry in men for both
mathematical problem solving [34]. Women do studies of CAH, Turner’s and Androgen from studies of Turner’s and XX males [131] verbal and non-verbal tasks [137,179]

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx


better on verbal fluency, articulation, and verbal Insensitivity Syndromes [130]
memory tests [28]
Play behavior–movement There are sex differences in choice of Testosterone influences juvenile play [17] Genetics sex seems to affect play behavior more
toys, gender of the play partner, social than prenatal hormone exposure [17]
play [147] and movement [5,78,248]
Parents and other socializing agents (i.e. peers,
community, and child’s own cognitive processes)
[28]
Prenatal androgen levels affect play behavior and Developmental experience [129], visual information [316] affect
movement [235,38] movement organization
Language Women perform better on episodic memory Estrogen influences word and Single nucleotide polymorphisms in the gene, brain derived Greater degrees of left hemispheric
[144] and verbal fluency tasks, men are better at declarative memory abilities in women neurotrophic factor (BDNF) affecting BDNF secretion rates, partly lateralization of brain for language in males and
visuospatial processing [178,195,292] [291,246,205,298,240,294,228,341,280] accounting for greater dependence of females on declarative memory the bilateral language processing in females
and the sex differences observed in language-related tasks [95] [171]
Greater dependence of females on declarative Testosterone influences word memory Faster development of hippocampal brain
memory and males on procedural memory in men [101] regions in girls, activation of certain brain
[320,139] regions such as hippocampus and
parahippocampal gyrus [245,29]
Prenatal testosterone levels relate to
language processing in girls [122]
Aggression Foul language, imitation of aggressive models, Estradiol and progesterone influencing Association between serotonin transporter gene polymorphisms and Low self-control, high impulsivity and negative
violence and physical aggression more common the serotonergic system [336,274] greater impulsivity in males but not females [207] emotionality [187]
in males [40]
Weak association between testosterone Polymorphisms in monoamine oxidase-A (MAOA) gene associated Sex-specific disparities in the neural circuitry of
and aggression in both sexes [8,49] with antisocial personality disorder and aggression in males [62] impulse control and emotion regulation, as well
as serotonergic systems [233]
High testosterone levels leading to Larger orbitofrontal cortexes in women [125]
increased verbal aggression and
impulsivity in women [61,325]

No. of Pages 20, Model 5G


5
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

6 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

197 power) and social rules (e.g., customs and traditions) may affect The life sciences have elucidated many factors that contribute 202
198 the ability for people to access educational resources or to engage to sex differences. In this section, we briefly review the classical 203
199 in certain behaviors [156,339]. However, social factors alone do not view that gonadal hormones contribute to most, if not all, sex 204
200 contribute to all differences seen between males and females— differences after gonadal differentiation. We will then present 205
201 especially regarding biological differences [72]. some findings that have challenged this view. 206

Table 4
Sex differences in neurological disease.

Disease Sex bias Evidence for the role of hormones Evidence for the role of genetics Other factors affecting sex
differences in disease
Alzheimer’s Women demonstrate higher AD Gonadal hormones implicated in APOE allele type [172,75] (i.e. less Greater degeneration in areas of
disease prevalence at older ages [7,110] gender-related cognitive deficits of and slower rate of amyloid plaque orbitofrontal cortex, middle and
(AD) AD but the interaction is complex formation in men due to APOE e2 posterior cingulate cortex,
[30] [164] hypothalamus, and mammillary
bodies in men, and anterior
thalamic in women [58]
Parkinson’s Overrepresented in males [79,321] Most women manifest PD after Linkage to X chromosome markers Environmental factors [87]
disease menopause [31] in 362 families, and to Xq28 in 443
(PD) discordant sibling pairs [208,241]
Age at onset is later in women Estrogen affecting BDNF secretion Val66met polymorphism in BDNF Anatomical and structural
[318] [108] in women [152] differences in dopaminergic
systems among males and females
[28]
Pathological symptoms of PD differ Early life estrogen decline seems to
among males and females be more important [35,210,258]
[19,104,267]
Autism There is a high male to female ratio Gonadal hormones affecting Single nucleotide polymorphisms Alterations in oxytocin or arginine
in the prevalence of autism [222] oxytocin (OT) and arginine in the OT receptor in the Chinese vasopressin activity, and
vasopressin (AVP) receptors Han [343] and American Caucasian differential processing of the
[83,330] population [161], SNPs in the oxytocin precursor [184,181,140]
vasopressin receptor (V1aR) gene
[347,331]
X-chromosome has effects on
cognition and social aspects
[209,322]
Addiction Drug addiction more frequent in Estradiol levels correlate with drug Genes encoded on sex Neuroanatomical differences in
men [28] induced reinforcing behavior chromosomes can affect sex- motivation systems among males
whereas progesterone levels are related differences in addiction and females [28]
negatively associated with (the four core genotype mice)
addiction [334,169,168] [256]
Higher relapse rates, faster Sex-related alterations in the
progression of compulsive drug cortico-limbic-striatal system that
abuse and dependence have in mediates reward processing [266]
women [51,185]
Depression Women are twice as likely as men Low estrogen levels in female rats Heritability rates estimated to be Maladaptive coping, pessimism,
to develop depression during mediated by influences on 70% [173] dependency, low self-esteem,
reproductive years [177] neurotransmitter levels [175] victimization, sexual abuse,
comorbid anxiety disorder more
common in depressed women
[121]
Low testosterone levels associate Polymorphisms in serotonin gene, Early life events increase
with risk for depression in young estrogen receptor 1 (ESR1) depression rates in adult women
and middle aged-men [70,285] polymorphism in the presence of [176]
Val/Val genotype of the Val158Met
polymorphism in the Catechol-O-
methyl transferase (COMT) gene,
longer CA repeats of human
estrogen receptor 2 (ESR2), short
CAG repeats in androgen receptor
gene [63]
Anxiety The rate of anxiety disorders is States of anxiety and panic have The Val158 allele of COMT is Animal studies indicate females
disorders higher in females [283]. The high been reported to be affected by the associated with panic disorder in undergo less neurobiological
co morbidity of these disorders menstrual cycle and pregnancy, Caucasian women but not men changes in response to stress
with major depression helps implicating a role for estrogen and [136]. In Asians, Met158 is compared to males [6]. It is
account for the sex difference in progesterone [283] associated with panic disorder in speculated that this indicates
depression [52] women but not men [136] increased adaptability in males and
hence lower prevalence of affective
illness [6]
Pregnancy and lactation seem to 5HTTLPR is a polymorphism
alter brain neurochemical system associated with anxiety in humans.
that affect anxiety and fear [6] The orthologous polymorphism in
rhesus macaques interacts with
early adversity in a sexually
dimorphic manner [27]

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 7

Table 4 (continued)
Disease Sex bias Evidence for the role of hormones Evidence for the role of genetics Other factors affecting sex
differences in disease
Schizophrenia More common in men than in This disease is not common before Eight ultra-rare variants in eight Anatomical and structural brain
women [160] adolescence and puberty [88] distinct miRNA genes in 4% of differences among males and
analyzed males with schizophrenia females [115]
[103]
Age at onset is later in women, Male schizophrenics have higher Relatives of females with Higher cortical levels in males as
another smaller peak of onset levels of luteinizing hormone (LH) schizophrenia demonstrate higher compared to females according to
during peri- and post-menopause and testosterone than healthy levels of the psychotic forms some studies [115]
[160,254] subjects, and female whereas relatives of schizophrenic
schizophrenics higher levels of LH men express lower rates of
and lower levels of estrogen [186] psychosis suggesting the presence
of genetic heterogeneity [116]
Pathological symptoms of Higher rate of CAG repeat Higher sensitivity of the dopamine
schizophrenia differ among males expansions among families of system in men as compared to
and females (males experience female patients and not male women (normal males produce
more negative symptoms, greater patients [224] more striatal dopamine in
decrease in emotion expression response to an amphetamine
and recognition, greater paranoid challenge as compared to females)
delusions in women) [115] [227]
Lower chances of full recovery, and
a poorer prognosis in men
[160,254]
Anatomical brain differences
between male and female patients

207 4.1. The role of gonadal hormones from the expression of X and Y genes within non-gonadal cells that 247
result in sex differences in the functions of those cells or target 248
208 Sexual development in mammals can be divided into two cells. Such direct genetic actions are wide-ranging and can include 249
209 main components: sex determination and sex differentiation effects of locally produced hormones or other non-hormonal mes- 250
210 [324]. ‘Sex determination’ is the process by which the bipotential senger molecules. For example, sex differences arising in the brain 251
211 gonad develops into either a testis or an ovary, which depends from differential paracrine secretion of neurosteroids would be 252
212 exclusively on genetics. ‘Sex differentiation’ is the development considered a direct genetic effect. The commonality among these 253
213 of other internal reproductive structures, the external genitalia, actions is that they are not dependent on mediation by hormones 254
214 and non-gonadal sex differences. Unlike sex determination, sex secreted by the gonads. In many cases, the identity of the messen- 255
215 differentiation is driven by gonadal hormones. It was widely be- ger molecules have yet to be identified. This review will now focus 256
216 lieved that sex differences that emerged after sex determination on examples in which sex differences in behaviors are unlikely to 257
217 were largely due to the actions of gonadal hormones. Examples be influenced by only the action of gonadal hormonal secretions 258
218 of this pervasive view include writings from Lillie in 1939 and may in fact be due to direct genetic effects. 259
219 (‘‘[T]he mechanism of sex differentiation is taken over by extra-
220 cellular agents, the male and female hormones.” [197]), Jost in 4.2. Exceptions to the classical view 260
221 1970 (‘‘The developmental analysis of the body sex characteristics
222 reveals a hormonal control.” [167]), Morris et al. in 2004 (‘‘[A] The idea that factors other than the gonadal hormone milieu 261
223 single factor—the steroid hormone testosterone—accounts for could account for sex differences first gained credence from re- 262
224 most, and perhaps all, of the known sex differences. . .” [225]) search performed on the zebra finch. In zebra finches, males exhi- 263
225 and Zhao et al. in 2010 (‘‘[T]he sexual phenotype of individuals bit courtship behaviors that are unique to their sex. Specifically, 264
226 is dependent on the gonad. . .” [349]). We will use the term they possess the ability to sing a distinct courtship song. This 265
227 ‘classical view’ to refer to this hypothesis. male-specific ability has been attributed to several brain regions 266
228 The classical view was based on decades of compelling re- that are larger in males compared to females [10,236]. Given the 267
229 search demonstrating the organizational and activation effects hypothesis that such differences must have been the result of 268
230 of gonadal hormones in vertebrates [196,12]. ‘Organizational ef- sex-specific hormones, several researchers unsuccessfully at- 269
231 fects’ refer to the permanent, irreversible changes during develop- tempted to alter the courtship behavior of finches by manipulating 270
232 ment that organize the body in either a male- or female-typical hormone levels [244]. For example, it was shown that castrated 271
233 pattern. For instance, the neonatal surge of testosterone in male male zebra finches were not significantly different from intact 272
234 rodents leads to life-long changes in the synaptic pattern of the male zebra finches in terms of song development [9]. Furthermore, 273
235 ventrolateral ventromedial hypothalamic nucleus [253]. ‘Activa- female zebra finches that developed testes continued to develop 274
236 tional effects’ refer to the short-term changes that occur in the feminine song circuitry and did not exhibit masculine song behav- 275
237 body depending on the presence or absence of specific hormones. ior [328,327]. 276
238 An example of this is the requirement for the presence of both Several other experimental manipulations led researchers to 277
239 estrogen and progesterone to induce or ‘‘activate” lordosis in question the role of hormones. For instance, Jacobs et al. treated fe- 278
240 female rats [251]. male zebra finches with estrogen at the beginning of hatching gi- 279
241 Recently, it was found that gonadal hormones might not be the ven that estrogen induces male sexual differentiation in the 280
242 sole contributor to male- and female-typical development. Genes zebra finch neural song system [126]. Interestingly, estrogen treat- 281
243 encoded on the sex chromosomes that directly act on the brain ment was not able to cause full masculinization of the neural cir- 282
244 to influence neural developmental and sex-specific behaviors have cuitry of the zebra finch song system (the song circuitry was still 283
245 been identified—an example of what we describe as direct genetic smaller compared to control males) [162,299] and supraphysiolog- 284
246 effects [113,84]. When we use this term, we refer to effects arising ical doses of estrogen were required for full masculinization [13]. 285

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

8 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

286 Similarly, it was shown that inhibiting the action of estrogen by 5. An expanded view on sex differences 352
287 using aromatase blockers in males did not completely prevent
288 the male differentiation pathway [10,91,150,11,16]. In light of scientific findings, such as with the zebra finch and 353
289 The discovery of a rare type of zebra finch provided further sup- the tammar wallaby presented above, the field of sex differences 354
290 port for a new hypothesis regarding sexual differentiation: The has now come to encompass studies that examine gonadal hor- 355
291 bilateral gynandromorphic finch has male-typical phenotypes on mone as well as genetic origins of these differences. One of the 356
292 one half of the body (e.g., plumage, testis, and song circuitry) and most significant challenges in studying the establishment of sex 357
293 female-typical phenotypes on the other half of the body. Each half differences in animal models has been the difficulty in separating 358
294 of such finches is either entirely genetically male or genetically fe- gonadal sex from chromosomal sex. These two parameters almost 359
295 male. Thus, each side contains the sex-specific genes necessary for always correlate in an animal. 360
296 the development of the corresponding sex-specific traits. In this In the following section, we highlight the ‘four core genotypes’ 361
297 model, while the gonadal hormonal actions in producing sex differ- model, which has proven to be a powerful tool in teasing out the 362
298 ences in the brain cannot be completely ruled out (both sides of the effects of gonadal versus chromosomal sex and enabling research- 363
299 neural song system were larger than that of normal females), their ers to overcome this confound. We then discuss in-depth sexual 364
300 influences cannot fully explain the differences observed between differentiation and sex differences in the midbrain dopaminergic 365
301 the left and right sides of the brain. Given this explanation, the system, focusing specifically on the role of Sry. We discuss the 366
302 most reasonable theory is that endogenous genetic differences in implications that these differences may have on the development 367
303 the brain cells themselves can also contribute to the unequal dif- of this system as well as neurological health implications. 368
304 ferentiation of the two sides producing sex differences through
305 their local action within the brain [2]. 5.1. The ‘four core genotypes’ model 369
306 Recent work on gynandromorphic chickens strengthens the
307 case that the classical view largely does not apply to sexual dif- A 2 ! 2 mouse-model was developed to separate the effects of 370
308 ferentiation in birds. Zhao et al. showed that the ‘sex identity’ gonadal sex from chromosomal sex. This model, known as the ‘four 371
309 (or the expression of sex-specific phenotypes) of somatic cells core genotypes’ (FCG), allows researchers to establish the relative 372
310 in birds is determined by the sex chromosome complement of contribution of sex chromosomes and hormones in sexual differen- 373
311 those cells and not the gonadal hormonal environment [349,2]. tiation as well as the interaction between the two. Arnold and Chen 374
312 In mammals, transplantation of somatic cells from one sex into recently reviewed this model [14]. Here we highlight some of the 375
313 the gonad of the other sex reverses the sex identity of the donor model’s basic concepts. 376
314 somatic cells. For example, XX cells can develop into functioning Fig. 3 depicts the effect of the presence or absence of Sry—a 377
315 Sertoli cells while XY cells can become functioning granulosa cells 12 kb region on the Y chromosome that is responsible for testis 378
316 [242,56]. However, this is not the case in the chicken as male determination—using the FCG model. An XY mouse should develop 379
317 donor cells introduced into the developing ovary continued to ex- testes; however, if Sry is deleted from the Y chromosome (symbol- 380
318 press a male-specific marker and were excluded from ‘functional’ ized by Y–) then the mouse will develop ovaries [124]. If the Sry 381
319 structures of the host gonad. The host and donor somatic cells gene is inserted into any chromosome of an XX mouse (symbolized 382
320 were exposed to the same hormones, but they responded differ- by XXSry), then the mouse will develop testes. Finally, if Sry is de- 383
321 ently based on their respective sex chromosome complement. leted from the Y chromosome of an XY mouse and then inserted 384
322 A second exception to the classical view that we highlight be- into one of its autosomes (symbolized by XY–Sry), then it will devel- 385
323 low concerns the development of the tammar wallaby. As with op testes. 386
324 brain development, gonadal hormones drive the sex-specific devel- XY–Sry mice are fully fertile because the presence of Sry pro- 387
325 opment of the external genitalia in most mammals. Specifically, motes testes development. XXSry mice lack some of the genes 388
326 androgens promote the development of male genitalia. However, required for sperm production, which are found on the Y chromo- 389
327 the formation of reproductive structures in the tammar wallaby some [214], and therefore do not appear to be fertile. However, 390
328 appears to be independent of gonadal hormone control and is so-
329 lely due to the effect of sex chromosome complement.
330 The tammar wallaby is a marsupial that is much smaller than
331 the kangaroo. During fetal development, the production of testos-
332 terone, which would typically masculinize mammalian fetuses,
333 does not occur in these marsupials until about the fourth or fifth
334 day after birth [293,260,261,337]. Yet, signs of sex-specific repro-
335 ductive structures (e.g., scrotum, mammary gland, and pouch
336 formation) can be observed as early as several days before birth.
337 In mammals, the development of male-specific structures, is
338 thought to be completely dependent on the action of androgens
339 [324]. Experiments that increased or decreased the action of tes-
340 tosterone or estrogen in the tammar wallaby had no significant
341 effect on the development of the external genitalia [199,290]. This
342 suggested that such differences were not under gonadal hormone
343 control.
344 A case similar to the gynandromorphic zebra finch has also been
345 reported in tammar wallabies: This consists of wallabies that are
346 XX on one side of the body and XY on the other side of the body.
347 Such wallabies develop a hemipouch on the XX side and a hemi-
348 scrotum on the XY side even after exposure to circulating gonadal
349 hormones [13,289,74]. As with the zebra finch, such cases chal-
350 lenged the view that all sex differences were due to hormones pro- Fig. 3. 2 ! 2 comparison in the four core genotypes model. In this comparison, the
351 duced by the gonads. factors are gonadal sex and sex chromosome complement.

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 9

391 they have small testes and are fully masculinized in terms of mea- 5.1.2. Lateral septum 456
392 sures of male copulatory behavior, social exploration behavior, and One clear example of the role of sex-chromosome genes in brain 457
393 sexually dimorphic neuroanatomical structures in the septum, and phenotypes can be found in the lateral septum. The lateral septum 458
394 lumbar spinal cord. is part of the limbic system and is involved in stress-related behav- 459
395 XY–Sry mice can be mated with XX females to produce the four iors. This nucleus is denser in male brains compared to female 460
396 types of offspring (XX, XY–, XXSry, and XY–Sry) that can then be used brains. However, it was found that the vasopressin fiber density 461
397 to assess the impact of a mouse’s chromosomal and gonadal sex on was greater in the lateral septum of XY–Sry and XY– mice compared 462
398 different phenotypes. That is, if there is a difference between mice to XX and XXSry mice [113]. In addition, an examination of vaso- 463
399 that carry the Sry gene (i.e., XXSry and XY–Sry) versus those that do pressin fiber densities in animals with the same sex chromosome 464
400 not (i.e., XX and XY–), then the observed difference can be attrib- complement indicated a role for the action of gonadal steroid hor- 465
401 uted to the gonadal type and/or presence of Sry. On the other hand, mones. No interaction was observed between gonadal sex and sex 466
402 if there is a difference between mice that have the Y chromosome chromosomes [84]. 467
403 (i.e., XY– and XY–Sry) versus those that do not (i.e., XX and XXSry),
404 then the observed difference can be attributed to complement of 5.1.3. Addiction 468
405 sex chromosomes (XX versus XY). On average, women use addictive drugs at lower levels than 469
men, but women become addicted to drugs more rapidly than 470
406 5.1.1. Limitations of the FCG model men [138]. Based on the FCG model, Quinn et al. showed that this 471
407 Some possible limits to the FCG model have been suggested. difference could be attributed to the differences in the comple- 472
408 One possible limit is that the size, morphology, and function of ment of the sex chromosomes and not to the gonadal secretions 473
409 the gonads are not exactly the same in XX and XY mice of the same and/or the expression of the Sry gene. XX mice developed habit- 474
410 gonadal type (e.g. XXsry versus XY-sry). Consequently, the level of ual behavior more rapidly than the XY animals independent of 475
411 gonadal hormone secretions in FCG mice may differ during critical their gonadal phenotype and even after gonadectomy. This im- 476
412 periods of development—a confound that has yet to be investi- plies that neither gonadal sex nor circulating steroid hormones 477
413 gated. Yet, numerous phenotypes that are responsive to the orga- exert major effects on the development of habit-driven behavior 478
414 nizational effects of gonadal hormones (including sexually in mice [256]. 479
415 dimorphic brain structures) do not differ in XX and XY mice of
416 same gonadal type [288,259,18,33], indicating that XX and XY mice 5.1.4. Aggression 480
417 of the same sex are likely experiencing similar levels of gonadal Males typically exhibit more aggressive behaviors compared to 481
418 secretions. For example, measurements of circulating testosterone females [229,69,311]. Recent reports have shown that aggression 482
419 in XX and XY males found no difference in testosterone levels be- latencies are strongly influenced by the simultaneous action of go- 483
420 tween the groups [15]. nadal hormones and sex chromosomes. Using the four core geno- 484
421 A second limit relates to the biochemical and molecular envi- types model, it was found that a significant interaction exists 485
422 ronment. That is, one cannot rule out the effect of prenatal hor- between the two variables. In this model, the XX females appeared 486
423 monal secretions, the influence of adult circulating hormones to be slower at displaying aggressive behavior on their first 487
424 produced by the gonads or other tissues, acute fluctuations in hor- encounter with an intruder compared to animals in all other 488
425 monal levels, and the influence of the Sry transgene (i.e., the poten- groups [113]. 489
426 tially higher expression-level of the Sry transgene in XY–Sry animals
427 versus XY mice). For example, the Sry transgene could hasten the 5.2. Direct role of sry in brain sex differences 490
428 early stages of testis organogenesis in XY–Sry males. Furthermore,
429 several phenotypes have been found to differ between XY and Sex differences in the brain may contribute to some of the psy- 491
430 XY–Sry males. However, it is not known whether these differences chological and behavioral differences we observe between the 492
431 are caused by the effect of Sry on androgen production or by some sexes. Furthermore, they may influence the susceptibility to differ- 493
432 other mechanisms that are not mediated through the action of go- ent diseases. For instance, Parkinson’s disease—a neurodegenera- 494
433 nadal hormones. tive disease that impairs motor function and speech—affects 495
434 To address these limits, it is best to rule out the effect of circu- more men than women. Research has established a link between 496
435 lating gonadal hormones. An effective approach would be to first Parkinson’s disease and a loss of dopaminergic neurons in the sub- 497
436 gonadectomize the mice followed by an administration of equiva- stantia nigra [114]. Such losses disrupt dopamine pathways, which 498
437 lent doses of gonadal steroid hormones. This is particularly impor- leads to many of the symptoms associated with Parkinson’s 499
438 tant in the case of XY– females since their level of ovarian steroid disease. 500
439 hormones differ from that in the XX wild type females [15]. Never- Robust sex differences have been observed in the development, 501
440 theless, a major limitation still remains: It will not be obvious activity, and number of dopaminergic neurons. The data described 502
441 whether the sex difference attributed to the complement of sex below represents a clear example of a sex difference in the brain 503
442 chromosomes within cells is caused by (a) gene or genes encoded that has a strong genetic component. 504
443 on the Y chromosome; (b) higher dosage of X genes particularly the
444 ones that escape X inactivation in XX animals [59]; or (c) the pater- 5.2.1. Dopaminergic neurons in rodents 505
445 nal imprint of the genes encoded on the X chromosome in XX ani- Sex differences in dopaminergic neurons have been found prior 506
446 mals, which changes the expression of these genes to exhibit a to exposure to gonadal steroid hormones. During in utero develop- 507
447 female-specific pattern [244,345]. If one determines that the sex ment, rat embryos are exposed to a plasma surge of hormones 508
448 difference in phenotype is due to the sex chromosome comple- around embryonic day 17 or 18 (E17 or E18). Yet, as early as 509
449 ment, then the next step would be to discover the nature of the E14, dissociated cell cultures of dopaminergic neurons obtained 510
450 gene or genes involved and identify whether those genes are en- from male and female rat brainstems were found to be fundamen- 511
451 coded on the X or Y chromosome and how and where they mediate tally different in their morphology and function prior to exposure 512
452 their role [84]. to gonadal steroid hormones [259]. Furthermore, females had 513
453 Notwithstanding these potential limitations, a variety of sex dif- higher numbers of dopaminergic, tyrosine hydroxylase-immuno- 514
454 ferences have been examined using the FCG model. We review reactive (TH-ir) cells in the midbrain; and their mesencepahlic 515
455 three of these. and diencepahlic neurons produced more dopamine when 516

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

10 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

517 compared to males. On the other hand, soma measurements of 5.2.3. Sry is a direct effector of TH expression 582
518 diencephalic neurons from male cultures contained larger dopami- Sry is the gene on the Y chromosome that directs the bipotential 583
519 nergic neurons. Although it is difficult to make accurate measure- mammalian gonad to develop as testes—hence, its name: Sex 584
520 ments of hormonal levels in the embryonic brain, it is unlikely that determining region on Y. Sry is the founding member of the Sox 585
521 there is a huge sex difference due to gonadal hormone exposure at family of proteins, which play a major role in a wide range of bio- 586
522 this stage as the rat gonad only begins to differentiate at this point. logical processes such as neurogenesis, hematopoiesis, and neural 587
523 Therefore, this suggests a contribution of sex chromosome comple- crest development [189]. Sry contains a high mobility group 588
524 ment and/or sex-specific gene expression. (HMG) box domain and shows little conservation from mouse to 589
525 These differences are not altered even when gonadal hormone human outside this stretch of about 80 amino acids [335]. 590
526 levels are manipulated. Specifically, treatment with estradiol and The HMG box forms a domain that induces a sharp bend in the 591
527 testosterone does not eliminate the observed sex differences in DNA [332]. It is proposed that this bending of the DNA enhances 592
528 number, size, or function of the dopaminergic cells. Similar find- recruitment of specific transcriptional factors. In line with this 593
529 ings were later replicated in a study using mesencephalic cultures hypothesis, Sry has two nuclear localization signals within the 594
530 from the NMRI strain of mice [295]. Collectively, these observa- HMG domain [307] and its ability to activate transcription 595
531 tions strongly support the idea that some of the sex-specific prop- in vitro has been demonstrated [93]. 596
532 erties of the dopaminergic neurons appear to be under the control Recently, researchers have used genome-wide surveys to iden- 597
533 of non-hormonal mechanisms. tify targets of Sry in the gonad. The most widely known target of 598
Sry is Sox9 [286]. Two other notable targets of Sry are Cbln4 [50], 599
534 5.2.2. The Y chromosome’s role in dopaminergic neuron development which codes for the cerebellin precursor; and MAO A, which codes 600
535 A study utilizing the four core genotype model further strength- for monoamine oxidase A [344]. Wu and colleagues also found that 601
536 ened the case that a genetic component largely accounts for these MAO A was upregulated by Sry in the BE(2)C neuroblastoma cell 602
537 sex differences. Carruth et al. cultured mesencephalic neurons line suggesting that MAO A may be a neural Sry target [344]. 603
538 from E14 animals representing each of the groups from the four Most studies on Sry expression have focused on the gonad and 604
539 core genotypes [60]. Cultures from XY– and XY–Sry animals devel- Sry’s subsequent effects on sex determination and differentiation 605
540 oped significantly more TH-ir neurons compared to the XX and [312]. In the developing mouse embryo, Sry is expressed between 606
541 XXSry animals. However, gonadal sex did have a small effect: Ani- E 10.5 and E 12.5 in the developing genital ridge, prior to overt tes- 607
542 mals that had Sry (and hence testes) were associated with a higher tis differentiation [128]. Until recently, it was thought that Sry had 608
543 number TH-ir cells compared to those without Sry. Due to the de- no role other than sex determination. However, Sry expression has 609
544 sign of the four core genotypes model, it is difficult to separate the been found in numerous tissues outside of the testis (see below) 610
545 direct effects of Sry from its indirect effects on dopaminergic neu- and this expression in the adult male rat is now known to have bio- 611
546 rons or their precursors (e.g., through testis determination and the logically significant effects. Sry’s crucial role in the regulation of the 612
547 subsequent hormonal secretions). catecholaminergic system is one of the best examples of a direct 613
548 The data pertaining to sex differences in dopaminergic neuron genetic regulator of a trait that differs between the sexes. 614
549 development show sex differences in distinct directions and so
550 are difficult to interpret. In cultures from E14 rats and NMRI mice, 5.2.4. SRY in the brain 615
551 the sex difference is the reverse of what was seen with the four Clépet et al. were the first to perform a survey of SRY expression 616
552 core genotypes. However, rather than invalidating the findings, in human tissue outside of the gonads [68]. In fetal tissue, SRY was 617
553 the conflicting information highlights the complex interactions be- seen in the brain, adrenal, heart, and pancreas. In adults, transcrip- 618
554 tween genetics and gonadal hormones in leading to the sex differ- tion was detected in the kidney, heart, and liver. This study also 619
555 ences that are observed. First, the differences between data from showed that SRY was expressed in the teratocarcinoma cell line 620
556 NMRI mice and the four core genotypes may be attributable to NT2/D1, which was derived from adult male tissue and which 621
557 strain differences. Data from the former study indicate that genetic can be used as a model for dopaminergic neurons. When NT2/D1 622
558 background can significantly affect whether a sex difference is ob- was induced to differentiate into neurons by retinoic acid, SRY 623
559 served [295]. Carruth et al. had outbred their mice onto the MF1 expression remained. 624
560 background [60]. In regards to the differences seen between cul- SRY expression in the human adult brain was not surveyed until 625
561 tures from rats and the four core genotypes, one possible explana- 1998. Mayer et al. showed that SRY mRNA was present in the hypo- 626
562 tion is that both androgens and the Y chromosome are needed to thalamus, frontal, and temporal cortex of only the adult male 627
563 lead to the number of dopaminergic neurons being higher in males. [212]. Sry mRNA is also found in the adult male mouse brain where 628
564 Support for this hypothesis comes from our own studies where we it can be detected in the midbrain (including the substantia nigra) 629
565 see that the number of dopaminergic neurons in the rat substantia and hypothalamus in all developmental stages [213]. 630
566 nigra is higher in adult males [86]. Additionally, the finding that the
567 presence of testis was associated with a higher number of these 5.2.5. SRY and the regulation of TH expression 631
568 neurons fits with our hypothesis. There are also important distinc- Sry has a biologically significant role in the brain in at least one 632
569 tions in the timing of the cultures in relation to gonadal develop- instance—the regulation of tyrosine hydroxylase (TH) [86,219]. In a 633
570 ment: while both studies cultured E14 neurons, the bipotential 2004 study, Milsted et al. found that Sry is a regulator of TH gene 634
571 gonad has already differentiated into a testis to a larger extent in transcription [219]. The study looked at Sry’s role in relation to 635
572 the mice [97,182] than in the rat at this gestational stage [204]. TH in both the brain and the adrenal medulla. They demonstrated 636
573 As such, the hormonal environment from which these cultures that Sry and TH mRNA were co-localized in the locus coeruleus, 637
574 were derived may not be the same, which could account for some substantia nigra, and ventral tegmental area of the male rat 638
575 of the disparity in the direction of the sex difference. An elabora- (Fig. 4A). They then used a luciferase reporter assay to show that 639
576 tion of the hypothesis presented above is that it is not just the tes- Sry’s ability to upregulate TH expression is dependent on the AP1 640
577 tes and androgens that are essential but also Sry, the gene that binding sites in the promoter of TH. 641
578 initiates testicular development. Using a rat model, our laboratory The in vivo significance of those findings was shown and 642
579 has found evidence that this may be the case and showed that Sry expanded upon by a study from our laboratory. By in situ hybrid- 643
580 has a direct effect on the expression of TH in the substantia nigra ization, we were able to determine the spatial distribution of Sry 644
581 [86]. mRNA within the rodent brain [86]. Specific labeling of Sry was 645

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 11

observed in the substantia nigra, medial mammillary bodies of the 646


hypothalamus, and the cortex of male rats only. These transcripts 647
were translated and co-localized with the TH protein—all neurons 648
in the substania nigra positive for Sry were also positive for TH. 649
Knocking down Sry expression in the male rat substania nigra led 650
to 38% fewer TH-immunoreactive neurons and introduced a signif- 651
icant asymmetry in limb use where the animals strongly favored 652
the usage of their ipsilateral limbs (Fig. 4B and C). The reduction 653
in TH-ir neurons was not due to neural degeneration and is most 654
likely due to a reduction in TH expression. There was also a 26% de- 655
crease in TH-ir cells in the striatum when Sry expression was 656
knocked down in that region. TH-ir neuron number was not af- 657
fected in females infused with the Sry antisense cocktail. 658
The nature of Sry’s modulation on TH remains unclear. The re- 659
sults of the study by Milsted et al. argue that the TH response to 660
Sry is likely an indirect one [219]. Our data indicate that there 661
may be both direct and indirect mechanisms [86]. 662
The identification of a specific function for Sry in the dopami- 663
nergic system, specifically, and the brain, generally, is still absent. 664
Additionally, comprehensive temporal and spatial expression stud- 665
ies need to be performed on brain Sry expression. Another largely 666
unanswered question concerns the identity of a female-specific 667
‘compensatory’ factor for Sry. 668
We have shown that the attenuation of Sry expression in males 669
results in detrimental motor effects and that females have lower 670
levels of TH neurons [86]. However, female rats do not go through 671
life exhibiting motor dysfunction. The higher susceptibility of men 672
to Parkinson’s disease also implies that this factor exists and might 673
have protective effects against the nigrostriatal degeneration that 674
is the hallmark of Parkinson’s [114]. Estrogens are a viable candi- 675
date for this factor – short-term injections of estradiol benzoate 676
lead to an increase in TH mRNA [287] and ovariectomy results in 677
loss of TH-positive neurons [190]. 678

6. Novel approaches to studying sex differences 679

Traditional animal models have played an invaluable role in 680


advancing our understanding of sex differences. In particular, sci- 681
entists are able to conduct experimental manipulations that would 682
be unethical on human subjects. However, research on specific 683
groups of people has addressed some complex questions. In this 684
section, we focus on research conducted on four such groups: peo- 685
ple with sex-chromosome variations, people with genetic muta- 686
tions of the sexual development pathway, people attracted to the 687
same sex, and those with cross-sex gender identity. For readers 688
interested in learning more about disorders of sex development, 689
several comprehensive resources exist [324,107]. 690

6.1. Genetic disorders of the sex chromosomes 691

The most obvious genetic difference between females and 692


males is their sex chromosome complement (i.e., XX and XY). Var- 693
ious human sex chromosome disorders exist, which might be con- 694
sidered a human model for sex chromosome effects similar to the 695
four core genotypes. The most common variants in men involve 696
additional X or Y chromosomes: Klinefelter’s Syndrome (47,XXY); 697
and 47,XYY Syndrome. In women, the most common variants en- 698
tail the addition or absence of X chromosomes including 47,XXX; 699
Fig. 4. Sry regulates tyrosine hydroxylase (TH) levels and motor behavior. (A) Sry
and TH colocalize in the locus coeruleus (LC), ventral tegmental area (VTA) and 48,XXXX; and Turner Syndrome (45,X). 700
substantia nigra pars compacta (SNc) [86,219]. (B) Knockdown of Sry expression in Chromosomal abnormalities can highlight the role that sex 701
the SNc leads to a reduction in the number of TH-immunoreactive (TH-ir) neurons. chromosomes play in the phenotypic differences typically seen be- 702
Unilateral infusion of antisense oligodeoxynucleotides (ODN) against Sry decreased tween 46,XY men and 46,XX women. For instance, adolescent girls 703
the number of TH-if neurons by 38% compared to the contralateral side infused
with sense ODN [86]. (C) Unilateral downregulation of TH expression by Sry leads to
with Turner Syndrome are more likely to have social difficulties 704
asymmetric limb use. Animals preferentially used the forelimb ipsilateral to the compared to 46,XX girls [215], which may be partly related to fa- 705
side of the antisense ODN infusion (preferred limb highlighted in yellow) [86]. cial and emotional-processing impairments [188]. Furthermore, 706

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

12 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

707 46,XX girls score better than boys on tests of social cognitive skills identification, which may be mediated by numerous social and 769
708 [300]. The fact that both 46,XY boys and 45,X girls experience more psychological factors [65,143,194]. Nevertheless, most people 770
709 social adjustment problems compared to 46,XX girls suggests the report primarily opposite-sex or heterosexual attractions. Yet, a 771
710 presence of a genetic locus involved in social cognitive skills on significant number of people (approximately 2–6%) report pre- 772
711 the X chromosome. Data from Skuse et al. [301], suggest that this dominantly homosexual attractions [89]. 773
712 locus may be subject to imprinting. Significant differences between The distribution of sexual behavior differs between men and 774
713 45,XpO Turner-syndrome girls (in which the X was of paternal ori- women. In men, the distribution is largely bimodal [133]. That is, 775
714 gin) and 45,XmO girls (in which the X was maternally derived) in men are either attracted to one sex or the other. Although there 776
715 terms of social skills have been reported. 45,XpO had superior so- is disagreement regarding bisexuality among men [342], physio- 777
716 cial competence and better social skills than 45,XmO girls suggest- logical research has found that very few men (even those who 778
717 ing that the genes in this locus are expressed only from the openly identify as bisexual) show comparable physical attraction 779
718 paternal X. This could potentially be one of the reasons why boys to both men and women [265]. The distribution is more complex 780
719 are more susceptible to disorders such as autism that affect social in women, in which the fraction of women that show exclusive 781
720 adjustment and social skills such as language. In boys the X is only same-sex attraction is lower than men (1–3%), but many more wo- 782
721 of maternal origin and therefore this locus would be silenced. men than men report erotic fantasies towards both sexes [154]. 783
722 In other cases, however, the characteristics of individuals with In this section we will highlight some of the biological research 784
723 sex chromosome abnormalities may augment the expected sex dif- that has focused on same-sex attraction. A more thorough review 785
724 ference. On average, men in the general population have better vis- is available for interested readers [278]. 786
725 uospatial skills than women, and women have better verbal skills
726 than men—which suggests that increased dosage of X chromosome 6.3.1. Neuroanatomy differences in sexual orientation 787
727 genes may contribute to these skills. However, women with Turner Neuroanatomical differences have been reported for three brain 788
728 Syndrome have impaired visuospatial abilities yet greater language regions based on sexual orientation in human males: the arginine 789
729 skill compared to control women [272,314]. vasopressin neuronal population of the suprachiasmatic nucleus, 790
730 The role of the Y chromosome in psychosexual differentiation is which was larger in gay men than in male and female controls 791
731 still unclear. Work by McCarty et al. indicates that genes on the Y [309]; the third interstitial nucleus of the anterior hypothalamus 792
732 chromosome outside of SRY and the pseudoautosomal region have (INAH-3), which is smaller in gay men and more similar in size 793
733 ‘‘no obvious role. . .on psychosexual differentiation in genetic to female controls [192]; and the anterior commissure, which is 794
734 males [214].” This is difficult to ascertain because (a) the incidence larger in gay men than in control males and females [3]. The most 795
735 of XY gonadal dysgenesis is extremely rare, estimated to be 1 in discussed anatomical finding was in INAH-3 [192]. Although sub- 796
736 20,000 [64]; and (b) these individuals have been poorly studied sequent researchers reported inconsistent findings [57], a compa- 797
737 in regards to sexual differentiation of the brain. rable difference was found in sheep [271]. 798
Approximately 8–10% of the domestic ram population has been 799
738 6.2. Androgen Insensitivity Syndrome found to sexually prefer other males. Unlike other animal models 800
showing atypical sexual behavior, these male-oriented rams 801
739 The role of the androgen receptor (AR) in brain sexual differen- mount and ejaculate on other males versus simply exhibiting a 802
740 tiation has been discussed in patients with Androgen Insensitivity passive stance (i.e., lordosis). Consequently, they are an ideal ani- 803
741 Syndrome (AIS). AIS is an X-linked recessive disorder that is seen in mal model of male homosexuality because their coital behavior 804
742 1 out of 20,400 live male births [250]. There are two forms of AIS: is masculine but their sexual partner preference is feminine. 805
743 Complete (cAIS) and Partial (pAIS). People with Complete AIS are An analog of the sexually dimorphic nucleus (ovine SDN or 806
744 genetically male (46,XY with undescended testes) but phenotypi- oSDN)—a hypothalamic nucleus thought to be involved in mate 807
745 cally female. However, individuals with Partial AIS typically have selection—was identified in the sheep brain [271]. The oSDN was 808
746 ambiguous genitalia. found to be larger in female-oriented rams compared to male-ori- 809
747 AIS is caused by mutations in the androgen receptor (AR) gene ented rams (MORs), and ewes; the latter two groups had oSDN’s 810
748 [255]. There is an important difference in the behavioral phenotype comparable in size. It was hypothesized that the oSDN corresponds 811
749 between humans and rats with Complete AIS. Humans with Com- with human INAH-3, which suggests that the relevant neuroana- 812
750 plete AIS are female-typical in their play behavior and sexual ori- tomical pathways are conserved between mammalian species. 813
751 entation [353]. In contrast, XY rats with AR mutations behave
752 sexually like wild-type males and have a male-typical partner pref- 6.3.2. The role of prenatal androgens 814
753 erence [353]. The reasons behind this difference remain unclear One of the main hypotheses on the determinants of sexual ori- 815
754 although the implication is that androgens play an important role entation was that same-sex attraction was the result of atypical 816
755 in masculinizing the human brain. However, we cannot completely sex-hormone levels during gestation. Studies in rodents and ferrets 817
756 discount the role of estradiol as the expression of aromatase showed that pre- or perinatal hormonal manipulation could lead to 818
757 (which converts testosterone to estradiol) is dependent on andro- changes in partner preference, sexual behavior, and coital perfor- 819
758 gen signaling via AR [26]. mance largely controlled by the hypothalamus [92,306]. Yet, 820
extending this hypothesis from animal research to humans is diffi- 821
759 Q2 6.3. Sexual orientation cult in our opinion. Atypical sexual behavior in rodents is hard to 822
equate to human sexuality. For example, the induction of lordosis 823
760 Of all behavioral differences between males and females, part- in male rats does not change their partner preference. Instead, what 824
761 ner choice is one of the most pronounced. With very few excep- changes is the rat’s entire sexual behavior, which is different from 825
762 tions in the Animal Kingdom, males typically choose females to sexual orientation. Rather, an animal that consistently chooses 826
763 mate with, and females typically choose males to mate with. same-sex partners—such as the above-mentioned ram whose adult 827
764 Although sexual selection is a driving force of evolution, little is hormone levels are within the male-typical range [268]—would be 828
765 known about the molecular basis of partner preference. a better model. Furthermore, the treatment necessary to change 829
766 Human sexual orientation is a complex phenotype to study. Part the sexual behavior of rodents goes far beyond any naturally occur- 830
767 of this difficulty comes from the accurate assessment of sexual ori- ring variation in androgen levels [247], and as such is unlikely to 831
768 entation [218,223], especially when researchers depend on self- reflect natural causes of human variation in sexual orientation. 832

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 13

833 Additionally, hormonal manipulations have failed to make male The inactive chromosome remains inactive in all resulting daugh- 898
834 animals mount other males. ter cells [54]. If inactivation is completely random, this means that 899
835 Case studies on humans with various genetic defects in the in a population of female cells the maternal X will be inactivated in 900
836 androgen pathway show only limited support for the hypothesis. 50% of the cells whereas the paternal X is inactive in the remaining 901
837 There are no reports showing an increase in attraction to men in half. If a particular X chromosome, whether maternal or paternal, is 902
838 hypovirilized XY individuals relative to the general population. inactivated in more than 90% of cells, that individual is considered 903
839 This implies that disruption of the androgen pathway does not to be extremely skewed in regards to X-inactivation. Mothers with 904
840 have a strong effect on male sexual orientation. The role of andro- gay sons were found to have extreme skewing of X-inactivation 905
841 gens in female sexual orientation appears more complex. Women when compared to mothers with no gay sons: Skewing in mothers 906
842 with congenital adrenal hyperplasia (CAH) experience abnormal with one gay son = 13/97 or 13%; skewing in mothers with two or 907
843 activity of the embryonic adrenal glands. This leads to a much more gay sons = 10/44 or 23% [47]. This suggested an involvement 908
844 higher exposure of female fetuses to androgens, greatly exceeding of the X chromosome in the molecular mechanisms of sexual ori- 909
845 female-typical levels. The exposure is often high enough to cause entation. Arguably, the effect of the X-chromosome gene(s) or 910
846 some degree of genital masculinization. Several studies have found mechanisms that influence sexual orientation in the sons is visible 911
847 that CAH women reported more same-sex sexual activity and that in the blood of their mothers. 912
848 more self-identified as homosexual compared to the general popu- A genome-wide linkage scan on gay-brother pairs showed sug- 913
849 lation, which suggests that typical female sexual development is gestive linkage to loci on chromosome 7 and 8 [230]. A maternal 914
850 disrupted by extreme prenatal androgen exposure [149]. It is origin effect was found near marker D10S217, located at 10q26, 915
851 important to note that while women with CAH reported more gen- with significant linkage for maternal meioses but no paternal con- 916
852 der atypical attitudes, interests, and behavior, the majority still tribution. This result suggested the presence of a maternally-ex- 917
853 identified as heterosexual. The role of androgens in the sexual ori- pressed, paternally-silenced imprinted gene for sexual 918
854 entation of lesbian women who have no genital masculinization is orientation in 10q26. The relatively small sample size (N = 456) 919
855 still unclear. likely underpowered this study. However, a larger linkage scan 920
856 Two studies looked at genetic variation in genes related to the on 1000 homosexual male sibling pairs is currently underway 921
857 steroid pathway. A candidate gene study on the human androgen (A.R. Sanders, personal communication). 922
858 receptor gene [202] and one on the aromatase gene (CYP19) [94] The presence of a possible imprinted gene on chromosome 10 is 923
859 found no evidence that variations in these genes play a role in vari- particularly interesting. Previously reported evidence of maternal 924
860 ations in human sexual orientation. A variety of anthropomorphic loading of sexual orientation transmission was initially used to 925
861 measures have been used as indirect measures of prenatal andro- implicate the X-chromosome in human sexual orientation, but it 926
862 gen exposure, but results have been inconsistent. A recent prospec- could just as well indicate epigenetic factors acting on autosomal 927
863 tive study showed no correlation between maternal circulating genes. A role for imprinted genes in human sexual orientation 928
864 androgen concentration at 18 and 34 weeks of gestation and digit was hypothesized earlier [46]. 929
865 ratio in girls [145]. An in-depth discussion of these studies and a One of the most replicated findings in sexual orientation re- 930
866 speculation on their widely varying results falls outside of the search is known as the ‘fraternal birth order effect’: Each older 931
867 scope of this manuscript. For a review on the often cited 2D:4D fin- brother increases the odds of male homosexuality by approxi- 932
868 ger-length ratio in sexual orientation, see McFadden et al. [216]. Fi- mately 33% [41,166]. This is relative to the baseline frequency of 933
869 nally, studies retrospectively examining the influence of stressful homosexuality, and the odds of being homosexual are about twice 934
870 events during pregnancy have been inconclusive [23,99]. Therefore as high for the fourth-born son relative to the first-born son. Yet, 935
871 we believe that there is little evidence that naturally occurring this finding is not so simple: The effect is only influenced by older 936
872 variations of prenatal circulating gonadal hormones within one brothers born via the same mother, it is not influenced by the num- 937
873 sex play a role in determining variants of sexual orientation ber of older sisters, and it only seems to be true for right-handed 938
874 although diverging views have been expressed on this topic. homosexual men [44]. The dominant hypothesis for this effect, 939
which lacks empirical support, is that each successive male preg- 940
875 6.3.3. The genetics of sexual orientation nancies increases the mother’s immunity against male-specific 941
876 Evidence is mounting that there is a strong genetic component antigens expressed by the fetus [42,43], and this immune response 942
877 influencing sexual orientation. Family studies [249,21,243,24] have affects any subsequent male fetuses. Although this phenomenon 943
878 found an increased rate of homosexuality among siblings and in does not directly implicate genetics, at the very least it demon- 944
879 the maternal uncles of gay men (a median rate of 9% for brothers strates a biological basis for human male sexual orientation and 945
880 of gay men) [22]. Although the concordance rates of homosexuality suggests the immune system as an alternative, gonadal hormone- 946
881 in monozygotic twins vary depending on ascertainment methods independent mechanism through which sex differences can be 947
882 [21,25,174,180], twin studies have found that there is a substantial mediated. 948
883 genetic component in the development of sexual orientation. Altogether, there is mounting evidence for a genetic role of 949
884 There has been limited molecular genetics research. In 1993, human sexual orientation. The overwhelming dominance of heter- 950
885 Hamer et al. reported that male homosexuality was more often osexual behavior in the animal kingdom points at a tight molecular 951
886 on the mother’s side of the family versus the father’s side [133]. regulation of this trait. 952
887 A linkage scan showed significant linkage of male homosexuality
888 to the X-chromosome region Xq28 [133]. This finding was subse-
889 quently replicated by two studies [154,279] but not by an indepen- 6.4. Gender identity 953
890 dent group [264]. However, a meta-analysis of the results across all
891 four studies yielded an estimated level of Xq28 allele sharing be- Gender identity, or our sense of maleness or femaleness, plays 954
892 tween gay brothers of 64% instead of the expected 50% [132]. Nev- an important role throughout our development affecting both 955
893 ertheless, the exact gene(s) involved has (have) yet to be identified. our sense of self and our relationships [278,277,282]. Our gender 956
894 A different method also implicated the role of the X-chromo- identity and the roles ascribed to that gender are heavily 957
895 some. Unlike male cells, female cells contain two X-chromosomes. influenced by social factors [e.g. 226]. Most people adopt a gender 958
896 Consequently, each female cell randomly inactivates one X-chro- identity congruent with the sex assigned at birth, which remains 959
897 mosome during embryogenesis to create dosage compensation: constant throughout life [339]. The best approach to study the 960

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

14 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

961 biological basis of gender identity is to study individuals who de- proposed that observed molecular sex differences may compensate 1024
962 velop a cross-gender identity—in particular transsexuals. for other sex differences rather than generating differences be- 1025
963 Zhou et al. were the first to describe a sex difference in the cen- tween males and females themselves [80]. It will be crucial to tease 1026
964 tral subdivision bed nucleus of the stria terminalis (BSTc) in hu- out whether small imbalances influence specific differences in 1027
965 mans and a potential biological marker for gender identity [350]. neuropsychological function. Finally, it will be essential to trans- 1028
966 The type and direction of the sex difference mirrored that of the late our knowledge of sex differences to improve the quality of 1029
967 rat: the volume of the BSTc is larger in men than in women. The medical and psychological care. 1030
968 study also found that the BSTc of male-to-female (MtF) transsexu- As science continues to advance our understanding of sex dif- 1031
969 als is female-sized but the interpretation of this finding is compli- ferences, a new field is emerging focused on better addressing 1032
970 cated. The MtF subjects used in the study had all received estrogen the needs of men and women: gender-based biology and medicine. 1033
971 therapy so it remains unclear if the sex difference is related to gen- The ultimate aim of this field is to translate scientific data into 1034
972 der identity or hormonal exposure since estrogens can modify the practical applications that are effective for each sex [158]. From 1035
973 structure of the brain. A second confound is the relatively small tailoring preventive screenings to treating sex-specific illnesses, 1036
974 size of the sample pool as the authors were only able to gain access this field recognizes that ‘‘one-size-fits-all” healthcare has its lim- 1037
975 to tissue from six MtF transsexuals. its. Rather, our biological sex is an important variable that must be 1038
976 The literature on the genetic basis of transsexualism is extre- considered in our mental and physical health. 1039
977 mely limited. Although there are reports of families where several
978 members identify as transsexuals [119], such reports are rare. Acknowledgment 1040
979 There are few twin case studies, and they have reported differing
980 concordance rates for transsexualism [284,111,157,276]. Since no This work was funded in part by the National Institutes of 1041
981 systematic twin study has been reported, it is impossible to sepa- Health. 1042
982 rate genetic from environmental influences. Consequently, there is
983 no clear support for a genetic basis of transsexualism at this point. References 1043
984 A number of chromosomal abnormalities have been reported in
[1] M.E. Addis, A.K. Mansfield, M.R. Syzdek, Is ‘‘masculinity” a problem?: framing 1044
985 transsexuals [127,55,313,302]. In all cases, sex chromosomes were in-
the effects of gendered social learning in men, Psychol. Men Masc. 11 (2010) 1045
986 volved. The most common association was with disomy-Y (47,XYY). 77–90. 1046
987 However, because of the relatively high frequency of sex chromosome [2] R.J. Agate, W. Grisham, J. Wade, S. Mann, J. Wingfield, C. Schanen, A. Palotie, 1047
A.P. Arnold, Neural, not gonadal, origin of brain sex differences in a 1048
988 aneuploidy (1 in 900 males for XYY; [232] a statistically significant
gynandromorphic finch, Proc. Natl. Acad. Sci. USA 100 (2003) 4873–4878. 1049
989 association with transsexualism has not been shown. [3] L.S. Allen, R.A. Gorski, Sexual orientation and the size of the anterior 1050
990 A small number of candidate genes have been studied for trans- commissure in the human brain, Proc. Natl. Acad. Sci. USA 89 (1992) 7199– 1051
991 sexualism. A recent study looked at a polymorphism in the gene 7202. 1052
[4] L.S. Allen, M. Hines, J.E. Shryne, R.A. Gorski, Two sexually dimorphic cell 1053
992 coding for 5-alpha reductase and found no association in a sample groups in the human brain, J. Neurosci. 9 (1989) 497–506. 1054
993 of MtF and female-to-male (FtM) transsexuals [36]. The same [5] C.R. Almli, R.H. Ball, M.E. Wheeler, Human fetal and neonatal movement 1055
994 group found a significant association between a single nucleotide patterns: gender differences and fetal-to-neonatal continuity, Dev. 1056
Psychobiol. 38 (2001) 252–273. 1057
995 polymorphism in the CYP17 gene (which encodes the 17a-hydrox- 1058
[6] M. Altemus, Sex differences in depression and anxiety disorders: potential
996 ylase enzyme) in FtM transsexuals but not MtF transsexuals [37]. biological determinants, Horm. Behav. 50 (2006) 534–538. 1059
997 However, their sample size was small and they reported a signifi- [7] K. Andersen, L.J. Launer, M.E. Dewey, L. Letenneur, A. Ott, J.R. Copeland, J.F. 1060
Dartigues, P. Kragh-Sorensen, M. Baldereschi, C. Brayne, A. Lobo, J.M. 1061
998 cant difference in allele distribution between male and female con- 1062
Martinez-Lage, T. Stijnen, A. Hofman, Gender differences in the incidence of
999 trols as well, shedding doubts on these results. AD and vascular dementia: the EURODEM studies. EURODEM incidence 1063
1000 A small study of MtF transsexuals in a Swedish population stud- research group, Neurology 53 (1999) 1992–1997. 1064
[8] J. Archer, N. Graham-Kevan, M. Davies, Testosterone and aggression: a 1065
1001 ied repeat sequences in or near the androgen receptor gene, the
reanalysis of book, Starzyk and Quinsey’s (2001) study, Aggress. Violent 1066
1002 estrogen receptor beta gene, and the aromatase gene. They found Behav. 10 (2005) 241–261. 1067
1003 an association between MtF transsexualism and a dinucleotide [9] A.P. Arnold, The effects of castration on song development in zebra finches 1068
(Poephila guttata), J. Exp. Zool. 191 (1975) 261–278. 1069
1004 CA polymorphism in the estrogen receptor beta (ERb) gene [142].
[10] A.P. Arnold, Genetically triggered sexual differentiation of brain and behavior, 1070
1005 However, a larger study of MtF transsexuals failed to replicate Horm. Behav. 30 (1996) 495–505. 1071
1006 the ERb results and instead found a significant association with [11] A.P. Arnold, Sexual differentiation of the zebra finch song system: positive 1072
1007 the androgen receptor repeat [135]. This, too, was not replicated evidence, negative evidence, null hypotheses, and a paradigm shift, J. 1073
Neurobiol. 33 (1997) 572–584. 1074
1008 by a subsequent research team [319]. Overall, the significance of [12] A.P. Arnold, S.M. Breedlove, Organizational and activational effects of sex 1075
1009 these genetics studies remains unclear. steroids on brain and behavior: a reanalysis, Horm. Behav. 19 (1985) 469–498. 1076
[13] A.P. Arnold, P.S. Burgoyne, Are XX and XY brain cells intrinsically different?, 1077
Trends Endocrinol Metab. 15 (2004) 6–11. 1078
[14] A.P. Arnold, X. Chen, What does the ‘‘four core genotypes” mouse model tell 1079
1010 7. Conclusion us about sex differences in the brain and other tissues?, Front 1080
Neuroendocrinol. 30 (2009) 1–9. 1081
[15] A.P. Arnold, X. Chen, What does the ‘‘four core genotypes” mouse model tell 1082
1011 There are many differences between men and women. In this 1083
us about sex differences in the brain and other tissues?, Front
1012 review, we have focused on brain sex differences because of the Neuroendocrinol. 30 (2009) 1–9. 1084
1013 role that they play in people’s health and behavior. Historically, [16] A.P. Arnold, J. Wade, W. Grisham, E.C. Jacobs, A.T. Campagnoni, Sexual 1085
differentiation of the brain in songbirds, Dev. Neurosci. 18 (1996) 124–136. 1086
1014 it was believed that such differences were solely due to gonadal
[17] B. Auyeung, S. Baron-Cohen, E. Ashwin, R. Knickmeyer, K. Taylor, G. Hackett, 1087
1015 hormone secretions. Yet, emerging research is also implicating di- M. Hines, Fetal testosterone predicts sexually differentiated childhood 1088
1016 rect genetic effects. behavior in girls and in boys, Psychol. Sci. 20 (2009) 144–148. 1089
[18] S. Avissar, Y. Egozi, M. Sokolovsky, Studies on muscarinic receptors in mouse 1090
1017 The next challenge will be to first elucidate the molecular
and rat hypothalamus: a comparison of sex and cyclical differences, 1091
1018 mechanisms by which these direct genetic effects on sex differ- Neuroendocrinology 32 (1981) 295–302. 1092
1019 ences arise. One way to address this question will be to manipulate [19] Y. Baba, J.D. Putzke, N.R. Whaley, Z.K. Wszolek, R.J. Uitti, Gender and the 1093
1020 gene dosage in specific tissues and at specific times of develop- Parkinson’s disease phenotype, J. Neurol. 252 (2005) 1201–1205. 1094
[20] D. Badre, A.D. Wagner, Semantic retrieval, mnemonic control, and prefrontal 1095
1021 ment by using targeted approaches in genetically modified ani- cortex, Behav. Cogn. Neurosci. Rev. 1 (2002) 206–218. 1096
1022 mals. Another critical issue will be to understand how [21] J.M. Bailey, R.C. Pillard, A genetic study of male sexual orientation, Arch. Gen. 1097
1023 compensation mechanisms between the sexes operate. de Vries Psychiatry 48 (1991) 1089–1096. 1098

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 15

1099 [22] J.M. Bailey, R.C. Pillard, Genetics of human sexual orientation, Annu. Rev. Sex [55] N. Buhrich, R. Barr, P.R. Lam-Po-Tang, Two transsexuals with 47-XYY 1185
1100 Res. (1995) 126–150. karyotype, Br. J. Psychiatry 133 (1978) 77–81. 1186
1101 [23] J.M. Bailey, L. Willerman, C. Parks, A test of the maternal stress theory of [56] P.S. Burgoyne, M. Buehr, A. McLaren, XY follicle cells in ovaries of XX–XY 1187
1102 human male homosexuality, Arch. Sex. Behav. 20 (1991) 277–293. female mouse chimaeras, Development 104 (1988) 683–688. 1188
1103 [24] J.M. Bailey, R.C. Pillard, K. Dawood, M.B. Miller, L.A. Farrer, S. Trivedi, R.L. [57] W. Byne, S. Tobet, L. Mattiace, M.S. Lasco, E. Kemether, M.A. Edgar, S. 1189
1104 Murphy, A family history study of male sexual orientation using three Morgello, M.S. Buchsbaum, L.B. Jones, The interstitial nuclei of the human 1190
1105 independent samples, Behav. Genet. 29 (1999) 79–86. anterior hypothalamus: an investigation of variation within sex, sexual 1191
1106 [25] J.M. Bailey, M.P. Dunne, N.G. Martin, Genetic and environmental influences on orientation and HIV status, Horm. Behav. 40 (2001) 86–92. 1192
1107 sexual orientation and its correlates in an Australian twin sample, J. Pers. Soc. [58] D.J. Callen, S.E. Black, C.B. Caldwell, C.L. Grady, The influence of sex on limbic 1193
1108 Psychol. 78 (2000) 524–536. volume and perfusion in AD, Neurobiol. Aging 25 (2004) 761–770. 1194
1109 [26] J. Balthazart, G.F. Ball, New insights into the regulation and function of brain [59] L. Carrel, A.A. Cottle, K.C. Goglin, H.F. Willard, A first-generation X- 1195
1110 estrogen synthase (aromatase), Trends Neurosci. 21 (1998) 243–249. inactivation profile of the human X chromosome, Proc. Natl. Acad. Sci. USA 1196
1111 [27] C.S. Barr, T.K. Newman, M. Schwandt, C. Shannon, R.L. Dvoskin, S.G. Lindell, J. 96 (1999) 14440–14444. 1197
1112 Taubman, B. Thompson, M. Champoux, K.P. Lesch, D. Goldman, S.J. Suomi, J.D. [60] L.L. Carruth, I. Reisert, A.P. Arnold, Sex chromosome genes directly affect brain 1198
1113 Higley, Sexual dichotomy of an interaction between early adversity and the sexual differentiation, Nat. Neurosci. 5 (2002) 933–934. 1199
1114 serotonin transporter gene promoter variant in rhesus macaques, Proc. Natl. [61] E. Cashdan, Hormones and competitive aggression in women, Aggress. Behav. 1200
1115 Acad. Sci. USA 101 (2004) 12358–12363. 29 (2003) 107–115. 1201
1116 Q3 [28] J.B. Becker, Sex Differences in the Brain: From Genes to Behavior, Oxford [62] A. Caspi, J. McClay, T.E. Moffitt, J. Mill, J. Martin, I.W. Craig, A. Taylor, R. 1202
1117 University Press, Oxford; New York, 2008. Poulton, Role of genotype in the cycle of violence in maltreated children, 1203
1118 [29] J.B. Becker, Sex Differences in the Brain: From Genes to Behavior, Oxford Science 297 (2002) 851–854. 1204
1119 University Press, Oxford; New York, 2008. p. 427. [63] A. Caspi, K. Sugden, T.E. Moffitt, A. Taylor, I.W. Craig, H. Harrington, J. McClay, 1205
1120 [30] J.B. Becker, Sex Differences in the Brain: From Genes to Behavior, Oxford J. Mill, J. Martin, A. Braithwaite, R. Poulton, Influence of life stress on 1206
1121 University Press, Oxford; New York, 2008. pp. 442–445. depression: moderation by a polymorphism in the 5-HTT gene, Science 301 1207
1122 [31] J.B. Becker, Sex Differences in the Brain: From Genes to Behavior, Oxford (2003) 386–389. 1208
1123 University Press, Oxford; New York, 2008. p. 458. [64] D. Chadwick, J. Goode, The Genetics and Biology of Sex Determination, J. 1209
1124 [32] C. Beltramino, S. Taleisnik, Dual action of electrochemical stimulation of the Wiley, Chichester, West Sussex, England, New York, NY, 2002. 1210
1125 bed nucleus of the stria terminalis on the release of LH, Neuroendocrinology [65] P.K.Y. Chow, S.T. Cheng, Shame, internalized heterosexism, lesbian identity, 1211
1126 30 (1980) 238–242. and coming out to others: a comparative study of lesbians in mainland China 1212
1127 [33] D. Ben Shalom, D. Poeppel, Functional anatomic models of language: and Hong Kong, J. Couns. Psychol. 57 (2010) 92–104. 1213
1128 assembling the pieces, Neuroscientist 14 (2008) 119–127. [66] P. Ciofi, D. Leroy, G. Tramu, Sexual dimorphism in the organization of the rat 1214
1129 [34] C.P. Benbow, D. Lubinski, D.L. Shea, H. Eftekhari-Sanjani, Sex differences in hypothalamic infundibular area, Neuroscience 141 (2006) 1731–1745. 1215
1130 mathematical reasoning ability at age 13: their status 20 years later, Psychol. [67] P. Ciofi, O.C. Lapirot, G. Tramu, An androgen-dependent sexual dimorphism 1216
1131 Sci. 11 (2000) 474–480. visible at puberty in the rat hypothalamus, Neuroscience 146 (2007) 630– 1217
1132 [35] M.D. Benedetti, D.M. Maraganore, J.H. Bower, S.K. McDonnell, B.J. Peterson, 642. 1218
1133 J.E. Ahlskog, D.J. Schaid, W.A. Rocca, Hysterectomy, menopause, and estrogen [68] C. Clepet, A.J. Schafer, A.H. Sinclair, M.S. Palmer, R. Lovell-Badge, P.N. 1219
1134 use preceding Parkinson’s disease: an exploratory case-control study, Mov. Goodfellow, The human SRY transcript, Hum. Mol. Genet. 2 (1993) 2007– 1220
1135 Disord. 16 (2001) 830–837. 2012. 1221
1136 [36] E.K. Bentz, C. Schneeberger, L.A. Hefler, M. van Trotsenburg, U. Kaufmann, J.C. [69] A.M. Cohn, A. Zeicher, A.L. Seibert, Labile affect as a risk factor for aggressive 1222
1137 Huber, C.B. Tempfer, A common polymorphism of the SRD5A2 gene and behavior in men, Psychol. Men Masc. 9 (2008) 29–39. 1223
1138 transsexualism, Reprod. Sci. 14 (2007) 705–709. [70] L.A. Colangelo, L. Sharp, P. Kopp, D. Scholtens, B.C. Chiu, K. Liu, S.M. Gapstur, 1224
1139 [37] E.K. Bentz, L.A. Hefler, U. Kaufmann, J.C. Huber, A. Kolbus, C.B. Tempfer, A Total testosterone, androgen receptor polymorphism, and depressive 1225
1140 polymorphism of the CYP17 gene related to sex steroid metabolism is symptoms in young black and white men: the CARDIA male hormone 1226
1141 associated with female-to-male but not male-to-female transsexualism, study, Psychoneuroendocrinology 32 (2007) 951–958. 1227
1142 Fertil. Steril. 90 (2008) 56–59. [71] A. Colciago, F. Celotti, A. Pravettoni, O. Mornati, L. Martini, P. Negri-Cesi, 1228
1143 [38] S.A. Berenbaum, S.C. Duck, K. Bryk, Behavioral effects of prenatal versus Dimorphic expression of testosterone metabolizing enzymes in the 1229
1144 postnatal androgen excess in children with 21-hydroxylase-deficient hypothalamic area of developing rats, Dev. Brain Res. 155 (2005) 107–116. 1230
1145 congenital adrenal hyperplasia, J. Clin. Endocrinol. Metab. 85 (2000) 727–733. [72] J.C. Confers, J.A. Easton, D.S. Fleischman, C.D. Goetz, D.M. Lewis, C. Perilloux, 1231
1146 [39] M. Berger, J.A. Gray, B.L. Roth, The expanded biology of serotonin, Annu. Rev. D.M. Buss, Evolutionary psychology: controversies, questions, prospects, and 1232
1147 Med. 60 (2009) 355–366. limitations, Am. Psychol. 65 (2010) 110–126. 1233
1148 [40] B.A. Bettencourt, N. Miller, Gender differences in aggression as a function of [73] B.M. Cooke, G. Tabibnia, S.M. Breedlove, A brain sexual dimorphism 1234
1149 provocation: a meta-analysis, Psychol. Bull. 119 (1996) 422–447. controlled by adult circulating androgens, Proc. Natl. Acad. Sci. USA 96 1235
1150 [41] R. Blanchard, Birth order and sibling sex ration in homosexual versus (1999) 7538–7540. 1236
1151 heterosexual males and females, Annu. Rev. Sex Res. 8 (1997) 27–67. [74] D.W. Cooper, The evolution of sex determination, sex chromosome 1237
1152 [42] R. Blanchard, A.F. Bogaert, Homosexuality in men and number of older dimorphism, and X-inactivation in Therian mammals: a comparison of 1238
1153 brothers, Am. J. Psychiatry 153 (1996) 27–31. metatherians (marsupials) and eutherians (‘‘placentals”), in: K. Reed, J.A.M. 1239
1154 [43] R. Blanchard, P. Klassen, HY antigen and homosexuality in men, J. Theor. Biol. Graves (Eds.), Sex Chromosomes and Sex-Determining Genes, Harwood 1240
1155 185 (1997) 373–378. Academic Publishers, Chur, Switzerland, Langhorne, PA, 1993, pp. 183–200. 1241
1156 [44] R. Blanchard, J.M. Cantor, A.F. Bogaert, S.M. Breedlove, L. Ellis, Interaction of [75] E.H. Corder, A.M. Saunders, W.J. Strittmatter, D.E. Schmechel, P.C. Gaskell, 1242
1157 fraternal birth order and handedness in the development of male G.W. Small, A.D. Roses, J.L. Haines, M.A. Pericak-Vance, Gene dose of 1243
1158 homosexuality, Horm. Behav. 49 (2006) 405–414. apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late 1244
1159 [45] R. Bleier, W. Byne, I. Siggelkow, Cytoarchitectonic sexual dimorphisms of the onset families, Science 261 (1993) 921–923. 1245
1160 medial preoptic and anterior hypothalamic areas in guinea pig, rat, hamster, [76] R.M. Craft, Sex differences in opioid analgesia: ‘‘from mouse to man”, Clin. J. 1246
1161 and mouse, J. Comp. Neurol. 212 (1982) 118–130. Pain 19 (2003) 175–186. 1247
1162 [46] S. Bocklandt, D.H. Hamer, Beyond hormones: a novel hypothesis for the [77] J.A. Dani, D. Bertrand, Nicotinic acetylcholine receptors and nicotinic 1248
1163 biological basis of male sexual orientation, J. Endocrinol. Invest. 26 (2003) 8– cholinergic mechanisms of the central nervous system, Annu. Rev. 1249
1164 12. Pharmacol. Toxicol. 47 (2007) 699–729. 1250
1165 [47] S. Bocklandt, S. Horvath, E. Vilain, D.H. Hamer, Extreme skewing of X [78] P.L. Davies, J.D. Rose, Motor skills of typically developing adolescents: 1251
1166 chromosome inactivation in mothers of homosexual men, Hum. Genet. 118 awkwardness or improvement?, Phys Occup. Ther. Pediatr. 20 (2000) 19–42. 1252
1167 (2006) 691–694. [79] L.M. de Lau, P.C. Giesbergen, M.C. de Rijk, A. Hofman, P.J. Koudstaal, M.M. 1253
1168 [48] L. Booij, A.J. Van der Does, W.J. Riedel, Monoamine depletion in psychiatric Breteler, Incidence of Parkinsonism and Parkinson disease in a general 1254
1169 and healthy populations: review, Mol. Psychiatry 8 (2003) 951–973. population: the Rotterdam study, Neurology 63 (2004) 1240–1244. Q5 1255
1170 [49] A.S. Book, K.B. Starzyk, V.L. Quinsey, The relationship between testosterone [80] G.J. De Vries, Minireview: sex differences in adult and developing brains: 1256
1171 and aggression: a meta-analysis, Aggress. Violent Behav. 6 (2001) 579–599. compensation, compensation, compensation, Endocrinology 145 (2004) 1257
1172 [50] S.T. Bradford, R. Hiramatsu, M.P. Maddugoda, P. Bernard, M.C. Chaboissier, A. 1063–1068. 1258
1173 Sinclair, A. Schedl, V. Harley, Y. Kanai, P. Koopman, D. Wilhelm, The cerebellin [81] G.J. De Vries, P.A. Boyle, Double duty for sex differences in the brain, Behav. 1259
1174 Q4 4 precursor gene is a direct target of SRY and SOX9 in mice, Biol. Reprod. Brain Res. 92 (1998) 205–213. 1260
1175 (2009). [82] G.J. de Vries, P. Sodersten, Sex differences in the brain: the relation between 1261
1176 [51] K.T. Brady, C.L. Randall, Gender differences in substance use disorders, structure and function, Horm. Behav. 55 (2009) 589–596. 1262
1177 Psychiatr. Clin. North Am. 22 (1999) 241–252. [83] G.J. De Vries, Z. Wang, N.A. Bullock, S. Numan, Sex differences in the 1263
1178 [52] N. Breslau, L. Schultz, E. Peterson, Sex differences in depression: a role for effects of testosterone and its metabolites on vasopressin messenger RNA 1264
1179 preexisting anxiety, Psychiatry Res. 58 (1995) 1–12. levels in the bed nucleus of the stria terminalis of rats, J. Neurosci. 14 1265
1180 [53] G.R. Brooks, Despite problems, ‘‘masculinity” is a vital construct, Psychol. (1994) 1789–1794. 1266
1181 Men Masc. 11 (2010) 107–108. [84] G.J. De Vries, E.F. Rissman, R.B. Simerly, L.Y. Yang, E.M. Scordalakes, C.J. Auger, 1267
1182 [54] C.J. Brown, W.P. Robinson, The causes and consequences of random and non- A. Swain, R. Lovell-Badge, P.S. Burgoyne, A.P. Arnold, A model system for 1268
1183 random X chromosome inactivation in humans, Clin. Genet. 58 (2000) study of sex chromosome effects on sexually dimorphic neural and 1269
1184 353–363. behavioral traits, J. Neurosci. 22 (2002) 9005–9014. 1270

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

16 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

1271 [85] A. del Abril, S. Segovia, A. Guillamon, The bed nucleus of the stria terminalis in [116] J.M. Goldstein, S.V. Faraone, W.J. Chen, M.T. Tsuang, Genetic heterogeneity 1357
1272 the rat: regional sex differences controlled by gonadal steroids early after may in part explain sex differences in the familial risk for schizophrenia, Biol. 1358
1273 birth, Brain Res. 429 (1987) 295–300. Psychiatry 38 (1995) 808–813. 1359
1274 [86] P. Dewing, C.W. Chiang, K. Sinchak, H. Sim, P.O. Fernagut, S. Kelly, M.F. [117] C.D. Good, I. Johnsrude, J. Ashburner, R.N.A. Henson, K.J. Friston, R.S.J. 1360
1275 Chesselet, P.E. Micevych, K.H. Albrecht, V.R. Harley, E. Vilain, Direct regulation Frackowiak, Cerebral asymmetry and the effects of sex and handedness on 1361
1276 of adult brain function by the male-specific factor SRY, Curr. Biol. 16 (2006) brain structure: a voxel-based morphometric analysis of 465 normal adult 1362
1277 415–420. human brains, NeuroImage 14 (2001) 685–700. 1363
1278 [87] D.A. Di Monte, M. Lavasani, A.B. Manning-Bog, Environmental factors in [118] R.A. Gorski, R.E. Harlan, C.D. Jacobson, J.E. Shryne, A.M. Southam, Evidence for 1364
1279 Parkinson’s disease, Neurotoxicology 23 (2002) 487–502. the existence of a sexually dimorphic nucleus in the preoptic area of the rat, J. 1365
1280 [88] T. Di Paolo, Modulation of brain dopamine transmission by sex steroids, Rev. Comp. Neurol. 193 (1980) 529–539. 1366
1281 Neurosci. 5 (1994) 27–41. [119] R. Green, Family coocurrence of ‘‘gender dysphoria”: ten siblings or parent– 1367
1282 [89] M. Diamond, Homosexuality and bisexuality in different populations, Arch. child pairs, Arch. Sex. Behav. 29 (2000) 499–507. 1368
1283 Sex. Behav. 22 (1993) 291–310. [120] W.T. Greenough, C.S. Carter, C. Steerman, T.J. DeVoogd, Sex differences in 1369
1284 [90] M.C. Diamond, G.A. Dowling, R.E. Johnson, Morphologic cerebral cortical dendritic patterns in hamster preoptic area, Brain Res. 126 (1977) 63–72. 1370
1285 asymmetry in male and female rats, Exp. Neurol. 71 (1981) 261–268. [121] S. Grigoriadis, G.E. Robinson, Gender issues in depression, Annu. Clin. 1371
1286 [91] F. Dittrich, Y. Feng, R. Metzdorf, M. Gahr, Estrogen-inducible, sex-specific Psychiatry 19 (2007) 247–255. 1372
1287 expression of brain-derived neurotrophic factor mRNA in a forebrain song [122] G.M. Grimshaw, M.P. Bryden, J.A.K. Finegan, Relations between prenatal 1373
1288 control nucleus of the juvenile zebra finch, Proc. Natl. Acad. Sci. USA 96 testosterone and cerebral lateralization in children, Neuropsychology 9 1374
1289 (1999) 8241–8246. (1995) 68–79. 1375
1290 [92] E. Dominguez-Salazar, W. Portillo, M.J. Baum, J. Bakker, R.G. Paredes, Effect of [123] H.J. Groenewegen, The basal ganglia and motor control, Neural Plast. 10 1376
1291 prenatal androgen receptor antagonist or aromatase inhibitor on sexual (2003) 107–120. 1377
1292 behavior, partner preference and neuronal Fos responses to estrous female [124] J. Gubbay, P. Koopman, J. Collignon, P. Burgoyne, R. Lovell-Badge, Normal 1378
1293 odors in the rat accessory olfactory system, Physiol. Behav. 75 (2002) 337–346. structure and expression of Zfy genes in XY female mice mutant in Tdy, 1379
1294 [93] R.A. Dubin, H. Ostrer, Sry is a transcriptional activator, Mol. Endocrinol. 8 Development 109 (1990) 647–653. 1380
1295 (1994) 1182–1192. [125] R.C. Gur, F. Gunning-Dixon, W.B. Bilker, R.E. Gur, Sex differences in temporo- 1381
1296 [94] M.G. DuPree, B.S. Mustanski, S. Bocklandt, C. Nievergelt, D.H. Hamer, A limbic and frontal brain volumes of healthy adults, Cereb. Cortex 12 (2002) 1382
1297 candidate gene study of CYP19 (aromatase) and male sexual orientation, 998–1003. 1383
1298 Behav. Genet. 34 (2004) 243–250. [126] M.E. Gurney, M. Konishi, Hormone-induced sexual differentiation of brain 1384
1299 [95] M.F. Egan, M. Kojima, J.H. Callicott, T.E. Goldberg, B.S. Kolachana, A. Bertolino, and behavior in zebra finches, Science 208 (1980) 1380–1383. 1385
1300 E. Zaitsev, B. Gold, D. Goldman, M. Dean, B. Lu, D.R. Weinberger, The BDNF [127] M. Haberman, F. Hollingsworth, A. Falek, R.P. Michael, Gender identity 1386
1301 val66met polymorphism affects activity-dependent secretion of BDNF and confusion, schizophrenia and a 47 XYY karyotype: a case report, 1387
1302 human memory and hippocampal function, Cell 112 (2003) 257–269. Psychoneuroendocrinology 1 (1975) 207–209. 1388
1303 [96] Y. Egozi, S. Avissar, M. Sokolovsky, Muscarinic mechanisms and sex hormone [128] A. Hacker, B. Capel, P. Goodfellow, R. Lovell-Badge, Expression of Sry, the 1389
1304 secretion in rat adenohypophysis and preoptic area, Neuroendocrinology 35 mouse sex determining gene, Development 121 (1995) 1603–1614. 1390
1305 (1982) 93–97. [129] J.A. Hall, D. Kimura, Sexual orientation and performance on sexually 1391
1306 [97] E.M. Eicher, L.L. Washburn, J.B. Whitney 3rd, K.E. Morrow, Mus poschiavinus dimorphic motor tasks, Arch. Sex. Behav. 24 (1995) 395–407. 1392
1307 Y chromosome in the C57BL/6J murine genome causes sex reversal, Science [130] D.F. Halpern, Sex Differences in Cognitive Abilities, Lawrence Erlbaum 1393
1308 217 (1982) 535–537. Associates, 2000. pp. 159–163. 1394
1309 [98] G. Einstein, Sex and the Brain, MIT Press, Cambridge, Mass., 2007. [131] D.F. Halpern, Sex Differences in Cognitive Abilities, Lawrence Erlbaum 1395
1310 [99] L. Ellis, S. Cole-Harding, The effects of prenatal stress, and of prenatal alcohol Associates, 2000. pp. 147–148. 1396
1311 and nicotine exposure, on human sexual orientation, Physiol. Behav. 74 [132] D. Hamer, Genetics and male sexual orientation, Science 285 (1999) 803. 1397
1312 (2001) 213–226. [133] D. Hamer, S. Hu, V. Magnuson, N. Hu, A.M. Pattatucci, A linkage between DNA 1398
1313 [100] D.E. Emery, B.D. Sachs, Copulatory behavior in male rats with lesions in the markers on the X chromosome and male sexual orientation, Science 261 1399
1314 bed nucleus of the stria terminalis, Physiol. Behav. 17 (1976) 803–806. (1993) 321–327. 1400
1315 [101] I.V. Estabrooke, K. Mordecai, P. Maki, M.T. Ullman, The effect of sex hormones [134] S.E. Hammack, K.J. Richey, L.R. Watkins, S.F. Maier, Chemical lesion of the bed 1401
1316 on language processing, Brain Lang. 83 (2002) 143–146. nucleus of the stria terminalis blocks the behavioral consequences of 1402
1317 [102] A.M. Etgen, J.C. Morales, Somatosensory stimuli evoke norepinephrine release uncontrollable stress, Behav. Neurosci. 118 (2004) 443–448. 1403
1318 in the anterior ventromedial hypothalamus of sexually receptive female rats, [135] L. Hare, P. Bernard, F.J. Sanchez, P.N. Baird, E. Vilain, T. Kennedy, V.R. Harley, 1404
1319 J. Neuroendocrinol. 14 (2002) 213–218. Androgen receptor repeat length polymorphism associated with male-to- 1405
1320 [103] J. Feng, G. Sun, J. Yan, K. Noltner, W. Li, C.H. Buzin, J. Longmate, L.L. Heston, J. female transsexualism, Biol. Psychiatry 65 (2009) 93–96. 1406
1321 Rossi, S.S. Sommer, Evidence for X-chromosomal schizophrenia associated [136] P.J. Harrison, E.M. Tunbridge, Catechol-O-methyltransferase (COMT): a gene 1407
1322 with microRNA alterations, PLoS ONE 4 (2009) e6121. contributing to sex differences in brain function, and to sexual dimorphism in 1408
1323 [104] H.H. Fernandez, K.L. Lapane, B.R. Ott, J.H. Friedman, Gender differences in the the predisposition to psychiatric disorders, Neuropsychopharmacology 33 1409
1324 frequency and treatment of behavior problems in Parkinson’s disease. SAGE (2008) 3037–3045. 1410
1325 study group. Systematic assessment and geriatric drug use via epidemiology, [137] R.A. Harshman, E. Hampson, S.A. Berenbaum, Individual differences in 1411
1326 Mov. Disord. 15 (2000) 490–496. cognitive abilities and brain organization. Part I: sex and handedness 1412
1327 [105] R.H. Fitch, A.S. Berrebi, P.E. Cowell, L.M. Schrott, V.H. Denenberg, Corpus differences in ability, Can. J. Psychol. 37 (1983) 144–192. 1413
1328 callosum: effects of neonatal hormones on sexual dimorphism in the rat, [138] C.L. Hart, G.D. Smith, Alcohol consumption and mortality and hospital 1414
1329 Brain Res. 515 (1990) 111–116. admissions in men from the Midspan collaborative cohort study, Addiction 1415
1330 [106] R.H. Fitch, P.E. Cowell, L.M. Schrott, V.H. Denenberg, Corpus callosum: ovarian 103 (2008) 1979–1986. 1416
1331 hormones and feminization, Brain Res. 542 (1991) 313–317. [139] J.K. Hartshorne, M.T. Ullman, Why girls say ‘holded’ more than boys, Dev. Sci. 1417
1332 [107] A. Fleming, E. Vilain, The endless quest for sex determination genes, Clin. 9 (2006) 21–32. 1418
1333 Genet. 67 (2005) 15–25. [140] M. Heinrichs, T. Baumgartner, C. Kirschbaum, U. Ehlert, Social support and 1419
1334 [108] T. Foltynie, S.G. Lewis, T.E. Goldberg, A.D. Blackwell, B.S. Kolachana, D.R. oxytocin interact to suppress cortisol and subjective responses to 1420
1335 Weinberger, T.W. Robbins, R.A. Barker, The BDNF Val66Met polymorphism psychosocial stress, Biol. Psychiatry 54 (2003) 1389–1398. 1421
1336 has a gender specific influence on planning ability in Parkinson’s disease, J. [141] A.C. Hennessey, K. Wallen, D.A. Edwards, Preoptic lesions increase the display 1422
1337 Neurol. 252 (2005) 833–838. of lordosis by male rats, Brain Res. 370 (1986) 21–28. 1423
1338 [109] N.G. Forger, G.J. Rosen, E.M. Waters, D. Jacob, R.B. Simerly, G.J. de Vries, [142] S. Henningsson, L. Westberg, S. Nilsson, B. Lundstrom, L. Ekselius, O. Bodlund, 1424
1339 Deletion of Bax eliminates sex differences in the mouse forebrain, Proc. Natl. E. Lindstrom, M. Hellstrand, R. Rosmond, E. Eriksson, M. Landen, Sex steroid- 1425
1340 Acad. Sci. USA 101 (2004) 13666–13671. related genes and male-to-female transsexualism, Psychoneuro- 1426
1341 [110] S. Gao, H.C. Hendrie, K.S. Hall, S. Hui, The relationships between age, sex, and endocrinology 30 (2005) 657–664. 1427
1342 the incidence of dementia and Alzheimer disease: a meta-analysis, Arch. Gen. [143] G.M. Herek, J.R. Gillis, J.C. Cogan, Internalized stigma among sexual minority 1428
1343 Psychiatry 55 (1998) 809–815. adults: insights from a social psychological perspective, J. Couns. Psychol. 56 1429
1344 [111] G.M. Garden, D.J. Rothery, A female monozygotic twin pair discordant for (2009) 32–43. 1430
1345 transsexualism. Some theoretical implications, Br. J. Psychiatry 161 (1992) [144] A. Herlitz, E. Airaksinen, E. Nordstrom, Sex differences in episodic memory: the 1431
1346 852–854. impact of verbal and visuospatial ability, Neuropsychology 13 (1999) 590–597. 1432
1347 [112] L.H. Gargaglioni, K.C. Bícego, L.G.S. Branco, Brain monoaminergic neurons and [145] M. Hickey, D.A. Doherty, R. Hart, R.J. Norman, E. Mattes, H.C. Atkinson, D.M. 1433
1348 Q6 ventilatory control in vertebrates. Resp. Physiol. Neurobiol., in press Sloboda, Maternal and umbilical cord androgen concentrations do not predict 1434
1349 (Corrected Proof 2008). digit ratio (2D:4D) in girls: a prospective cohort study, Psychoneuro- 1435
1350 [113] J.D. Gatewood, A. Wills, S. Shetty, J. Xu, A.P. Arnold, P.S. Burgoyne, E.F. endocrinology 35 (2010) 1235–1244. 1436
1351 Rissman, Sex chromosome complement and gonadal sex influence aggressive [146] M. Hines, Brain Gender, Oxford University Press, Oxford; New York, 2004. pp. 1437
1352 and parental behaviors in mice, J. Neurosci. 26 (2006) 2335–2342. 191–197. 1438
1353 [114] W.R. Gibb, Neuropathology of Parkinson’s disease and related syndromes, [147] M. Hines, Brain Gender, Oxford University Press, Oxford; New York, 2004. 1439
1354 Neurol. Clin. 10 (1992) 361–376. [148] M. Hines, L.S. Allen, R.A. Gorski, Sex differences in subregions of the medial 1440
1355 [115] J.M. Goldstein, Sex, hormones and affective arousal circuitry dysfunction in nucleus of the amygdala and the bed nucleus of the stria terminalis of the rat, 1441
1356 schizophrenia, Horm. Behav. 50 (2006) 612–622. Brain Res. 579 (1992) 321–326. 1442

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 17

1443 [149] M. Hines, C. Brook, G.S. Conway, Androgen and psychosexual development: [179] D. Kimura, R.A. Harshman, Sex differences in brain organization for verbal 1529
1444 core gender identity, sexual orientation and recalled childhood gender role and non-verbal functions, Prog. Brain Res. 61 (1984) 423–441. 1530
1445 behavior in women and men with congenital adrenal hyperplasia (CAH), J. [180] K.M. Kirk, J.M. Bailey, M.P. Dunne, N.G. Martin, Measurement models for 1531
1446 Sex Res. 41 (2004) 75–81. sexual orientation in a community twin sample, Behav. Genet. 30 (2000) 1532
1447 [150] C.C. Holloway, D.F. Clayton, Estrogen synthesis in the male brain triggers 345–356. 1533
1448 development of the avian song control pathway in vitro, Nat. Neurosci. 4 [181] P. Kirsch, C. Esslinger, Q. Chen, D. Mier, S. Lis, S. Siddhanti, H. Gruppe, V.S. 1534
1449 (2001) 170–175. Mattay, B. Gallhofer, A. Meyer-Lindenberg, Oxytocin modulates neural 1535
1450 [151] M.M. Holmes, B.D. Goldman, S.L. Goldman, M.L. Seney, N.G. Forger, circuitry for social cognition and fear in humans, J. Neurosci. 25 (2005) 1536
1451 Neuroendocrinology and sexual differentiation in eusocial mammals, Front. 11489–11493. 1537
1452 Neuroendocrinol. 30 (2009) 519–533. [182] P. Koopman, Gonad development: signals for sex, Curr. Biol. 11 (2001) R481– 1538
1453 [152] C.J. Hong, H.C. Liu, T.Y. Liu, C.H. Lin, C.Y. Cheng, S.J. Tsai, Brain-derived R483. 1539
1454 neurotrophic factor (BDNF) Val66Met polymorphisms in Parkinson’s disease [183] C. Kordon, J. Glowinski, Role of hypothalamic monoaminergic neurones in the 1540
1455 and age of onset, Neurosci. Lett. 353 (2003) 75–77. gonadotrophin release-regulating mechanisms, Neuropharmacology 11 1541
1456 [153] H. Hörtnagl, L. Hansen, G. Kindel, B. Schneider, A.E. Tamert, I. Hanin, Sex (1972) 153–162. 1542
1457 differences and estrous cycle-variations in the AF64A-induced cholinergic [184] M. Kosfeld, M. Heinrichs, P.J. Zak, U. Fischbacher, E. Fehr, Oxytocin increases 1543
1458 deficit in the rat hippocampus, Brain Res. Bull. 31 (1993) 129–134. trust in humans, Nature 435 (2005) 673–676. 1544
1459 [154] S. Hu, A. Pattatucci, C. Patterson, L. Li, D. Fulker, S. Cherny, L. Kruglyak, D. [185] T.R. Kosten, B.J. Rounsaville, H.D. Kleber, Ethnic and gender differences 1545
1460 Hamer, Linkage between sexual orientation and chromosome Xq28 in males among opiate addicts, Int. J. Addict. 20 (1985) 1143–1162. 1546
1461 but not females, Nat. Genet. 11 (1995) 248–256. [186] J. Kulkarni, A. de Castella, D. Smith, J. Taffe, N. Keks, D. Copolov, A clinical trial 1547
1462 [155] J.S. Hyde, The gender similarities hypothesis, Am. Psychol. 60 (2005) 581– of the effects of estrogen in acutely psychotic women, Schizophr. Res. 20 1548
1463 592. (1996) 247–252. 1549
1464 [156] J.S. Hyde, New directions in the study of gender similarities and differences, [187] T.C. LaGrange, R.A. Silverman, Low self-control and opportunity: testing the 1550
1465 Curr. Dir. Psychol. Sci. 16 (2007) 259–263. general theory of crime as an explanation for gender differences in 1551
1466 [157] C. Hyde, J.C. Kenna, A male MZ twin pair, concordant for delinquency, Criminology 37 (1999) 41–72. 1552
1467 transsexualism, discordant for schizophrenia, Acta Psychiatr. Scand. 56 [188] K. Lawrence, J. Kuntsi, M. Coleman, R. Campbell, D. Skuse, Face and emotion 1553
1468 (1977) 265–275. recognition deficits in Turner syndrome: a possible role for X-linked genes in 1554
1469 [158] Institute of Medicine (US), Committee on understanding the biology of sex amygdala development, Neuropsychology 17 (2003) 39–49. 1555
1470 and gender differences, in: T.M. Wizemann, M.L. Pardue (Eds.), Exploring the [189] V. Lefebvre, B. Dumitriu, A. Penzo-Mendez, Y. Han, B. Pallavi, Control of cell 1556
1471 Biological Contributions to Human Health: Does Sex Matter?, National fate and differentiation by Sry-related high-mobility-group box (Sox) 1557
1472 Academy Press, Washington, DC, 2001 transcription factors, Int. J. Biochem. Cell Biol. 39 (2007) 2195–2214. 1558
1473 [159] S. Ito, S. Murakami, K. Yamanouchi, Y. Arai, Prenatal androgen exposure, [190] C. Leranth, R.H. Roth, J.D. Elsworth, F. Naftolin, T.L. Horvath, D.E. Redmond Jr., 1559
1474 preoptic area and reproductive functions in the female rat, Brain Dev. 8 Estrogen is essential for maintaining nigrostriatal dopamine neurons in 1560
1475 (1986) 463–468. primates: implications for Parkinson’s disease and memory, J. Neurosci. 20 1561
1476 [160] A. Jablensky, Epidemiology of schizophrenia: the global burden of disease and (2000) 8604–8609. 1562
1477 disability, Eur. Arch. Psychiatry Clin. Neurosci. 250 (2000) 274–285. [191] R.F. Levant, R.J. Hall, C.M. Williams, N.T. Hasan, Gender differences in 1563
1478 [161] S. Jacob, C.W. Brune, C.S. Carter, B.L. Leventhal, C. Lord, E.H. Cook Jr., alexithymia, Psychol. Men Masc. 10 (2009) 190–203. 1564
1479 Association of the oxytocin receptor gene (OXTR) in Caucasian children and [192] S. LeVay, A difference in hypothalamic structure between heterosexual and 1565
1480 adolescents with autism, Neurosci. Lett. 417 (2007) 6–9. homosexual men, Science 253 (1991) 1034–1037. 1566
1481 [162] E.C. Jacobs, W. Grisham, A.P. Arnold, Lack of a synergistic effect between [193] M. Levitt, S. Spector, A. Sjoerdsma, S. Udenfriend, Elucidation of the rate- 1567
1482 estradiol and dihydrotestosterone in the masculinization of the zebra finch limiting step in norepinephrine biosynthesis in the perfused guinea-pig 1568
1483 song system, J. Neurobiol. 27 (1995) 513–519. heart, J. Pharmacol. Exp. Ther. 148 (1965) 1–8. 1569
1484 [163] S.m.R.L.o. Joca, F.R.r. Ferreira, F.S. Guimarães, Modulation of stress [194] H.M. Levitt, E. Ovrebo, M.B. Anderson-Cleveland, C. Leone, J.Y. Jeong, J.R. Arm, 1570
1485 consequences by hippocampal monoaminergic, glutamatergic and nitrergic B.P. Bonin, J. Cicala, R. Coleman, A. Laurie, J.M. Vardaman, S.G. Horne, 1571
1486 neurotransmitter systems, Stress 10 (2007) 227–249. Balancing dangers: GLBT experience in a time of anti-GLBT legislation, J. 1572
1487 [164] J.K. Johnson, R. McCleary, M.H. Oshita, C.W. Cotman, Initiation and Couns. Psychol. 56 (2009) 67–81. 1573
1488 propagation stages of beta-amyloid are associated with distinctive [195] C. Lewin, G. Wolgers, A. Herlitz, Sex differences favoring women in verbal but 1574
1489 apolipoprotein E, age, and gender profiles, Brain Res. 798 (1998) 18–24. not in visuospatial episodic memory, Neuropsychology 15 (2001) 165–173. 1575
1490 [165] B.E. Jones, From waking to sleeping: neuronal and chemical substrates, [196] F.R. Lillie, The theory of the free-martin, Science 43 (1916) 611–613. 1576
1491 Trends Pharmacol. Sci. 26 (2005) 578–586. [197] F.R. Lillie, General Biological Introduction, Williams and Wilkins Co., 1577
1492 [166] M.B. Jones, R. Blanchard, Birth order and male homosexuality: extension of Baltimore, 1939. 1578
1493 Slater’s index, Hum. Biol. 70 (1998) 775–787. [198] R.A. Lippa, Sex differences in sex drive, sociosexuality, and height across 53 1579
1494 [167] A. Jost, Hormonal factors in the sex differentiation of the mammalian foetus, nations: testing evolutionary and social structural theories, Arch. Sex. Behav. 1580
1495 Philos. Trans. R. Soc. B – Biol. Sci. 259 (1970) 119–130. 38 (2009) 631–651. 1581
1496 [168] A.J. Justice, H. de Wit, Acute effects of d-amphetamine during the follicular [199] J.C. Lucas, M.B. Renfree, G. Shaw, C.M. Butler, The influence of the anti- 1582
1497 and luteal phases of the menstrual cycle in women, Psychopharmacology androgen flutamide on early sexual differentiation of the marsupial male, J. 1583
1498 (Berlin) 145 (1999) 67–75. Reprod. Fertil. 109 (1997) 205–212. 1584
1499 [169] A.J. Justice, H. De Wit, Acute effects of d-amphetamine during the early and [200] E. Luders, C. Gaser, K.L. Narr, A.W. Toga, Why sex matters: brain size 1585
1500 late follicular phases of the menstrual cycle in women, Pharmacol. Biochem. independent differences in gray matter distributions between men and 1586
1501 Behav. 66 (2000) 509–515. women, J. Neurosci. 29 (2009) 14265–14270. 1587
1502 [170] S.P. Kalra, P.S. Kalra, Neural regulation of luteinizing hormone secretion in the [201] V. Luine, Sex differences in chronic stress effects on memory in rats, Stress 5 1588
1503 rat, Endocr. Rev. 4 (1983) 311–351. (2002) 205–216. 1589
1504 [171] K. Kansaku, A. Yamaura, S. Kitazawa, Sex differences in lateralization revealed [202] J.P. Macke, N. Hu, S. Hu, M. Bailey, V.L. King, T. Brown, D. Hamer, J. Nathans, 1590
1505 in the posterior language areas, Cereb. Cortex 10 (2000) 866–872. Sequence variation in the androgen receptor gene is not a common 1591
1506 [172] P. Kehoe, F. Wavrant-De Vrieze, R. Crook, W.S. Wu, P. Holmans, I. Fenton, G. determinant of male sexual orientation, Am. J. Hum. Genet. 53 (1993) 844– 1592
1507 Spurlock, N. Norton, H. Williams, N. Williams, S. Lovestone, J. Perez-Tur, M. 852. 1593
1508 Hutton, M.C. Chartier-Harlin, S. Shears, K. Roehl, J. Booth, W. Van Voorst, D. [203] K. Maeda, S. Adachi, K. Inoue, S. Ohkura, H. Tsukamura, Metastin/kisspeptin 1594
1509 Ramic, J. Williams, A. Goate, J. Hardy, M.J. Owen, A full genome scan for late and control of estrous cycle in rats, Rev. Endocr. Metab. Disord. 8 (2007) 21– 1595
1510 onset Alzheimer’s disease, Hum. Mol. Genet. 8 (1999) 237–245. 29. 1596
1511 [173] K.S. Kendler, M.C. Neale, R.C. Kessler, A.C. Heath, L.J. Eaves, A longitudinal [204] S. Magre, A. Jost, Sertoli cells and testicular differentiation in the rat fetus, J. 1597
1512 twin study of 1-year prevalence of major depression in women, Arch. Gen. Electron. Microsc. Tech. 19 (1991) 172–188. 1598
1513 Psychiatry 50 (1993) 843–852. [205] P.M. Maki, S.M. Resnick, Longitudinal effects of estrogen replacement therapy 1599
1514 [174] K.S. Kendler, L.M. Thornton, S.E. Gilman, R.C. Kessler, Sexual orientation in a on PET cerebral blood flow and cognition, Neurobiol. Aging 21 (2000) 373– 1600
1515 US national sample of twin and nontwin sibling Pairs, Am. J. Psychiatry 157 383. 1601
1516 (2000) 1843–1846. [206] A. Mansour, C.A. Fox, H. Akil, S.J. Watson, Opioid-receptor mRNA expression 1602
1517 [175] G.A. Kennett, F. Chaouloff, M. Marcou, G. Curzon, Female rats are more in the rat CNS: anatomical and functional implications, Trends Neurosci. 18 1603
1518 vulnerable than males in an animal model of depression: the possible role of (1995) 22–29. 1604
1519 serotonin, Brain Res. 382 (1986) 416–421. [207] S.B. Manuck, J.D. Flory, R.E. Ferrell, J.J. Mann, M.F. Muldoon, A regulatory 1605
1520 [176] R. Kessler, Gender differences in major depression: epidemiological findings, polymorphism of the monoamine oxidase-A gene may be associated with 1606
1521 in: E. Frank (Ed.), Gender and its Effects on Psychopathology, American variability in aggression, impulsivity, and central nervous system 1607
1522 Psychiatric Press, Washington, 2000, pp. 61–84. serotonergic responsivity, Psychiatry Res. 95 (2000) 9–23. 1608
1523 [177] R.C. Kessler, K.A. McGonagle, S. Zhao, C.B. Nelson, M. Hughes, S. Eshleman, [208] D.M. Maraganore, M. de Andrade, T.G. Lesnick, K.J. Strain, M.J. Farrer, W.A. 1609
1524 H.U. Wittchen, K.S. Kendler, Lifetime and 12-month prevalence of DSM-III-R Rocca, P.V. Pant, K.A. Frazer, D.R. Cox, D.G. Ballinger, High-resolution whole- 1610
1525 psychiatric disorders in the United States. Results from the national genome association study of Parkinson disease, Am. J. Hum. Genet. 77 (2005) 1611
1526 comorbidity survey, Arch. Gen. Psychiatry 51 (1994) 8–19. 685–693. 1612
1527 [178] D. Kimura, Sex, sexual orientation and sex hormones influence human [209] E.J. Marco, D.H. Skuse, Autism-lessons from the X chromosome, Soc. Cogn. 1613
1528 cognitive function, Curr. Opin. Neurobiol. 6 (1996) 259–263. Affect. Neurosci. 1 (2006) 183–193. 1614

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

18 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

1615 [210] E. Martignoni, R.E. Nappi, A. Citterio, D. Calandrella, E. Corengia, A. Fignon, R. [238] J.M. O’neil, Is criticism of generic masculinity, essentialism, and positive- 1700
1616 Zangaglia, G. Riboldazzi, C. Pacchetti, G. Nappi, Parkinson’s disease and healthy-masculinity a problem for the psychology of men?, Psychol Men 1701
1617 reproductive life events, Neurol. Sci. 23 (Suppl. 2) (2002) S85–S86. Masc. 11 (2010) 98–106. 1702
1618 [211] A. Matsumoto, Y. Arai, Male-female difference in synaptic organization of the [239] M. Ota, J.T. Crofton, L. Share, Hemorrhage-induced vasopressin release in the 1703
1619 ventromedial nucleus of the hypothalamus in the rat, Neuroendocrinology 42 paraventricular nucleus measured by in vivo microdialysis, Brain Res. 658 1704
1620 (1986) 232–236. (1994) 49–54. 1705
1621 [212] A. Mayer, G. Lahr, D.F. Swaab, C. Pilgrim, I. Reisert, The Y-chromosomal genes [240] M.G. Packard, Posttraining estrogen and memory modulation, Horm. Behav. 1706
1622 SRY and ZFY are transcribed in adult human brain, Neurogenetics 1 (1998) 34 (1998) 126–139. 1707
1623 281–288. [241] N. Pankratz, W.C. Nichols, S.K. Uniacke, C. Halter, J. Murrell, A. Rudolph, C.W. 1708
1624 [213] A. Mayer, G. Mosler, W. Just, C. Pilgrim, I. Reisert, Developmental profile of Sry Shults, P.M. Conneally, T. Foroud, Genome-wide linkage analysis and 1709
1625 transcripts in mouse brain, Neurogenetics 3 (2000) 25–30. evidence of gene-by-gene interactions in a sample of 362 multiplex 1710
1626 [214] B.M. McCarty, C.J. Migeon, H.F. Meyer-Bahlburg, H. Zacur, A.B. Wisniewski, Parkinson disease families, Hum. Mol. Genet. 12 (2003) 2599–2608. 1711
1627 Medical and psychosexual outcome in women affected by complete gonadal [242] C.E. Patek, J.B. Kerr, R.G. Gosden, K.W. Jones, K. Hardy, A.L. Muggleton-Harris, 1712
1628 dysgenesis, J. Pediatr. Endocrinol. Metab. 19 (2006) 873–877. A.H. Handyside, D.G. Whittingham, M.L. Hooper, Sex chimaerism, fertility and 1713
1629 [215] E. McCauley, P. Feuillan, H. Kushner, J.L. Ross, Psychosocial development sex determination in the mouse, Development 113 (1991) 311–325. 1714
1630 in adolescents with Turner syndrome, J. Dev. Behav. Pediatr. 22 (2001) [243] A.M.L. Pattatucci, D.H. Hamer, Development and familiarity of sexual 1715
1631 360–365. orientation in females, Behav. Genet. 25 (1995) 407–420. 1716
1632 [216] D. McFadden, J.C. Loehlin, S.M. Breedlove, R.A. Lippa, J.T. Manning, Q. Rahman, [244] D.W. Pfaff, Hormones, Brain, and Behavior, Academic Press, Amsterdam, 1717
1633 A reanalysis of five studies on sexual orientation and the relative length of Boston, 2002. 1718
1634 the 2nd and 4th fingers (the 2D:4D ratio), Arch. Sex. Behav. 34 (2005) 341– [245] T. Pfluger, S. Weil, S. Weis, C. Vollmar, D. Heiss, J. Egger, R. Scheck, K. Hahn, 1719
1635 356. Normative volumetric data of the developing hippocampus in children based 1720
1636 [217] B. Meister, Neurotransmitters in key neurons of the hypothalamus that on magnetic resonance imaging, Epilepsia 40 (1999) 414–423. 1721
1637 regulate feeding behavior and body weight, Physiol. Behav. 92 (2007) 263– [246] S.M. Phillips, B.B. Sherwin, Effects of estrogen on memory function in 1722
1638 271. surgically menopausal women, Psychoneuroendocrinology 17 (1992) 485– 1723
1639 [218] I.H. Meyer, P.A. Wilson, Sampling lesbian, gay, and bisexual populations, J. 495. 1724
1640 Couns. Psychol. 56 (2009) 23–31. [247] C.H. Phoenix, R.W. Goy, A.A. Gerall, W.C. Young, Organizing action of 1725
1641 [219] A. Milsted, L. Serova, E.L. Sabban, G. Dunphy, M.E. Turner, D.L. Ely, Regulation prenatally administered testosterone propionate on the tissues 1726
1642 of tyrosine hydroxylase gene transcription by Sry, Neurosci. Lett. 369 (2004) mediating mating behavior in the female guinea pig, Endocrinology 1727
1643 203–207. 65 (1959) 369–382. 1728
1644 [220] D. Mitsushima, K. Yamada, K. Takase, T. Funabashi, F. Kimura, Sex differences [248] J.P. Piek, N. Gasson, N. Barrett, I. Case, Limb and gender differences in the 1729
1645 in the basolateral amygdala: the extracellular levels of serotonin and development of coordination in early infancy, Hum. Mov. Sci. 21 (2002) 621– 1730
1646 dopamine, and their responses to restraint stress in rats, Eur. J. NeuroSci. 639. 1731
1647 24 (2006) 3245–3254. [249] R.C. Pillard, J.D. Weinrich, Evidence of familial nature of male homosexuality, 1732
1648 [221] S. Mizukami, M. Nishizuka, Y. Arai, Sexual difference in nuclear volume and Arch. Gen. Psychiatry 43 (1986) 808–812. 1733
1649 its ontogeny in the rat amygdala, Exp. Neurol. 79 (1983) 569–575. [250] A. Poletti, P. Negri-Cesi, L. Martini, Reflections on the diseases linked to 1734
1650 [222] C. Modahl, L. Green, D. Fein, M. Morris, L. Waterhouse, C. Feinstein, H. Levin, mutations of the androgen receptor, Endocrine 28 (2005) 243–262. 1735
1651 Plasma oxytocin levels in autistic children, Biol. Psychiatry 43 (1998) 270– [251] J.B. Powers, Hormonal control of sexual receptivity during the estrous cycle of 1736
1652 277. the rat, Physiol. Behav. 5 (1970) 831–835. 1737
1653 [223] B. Moradi, J.J. Mohr, R.L. Worthington, R.E. Fassinger, Counseling psychology [252] B. Powers, E.S. Valenstein, Sexual receptivity: facilitation by medial preoptic 1738
1654 research on sexual (orientation) minority issues: conceptual and lesions in female rats, Science 175 (1972) 1003–1005. 1739
1655 methodological challenges and opportunities, J. Couns. Psychol. 56 (2009) [253] M.L.D. Pozzo, A. Aoki, Stereological analysis of the hypothalamic 1740
1656 5–22. ventromedial nucleus: II. Hormone-induced changes in the synaptogenic 1741
1657 [224] A.G. Morris, E. Gaitonde, P.J. McKenna, J.D. Mollon, D.M. Hunt, CAG repeat pattern, Brain Res. Dev. Brain Res. 61 (1991) 189–196. 1742
1658 expansions and schizophrenia: association with disease in females and with [254] N.J. Preston, K.G. Orr, R. Date, L. Nolan, D.J. Castle, Gender differences in 1743
1659 early age-at-onset, Hum. Mol. Genet. 4 (1995) 1957–1961. premorbid adjustment of patients with first episode psychosis, Schizophr. 1744
1660 [225] J.A. Morris, C.L. Jordan, S.M. Breedlove, Sexual differentiation of the Res. 55 (2002) 285–290. 1745
1661 vertebrate nervous system, Nat. Neurosci. 7 (2004) 1034–1039. [255] C.A. Quigley, A. De Bellis, K.B. Marschke, M.K. el-Awady, E.M. Wilson, F.S. 1746
1662 [226] C.A. Moss-Racusin, J.E. Phelan, L.A. Rudman, When men break the gender French, Androgen receptor defects: historical, clinical, and molecular 1747
1663 rules: status incongruity and backlash against modest men, Psychol. Men perspectives, Endocr. Rev. 16 (1995) 271–321. 1748
1664 Masc. 11 (2010) 140–151. [256] J.J. Quinn, P.K. Hitchcott, E.A. Umeda, A.P. Arnold, J.R. Taylor, Sex chromosome 1749
1665 [227] C.A. Munro, M.E. McCaul, D.F. Wong, L.M. Oswald, Y. Zhou, J. Brasic, H. complement regulates habit formation, Nat. Neurosci. 10 (2007) 1398–1400. 1750
1666 Kuwabara, A. Kumar, M. Alexander, W. Ye, G.S. Wand, Sex differences in [257] H. Raab, C. Pilgrim, I. Reisert, Effects of sex and estrogen on tyrosine 1751
1667 striatal dopamine release in healthy adults, Biol. Psychiatry 59 (2006) 966– hydroxylase mRNA in cultured embryonic rat mesencephalon, Brain Res. Mol. 1752
1668 974. Brain Res. 33 (1995) 157–164. 1753
1669 [228] D.D. Murphy, N.B. Cole, M. Segal, Brain-derived neurotrophic factor mediates [258] P. Ragonese, M. D’Amelio, G. Salemi, P. Aridon, M. Gammino, A. Epifanio, L. 1754
1670 estradiol-induced dendritic spine formation in hippocampal neurons, Proc. Morgante, G. Savettieri, Risk of Parkinson disease in women: effect of 1755
1671 Natl. Acad. Sci. USA 95 (1998) 11412–11417. reproductive characteristics, Neurology 62 (2004) 2010–2014. 1756
1672 [229] C.M. Murphy, C.T. Taft, C.I. Eckhardt, Anger problem profiles among partner [259] I. Reisert, C. Pilgrim, Sexual differentiation of monoaminergic neurons – 1757
1673 violent men: differences in clinical presentation and treatment outcome, J. genetic or epigenetic?, Trends Neurosci 14 (1991) 468–473. 1758
1674 Couns. Psychol. 54 (2007) 189–200. [260] M.B. Renfree, R.V. Short, Sex determination in marsupials: evidence for a 1759
1675 [230] B.S. Mustanski, M.G. Dupree, C.M. Nievergelt, S. Bocklandt, N.J. Schork, D.H. marsupial-eutherian dichotomy, Philos. Trans. R. Soc. B – Biol. Sci. 322 (1988) 1760
1676 Hamer, A genomewide scan of male sexual orientation, Hum. Genet. 116 41–53. 1761
1677 (2005) 272–278. [261] M.B. Renfree, W.S.O.R.V. Short, G. Shaw, Sexual differentiation of the 1762
1678 [231] F. Naftolin, Brain aromatization of androgens, J. Reprod. Med. 39 (1994) 257– urogenital system of the fetal and neonatal tammar wallaby, Macropus 1763
1679 261. eugenii, Anat. Embryol. (Berl.) 194 (1996) 111–134. 1764
1680 [232] J. Nielsen, M. Wohlert, Sex chromosome abnormalities found among 34,910 [262] R.W. Rhees, H.N. Al-Saleh, E.W. Kinghorn, D.E. Fleming, E.D. Lephart, 1765
1681 newborn children: results from a 13-year incidence study in Arhus, Denmark, Relationship between sexual behavior and sexually dimorphic structures in 1766
1682 Birth Defects – Orig. Artic. Ser. 26 (1990) 209–223. the anterior hypothalamus in control and prenatally stressed male rats, Brain 1767
1683 [233] S. Nishizawa, C. Benkelfat, S.N. Young, M. Leyton, S. Mzengeza, C. Res. Bull. 50 (1999) 193–199. 1768
1684 DeMontigny, P. Blier, M. Diksic, Differences between males and females in [263] M.E. Rhodes, R.T. Rubin, Functional sex differences (‘sexual diergism’) of 1769
1685 rates of serotonin synthesis in human brain, Proc. Natl. Acad. Sci. USA 94 central nervous system cholinergic systems, vasopressin, and hypothalamic– 1770
1686 (1997) 5308–5313. pituitary–adrenal axis activity in mammals: a selective review, Brain Res. 1771
1687 [234] S. Nishizawa, C. Benkelfat, S.N. Young, M. Leyton, S. Mzengeza, C. De Rev. 30 (1999) 135–152. 1772
1688 Montigny, P. Blier, M. Diksic, Differences between males and females in rates [264] G. Rice, C. Anderson, N. Risch, G. Ebers, Male homosexuality: absence of 1773
1689 of serotonin synthesis human brain, Proc. Natl. Acad. Sci. USA 94 (1997) linkage to microsatellite markers at Xq28, Science 284 (1999) 665–667. 1774
1690 5308–5313. [265] G. Rieger, M.L. Chivers, J.M. Bailey, Sexual arousal patterns of bisexual men, 1775
1691 [235] A. Nordenstrom, A. Servin, G. Bohlin, A. Larsson, A. Wedell, Sex-typed toy play Psychol. Sci. 16 (2005) 579–584. 1776
1692 behavior correlates with the degree of prenatal androgen exposure assessed [266] S.J. Robbins, R.N. Ehrman, A.R. Childress, C.P. O’Brien, Comparing levels of 1777
1693 by CYP21 genotype in girls with congenital adrenal hyperplasia, J. Clin. cocaine cue reactivity in male and female outpatients, Drug Alcohol Depend. 1778
1694 Endocrinol. Metab. 87 (2002) 5119–5124. 53 (1999) 223–230. 1779
1695 [236] F. Nottebohm, A.P. Arnold, Sexual dimorphism in vocal control areas of the [267] A. Rojo, M. Aguilar, M.T. Garolera, E. Cubo, I. Navas, S. Quintana, Depression in 1780
1696 songbird brain, Science 194 (1976) 211–213. Parkinson’s disease: clinical correlates and outcome, Parkinsonism Relat. 1781
1697 [237] K.N. Ochsner, The social-emotional processing stream: five core constructs Disord. 10 (2003) 23–28. 1782
1698 and their translational potential for schizophrenia and beyond, Biol. [268] C.E. Roselli, F. Stormshak, The neurobiology of sexual partner preferences in 1783
1699 Psychiatry 64 (2008) 48–61. rams, Horm. Behav. 55 (2009) 611–620. 1784

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx 19

1785 [269] C.E. Roselli, L.E. Horton, J.A. Resko, Distribution and regulation of aromatase [299] H.B. Simpson, D.S. Vicario, Early estrogen treatment of female zebra finches 1871
1786 activity in the rat hypothalamus and limbic system, Endocrinology 117 masculinizes the brain pathway for learned vocalizations, J. Neurobiol. 22 1872
1787 (1985) 2471–2477. (1991) 777–793. 1873
1788 [270] C.E. Roselli, S.E. Abdelgadir, J.A. Resko, Regulation of aromatase gene [300] D.H. Skuse, R.S. James, D.V. Bishop, B. Coppin, P. Dalton, r.G. Aamodt-Leepe, 1874
1789 expression in the adult rat brain, Brain Res. Bull. 44 (1997) 351–357. M. Bacarese-Hamilton, C. Creswell, R. McGurk, P.A. Jacobs, Evidence from 1875
1790 [271] C.E. Roselli, K. Larkin, J.A. Resko, J.N. Stellflug, F. Stormshak, The volume of a Turner’s syndrome of an imprinted X-linked locus affecting cognitive 1876
1791 sexually dimorphic nucleus in the ovine medial preoptic area/anterior function, Nature 387 (1997). 1877
1792 hypothalamus varies with sexual partner preference, Endocrinology 145 [301] D.H. Skuse, R.S. James, D.V. Bishop, B. Coppin, P. Dalton, G. Aamodt-Leeper, M. 1878
1793 (2004) 478–483. Bacarese-Hamilton, C. Creswell, R. McGurk, P.A. Jacobs, Evidence from 1879
1794 [272] J.L. Ross, G.A. Stefanatos, H. Kushner, C. Bondy, L. Nelson, A. Zinn, D. Roeltgen, Turner’s syndrome of an imprinted X-linked locus affecting cognitive 1880
1795 The effect of genetic differences and ovarian failure: intact cognitive function function, Nature 387 (1997) 705–708. 1881
1796 in adult women with premature ovarian failure versus Turner syndrome, J. [302] R.P. Snaith, S. Penhale, P. Horsfield, Male-to-female transsexual with XYY 1882
1797 Clin. Endocrinol. Metab. 89 (2004) 1817–1822. karyotype, Lancet 337 (1991) 557–558. 1883
1798 [273] R.T. Rubin, L.K. Sekula, S. O’Toole, M.E. Rhodes, R.K. Czambel, Pituitary– [303] M.B. Solomon, J.P. Herman, Sex differences in psychopathology: of gonads, 1884
1799 adrenal cortical responses to low-dose physostigmine and arginine adrenals and mental illness, Physiol. Behav. 97 (2009) 250–258. 1885
1800 vasopressin administration in normal women and men, [304] R. Sperry, Some effects of disconnecting the cerebral hemispheres, Science 1886
1801 Neuropsychopharmacology 20 (1999) 434–446. 217 (1982) 1223–1226. 1887
1802 [274] D.R. Rubinow, P.J. Schmidt, C.A. Roca, Estrogen–serotonin interactions: [305] J.K. Staley, S. Krishnan-Sarin, S. Zoghbi, G. Tamagnan, M. Fujita, J.P. Seibyl, P.K. 1888
1803 implications for affective regulation, Biol. Psychiatry 44 (1998) 839–850. Maciejewski, S. O’Malley, R.B. Innis, Sex differences in [123I]beta-CIT SPECT 1889
1804 [275] H.G. Ruhe, N.S. Mason, A.H. Schene, Mood is indirectly related to serotonin, measures of dopamine and serotonin transporter availability in healthy 1890
1805 norepinephrine and dopamine levels in humans: a meta-analysis of smokers and nonsmokers, Synapse 41 (2001) 275–284. 1891
1806 monoamine depletion studies, Mol. Psychiatry 12 (2007) 331–359. [306] E.R. Stockman, R.S. Callaghan, M.J. Baum, Effects of neonatal castration and 1892
1807 [276] M. Sadeghi, A. Fakhrai, Transsexualism in female monozygotic twins: a case testosterone treatment on sexual partner preference in the ferret, Physiol. 1893
1808 report, Aust. NZ J. Psychiatry 34 (2000) 862–864. Behav. 34 (1985) 409–414. 1894
1809 [277] F.J. Sanchez, E. Vilain, Collective self-esteem as a coping resource for male-to- [307] P. Sudbeck, G. Scherer, Two independent nuclear localization signals are 1895
1810 female transsexuals, J. Couns. Psychol. 56 (2009) 202–209. present in the DNA-binding high-mobility group domains of SRY and SOX9, J. 1896
1811 [278] F.J. Sánchez, S. Bocklandt, E. Vilain, The biology of sexual orientation and Biol. Chem. 272 (1997) 27848–27852. 1897
1812 gender identity, in: D.W. Pfaff, A.P. Arnold, A.M. Etgen, S.E. Fahrbach, R.T. [308] H. Sumida, M. Nishizuka, Y. Kano, Y. Arai, Sex differences in the anteroventral 1898
1813 Rubin (Eds.), Hormones, Brain and Behavior, Academic Press, San Diego, CA, periventricular nucleus of the preoptic area and in the related effects of 1899
1814 2009, pp. 1911–1929. androgen in prenatal rats, Neurosci. Lett. 151 (1993) 41–44. 1900
1815 [279] A.R. Sanders, K. Dawood, Nature Encyclopedia of Life Sciences, Nature [309] D.F. Swaab, J.N. Zhou, M. Fodor, M.A. Hofman, Sexual differentiation of the 1901
1816 Publishing Group, London, 2003. human hypothalamus: differences according to sex, sexual orientation, and 1902
1817 [280] H.E. Scharfman, N.J. Maclusky, Similarities between actions of estrogen and transsexuality, in: Ellis Lee, Ebertz Linda (Eds.), Sexual Orientation: Toward 1903
1818 BDNF in the hippocampus: coincidence or clue?, Trends Neurosci 28 (2005) Biological Understanding, Praeger Publishers/Greenwood Publishing Group, 1904
1819 79–85. Inc., Westport, CT, US, 1997, pp. 129–150 (xxii, 276 pp. see book). 1905
1820 [281] D.P. Schmitt, Sociosexuality from Argentina to Zimbabwe: a 48-nation study [310] M. Sylvester, S.C. Hayes, Unpacking masculinity as a construct: ontology, 1906
1821 of sex, culture, and strategies of human mating, Behav. Brain. Sci. 28 (2005) pragmatism, and an analysis of language, Psychol. Men Masc. 11 (2010) 91–97. 1907
1822 247–275 (Discussion 275–311). [311] D. Tager, G.E. Good, S. Brammer, ‘‘Walking over ‘Em”: an exploration of relations 1908
1823 [282] K. Schweizer, F. Brunner, K. Schutzmann, V. Schonbucher, H. Richter-Appelt, between emotion dysregulation, masculine norms, and intimate partner abuse 1909
1824 Gender identity and coping in female 46, XY adults with androgen in a clinical sample of men, Psychol. Men Masc. 11 (2010) 233–239. 1910
1825 biosynthesis deficiency (intersexuality/DSD), J. Couns. Psychol. 56 (2009) [312] T. Taketo, C.H. Lee, J. Zhang, Y. Li, C.Y. Lee, Y.F. Lau, Expression of SRY proteins 1911
1826 189–201. in both normal and sex-reversed XY fetal mouse gonads, Dev. Dyn. 233 1912
1827 [283] M.V. Seeman, Psychopathology in women and men: focus on female (2005) 612–622. 1913
1828 hormones, Am. J. Psychiatry 154 (1997) 1641–1647. [313] N. Taneja, A.C. Ammini, I. Mohapatra, S. Saxena, K. Kucheria, A transsexual 1914
1829 [284] N.L. Segal, Two monozygotic twin pairs discordant for female-to-male male with 47,XYY karyotype, Br. J. Psychiatry 161 (1992) 698–699. 1915
1830 transsexualism, Arch. Sex. Behav. 35 (2006) 347–358. [314] C.M. Temple, Oral fluency and narrative production in children with Turner’s 1916
1831 [285] S.N. Seidman, A.B. Araujo, S.P. Roose, J.B. McKinlay, Testosterone level, syndrome, Neuropsychologia 40 (2002) 1419–1427. 1917
1832 androgen receptor polymorphism, and depressive symptoms in middle-aged [315] A.J. Tilbrook, A.I. Turner, I.J. Clarke, Stress and reproduction: central 1918
1833 men, Biol. Psychiatry 50 (2001) 371–376. mechanisms and sex differences in non-rodent species, Stress 5 (2002) 83–100. 1919
1834 [286] R. Sekido, R. Lovell-Badge, Sex determination involves synergistic action [316] L.S. Tottenham, D.M. Saucier, Throwing accuracy during prism adaptation: 1920
1835 of SRY and SF1 on a specific Sox9 enhancer, Nature 453 (2008) 930– male advantage for throwing accuracy is independent of prism adaptation 1921
1836 934. rate, Percept. Mot. Skills 98 (2004) 1449–1455. 1922
1837 [287] L.I. Serova, S. Maharjan, A. Huang, D. Sun, G. Kaley, E.L. Sabban, Response of [317] Shinji Tsukahara, Sex differences and roles of sex steroids in apoptosis of 1923
1838 tyrosine hydroxylase and GTP cyclohydrolase I gene expression to estrogen in sexually dimorphic nuclei of preoptic area in postnatal rats, J. 1924
1839 brain catecholaminergic regions varies with mode of administration, Brain Neuroendocrinol. 9999 (2009). 1925
1840 Res. 1015 (2004) 1–8. [318] D. Twelves, K.S. Perkins, C. Counsell, Systematic review of incidence studies of 1926
1841 [288] B. Seymour, R. Dolan, Emotion, decision making, and the amygdala, Neuron Parkinson’s disease, Mov. Disord. 18 (2003) 19–31. 1927
1842 58 (2008) 662–671. [319] H. Ujike, K. Otani, M. Nakatsuka, K. Ishii, A. Sasaki, T. Oishi, T. Sato, Y. Okahisa, 1928
1843 [289] G.B. Sharman, The chromosomal basis of sex differentiation in marsupials, Y. Matsumoto, Y. Namba, Y. Kimata, S. Kuroda, Association study of gender 1929
1844 Aust. J. Zool. 37 (1990) 451–466. identity disorder and sex hormone-related genes, Prog. Neuropsycho- 1930
1845 [290] G. Shaw, M.B. Renfree, R.V. Short, W.S. O, Experimental manipulation of pharmacol. Biol. Psychiatry 33 (2009) 1241–1244. 1931
1846 sexual differentiation in wallaby pouch young treated with exogenous [320] M.T. Ullman, Contributions of memory circuits to language: the declarative/ 1932
1847 steroids, Development 104 (1988) 689–701. procedural model, Cognition 92 (2004) 231–270. 1933
1848 [291] B.B. Sherwin, Estrogen and cognitive functioning in women, Proc. Soc. Exp. [321] S.K. Van Den Eeden, C.M. Tanner, A.L. Bernstein, R.D. Fross, A. Leimpeter, D.A. 1934
1849 Biol. Med. 217 (1998) 17–22. Bloch, L.M. Nelson, Incidence of Parkinson’s disease: variation by age, gender, 1935
1850 [292] B.B. Sherwin, Estrogen and cognitive functioning in women, Endocr. Rev. 24 and race/ethnicity, Am. J. Epidemiol. 157 (2003) 1015–1022. 1936
1851 (2003) 133–151. [322] S. van Rijn, H. Swaab, A. Aleman, R.S. Kahn, X Chromosomal effects on social 1937
1852 [293] W.S.O.R.V. Short, M.B. Renfree, G. Shaw, Primary genetic control of somatic cognitive processing and emotion regulation: a study with Klinefelter men 1938
1853 sexual differentiation in a mammal, Nature 331 (1988) 716–717. (47,XXY), Schizophr. Res. 84 (2006) 194–203. 1939
1854 [294] P.J. Shughrue, P.J. Scrimo, I. Merchenthaler, Estrogen binding and estrogen [323] A.H. van Stegeren, R. Goekoop, W. Everaerd, P. Scheltens, F. Barkhof, J.P.A. 1940
1855 receptor characterization (ERalpha and ERbeta) in the cholinergic neurons of Kuijer, S.A.R.B. Rombouts, Noradrenaline mediates amygdala activation in 1941
1856 the rat basal forebrain, Neuroscience 96 (2000) 41–49. men and women during encoding of emotional material, NeuroImage 24 1942
1857 [295] R. Sibug, E. Kuppers, C. Beyer, S.C. Maxson, C. Pilgrim, I. Reisert, (2005) 898–909. 1943
1858 Genotype-dependent sex differentiation of dopaminergic neurons in [324] E. Vilain, Genetics of intersexuality, J. Gay Lesbian Psychother. 10 (2006) 9– 1944
1859 primary cultures of embryonic mouse brain, Brain Res. Dev. Brain Res. 26. 1945
1860 93 (1996) 136–142. [325] B. von der Pahlen, R. Lindman, T. Sarkola, H. Makisalo, C.J.P. Eriksson, An 1946
1861 [296] A. Siddiqui, B.H. Shah, Neonatal androgen manipulation differentially affects exploratory study on self-evaluated aggression and androgens in women, 1947
1862 the development of monoamine systems in rat cerebral cortex, amygdala and Aggress. Behav. 28 (2002) 273–280. 1948
1863 hypothalamus, Brain Res. Dev. Brain Res. 98 (1997) 247–252. [326] D. Voyer, S. Voyer, M.P. Bryden, Magnitude of sex differences in spatial 1949
1864 [297] L.W.S.R.B. Simerly, C. Chang, M. Muramatsu, Distribution of androgen and abilities: a meta-analysis and consideration of critical variables, Psychol. Bull. 1950
1865 estrogen receptor mRNA-containing cells in the rat brain: an in situ 117 (1995) 250–270. 1951
1866 hybridization study, J. Comp. Neurol. 294 (1990) 76–95. [327] J. Wade, Zebra finch sexual differentiation: the aromatization hypothesis 1952
1867 [298] J.W. Simpkins, P.S. Green, K.E. Gridley, M. Singh, N.C. de Fiebre, G. Rajakumar, revisited, Microsc. Res. Tech. 54 (2001) 354–363. 1953
1868 Role of estrogen replacement therapy in memory enhancement and the [328] J. Wade, A.P. Arnold, Functional testicular tissue does not masculinize 1954
1869 prevention of neuronal loss associated with Alzheimer’s disease, Am. J. Med. development of the zebra finch song system, Proc. Natl. Acad. Sci. USA 93 1955
1870 103 (1997) 19S–25S. (1996) 5264–5268. 1956

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001
YFRNE 451 No. of Pages 20, Model 5G
21 October 2010

20 T.C. Ngun et al. / Frontiers in Neuroendocrinology xxx (2010) xxx–xxx

1957 [329] D.L. Walker, D.J. Toufexis, M. Davis, Role of the bed nucleus of the stria [342] R.L. Worthington, A.L. Reynolds, Within-group differences in sexual 1988
1958 terminalis versus the amygdala in fear, stress, and anxiety, Eur. J. Pharmacol. orientation and identity, J. Couns. Psychol. 56 (2009) 44–55. 1989
1959 463 (2003) 199–216. [343] S. Wu, M. Jia, Y. Ruan, J. Liu, Y. Guo, M. Shuang, X. Gong, Y. Zhang, X. Yang, D. 1990
1960 [330] Z. Wang, N.A. Bullock, G.J. De Vries, Sexual differentiation of vasopressin Zhang, Positive association of the oxytocin receptor gene (OXTR) with autism 1991
1961 projections of the bed nucleus of the stria terminals and medial amygdaloid in the Chinese Han population, Biol. Psychiatry 58 (2005) 74–77. 1992
1962 nucleus in rats, Endocrinology 132 (1993) 2299–2306. [344] ..B. Wu, K. Chen, Y. Li, Y.-F.C. Lau, J.C. Shih, Regulation of monoamine oxidase 1993
1963 [331] T.H. Wassink, J. Piven, V.J. Vieland, J. Pietila, R.J. Goedken, S.E. Folstein, V.C. A by the SRY gene on the Y chromosome, FASEB J. (2009) (fj.09-139097). 1994
1964 Sheffield, Examination of AVPR1a as an autism susceptibility gene, Mol. [345] J. Xu, P.S. Burgoyne, A.P. Arnold, Sex differences in sex chromosome gene 1995
1965 Psychiatry 9 (2004) 968–972. expression in mouse brain, Hum. Mol. Genet. 11 (2002) 1409–1419. 1996
1966 [332] M.A. Weiss, Floppy SOX: mutual induced fit in HMG (high-mobility group) [346] T.L. Yaksh, Opioid receptor systems and the endorphins: a review of their 1997
1967 box-DNA recognition, Mol. Endocrinol. 15 (2001) 353–362. spinal organization, J. Neurosurg. 67 (1987) 157–176. 1998
1968 [333] L.A. Weiss, M. Abney, E.H. Cook Jr., C. Ober, Sex-specific genetic architecture [347] N. Yirmiya, C. Rosenberg, S. Levi, S. Salomon, C. Shulman, L. Nemanov, C. Dina, 1999
1969 of whole blood serotonin levels, Am. J. Hum. Genet. 76 (2005) 33–41. R.P. Ebstein, Association between the arginine vasopressin 1a receptor 2000
1970 [334] F.J. White, A behavioral/systems approach to the neuroscience of drug (AVPR1a) gene and autism in a family-based study: mediation by 2001
1971 addiction, J. Neurosci. 22 (2002) 3303–3305. socialization skills, Mol. Psychiatry 11 (2006) 488–494. 2002
1972 [335] L.S. Whitfield, R. Lovell-Badge, P.N. Goodfellow, Rapid sequence evolution of [348] L. Zhang, W. Ma, J.L. Barker, D.R. Rubinow, Sex differences in expression of 2003
1973 the mammalian sex-determining gene SRY, Nature 364 (1993) 713–715. serotonin receptors (subtypes 1A and 2A) in rat brain: a possible role of 2004
1974 [336] A.C. Wihlback, I.S. Poromaa, M. Bixo, P. Allard, T. Mjorndal, O. Spigset, Influence testosterone, Neuroscience 94 (1999) 251–259. 2005
1975 of menstrual cycle on platelet serotonin uptake site and serotonin(2A) receptor [349] D. Zhao, D. McBride, S. Nandi, H.A. McQueen, M.J. McGrew, P.M. Hocking, P.D. 2006
1976 binding, Psychoneuroendocrinology 29 (2004) 757–766. Lewis, H.M. Sang, M. Clinton, Somatic sex identity is cell autonomous in the 2007
1977 [337] J.D. Wilson, F.W. George, M.B. Renfree, The endocrine role in mammalian chicken, Nature 464 (2010) 237–242. 2008
1978 sexual differentiation, Recent Prog. Horm. Res. 50 (1995) 349–364. [350] J.N. Zhou, M.A. Hofman, L.J. Gooren, D.F. Swaab, A sex difference in the human 2009
1979 [338] S.F. Witelson, Glezer II, D.L. Kigar, Women have greater density of neurons in brain and its relation to transsexuality, Nature 378 (1995) 68–70. 2010
1980 posterior temporal cortex, J. Neurosci. 15 (1995) 3418–3428. [351] B. Zimmerberg, L.V. Scalzi, Commissural size in neonatal rats: effects of sex 2011
1981 [339] W. Wood, A.H. Eagly, A cross-cultural analysis of the behavior of women and and prenatal alcohol exposure, Int. J. Dev. Neurosci. 7 (1989) 81–86. 2012
1982 men: implications for the origins of sex differences, Psychol. Bull. 128 (2002) [352] J.-K. Zubieta, R.F. Dannals, J.J. Frost, Gender and age influences on human 2013
1983 699–727. brain mu-opioid receptor binding measured by PET, Am. J. Psychiatry 156 2014
1984 [340] S. Woody, P.D. McLean, S. Taylor, W.J. Koch, Treatment of major depression in (1999) 842–848. 2015
1985 the context of panic disorder, J. Affect. Disord. 53 (1999) 163–174. [353] D.G. Zuloaga, D.A. Puts, C.L. Jordan, S.M. Breedlove, The role of androgen 2016
1986 [341] C.S. Woolley, Effects of estrogen in the CNS, Curr. Opin. Neurobiol. 9 (1999) receptors in the masculinization of brain and behavior: what we’ve learned 2017
1987 349–354. from the testicular feminization mutation, Horm. Behav. 53 (2008) 613–626. 2018
2019

Please cite this article in press as: T.C. Ngun et al., The genetics of sex differences in brain and behavior, Front. Neuroendocrinol. (2010), doi:10.1016/
j.yfrne.2010.10.001

Vous aimerez peut-être aussi