Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Dec 1.
Published in final edited form as: Nat Rev Genet. 2008 Dec;9(12):911–922. doi: 10.1038/nrg2415

Sex-Specific Genetic Architecture of Human Disease

Carole Ober 1, Dagan A Loisel 1, Yoav Gilad 1
PMCID: PMC2694620  NIHMSID: NIHMS111210  PMID: 19002143

Abstract

Sexual dimorphism in anatomical, physiological, and behavioural traits characterize many vertebrate species. In humans, sexual dimorphism is also observed in the prevalence, course, and severity of many common diseases, including cardiovascular diseases, autoimmune diseases, and asthma. Although sex differences in the endocrine and immune systems probably contribute to these observations, recent studies suggest that sex-specific genetic architecture also influences human phenotypes, including reproductive, physiological, and disease traits. It is likely that an underlying mechanism is differential gene regulation in males and females, particularly in sex steroid responsive genes. Genetic studies that ignore sex-specific effects in their design and interpretation could fail to identify a significant proportion of the genes that contribute to risk for complex diseases.


Differences between males and females in anatomical, physiological, and behavioral traits characterize many vertebrate species, including humans. Although some may be apparent at birth, striking differences between the sexes most often emerge at or around the time of sexual maturation. It is thought that these are, in large part, due to sex hormone levels that differ in males and females beginning in utero and continuing throughout life1 (Figure 1). The genetic contribution to sexual dimorphism was, until recently, less studied. Indeed, whereas genes on sex chromosomes contribute to many sexually dimorphic traits, the autosomal genome is generally assumed to be similar among the males and females of a species. Mechanisms for dosage compensation in heterogametic species further assure that genetic contributions from the shared sex chromosome (X chromosome in mammals) is equivalent among males and females, at least for most genes2.

Figure 1. Approximate mean sex steroid levels in plasma in males and females.

Figure 1

Variation in steroid levels is shown as percent of the maximum mean testosterone (T) in males and the maximum mean estradiol (E) in females across the life stages. The figure does not show diurnal, cyclic (female), or possible seasonal fluctuations. Female estradiol levels refer to the mean for the mid-follicular phase of the menstrual cycle; estradiol production transiently increases about 5-fold during the pre-ovulatory and luteal phases of the menstrual cycle. Note the drop in levels of all sex steroids at birth and the transient ’minipuberty’ in early infancy. Free testosterone falls more with aging (to approximately 50% of the maximum in 80 year old men) than the total testosterone134, which is shown here. Modified from Alonso and Rosenfield 20021, Khosla et al. 1998135, Winters et al. 2000136.

Recent studies have challenged this paradigm, however, suggesting that natural variation within the autosomal genomes of many species also affects anatomical, physiological, and behavioral traits differently in males and females3-5. In this context, sex can be considered an ‘environmental’ variable that includes the cellular, metabolic, physiological, anatomical, and even behavioral differences between boys and girls (in childhood) or between men and women (in adulthood). Sex, then, may interact with genotype in a manner similar to other environmental factors (Figure 2). However, unlike most other environmental factors, sex is easily observable and (usually) unambiguous. Such sex-specific genetic architecture suggests new models of susceptibility for common diseases and sheds light on potential mechanisms of sexual dimorphism [Box 1] in human phenotypes.

Figure 2. Models of genotype-sex interactions reflecting genotype effects that differ between males and females. [I'll arrange for ‘B’ to be changed to ‘a’ in this figure].

Figure 2

For any measured phenotype or disease risk (y axes), the genotypic effects may be apparent only in females (red; panels a, d), only in males (blue; panels b, e), or be present in both sexes but with opposite directions of effects (panels c, f). The genotype effects can be additive (panels a-c) or recessive (panels d-f). Other models (e.g., dominant) or interactions (e.g., same direction of effect but differences in magnitude of effect) are not shown. Examples discussed in this review illustrate panel e (relationship between the DD genotype of the angiotensinogen converting enzyme (ACE) and hypertension), panel b (relationship between the DD genotype of ACE and blood pressure), panel d (relationship between the reelin (RELN) rs7341475-GG genotype and schizophrenia), and panel c (relationship between chromosome 4p16.3 SNPs rs3796619 and rs1670533SNPs and recombination rate). Red lines track phenotypic values by genotype in females; blue lines track phenotypic values by genotype in males.

In this review, we argue that sex-specific genetic architecture is common in humans and that genotype–sex interactions contribute to differences in the prevalence, course, and severity of diseases as well as to other quantitative phenotypes. We provide recent examples of genotype–sex interactions as evidence to support this argument and illustrate how patterns of tissue-specific gene expression differ markedly between males and females. Lastly, we discuss the importance of considering sex in the design and analysis of genetic studies.

Evidence of Sex Effects

Accumulating evidence suggests that nearly all human diseases have sex-specific differences in prevalence, age of onset, and/or severity. Classic examples include the predominance of men with cardiovascular disease throughout adult life but a higher rate of occurrence in post-menopausal women compared to men6, the higher prevalence of asthma among boys in childhood and higher occurrence of new cases among girls around and following puberty7, and the increased prevalence of autoimmune diseases in women throughout life but particularly for diseases that onset during or immediately following the reproductive years8 (Figure 3). In addition to those diseases highlighted in Figure 3, significant sex differences have been described for many common birth defects, neurological and psychiatric disorders, as well as for some common cancers. For example, in infancy or childhood, neural tube defects, congenital dislocation of the hip, and scoliosis are more common among girls whereas autism, stuttering, and pyloric stenosis are more common among boys9. In adulthood, major depression and Alzheimer disease are more common in women10,11 whereas schizophrenia, Parkinson disease, and colorectal cancer are more common in men12-14.

Figure 3. Sex-specific prevalence rates, age of onset, and sex ratios for common sex-skewed diseases.

Figure 3

The key for background colors is shown in Figure 1 [I'll ask for the key to be copied to this figure]. a) Cardiovascular disease in the U.S. (from the National Health and Nutrition Examination Survey (NHANES) III 1988−1994)6. Note the increase in female prevalence rates in the post-menopausal period. b) Asthma in the U.S. from 1998−2006 (Center for Disease Control National Health Interview Survey (CDC NHIS)). Note the increase in female prevalence rates during and following puberty. c) Sex ratios (%female) by mean or median age of onset for autoimmune diseases in the U.S. and Europe137,138. Note the female skewing at all ages, with the largest skew and number of diseases with onset during and immediately following the reproductive years. T1D, type 1 diabetes; JIA, juvenile idiopathic arthritis; JDM, juvenile dermatomyositis; MS, multiple sclerosis; MG, myasthenia gravis; GD, Grave's disease; SLE, systemic lupus erythematosus; SSc, systemic sclerosis (scleroderma); AD, Addison disease; DM, dermatomyositis/polymyositis; RA, rheumatoid arthritis; TH, thyroiditis; SS, Sjögren's disease.

It should be noted that differences in prevalence rates or age of onset do not necessarily imply that genetic variation leads to different effects in males and females15, as many of these differences could be due to hormonal profiles, particularly with regard to sex steroids (Figure 1), or to behaviors that differ between the sexes (e.g., exposure to cigarette smoke)16. For example, the consistent associations between increased risk for disease among females during and following puberty (asthma), during the reproductive years (autoimmune disease), or post-menopausal (cardiovascular disease) have implicated sex hormones as important mediators of disease pathogenesis and contributors to sex differences in prevalence rates and progression.

Importantly, differences in the immune systems of males and females have been observed as early as in the first few years of life, suggesting a developmental component to sex-specific differences in disease risk17. Such differences could result in sex-specific thresholds of susceptibility to immune-mediated diseases throughout life. Interestingly, immune responses may be modulated by sex hormones18,19. In fact, the transient rise in sex steroid levels (‘minipuberty’) that occurs in early infancy1 (Figure 1), could pattern immune cells differently in boys and girls. Thus, both the immune and endocrine systems likely contribute to sexual dimorphism in the epidemiology of many common diseases. However, recent evidence suggests that some of the differences between males and females may also be due to differences in genetic architecture. The review henceforth will focus on such sex-specific genetic effects.

Sex Effects on Disease Risk through Gene Regulation

Contribution of sex chromosomes

The contributions of the sex chromosomes to sex-specific genetic architecture of human disease has long been appreciated. For example, an excess of boys express X-chromosome-linked recessive diseases, and skewed patterns of X chromosome inactivation resulting in varied expression of disease phenotypes are seen in female carriers of X-linked mutations20. More generally, dosage differences in X-linked genes between the sexes probably account for some of the sex-specific genetic architecture of common diseases and phenotypes. In turn, the Y chromosome in males harbors relatively few genes, most of which are expressed exclusively in the testes and others that are typically thought of as ‘housekeeping’ genes (of the latter, most have X chromosome homologues that escape X inactivation)21. Thus, it is perhaps unlikely that Y-linked genes per se directly affect disease risk, other than constituting major contributors to genetic causes of male infertility22. However, Y-linked genes may interact with autosomal genes to differentially affect disease risk in males and females.

Contribution of autosomes

In contrast to the sex chromosomes, the autosomal genome is shared by both sexes. However, although the DNA sequence, gene structure, and frequency of polymorphisms on the autosomes do not differ between males and females, the regulatory genome is sexually dimorphic23-26. That is, sex-specific differences in gene regulation (rather than gene content) probably underlie most phenotypic sexual dimorphism, including sex-specific effects on human diseases. Indeed, at the mRNA level, sexually dimorphic gene expression has been observed in a wide range of organisms, including worms27, flies28,29, fish30, rodents 25,31, and primates23. Although genes with sex-biased expression are enriched on the sex chromosomes, thousands of sex-biased genes are also found on the autosomes.

Sexually dimorphic gene expression patterns are conserved

Interestingly, genes with sex-biased expression patterns tend to evolve rapidly at the coding region level26. This observation is consistent with the notion that many differences in gene expression between the sexes are the result of sexual selection (Box 1). The evolution of sex-biased genes was recently reviewed by Ellegren and Parsch26 and will not be discussed in detail here. However, it is relevant to note that although sex-biased genes often evolve rapidly at the protein coding level, differences in gene regulation between the sexes are often conserved in evolution. For example, Zhang and colleagues showed that sexually dimorphic expression patterns of a large number of genes are conserved across seven Drosophila species32. Similarly, Reinius et al. found a signature of evolutionary conserved sexually dimorphic gene expression in the brain of three primate species, including humans23. Specifically, they compared gene expression profiles in the occipital cortex of male and female humans and cynomolgus macaques (Macaca fascicularis), and identified hundreds of genes with sex-biased expression patterns in both species.

Phenotypic consequences of sexually dimorphic gene expression patterns

The observations of conserved sex-specific regulation suggest that at least a subset of the sexual dimorphism in gene expression underlie important phenotypic differences (developmental, physiological and/or behavioural) between the sexes. These conserved sexually dimorphic gene expression patterns suggest the existence of constant regulatory differences between males and females, which may be beneficial to each sex but can also contribute to different gene–environment interactions in the two sexes. In turn, such differences may result in sex-specific susceptibility to disease. For example, potential sexual dimorphism in the regulation of oxidative stress response pathways could differentially affect susceptibility to cardiovascular diseases in males and females33.

A second interesting observation is that sexually dimorphic gene expression patterns are often tissue-specific25, whereby a gene may be differentially expressed between the sexes in some but not in other tissues. This important observation suggests that a different architecture of regulatory interactions may underlie gene expression patterns in males and females in different tissues. Hence, it is likely that entire regulatory networks may differ between the sexes, interacting with functional genetic variation (such as expression quantitative traits) in a sex-specific manner. Such differences in gene regulation between the sexes may account for genotype–sex interactions that affect other measurable phenotypes as well as disease risk. A clear example of a sex-specific response to an environmental variable was recently provided by Zammaretti et al., who investigated the effects of long-term moderate/high fat diet on mice. They found phenotypic differences between males and females, including differences in gene regulation, following the application of identical diet in the two sexes34.

Bhasin et al. provided additional support for this hypothesis by mapping sex-specific expression qtl (eQTL) in mice35. They identified SNPs in putative cis regulatory elements that were associated with variation in gene expression within individuals from one sex, but not the other, indicating that some loci have a regulatory role in males but not in females, or in females but not in males. Because the SNPs are shared among the sexes, differences in the use of cis regulatory elements between the sexes indicate sex-specific differences in trans elements (e.g., transcription factors and co-factors). Sex steroid receptors may be one example of sex-specific trans regulatory elements 36. Similar analyses of human eQTL data have not been performed to date, yet the findings of Bhasin et al.35 are consistent with a growing number of observations23,25,28,30,32 suggesting that ignoring sex in studies of gene expression will underestimate, perhaps quite dramatically, the affect of genetic variations on gene regulation and mRNA abundance.

Genetic mechanisms other than gene regulation may also contribute to sex-specific disease risk or sexual dimorphism in quantitative phenotypes (Box 2, Box 3). However, regardless of the mechanism, abundant evidence now exists for a significant role of sex-specific genetic architecture.

Evidence of Sex-Specific Genetic Architecture in Humans

Estimating heritability

One way to estimate the relative contribution of genes to a trait is through ‘variance component analysis’ in related individuals. In this approach, the total variance in a quantitative, or measured, phenotype is divided into its genetic and environmental components. The proportion of the total phenotypic variance attributed to genetic factors (i.e., genetic differences between individuals) is referred to as the heritability of the trait. The genetic variance can be further divided into the variance due to additive genetic effects, to non-additive genetic effects (e.g., dominance, recessiveness, epistasis), as well as be assigned to autosomes or sex chromosomes. The proportion of the variance due to additive genetic effects is referred to as narrow heritability (h2); the overall proportion of genetic variance is referred to as broad heritability (H2). The theoretical basis for heritability estimates and derivation of the individual variance components has recently been reviewed37. The heritabilities of many human traits have been estimated, although most studies are limited to estimates of narrow heritabilities in combined samples of males and females (for examples, see refs.38-40).

Sex-specific genetic architecture of human quantitative traits: a case study

Recently, the sex-specific genetic architecture of 19 human quantitative traits in males and females, many of which are associated with common diseases, was investigated in a large multigenerational pedigree comprised of >500 members of the Hutterites, a founder population that practices a communal lifestyle41,42. Because of the remarkably uniform environment and lifestyle between individuals of both sexes in this community, the authors argued that sex-specific genetic architecture might be easier to detect. For example, smoking is prohibited and rare, meals are eaten and prepared in a communal kitchen, and large families desired43. Moreover, because all relative pairs in the extended pedigree are considered in the analysis, it was possible to estimate both additive and dominance variance components44.

In this population, sex was a significant predictor of the trait value in a linear regression model for 16 of the 19 phenotypes, including cardiovascular disease-associated traits (HDL cholesterol, lipoprotein[a], triglycerides, diastolic and systolic blood pressure), asthma-associated traits (forced expiratory flow at 1 second, (FEV1], the ratio of FEV1 to forced vital capacity [FVC], eosinophil count, total serum IgE level, percent lymphocytes), anthropometrics (body mass index, percent fat, fat free mass, adult height), and signaling molecules (morning serum cortisol, whole blood serotonin). Sex was not a significant predictor of three phenotypes (LDL cholesterol, lymphocyte count, fasting insulin).

The narrow and broad heritabilities of each of these traits were estimated in a unified model. Five traits had significant X chromosome variance components either in males only (systolic blood pressure, adult height, triglycerides) or in both sexes (lipoprotein[a], whole blood serotonin). Interestingly, four traits had significant non-X sex interactions in which either the estimates of heritability were significantly different between males and females (LDL-cholesterol, FEV1:FVC) or the best-fitting heritability model was different between males and females (HDL-cholesterol, fat free mass). Thus, the genetic architecture of nine (of 19) common phenotypes had significant sex-specific genetic architecture. The best-fitting heritability model for six representative traits with sex-specific architecture is shown separately for males and females in Figure 4.

Figure 4. Sex-specific heritabilities in males and females (data from Pan et al. 200741).

Figure 4

Six quantitative traits with significant sex-specific genetic architecture show differences between males and females in the overall estimates of H2 (e.g., LDL cholesterol, lipoprotein[a], systolic blood pressure) and/or with respect to the best-fitting model (triglycerides, HDL cholesterol, systolic blood pressure, height) are shown.

Taken together, these data suggest that the genetic architecture (additive, dominant, X-linked) and/or the overall genetic contribution (heritability) significantly differs between males and females for a large number of quantitative phenotypes, many of which are risk factors for common diseases, consistent with other studies of sex-specific heritabilities of common disease-associated quantitative phenotypes45,46. Although this data set is limited to only 19 quantitative traits, it further suggests that X chromosome genes may contribute disproportionately more to common phenotypes and quantitative trait variation in males than in females, not unlike Mendelian disease genes. Indeed, subsequent studies supported these conclusions, demonstrating significant sex differences in estimates of the autosomal narrow heritability for 13 (of 539) cardiovascular disease associated quantitative traits in French Canadian families45, and for bone mineral density in a number of recent studies (reviewed in Karasik and Ferrari46).

Thus, standing natural variation in the human genome contributes to quantitative phenotypes in a sex-specific manner. That many of these phenotypes are also risk factors for common diseases further suggests that significant sex-specific genetic architecture contributes to risk for common diseases.

Accumulating Evidence for Genotype-Sex Interactions

Demonstrating genotype–sex interaction effects on human diseases has been challenging because, until recently, most study designs did not allow a systematic search for sex-specific genetic contribution to quantitative variation or disease risk47. Moreover, in most linkage and association studies that address sex-specific architecture, analyses are performed in each sex separately (usually in addition to studies in the combined sample), adding to the number of statistical tests and increasing the likelihood of a type i error if multiple testing is not properly taken into account when assessing significance. On the other hand, the roughly halving of the sample size to conduct sex-specific analysis reduces the power to detect an effect. For example, a study with 80% power for a main effect will have only 29% power to detect an interaction of the same magnitude48, making replication of genotype–sex interactions particularly challenging.

It is, therefore, not surprising that a recent meta-analysis of 188 genetic association studies claiming sex effects in their title found only one association that was consistently replicated in at least two studies15. Among 188 claims of a sex difference, 83 were significant (P<0.05), although 44 of those had modest p-values between 0.01 and 0.05 (unadjusted for multiple testing). Sixty of those claims were judged to have good internal validity, including the one association that was replicated. This was the association between the deletion/insertion (D/I) polymorphism in the angiotensinogen converting enzyme (ACE) gene with hypertension in men only49-52 (discussed below).

Despite these limitations, a number of recent studies suggest the importance of genotype–sex interactions in the genetic architecture of quantitative phenotypes and common diseases, which should motivate the development of robust methods for both assessing and routine testing of genotype–sex interactions in genetic studies. It should be noted, however, that while many linkage studies have reported sex effects, only few have shown that increased lod scores in one sex are not due to chance findings resulting from splitting samples and performing multiple tests, or have explicitly tested for genotype– sex interactions (see refs.45,53 for exceptions). As a result, linkage studies will not be reviewed here. Instead, we first review evidence for genotype–sex interactions in model organisms, and then highlight three recent examples of genotype–sex interactions in human association studies.

Genotype-Sex Interaction Effects in Model Systems

The most compelling and consistent evidence for genotype–sex interaction effects comes from studies of physiological, anatomical, and behavioral traits in model organisms, including fruitflies54,55, mice56-59, and rats60. For example, sex-specific effects in which QTLs have significantly different effects in males and females are a near-ubiquitous characteristic of the genetic architecture of complex traits in the Drosophila genus (reviewed in Mackay and Anholt55).

In mice, studies of sex specific effects include alcohol preference, which in the C57BL/6 strain has been shown to be nearly entirely controlled by sex-specific effects56,57. Other examples include knocking out the cytochrome P2J5 gene in C57BL/6 mice, and the resulting sex-specific effects on blood pressure and renal phenotypes58, and studies of consomic strains of mice that revealed sex-specific effects of individual chromosomes on fear conditioning59. Similarly, in consomic rat strains, sex–specific effects on phenotypes related to hypertension and kidney disease predominate60.

It should be noted that many of the mapping studies claiming sex effects in model organisms suffer from the same limitations as those described above for human studies. However, the experimental toolbox available for studies of model organisms allows for a more thorough dissection of sex-specific genetic architecture, which has, in many cases, directly implicated specific genes or chromosomes in genotype–sex interactions. Overall, genotype–sex interaction effects on diverse biological processes are common in model organisms and often account for a significant proportion of the phenotypic variability. The extent of sex-specific genetic architecture in the human genome has yet to be determined, although we predict that humans are similar to other organisms in this respect.

Genotype-Sex Interaction Effects in Humans

Example 1: Hypertension and Blood Pressure

Hypertension is a major risk factor for cardiovascular disease, stroke, and end-stage renal disease61. In 2005 the prevalence of hypertension of the adult population worldwide was 26% 62. Blood pressure is higher in men compared to women among adults under the age of 45, but this trend switches and at 70−79 years of age women have higher blood pressure than men63,64, similar to overall trends for cardiovascular disease (Figure 1). Genes involved in the renin-angiotensinogen system are functional candidates for blood pressure regulation and hypertension, and have been associated with these phenotypes with varying success (reviewed in Kato et al.65). A 250 bp deletion/insertion (D/I) polymorphism in intron 16 of the angiotensin converting enzyme (ACE) gene accounts for approximately 47% of the variance in plasma ACE protein levels, with each copy of the D allele associated with an approximately 30% increase in ACE levels66. ACE was considered a candidate gene for blood pressure and hypertension, but results of case–control association studies with blood pressure were conflicting and family-based studies failed to demonstrate linkage between the ACE locus and hypertension (ref.50 and references therein).

Studies in a rat model suggested genotype–sex interactions67. Both male and female rats heterozygous for an inactivating mutation in Ace had lower Ace protein levels compared to wild type animals (23% reduction in males and 35% reduction in females). However, only heterozygous males had a reduced blood pressure compared to the wild type males; heterozygous females had blood pressures similar to wild type females. Therefore, low Ace levels due to an inactivating mutation in the Ace gene did not affect blood pressure in female rats but protected against hypertension in male rats. The authors suggested that interactions with sex should be evaluated in genetic studies of the human ACE gene. Indeed, subsequent studies in humans have replicated this interaction49-51 (Table 1).

Table 1. ACE D/I genotype–sex interaction on hypertension.

In three independent studies, the D allele at the angiotensin converting enzyme (ACE) locus was associated with risk for hypertension in men but not in women. Odd ratios (ORs) and confidence intervals (CIs) from multivariate model adjusted for other covariates.

Sample Size (Cases/Controls) OR (95% CI) Relative to Genotype II
Sample DD DI Reference
Men
U.S. Caucasian 689/755 1.59 (1.13, 2,23) 1.18 (0.87,1.62) O'Donnell50
Japanese 604/1736 1.75 (1.21, 2.53) 1.14 (0.87, 1.51) Higaki49
Serbian 98/112 +2.05 (1.07, 3.91) NA Stanković51
Women
U.S. Caucasian 705/945 1.00 (0.70, 1.44) 0.78 (0.56, 1.09) O'Donnell50
Japanese 596/2079 1.17 (0.79, 1.72) 0.87 (0.65, 1.17) Higaki49
Serbian 77/98 +0.72 (0.33, 1.60) NA Stanković51
+

Relative to genotype II.

NA, information not available.

Collectively these studies provide convincing evidence for an ACE genotype–sex interaction effect on hypertension and possibly on blood pressure, although the mechanisms for these effects are still unknown. Moreover, these studies demonstrate that even in the absence of a genotype–sex interaction in quantitative trait variation (in this example, ACE protein levels66), a genotype–sex interaction can still occur with respect to an associated physiological trait (e.g., blood pressure)50,52 and a disease phenotype (e.g., hypertension)49-51. Lastly, these studies further provide an example in which the genetic model underlying the interaction can differ between the physiological trait and the disease: in males, the effect of the D allele of ACE is additive on blood pressure, but recessive on hypertension (models B and E, respectively, in Figure 2), suggesting a quantitative (blood pressure) threshold effect for expression, or penetrance, of a common disease (hypertension) that is sex-specific.

Example 2: Schizophrenia

Schizophrenia is a common psychiatric disorder with significant sex differences in prevalence, age of onset and morbidity68. For example, most cases occur between the ages of 16 and 25 years in men and between the ages of 25 and 30 years among women. Overall, the male to female sex ratio is 1.412,69. Estimates of heritability for this complex disease is approximately 0.8070, indicating that a significant proportion of disease risk is attributable to genetic variation. A number of sex-specific genetic associations with schizophrenia risk have been reported, but none has been consistently replicated15.

Shifman and colleagues conducted a genome-wide association study for schizophrenia using a novel DNA pooling strategy71. One hundred ninety four SNPs were selecting for further studies based on their ranking and statistical significance in the studies in pooled DNA, and their biological plausibility71. These SNPs were then individually typed in 745 patients and 759 controls from the Ashkenazi Jewish population. The smallest P-value corresponded to SNP rs7341475 (G→A), for which the frequency of the GG genotype was 0.76 in female patients compared to 0.59 in female controls (P = 9.8 × 10−5). There was no association in males (P = 0.47), yielding a significant genotype-sex interaction (P = 0.0053) (Table 2). This SNP is located on chromosome 7 in intron 4 of the Reelin gene (RELN), which had previously been studied as a candidate for schizophrenia or related phenotypes (Ref72 and references therein). In the Ashkenazi Jewish sample, rs7341475 showed high linkage disequilibrium (LD) with other SNPs in the third and fourth intron of the RELN gene, but the LD did not extend to neighboring genes, suggesting that the association with schizophrenia is with variation in the RELN gene.

Table 2. Genotype-sex interaction effects of the Reelin SNP rs75341475 (G→A) on schizophrenia.

Pinteraction for all samples combined = 1.6 × 10−5 (from ref.71).

Sample Sample Size (Cases/Controls) Freq. GG (Cases/Controls OR* (95% CI) GG Relative to GA+AA
Men
Ashkenazi 470/1988 0.606/0.619 0.95 (0.77,1.17)
U.K. 320/1439 0.709/0.725 0.93 (0.71, 1.21)
U.S. 295/202 0.692/0.698 0.97 (0.66, 1.43)
Irish 669/337 0.750/0.733 1.10 (0.81, 1.48)
Chinese 222/229 0.806/0.830 0.85 (0.53, 1.38)
Combined 1976/4195 -- 0.96 (0.85, 1.10)
Women
Ashkenazi 265/656 0.755/0.610 1.97 (1.43, 2.71)
U.K. 155/1488 0.813/0.702 1.85 (1.22, 2.81)
U.S. 109/232 0.725/0.638 1.50 (0.91, 2.46)
Irish 311/245 0.762/0.731 1.18 (0.80, 1.73)
Chinese 193/229 0.845/0.825 1.15 (0.69, 1.93)
Combined 1033/2850 -- 1.58 (1.31−1.89)

To confirm that rs7341475 is a female-specific risk factor for schizophrenia, the investigators assessed whether the GG genotype was increased in women with schizophrenia in four other samples from the U.K., U.S., Ireland, and China The predicted direction of effect was present in all the samples, but differences were only significant in the U.K. sample (Table 2). In the combined samples (with and without the primary Ashkenazi Jewish sample), the recessive (GG) genotype was a significant risk factor for schizophrenia in females only (Figure 2, panel D).

Although the association with rs7341475 did not meet criteria for genome-wide significance (i.e., corrected for multiple testing), the supportive data from four replication samples and the biological plausibility of the involvement of RELN in brain abnormalities73 make these results particularly intriguing. However, mechanistic studies demonstrating functionality of the associated intronic SNP, or a SNP in LD with rs7341475, are still needed. Interestingly, however, higher expression of the RELN gene (in layer I neurons) in women compared to men and a reduction of RELN expression (in the superficial interstitial white matter neurons) in men with schizophrenia but not in females with schizophrenia74, suggests sex-specific gene regulation. Whether the schizophrenia-associated variation is also associated RELN expression differences in women remains to be determined.

Example 3: Recombination Rate

Meiotic recombination is one of the most fundamental biological mechanisms to ensure normal embryonic development. Because too few recombination events can result in nondisjunction and aneuploidy, and ectopic exchange can result in chromosomal rearrangements75,76, it is likely that this process is highly regulated77. Recently, the rate of recombination78 and location of recombination79 were shown to be heritable phenotypes in human pedigrees.

Recombination rate itself is a sexually dimorphic trait, with overall higher rates in female germ cells in humans, except at the telomeres of chromosomes where male recombination rates exceed those of females80,81. A recent genome-wide association study of recombination rates in 1887 Icelandic men and 1702 Icelandic women identified a locus that showed significant sex-specific effects78. Three SNPs in a block of LD spanning 200 kb on chromosome 4p16.3 showed genome-wide significant evidence of association in men (P<10−10) and two of those SNPs were also genome-wide significant in women (P<10−7). Surprisingly, the combination of alleles associated with low recombination rates in men (allele C at rs3796619 and allele T at rs1670533) were associated with high recombination rates in women. The opposite effect of these SNPs on male versus female recombination was replicated in a second sample of 3135 men and 3365 women from Iceland (P<10−8 in men and P<10−4 in women). Relative to the average recombination rate in the population, each copy of rs3796619 decreased recombination rate by 2.62% in men whereas each copy of the rs1670533 T allele increased recombination by 1.8% in women. The former allele explained 3.5% of the variance in recombination rate in men and the latter allele explained 1.7% of the recombination rate in women.

The associated SNPs were in an ‘LD block’ that included two genes, spondin 2 (SPON2) and ring finger protein 212 (RNF212). The authors suggested that RNF212 is an excellent candidate for a human recombination gene because it is homologous to a gene involved in recombination in yeast, although further studies are required to determine which SNP and which gene influence recombination rates as well as the exact mechanism for the sex-specific effect. Nonetheless, these results illustrate a genotype-sex interaction of alleles with additive and opposite effects in males and females (Figure 2, Panel C). Loci with this type of genotype–sex interaction effects would never be detected in a genome-wide association study in a combined sample of men and women, where the opposite nature of the association in the two sexes would cancel out any observable effect in combined samples, similar to other genotype–environment interactions54,82-84.

Summary and Future Directions

Significant sexual dimorphism in prevalence, age of onset, severity, or genetic risk is observed for most common human diseases. Elucidating the underlying mechanisms for these observations remains challenging, but represents an important area for future research. Because it is unlikely that sexually dimorphic traits are due to differences in the structure of genes in males and females (with the possible exception of genes on the Y chromosome), the importance of the regulatory genome in this context becomes central to understanding mechanism. Standing variation in regulatory elements that contribute to sexually dimorphic traits could result in sex-specific gene–environment interactions. In addition, sexually dimorphic developmental processes, such as sex-specific changes in gene regulation with age85, can result in shifting differences in disease susceptibility between the sexes, for example, as has been observed for asthma (Figure 3).

To date, studies of genotype–sex interactions have interrogated genetic associations that have different effects in males and females on physiological or disease traits. These studies have had varying success, as discussed above15. Even among the more prominent examples discussed in this review, it has not yet been possible to relate the associated polymorphisms to sex-specific differences in the regulation of gene expression. Moreover, a large number of studies of diseases or QTLs in families have reported sex–specific linkages. However, as mentioned above, few have demonstrated that differences in lod scores between males and females are significant or tested directly for interactions. On the other hand, animal model studies suggest that genotype–sex interactions are widespread and that many important genes will be missed if such interactions are ignored. In that context, we favor testing for genotype–sex interactions in association studies, particularly for sexually dimorphic phenotypes, although appropriate significance testing is required to avoid type I errors. For example, it is striking that no pharmacogenetic study to date has looked for genotype–sex interactions on drug response, although sex-specific responses to drugs are well known86-88, or that genotype– sex interactions on development has not been explored.

An alternative approach for discovering genotype–sex interactions, in particular in the context of gene regulation, is to directly study gene expression as a quantitative phenotype, and identify genetic variation that is associated with expression levels differently in males and females. Because thousands of gene expression phenotypes can be measured simultaneously, it is likely that genotype–sex interaction effects on gene expression will be easier to detect than studies of physiological and disease traits, and that all types of interactions will be present (e.g., Figure 2). With the availability of many data sets with both dense SNP typing and measurements of global gene expression in the same individuals89-92, it should be possible to directly assess sex-specific genotype effects on heritable variation in mRNA abundance using eqtl mapping approaches90,93. Moreover, because the ultimate goal of eQTL mapping is to identify regulatory variation that results in physiological or disease phenotypes94, this approach can be extended to study the sex-specific architecture of these phenotypes (Figure 5). Traits or diseases with sex-specific genetic architecture, such as those shown in Figures 3 and 4, would be excellent candidates for these studies.

Figure 5. Strategy for discovering sex-specific eQTLs contributing to sexual dimorphism in disease risk.

Figure 5

Red symbols are results for females and blue symbols are results for males. a) Results of genome-wide association study for a disease-associated QTL (such as those shown in Figure 4). Analyses in sex-stratified samples identify an association with SNPs spanning a 50 kb region in females but not in males. b) mRNA expression level by genotype. Using publicly available expression data89-92, eQTL that reside within the 50 kb region with sex-specific effects on expression levels can be identified. Each copy of the T allele at this eQTL is associated with increased expression in females but has no effect on expression in males (Figure 2a). c) Odds ratios for disease risk by genotype. Validation of a role for the eQTL on disease risk is determined by directly demonstrating a genotype-specific risk for disease in one sex only, in a direction that is consistent with the patterns observed with the associated QTL and eQTL. In this example, each copy of the T allele is associated with increased risk for disease in females. The SNP is not associated with disease risk in males. This model of association is also represented in Figure 2a.

Understanding genotype-sex interactions at the level of gene expression would not only shed light on mechanism, but may also identify “signatures” for variations that participate in sex-specific gene regulation. Such knowledge may also inform studies of physiological and disease traits by allowing variants to be categorized as more or less likely to participate in the sex-specific regulatory genome, and by identifying genes that are differentially regulated as candidates for sexually dimorphic traits.

Box 1. Sexual Dimorphism.

Following Darwin's (1859) observation that males and females may have the same “general habits of life” but “differ in structure, color, or ornament”95, research on sexual dimorphism progressed gradually from qualitative descriptions of conspicuous anatomical and behavioral traits in animals96 to elegant experiments probing the sex-specific neural circuitry of reproductive behavior in flies97,98 and mice99. The results of this century-and-a-half of research demonstrated that sexual dimorphism was taxonomically widespread and remarkably variable in the magnitude and form of its expression100,101. It is now quite obvious that sex-specific differences occur not only in conspicuous morphological traits (i.e. size, shape, and coloration) but also in a diverse suite of behavioral97-99,102, psychological102,103, biochemical69,103, and gene expression23,24,26 phenotypes.

Variation in the magnitude of sexual dimorphism among closely related species, and sometimes within a species, motivated biologists to test Darwin's (1859) hypothesis that sex-specific differences were largely due to sexual selection, particularly male–male competition, in dozens of different taxa101. The results of these studies consistently reaffirmed the importance of sexual selection (via male–male competition and/or female choice) as a major driver of sexual dimorphism, but also suggested a significant role for natural selection and non-selective forces, i.e. genetic, ecological, and developmental pressures and constraints, in the evolution of sex-specific phenotypic divergence100,104. Indeed, future research into the nature and consequences of intersexual genetic correlations105 and intersexual ontogenetic conflict106 will lead to a more sophisticated understanding of the evolution and expression of sexual dimorphism.

Box 2. Microchimerism and Disease.

As a result of bi-directional cell trafficking between the mother and fetus during pregnancy, mothers may harbor cells from their children and children may harbor cells from their mother will into adulthood. This mixture of a small amount of cells from a genetically disparate individual is referred to as microchimerism107. The persistence of maternal cells in her children, called maternal microchimerism, has been detected in the peripheral blood mononuclear cells in approximately 22% of healthy individuals108-111. The persistence of fetal cells in the mother is called fetal microchimerism, which has been detected in peripheral blood mononuclear cells in 30% to 55% of healthy women, depending on the outcome of the pregnancy112. Maternal microchimerism is found less often than fetal microchimerism in unselected peripheral blood mononuclear cells as well as in cellular subsets, such as T and B lymphocytes, monocyte/macrophages, and natural killer cells113. Moreover, microchimerism has been found in many human tissues and has the capacity to differentiate into tissue-specific cells, including myocytes, hepatocytes, and other cell types114-116. While some studies have reported differences in the prevalence of microchimerism between healthy individuals and patients with autoimmune disease, a more striking difference has often been an increase in the quantity of microchimerism in patients with autoimmune disease110,117-124, including most of the diseases shown in Figure 3c. The idea that the destructive immune response causing disease may be directed at the chimeric cells raised the suggestion that some autoimmune diseases may in fact be alloimmune125. Lastly, the onset of many autoimmune diseases in women during and immediately following the reproductive years has been attributed to microchimerism126, suggesting that exposure to fetal cells during pregnancy is a sex-specific risk factor for autoimmune disease.

Box 3. Genetic Imprinting and Parent-of-Origin Effects.

One mechanism for sex-specific transmission of disease or quantitative phenotypes is genomic imprinting, which refers to the transcriptional silencing of a gene in the gamete inherited from either the mother or the father, but not both (i.e., allele-specific silencing). The best studied silencing mechanism is methylation, and differential methylation between alleles is considered the hallmark feature of an imprinted locus127. The cellular mechanisms for sex-specific gene silencing and the impact of such parent-of-origin effects on human disease and gene evolution have been previously reviewed127-130.

In roughly half of imprinted genes the maternally-inherited allele is silenced (i.e., imprinted) and in the other half the paternally-inherited allele is silenced. In a few interesting cases, the imprinting itself is polymorphic so that both bi-allelic and monoallelic expression is observed between individuals131,132. Mutations in or deletions of the expressed allele at imprinted loci in humans or mice have a wide range of phenotypic consequences, including effects on growth and development, behavior and learning, and carcinogenesis127,128.

A census of imprinted genes in 2005 suggested that approximately 41 genes in 16 chromosomal regions are imprinted in humans, compared to 71 genes in 22 chromosomal regions in mice (29 of the same genes are imprinted in both humans and mice)130. The authors speculated that the total numbers of imprinted genes are probably not much greater than these estimates, although they acknowledged the possibility that additional imprinted genes with more subtle phenotypic effects probably exist. They cite in support of the latter the large number of complex diseases with parent-of-origin effects, including asthma, autism, type I and type II diabetes, Alzheimer disease, and schizophrenia. For these diseases, the risk for disease in the child differs depending on whether the mother or father is likewise affected, or whether a particular risk allele is inherited from the mother or from the father. Some of these effects may reflect as yet unidentified imprinted loci. In fact, a recent genome-wide analysis of genomic imprinting in mice revealed evidence for parent-of-origin effects due to genomic imprinting on a wide range of quantitative phenotypes related to body size and growth rates, and for imprinting effects that varied over time and which arose or persisted into adulthood133. Therefore, some of the sex-specific parent-of-origin effects observed in complex human diseases, such as those mentioned above, may be attributable to genomic imprinting.

Acknowledgements

The authors are grateful to Robert Rosenfield and J. Lee Nelson for helpful discussions, and to Sundeep Khosla and Elizabeth Atkinson for providing primary data on sex steroid levels in adults. The authors are supported in part by National Institutes of Health (NIH) grants HD021244, HL070831, and HL085197 (C.O.), NIH GM077959 and the Sloan foundation (Y.G.), and NIH T32 HL07605 (D.A.L.).

Glossary

Heterogametic (species)

Refers to species that produce gametes that differ with respect to sex chromosomes. In mammals, males are the heterogametic sex (XY) and females are homogametic (XX), whereas in birds, females are heterogametic (ZW).

Pyloric stenosis (infantile hypertrophic pyloric stenosis)

A common birth defect that results from the narrowing of the pylorus (lower part of the stomach), which prevents food and other stomach contents from passing into the intestine. This condition causes severe vomiting in infancy.

Genetic architecture

Refers to the underlying genetic basis for a trait.

Regulatory genome

The total set of different DNA molecules of an organelle, cell, or organism that are involved in the regulation of gene expression.

Sexual selection

Differential reproductive success resulting from the competition for fertilization. Competition for fertilization can occur through competition among the same sex (mate competition) or through attraction to the opposite sex (mate choice).

Heritability

The proportion of the total phenotypic variance for a given trait that can be attributed to genetic variation among individuals.

Forced expiratory volume at 1 second (FEV1)

The volume exhaled in the first second of a forced expiratory maneuver. This index is used to assess airway obstruction, bronchoconstriction, or bronchodilation.

Type I error

The probability of rejecting the null hypothesis when it is true, also referred to as a false positive.

Multiple testing

When multiple independent hypotheses are tested, the combined probability of type I error increases in an unadjusted analysis.

Consomic strain

Inbred strain in which a chromosome has been replaced by a homologous chromosome from another inbred strain.

Penetrance

The probability of observing a specific phenotype in individuals carrying a particular genotype.

Linkage disequilibrium

The nonrandom association of alleles at two or more loci. The pattern of linkage disequilibrium in a given genomic region reflects the history of natural selection, mutation, recombination, genetic drift, and other demographic and evolutionary forces.

Nondisjunction

The failure of chromosomes to separate at anaphase.

Aneuploidy

The presence of an abnormal number of chromosomes (either more or less than the diploid number).

Ectopic exchange

Homologous recombination between non-allelic chromosomal regions.

eQTL

Loci at which genetic allelic variation is associated with variation in gene expression.

Ontogenetic conflict

occurs when an allele is advantageous at one stage of development and disadvantageous at another stage, or when it is advantageous in one sex and disadvantageous in the other sex, so that the allelic effects are antagonistic with respect to fitness.

Alloimmune

An immune reaction against cells from another individual of the same species. Alloimmunity can occur during transfusion or transplantation, or during pregnancy.

Odds ratio (OR)

Compares the likelihood of an outcome (e.g., a disease) between two groups (e.g., cases and controls). It is measured as the ratio of the odds in one group to the odds in the second group and can be calculated by the following formula: p(1-q)/q(1-p), where p is the probability of the event occurring for the first group and q the probability for the second group.

References

  • 1.Alonso LC, Rosenfield RL. Oestrogens and puberty. Best Pract Res Clin Endocrinol Metab. 2002;16:13–30. doi: 10.1053/beem.2002.0177. [DOI] [PubMed] [Google Scholar]
  • 2.Carrel L, Willard HF. X-inactivation profile reveals extensive variability in X-linked gene expression in females. Nature. 2005;434:400–404. doi: 10.1038/nature03479. [DOI] [PubMed] [Google Scholar]
  • 3.Korstanje R, et al. Influence of sex and diet on quantitative trait loci for HDL cholesterol levels in an SM/J by NZB/BlNJ intercross population. J Lipid Res. 2004;45:881–888. doi: 10.1194/jlr.M300460-JLR200. [DOI] [PubMed] [Google Scholar]
  • 4****.Mackay TF. The genetic architecture of quantitative traits: lessons from Drosophila. Curr Opin Genet Dev. 2004;14:253–257. doi: 10.1016/j.gde.2004.04.003. [A classic review of gene–environment (including genotype–sex) interactions in Drosophila.] [DOI] [PubMed] [Google Scholar]
  • 5.Ueno T, et al. Rat model of familial combined hyperlipidemia as a result of comparative mapping. Physiol Genomics. 2004;17:38–47. doi: 10.1152/physiolgenomics.00043.2003. [DOI] [PubMed] [Google Scholar]
  • 6.Choi BG, McLaughlin MA. Why men's hearts break: cardiovascular effects of sex steroids. Endocrinol Metab Clin North Am. 2007;36:365–377. doi: 10.1016/j.ecl.2007.03.011. [DOI] [PubMed] [Google Scholar]
  • 7.Postma DS. Gender differences in asthma development and progression. Gend Med. 2007;4(Suppl B):S133–146. doi: 10.1016/s1550-8579(07)80054-4. [DOI] [PubMed] [Google Scholar]
  • 8.Lockshin MD. Sex differences in autoimmune disease. Lupus. 2006;15:753–756. doi: 10.1177/0961203306069353. [DOI] [PubMed] [Google Scholar]
  • 9.Harper PS. Practical Genetic Counseling. 5th Edition Reed Educational and Professional Publishing; Oxford: 1998. [Google Scholar]
  • 10.Gater R, et al. Sex differences in the prevalence and detection of depressive and anxiety disorders in general health care settings: report from the World Health Organization Collaborative Study on Psychological Problems in General Health Care. Arch Gen Psychiatry. 1998;55:405–413. doi: 10.1001/archpsyc.55.5.405. [DOI] [PubMed] [Google Scholar]
  • 11.Andersen K, et al. Gender differences in the incidence of AD and vascular dementia: The EURODEM Studies. EURODEM Incidence Research Group. Neurology. 1999;53:1992–1997. doi: 10.1212/wnl.53.9.1992. [DOI] [PubMed] [Google Scholar]
  • 12.Aleman A, Kahn RS, Selten JP. Sex differences in the risk of schizophrenia: evidence from meta-analysis. Arch Gen Psychiatry. 2003;60:565–571. doi: 10.1001/archpsyc.60.6.565. [DOI] [PubMed] [Google Scholar]
  • 13.Wooten GF, Currie LJ, Bovbjerg VE, Lee JK, Patrie J. Are men at greater risk for Parkinson's disease than women? J Neurol Neurosurg Psychiatry. 2004;75:637–639. doi: 10.1136/jnnp.2003.020982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Matanoski G, Tao XG, Almon L, Adade AA, Davies-Cole JO. Demographics and tumor characteristics of colorectal cancers in the United States, 1998−2001. Cancer. 2006;107:1112–1120. doi: 10.1002/cncr.22008. [DOI] [PubMed] [Google Scholar]
  • 15****.Patsopoulos NA, Tatsioni A, Ioannidis JP. Claims of sex differences: an empirical assessment in genetic associations. Jama. 2007;298:880–893. doi: 10.1001/jama.298.8.880. [A comprehensive review and critique of the evidence for sex-specific genetic effects on risk for common diseases.] [DOI] [PubMed] [Google Scholar]
  • 16.Barrett-Connor E. Commentary: Masculinity, femininity and heart disease. Int J Epidemiol. 2007;36:621–622. doi: 10.1093/ije/dym069. [DOI] [PubMed] [Google Scholar]
  • 17.Uekert SJ, et al. Sex-related differences in immune development and the expression of atopy in early childhood. J Allergy Clin Immunol. 2006;118:1375–1381. doi: 10.1016/j.jaci.2006.09.008. [DOI] [PubMed] [Google Scholar]
  • 18.Whitacre CC, Reingold SC, O'Looney PA. A gender gap in autoimmunity. Science. 1999;283:1277–1278. doi: 10.1126/science.283.5406.1277. [DOI] [PubMed] [Google Scholar]
  • 19.Straub RH. The complex role of estrogens in inflammation. Endocr Rev. 2007;28:521–574. doi: 10.1210/er.2007-0001. [DOI] [PubMed] [Google Scholar]
  • 20.Dobyns WB, et al. Inheritance of most X-linked traits is not dominant or recessive, just X-linked. Am J Med Genet A. 2004;129A:136–143. doi: 10.1002/ajmg.a.30123. [DOI] [PubMed] [Google Scholar]
  • 21.Lahn BT, Page DC. Functional coherence of the human Y chromosome. Science. 1997;278:675–680. doi: 10.1126/science.278.5338.675. [DOI] [PubMed] [Google Scholar]
  • 22.Lange J, Skaletsky H, Bell GW, Page DC. MSY Breakpoint Mapper, a database of sequence-tagged sites useful in defining naturally occurring deletions in the human Y chromosome. Nucleic Acids Res. 2008;36:D809–814. doi: 10.1093/nar/gkm849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23****.Reinius B, et al. An evolutionarily conserved sexual signature in the primate brain. PLoS Genet. 2008;4:e1000100. doi: 10.1371/journal.pgen.1000100. [An elegant demonstration of the conserved evolution of sexual dimorphism in gene expression patterns in the brain of primates.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24****.Rinn JL, Snyder M. Sexual dimorphism in mammalian gene expression. Trends Genet. 2005;21:298–305. doi: 10.1016/j.tig.2005.03.005. [A modern overview of the evolution of sexual dimorphism in gene expression.] [DOI] [PubMed] [Google Scholar]
  • 25.Yang X, et al. Tissue-specific expression and regulation of sexually dimorphic genes in mice. Genome Res. 2006;16:995–1004. doi: 10.1101/gr.5217506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26****.Ellegren H, Parsch J. The evolution of sex-biased genes and sex-biased gene expression. Nat Rev Genet. 2007;8:689–698. doi: 10.1038/nrg2167. [A comprehensive review of sexual dimorphism in the regulatory genome.] [DOI] [PubMed] [Google Scholar]
  • 27.Reinke V, Gil IS, Ward S, Kazmer K. Genome-wide germline-enriched and sex-biased expression profiles in Caenorhabditis elegans. Development. 2004;131:311–323. doi: 10.1242/dev.00914. [DOI] [PubMed] [Google Scholar]
  • 28****.Ranz JM, Castillo-Davis CI, Meiklejohn CD, Hartl DL. Sex-dependent gene expression and evolution of the Drosophila transcriptome. Science. 2003;300:1742–1745. doi: 10.1126/science.1085881. [One of the first genome-wide characterizations of sex-biased gene expression patterns in more than one species of Drosophila.] [DOI] [PubMed] [Google Scholar]
  • 29.Baker DA, Meadows LA, Wang J, Dow JA, Russell S. Variable sexually dimorphic gene expression in laboratory strains of Drosophila melanogaster. BMC Genomics. 2007;8:454. doi: 10.1186/1471-2164-8-454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Santos EM, Kille P, Workman VL, Paull GC, Tyler CR. Sexually dimorphic gene expression in the brains of mature zebrafish. Comp Biochem Physiol A Mol Integr Physiol. 2008;149:314–324. doi: 10.1016/j.cbpa.2008.01.010. [DOI] [PubMed] [Google Scholar]
  • 31.Nishida Y, Yoshioka M, St-Amand J. Sexually dimorphic gene expression in the hypothalamus, pituitary gland, and cortex. Genomics. 2005;85:679–687. doi: 10.1016/j.ygeno.2005.02.013. [DOI] [PubMed] [Google Scholar]
  • 32.Zhang Y, Sturgill D, Parisi M, Kumar S, Oliver B. Constraint and turnover in sex-biased gene expression in the genus Drosophila. Nature. 2007;450:233–237. doi: 10.1038/nature06323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Sartori-Valinotti JC, Iliescu R, Fortepiani LA, Yanes LL, Reckelhoff JF. Sex differences in oxidative stress and the impact on blood pressure control and cardiovascular disease. Clin Exp Pharmacol Physiol. 2007;34:938–945. doi: 10.1111/j.1440-1681.2007.04643.x. [DOI] [PubMed] [Google Scholar]
  • 34.Zammaretti F, Panzica G, Eva C. Sex-dependent regulation of hypothalamic neuropeptide Y-Y1 receptor gene expression in moderate/high fat, high-energy dietfed mice. J Physiol. 2007;583:445–454. doi: 10.1113/jphysiol.2007.133470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35****.Bhasin JM, et al. Sex specific gene regulation and expression QTLs in mouse macrophages from a strain intercross. PLoS ONE. 2008;3:e1435. doi: 10.1371/journal.pone.0001435. [The first explicit study of genetic variation that affects sex-specific variation in gene expression and of sex-specific eQTLs in a mammalian species (mice).] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Angelopoulou R, Lavranos G, Manolakou P. Establishing sexual dimorphism in humans. Coll Antropol. 2006;30:653–658. [PubMed] [Google Scholar]
  • 37****.Visscher PM, Hill WG, Wray NR. Heritability in the genomics era--concepts and misconceptions. Nat Rev Genet. 2008;9:255–266. doi: 10.1038/nrg2322. [An excellent review of the use and mis-use of the concept of heritability and its role in modern genetic studies.] [DOI] [PubMed] [Google Scholar]
  • 38.Shea MK, et al. Genetic and non-genetic correlates of vitamins K and D. Eur J Clin Nutr. 2007 doi: 10.1038/sj.ejcn.1602959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Santamaria A, et al. Quantitative trait locus on chromosome 12q14.1 influences variation in plasma plasminogen levels in the San Antonio Family Heart Study. Hum Biol. 2007;79:515–523. doi: 10.1353/hub.2008.0008. [DOI] [PubMed] [Google Scholar]
  • 40.de Simone G, et al. Assessment of the interaction of heritability of volume load and left ventricular mass: the HyperGEN offspring study. J Hypertens. 2007;25:1397–1402. doi: 10.1097/HJH.0b013e328126851e. [DOI] [PubMed] [Google Scholar]
  • 41.Pan L, Ober C, Abney M. Heritability estimation of sex-specific effects on human quantitative traits. Genet Epidemiol. 2007;31:338–347. doi: 10.1002/gepi.20214. [DOI] [PubMed] [Google Scholar]
  • 42.Weiss LA, Pan L, Abney M, Ober C. The sex-specific genetic architecture of quantitative traits in humans. Nat Genet. 2006;38:218–222. doi: 10.1038/ng1726. [DOI] [PubMed] [Google Scholar]
  • 43.Ober C, Abney M, McPeek MS. The genetic dissection of complex traits in a founder population. Am J Hum Genet. 2001;69:1068–1079. doi: 10.1086/324025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Abney M, McPeek MS, Ober C. Heritabilities of quantitative traits in a founder population. Am J Hum Genet. 2001;68:1302–1307. doi: 10.1086/320112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Seda O, et al. Systematic, genome-wide, sex-specific linkage of cardiovascular traits in French Canadians. Hypertension. 2008;51:1156–1162. doi: 10.1161/HYPERTENSIONAHA.107.105247. [DOI] [PubMed] [Google Scholar]
  • 46.Karasik D, Ferrari SL. Contribution of Gender-Specific Genetic Factors to Osteoporosis Risk. Ann Hum Genet. 2008;72:696–714. doi: 10.1111/j.1469-1809.2008.00447.x. [DOI] [PubMed] [Google Scholar]
  • 47.Wang C, et al. A computational model for sex-specific genetic architecture of complex traits in humans: implications for mapping pain sensitivity. Mol Pain. 2008;4:13. doi: 10.1186/1744-8069-4-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Brookes ST, et al. Subgroup analyses in randomized trials: risks of subgroup-specific analyses; power and sample size for the interaction test. J Clin Epidemiol. 2004;57:229–236. doi: 10.1016/j.jclinepi.2003.08.009. [DOI] [PubMed] [Google Scholar]
  • 49.Higaki J, et al. Deletion allele of angiotensin-converting enzyme gene increases risk of essential hypertension in Japanese men : the Suita Study. Circulation. 2000;101:2060–2065. doi: 10.1161/01.cir.101.17.2060. [DOI] [PubMed] [Google Scholar]
  • 50.O'Donnell CJ, et al. Evidence for association and genetic linkage of the angiotensin-converting enzyme locus with hypertension and blood pressure in men but not women in the Framingham Heart Study. Circulation. 1998;97:1766–1772. doi: 10.1161/01.cir.97.18.1766. [DOI] [PubMed] [Google Scholar]
  • 51.Stankovic A, Zivkovic M, Alavantic D. Angiotensin I-converting enzyme gene polymorphism in a Serbian population: a gender-specific association with hypertension. Scand J Clin Lab Invest. 2002;62:469–475. doi: 10.1080/00365510260390028. [DOI] [PubMed] [Google Scholar]
  • 52.Fornage M, et al. Variation in the region of the angiotensin-converting enzyme gene influences interindividual differences in blood pressure levels in young white males. Circulation. 1998;97:1773–1779. doi: 10.1161/01.cir.97.18.1773. [DOI] [PubMed] [Google Scholar]
  • 53.Weiss LA, Abney M, Cook EH, Jr., Ober C. Sex-specific genetic architecture of whole blood serotonin levels. Am J Hum Genet. 2005;76:33–41. doi: 10.1086/426697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Anholt RR, Mackay TF. Quantitative genetic analyses of complex behaviours in Drosophila. Nat Rev Genet. 2004;5:838–849. doi: 10.1038/nrg1472. [DOI] [PubMed] [Google Scholar]
  • 55.Mackay TF, Anholt RR. Of flies and man: Drosophila as a model for human complex traits. Annu Rev Genomics Hum Genet. 2006;7:339–367. doi: 10.1146/annurev.genom.7.080505.115758. [DOI] [PubMed] [Google Scholar]
  • 56.Melo JA, Shendure J, Pociask K, Silver LM. Identification of sex-specific quantitative trait loci controlling alcohol preference in C57BL/ 6 mice. Nat Genet. 1996;13:147–153. doi: 10.1038/ng0696-147. [DOI] [PubMed] [Google Scholar]
  • 57.Peirce JL, Derr R, Shendure J, Kolata T, Silver LM. A major influence of sex-specific loci on alcohol preference in C57Bl/6 and DBA/2 inbred mice. Mamm Genome. 1998;9:942–948. doi: 10.1007/s003359900904. [DOI] [PubMed] [Google Scholar]
  • 58.Athirakul K, et al. Increased blood pressure in mice lacking cytochrome P450 2J5. Faseb J. 2008 Aug;20 doi: 10.1096/fj.08-114413. (fj.08−114413v1) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Ponder CA, Munoz M, Gilliam TC, Palmer AA. Genetic architecture of fear conditioning in chromosome substitution strains: relationship to measures of innate (unlearned) anxiety-like behavior. Mamm Genome. 2007;18:221–228. doi: 10.1007/s00335-007-9013-9. [DOI] [PubMed] [Google Scholar]
  • 60.Mattson DL, et al. Chromosomal mapping of the genetic basis of hypertension and renal disease in FHH rats. Am J Physiol Renal Physiol. 2007;293:F1905–1914. doi: 10.1152/ajprenal.00012.2007. [DOI] [PubMed] [Google Scholar]
  • 61.Tu K, Chen Z, Lipscombe LL. Prevalence and incidence of hypertension from 1995 to 2005: a population-based study. Cmaj. 2008;178:1429–1435. doi: 10.1503/cmaj.071283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kearney PM, et al. Global burden of hypertension: analysis of worldwide data. Lancet. 2005;365:217–223. doi: 10.1016/S0140-6736(05)17741-1. [DOI] [PubMed] [Google Scholar]
  • 63.Burt VL, et al. Prevalence of hypertension in the US adult population. Results from the Third National Health and Nutrition Examination Survey, 1988−1991. Hypertension. 1995;25:305–313. doi: 10.1161/01.hyp.25.3.305. [DOI] [PubMed] [Google Scholar]
  • 64.Martins D, Nelson K, Pan D, Tareen N, Norris K. The effect of gender on age-related blood pressure changes and the prevalence of isolated systolic hypertension among older adults: data from NHANES III. J Gend Specif Med. 2001;4:10–13. 20. [PubMed] [Google Scholar]
  • 65.Kato N, et al. Comprehensive analysis of the renin-angiotensin gene polymorphisms with relation to hypertension in the Japanese. J Hypertens. 2000;18:1025–1032. doi: 10.1097/00004872-200018080-00006. [DOI] [PubMed] [Google Scholar]
  • 66.Rigat B, et al. An insertion/deletion polymorphism in the angiotensin I-converting enzyme gene accounting for half the variance of serum enzyme levels. J Clin Invest. 1990;86:1343–1346. doi: 10.1172/JCI114844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Krege JH, et al. Male-female differences in fertility and blood pressure in ACE-deficient mice. Nature. 1995;375:146–148. doi: 10.1038/375146a0. [DOI] [PubMed] [Google Scholar]
  • 68.Leung A, Chue P. Sex differences in schizophrenia, a review of the literature. Acta Psychiatr Scand Suppl. 2000;401:3–38. doi: 10.1111/j.0065-1591.2000.0ap25.x. [DOI] [PubMed] [Google Scholar]
  • 69.McGrath J, et al. A systematic review of the incidence of schizophrenia: the distribution of rates and the influence of sex, urbanicity, migrant status and methodology. BMC Med. 2004;2:13. doi: 10.1186/1741-7015-2-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Cardno AG, Gottesman II. Twin studies of schizophrenia: from bow-and-arrow concordances to star wars Mx and functional genomics. Am J Med Genet. 2000;97:12–17. [PubMed] [Google Scholar]
  • 71.Shifman S, et al. Genome-wide association identifies a common variant in the reelin gene that increases the risk of schizophrenia only in women. PLoS Genet. 2008;4:e28. doi: 10.1371/journal.pgen.0040028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Wedenoja J, et al. Replication of linkage on chromosome 7q22 and association of the regional Reelin gene with working memory in schizophrenia families. Mol Psychiatry. 2008;13:673–684. doi: 10.1038/sj.mp.4002047. [DOI] [PubMed] [Google Scholar]
  • 73.Hong SE, et al. Autosomal recessive lissencephaly with cerebellar hypoplasia is associated with human RELN mutations. Nat Genet. 2000;26:93–96. doi: 10.1038/79246. [DOI] [PubMed] [Google Scholar]
  • 74.Eastwood SL, Harrison PJ. Interstitial white matter neurons express less reelin and are abnormally distributed in schizophrenia: towards an integration of molecular and morphologic aspects of the neurodevelopmental hypothesis. Mol Psychiatry. 2003;8:769, 821–731. doi: 10.1038/sj.mp.4001399. [DOI] [PubMed] [Google Scholar]
  • 75.Hassold T, Hunt P. To err (meiotically) is human: the genesis of human aneuploidy. Nat Rev Genet. 2001;2:280–291. doi: 10.1038/35066065. [DOI] [PubMed] [Google Scholar]
  • 76.Baker BS, Carpenter AT, Esposito MS, Esposito RE, Sandler L. The genetic control of meiosis. Annu Rev Genet. 1976;10:53–134. doi: 10.1146/annurev.ge.10.120176.000413. [DOI] [PubMed] [Google Scholar]
  • 77.Coop G, Przeworski M. An evolutionary view of human recombination. Nat Rev Genet. 2007;8:23–34. doi: 10.1038/nrg1947. [DOI] [PubMed] [Google Scholar]
  • 78.Kong A, et al. Sequence variants in the RNF212 gene associate with genome-wide recombination rate. Science. 2008;319:1398–1401. doi: 10.1126/science.1152422. [DOI] [PubMed] [Google Scholar]
  • 79.Coop G, Wen X, Ober C, Pritchard JK, Przeworski M. High-resolution mapping of crossovers reveals extensive variation in fine-scale recombination patterns among humans. Science. 2008;319:1395–1398. doi: 10.1126/science.1151851. [DOI] [PubMed] [Google Scholar]
  • 80.Broman KW, Murray JC, Sheffield VC, White RL, Weber JL. Comprehensive human genetic maps: individual and sex-specific variation in recombination. Am J Hum Genet. 1998;63:861–869. doi: 10.1086/302011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Cheung VG, Burdick JT, Hirschmann D, Morley M. Polymorphic variation in human meiotic recombination. Am J Hum Genet. 2007;80:526–530. doi: 10.1086/512131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Martinez FD. CD14, endotoxin, and asthma risk: actions and interactions. Proc Am Thorac Soc. 2007;4:221–225. doi: 10.1513/pats.200702-035AW. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Zambelli-Weiner A, et al. Evaluation of the CD14/−260 polymorphism and house dust endotoxin exposure in the Barbados Asthma Genetics Study. J Allergy Clin Immunol. 2005;115:1203–1209. doi: 10.1016/j.jaci.2005.03.001. [DOI] [PubMed] [Google Scholar]
  • 84.Simpson A, et al. Endotoxin exposure, CD14, and allergic disease: an interaction between genes and the environment. Am J Respir Crit Care Med. 2006;174:386–392. doi: 10.1164/rccm.200509-1380OC. [DOI] [PubMed] [Google Scholar]
  • 85.Berchtold NC, et al. Gene expression changes in the course of normal brain aging are sexually dimorphic. Proc Natl Acad Sci U S A. 2008 doi: 10.1073/pnas.0806883105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Gupta V, Singh SM. Sex dimorphism in antitumor response of chemotherapeutic drug cisplatin in a murine host-bearing a T-cell lymphoma. Anticancer Drugs. 2008;19:583–592. doi: 10.1097/CAD.0b013e3282fb97bf. [DOI] [PubMed] [Google Scholar]
  • 87.Jackson A, Stephens D, Duka T. Gender differences in response to lorazepam in a human drug discrimination study. J Psychopharmacol. 2005;19:614–619. doi: 10.1177/0269881105056659. [DOI] [PubMed] [Google Scholar]
  • 88.Klein W. Gender differences in clinical trials in coronary heart disease: response to drug therapy. Eur Heart J. 1996;17:1786–1790. doi: 10.1093/oxfordjournals.eurheartj.a014792. [DOI] [PubMed] [Google Scholar]
  • 89.Dixon AL, et al. A genome-wide association study of global gene expression. Nat Genet. 2007;39:1202–1207. doi: 10.1038/ng2109. [DOI] [PubMed] [Google Scholar]
  • 90.Stranger BE, et al. Relative impact of nucleotide and copy number variation on gene expression phenotypes. Science. 2007;315:848–853. doi: 10.1126/science.1136678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Zhang W, et al. Evaluation of genetic variation contributing to differences in gene expression between populations. Am J Hum Genet. 2008;82:631–640. doi: 10.1016/j.ajhg.2007.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Morley M, et al. Genetic analysis of genome-wide variation in human gene expression. Nature. 2004;430:743–747. doi: 10.1038/nature02797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Storey JD. A direct approach to false discovery rates. J Royal Stat Soc B. 2002;64:479–198. [Google Scholar]
  • 94.Gilad Y, Rifkin SA, Pritchard JK. Revealing the architecture of gene regulation: the promise of eQTL studies. Trends Genet. 2008;24:408–415. doi: 10.1016/j.tig.2008.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Darwin C. On the Origins of Species by Means of Natural Selection. John Murray; London: 1859. [Google Scholar]
  • 96.Cunningham JT. Sexual dimorphism in the animal kingdom. Adam and Charles Black; London: 1900. [Google Scholar]
  • 97.Kimura KI, Ote M, Tazawa T, Yamamoto D. Fruitless specifies sexually dimorphic neural circuitry in the Drosophila brain. Nature. 2005;438:229–233. doi: 10.1038/nature04229. [DOI] [PubMed] [Google Scholar]
  • 98.Stockinger P, Kvitsiani D, Rotkopf S, Tirian L, Dickson BJ. Neural circuitry that governs Drosophila male courtship behavior. Cell. 2005;121:795–807. doi: 10.1016/j.cell.2005.04.026. [DOI] [PubMed] [Google Scholar]
  • 99.Kimchi T, Xu J, Dulac C. A functional circuit underlying male sexual behaviour in the female mouse brain. Nature. 2007;448:1009–1014. doi: 10.1038/nature06089. [DOI] [PubMed] [Google Scholar]
  • 100.Plavcan JM. Sexual dimorphism in primate evolution. Yearbook of Physical Anthropology, Vol 44. 2001;44:25–53. doi: 10.1002/ajpa.10011.abs. [DOI] [PubMed] [Google Scholar]
  • 101****.Andersson M. Sexual Selection. Princeton University Press; Princeton, New Jersey: 1994. [Comprehensive review and synthesis of essential topics in sexual selection, providing insight into the evolution of sex differences in nature and the role of selection and constraint in the development of secondary sexual traits.] [Google Scholar]
  • 102.Geary DC. Male, Female: The Evolution of Human Sex Differences. American Psychological Association; Washington, D.C.: 1998. [Google Scholar]
  • 103.Wizeman TM, Pardue M-L, editors. Exploring the biological contributions to human health: Does sex matter? National Academy Press; Washington, D.C.: 2001. [PubMed] [Google Scholar]
  • 104.Badyaev AV, Hill GE. Avian sexual dichromatism in relation to phylogeny and ecology. Annual Review of Ecology Evolution and Systematics. 2003;34:27–49. [Google Scholar]
  • 105****.Lande R. Sexual Dimorphism, Sexual Selection, and Adaptation in Polygenic Characters. Evolution. 1980;34:292–305. doi: 10.1111/j.1558-5646.1980.tb04817.x. [A classic study that used quantitative population genetic models to show that genetic correlations influence the expression and evolution of sexually dimorphic traits.] [DOI] [PubMed] [Google Scholar]
  • 106.Rice WR, Chippindale AK. Intersexual ontogenetic conflict. Journal of Evolutionary Biology. 2001;14:685–693. [Google Scholar]
  • 107.Nelson JL. Microchimerism in human health and disease. Autoimmunity. 2003;36:5–9. doi: 10.1080/0891693031000067304. [DOI] [PubMed] [Google Scholar]
  • 108.Lambert NC, et al. Quantification of maternal microchimerism by HLA-specific real-time polymerase chain reaction: studies of healthy women and women with scleroderma. Arthritis Rheum. 2004;50:906–914. doi: 10.1002/art.20200. [DOI] [PubMed] [Google Scholar]
  • 109.Maloney S, et al. Microchimerism of maternal origin persists into adult life. J Clin Invest. 1999;104:41–47. doi: 10.1172/JCI6611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Reed AM, Picornell YJ, Harwood A, Kredich DW. Chimerism in children with juvenile dermatomyositis. Lancet. 2000;356:2156–2157. doi: 10.1016/S0140-6736(00)03500-5. [DOI] [PubMed] [Google Scholar]
  • 111.Reed AM, McNallan K, Wettstein P, Vehe R, Ober C. Does HLA-dependent chimerism underlie the pathogenesis of juvenile dermatomyositis? J Immunol. 2004;172:5041–5046. doi: 10.4049/jimmunol.172.8.5041. [DOI] [PubMed] [Google Scholar]
  • 112.Yan Z, et al. Male microchimerism in women without sons: quantitative assessment and correlation with pregnancy history. Am J Med. 2005;118:899–906. doi: 10.1016/j.amjmed.2005.03.037. [DOI] [PubMed] [Google Scholar]
  • 113.Loubiere LS, et al. Maternal microchimerism in healthy adults in lymphocytes, monocyte/macrophages and NK cells. Lab Invest. 2006;86:1185–1192. doi: 10.1038/labinvest.3700471. [DOI] [PubMed] [Google Scholar]
  • 114.Stevens AM, Hermes HM, Rutledge JC, Buyon JP, Nelson JL. Myocardial-tissue-specific phenotype of maternal microchimerism in neonatal lupus congenital heart block. Lancet. 2003;362:1617–1623. doi: 10.1016/S0140-6736(03)14795-2. [DOI] [PubMed] [Google Scholar]
  • 115.Stevens AM, et al. Liver biopsies from human females contain male hepatocytes in the absence of transplantation. Lab Invest. 2004;84:1603–1609. doi: 10.1038/labinvest.3700193. [DOI] [PubMed] [Google Scholar]
  • 116.Khosrotehrani K, Johnson KL, Cha DH, Salomon RN, Bianchi DW. Transfer of fetal cells with multilineage potential to maternal tissue. Jama. 2004;292:75–80. doi: 10.1001/jama.292.1.75. [DOI] [PubMed] [Google Scholar]
  • 117.Nelson JL, et al. Microchimerism and HLA-compatible relationships of pregnancy in scleroderma. Lancet. 1998;351:559–562. doi: 10.1016/S0140-6736(97)08357-8. [DOI] [PubMed] [Google Scholar]
  • 118.Johnson KL, et al. Fetal cell microchimerism in tissue from multiple sites in women with systemic sclerosis. Arthritis Rheum. 2001;44:1848–1854. doi: 10.1002/1529-0131(200108)44:8<1848::AID-ART323>3.0.CO;2-L. [DOI] [PubMed] [Google Scholar]
  • 119.Ando T, Imaizumi M, Graves PN, Unger P, Davies TF. Intrathyroidal fetal microchimerism in Graves’ disease. J Clin Endocrinol Metab. 2002;87:3315–3320. doi: 10.1210/jcem.87.7.8656. [DOI] [PubMed] [Google Scholar]
  • 120.Nelson JL, et al. Maternal microchimerism in peripheral blood in type 1 diabetes and pancreatic islet beta cell microchimerism. Proc Natl Acad Sci U S A. 2007;104:1637–1642. doi: 10.1073/pnas.0606169104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Klintschar M, Schwaiger P, Mannweiler S, Regauer S, Kleiber M. Evidence of fetal microchimerism in Hashimoto's thyroiditis. J Clin Endocrinol Metab. 2001;86:2494–2498. doi: 10.1210/jcem.86.6.7540. [DOI] [PubMed] [Google Scholar]
  • 122.Artlett CM, et al. Chimeric cells of maternal origin in juvenile idiopathic inflammatory myopathies. Childhood Myositis Heterogeneity Collaborative Group. Lancet. 2000;356:2155–2156. doi: 10.1016/s0140-6736(00)03499-1. [DOI] [PubMed] [Google Scholar]
  • 123.Miyashita Y, Ono M, Ono M, Ueki H, Kurasawa K. Y chromosome microchimerism in rheumatic autoimmune disease. Ann Rheum Dis. 2000;59:655–656. doi: 10.1136/ard.59.8.654b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Badenhoop K. Intrathyroidal microchimerism in Graves’ disease or Hashimoto's thyroiditis: regulation of tolerance or alloimmunity by fetal-maternal immune interactions? Eur J Endocrinol. 2004;150:421–423. doi: 10.1530/eje.0.1500421. [DOI] [PubMed] [Google Scholar]
  • 125.Nelson JL. Maternal-fetal immunology and autoimmune disease: is some autoimmune disease auto-alloimmune or allo-autoimmune? Arthritis Rheum. 1996;39:191–194. doi: 10.1002/art.1780390203. [DOI] [PubMed] [Google Scholar]
  • 126.Gammill HS, Nelson JL. Naturally acquired microchimerism. Int J Dev Biol. 2008 doi: 10.1387/ijdb.082767hg. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Reik W, Walter J. Genomic imprinting: parental influence on the genome. Nat Rev Genet. 2001;2:21–32. doi: 10.1038/35047554. [DOI] [PubMed] [Google Scholar]
  • 128.Falls JG, Pulford DJ, Wylie AA, Jirtle RL. Genomic imprinting: implications for human disease. Am J Pathol. 1999;154:635–647. doi: 10.1016/S0002-9440(10)65309-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Wilkins JF, Haig D. What good is genomic imprinting: the function of parent-specific gene expression. Nat Rev Genet. 2003;4:359–368. doi: 10.1038/nrg1062. [DOI] [PubMed] [Google Scholar]
  • 130.Morison IM, Ramsay JP, Spencer HG. A census of mammalian imprinting. Trends Genet. 2005;21:457–465. doi: 10.1016/j.tig.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 131.Bunzel R, et al. Polymorphic imprinting of the serotonin-2A (5-HT2A) receptor gene in human adult brain. Brain Res Mol Brain Res. 1998;59:90–92. doi: 10.1016/s0169-328x(98)00146-6. [DOI] [PubMed] [Google Scholar]
  • 132.Giannoukakis N, Deal C, Paquette J, Kukuvitis A, Polychronakos C. Polymorphic functional imprinting of the human IGF2 gene among individuals, in blood cells, is associated with H19 expression. Biochem Biophys Res Commun. 1996;220:1014–1019. doi: 10.1006/bbrc.1996.0524. [DOI] [PubMed] [Google Scholar]
  • 133.Wolf JB, Cheverud JM, Roseman C, Hager R. Genome-wide analysis reveals a complex pattern of genomic imprinting in mice. PLoS Genet. 2008;4:e1000091. doi: 10.1371/journal.pgen.1000091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Kaufman JM, Vermeulen A. The decline of androgen levels in elderly men and its clinical and therapeutic implications. Endocr Rev. 2005;26:833–876. doi: 10.1210/er.2004-0013. [DOI] [PubMed] [Google Scholar]
  • 135.Khosla S, et al. Relationship of serum sex steroid levels and bone turnover markers with bone mineral density in men and women: a key role for bioavailable estrogen. J Clin Endocrinol Metab. 1998;83:2266–2274. doi: 10.1210/jcem.83.7.4924. [DOI] [PubMed] [Google Scholar]
  • 136.Winters SJ, Talbott E, Guzick DS, Zborowski J, McHugh KP. Serum testosterone levels decrease in middle age in women with the polycystic ovary syndrome. Fertil Steril. 2000;73:724–729. doi: 10.1016/s0015-0282(99)00641-x. [DOI] [PubMed] [Google Scholar]
  • 137.Cooper GS, Stroehla BC. The epidemiology of autoimmune diseases. Autoimmun Rev. 2003;2:119–125. doi: 10.1016/s1568-9972(03)00006-5. [DOI] [PubMed] [Google Scholar]
  • 138.Mendez EP, et al. US incidence of juvenile dermatomyositis, 1995−1998: results from the National Institute of Arthritis and Musculoskeletal and Skin Diseases Registry. Arthritis Rheum. 2003;49:300–305. doi: 10.1002/art.11122. [DOI] [PubMed] [Google Scholar]

RESOURCES