2,277
Views
149
CrossRef citations to date
0
Altmetric
Review Article

Complex disease, gender and epigenetics

, &
Pages 530-544 | Published online: 08 Jul 2009

Abstract

Gender differences in susceptibility to complex disease such as asthma, diabetes, lupus, autism and major depression, among numerous other disorders, represent one of the hallmarks of non‐Mendelian biology. It has been generally accepted that endocrinological differences are involved in the sexual dimorphism of complex disease; however, specific molecular mechanisms of such hormonal effects have not been elucidated yet. This paper will review evidence that sex hormone action may be mediated via gene‐specific epigenetic modifications of DNA and histones. The epigenetic modifications can explain sex effects at DNA sequence polymorphisms and haplotypes identified in gender‐stratified genetic linkage and association studies. Hormone‐induced DNA methylation and histone modification changes at specific gene regulatory regions may increase or reduce the risk of a disease. The epigenetic interpretation of sexual dimorphism fits well into the epigenetic theory of complex disease, which argues for the primary pathogenic role of inherited and/or acquired epigenetic misregulation rather than DNA sequence variation. The new experimental strategies, especially the high throughput microarray‐based epigenetic profiling, can be used for testing the epigenetic hypothesis of gender effects in complex diseases.

Simple and complex genetic diseases

In the past 20 years, there has been a major effort into the development of experimental and computational strategies for the identification of molecular causes of human complex disease. Unlike simple Mendelian disorders, for which the cloning of disease genes has become a routine procedure, identification of the molecular genetic basis of complex disease represents a major challenge to human biologists. Although the genes for some rare early onset and familial cases of complex diseases (e.g. colon cancer, breast cancer and Alzheimer's disease) have been identified, the overwhelming proportion of non‐Mendelian pathology remains unexplained. Research in complex non‐Mendelian diseases is reaching an unprecedented level of sophistication and scale: from building massive databases of DNA sequence variants (single nucleotide polymorphisms (SNPs), haplotype maps (HapMap)) to screening thousands of polymorphisms in thousands of affected individuals and controls. Why is it that complex diseases are so resistant to the discovery of the etiological factors? In an attempt to address this question, it is important to note that complex diseases, in comparison to simple ones, exhibit numerous epidemiological, clinical and molecular differences. Complex diseases as a rule are common (more than 1 case per 1000 individuals), with sporadic cases dominating over the familial ones, while simple ones are rare (less than 0.1%) and predominantly familial. Phenotypic differences (discordance) in identical twins have been one of the hallmarks of complex non‐Mendelian disease. Concordance of monozygotic twins reaches only ∼15% in breast cancer, 20% in ulcerative colitis, 25%–30% in multiple sclerosis, 25%–45% in diabetes, 50% in schizophrenia and 40%–70% for Alzheimer's disease Citation1. Complex diseases usually exhibit a significantly later age at onset, while the overwhelming majority (∼90%) of Mendelian diseases have an onset before puberty and only 1% after 40 years of age Citation2. Major psychosis, asthma, multiple sclerosis, inflammatory bowel disease, rheumatoid arthritis and numerous other complex diseases exhibit major fluctuations in disease severity, and in some cases age‐dependent decline of symptoms and even full recovery from a disease. Epidemiological studies have revealed that quite often the risk of developing complex disease in offspring depends on the sex of affected parent. For example, asthma, bipolar disorder and epilepsy are more often transmitted from the affected mother, while type 1 diabetes seems to be more often transmitted from the affected father. Finally, one of the most common and well known non‐Mendelian features of complex disease is differential susceptibility to a disease in males and females, or sexual dimorphism Citation2. This article will review epidemiological and genetic findings of gender effects in complex disease. The other objectives are: 1) to introduce the concept of epigenetics and demonstrate that gender effects are consistent with the epigenetic interpretation of complex disease; and 2) to describe experimental strategies for the identification of the molecular substrate of gender effects, which may contribute to uncovering the fundamental molecular mechanisms of complex disease.

Key messages

  • Sexual dimorphism in complex diseases presents with an uneven disease frequency in men and women: although both genders can be affected, one is more susceptible.

  • It can be hypothesized that differential susceptibility to complex disease in males and females is mediated by sex hormone‐induced differences in the epigenetic regulation of genes.

  • The epidemiological and molecular evidence of gender effects in various complex diseases warrants dedicated molecular studies that would uncover underlying epigenetic mechanisms and genomic sites that are the primary targets of sex hormone action.

Gender and complex disease: epidemiological and molecular genetic findings

Sexual dimorphism in complex diseases presents with an uneven disease frequency in men and women: although both genders can be affected, one is more susceptible. Multiple sclerosis, rheumatoid arthritis, Crohn's disease, panic disorder, structural heart disease and hyperthyroidism are more common in females, while males are more often affected with autism, Hirschsprung's disease, ulcerative colitis, Parkinson's disease, alcoholism, allergies and asthma (especially at young age) Citation3 (Figure ). In psychiatric diseases such as Alzheimer's disease, schizophrenia, alcoholism, and mood and anxiety disorders, psychopathology exhibits a number of differences between the sexes in rates of illness as well as the course of illness Citation4. It is important to note that sex effects in complex diseases should be differentiated from those in single‐gene disorders where sex chromosome‐linked genes are known to be the cause of sexual dimorphism. In simple Mendelian disorders sexual dimorphism is clear‐cut and usually affects only one of two sexes. For example, only males (excluding some rare cases of e.g. skewed X inactivation) are affected with hemophilia and Duchenne's muscular dystrophy, but only females can have Rett's syndrome (again, with some rare exceptions).

Figure 1 Complex non‐Mendelian diseases displaying differential susceptibility to complex diseases (based on epidemiological data obtained from Harrison's ‘Principles of Internal Medicine’ 15th edition, 2001) Citation111.

Figure 1 Complex non‐Mendelian diseases displaying differential susceptibility to complex diseases (based on epidemiological data obtained from Harrison's ‘Principles of Internal Medicine’ 15th edition, 2001) Citation111.

Sex hormones have been the usual ‘culprit’ in the explanation of gender effects in various morphological and physiological differences between males and females Citation5. Differences in endocrine mediated development and maintenance throughout the lives of males and females are the second factor next to the sex chromosomes that define the sexes and are likely to drive a dynamic divergence of epigenetic patterns between the sexes. From brain studies it is known that differential effects of androgens and estrogens contribute to neural development, affecting programmed cell death, cellular migration, synaptogenesis and axonal migration, and formation of sexually distinct neuronal circuits Citation6, Citation7. Other studies have shown that androgens are directly related to neurite arborization, while estrogens are responsible for synapse formation and initiation of cellular communication Citation8. Differences in the circulating hormone concentrations between sexes orchestrate the divergence of behavior and cognitive development in males through two related processes known as masculization and defeminization mediated by the estrogen receptor α and β, respectively Citation9. Additionally, gonadal hormones are implicated in sex differences observed in neural protection to trauma or ischemic brain injury, with clinical reports showing a higher recovery rate in females than in males, and progesterone and estrogen treatments demonstrating a neuroprotective effect Citation5. Numerous other disease‐associated traits are differentially affected by the sex hormones. For example, in obesity, which is a risk factor for coronary artery disease, diabetes, osteoarthritis, and some metabolic disorders Citation10, Citation11, hormones have been implicated in the control of adipose tissue proliferation and estrogen and testosterone replacement therapy in older women and men resulting in a reduction of obesity Citation11. Estrogen replacement also attenuates Alzheimer's disease (AD) symptoms in postmenopausal women Citation12. Additionally, AD‐affected men have significantly lower levels of testosterone in the brain compared to controls Citation13. Animal models of disease offer a valuable resource for studying the hormone‐associated aspects of susceptibility to complex disease. Female nonobese diabetic (NOD) mice are more often affected by type I diabetes than the male mice of the same strain Citation14. Neonatal sex hormone manipulation demonstrated that disease phenotype in castrated males and ovariectomized females increased and decreased, respectively Citation15.

It is interesting to note that gender effects have also been detected in genetic linkage and association studies. There is now an increasing list of molecular genetic findings indicating that genetic markers on autosomes also can be gender dependent (Table ). For example, genetic linkage studies in autism revealed that 17q11 at markers D17S1294 and D17S798 reached genome‐wide significance with the linkage stratification on the basis of gender (P = 0.008) Citation16. Another group performed a search for susceptibility loci stratified for the sex of the affected proband using a number of markers on chromosomes 2q, 7q, 9p, 15q and 16p Citation17. Male‐female and female‐female pairs showed evidence for linkage to chromosome 15q markers (D15S117D15S125, P = 0.0011), while male pairs only demonstrated significance on 16p (D16S407D16S497, P = 0.026) with a trend towards significance observed on 7q (D7S480D7S530, P = 0.075) Citation17. In major depression studies, a genome‐wide linkage study in a sample of 110 Utah pedigrees with a history of major depression identified significant logarithm of odd scores (LOD) signifying linkage at 12q22‐q23.2 markers specific only to males (D12S1300, LOD = 4.6, P = 0.00003; D12S1706 LOD = 6.1, P = 0.0000007) Citation18. Another genome scan revealed a number of putative loci of linkage to major depression females only, with the most prominent linkage on 2q33‐q35 displaying LOD scores of 6.3 and 6.9 at D2S2321 and D2S2208, respectively Citation19. A haplotype analysis was performed on the TRAX/DISC1 region previously associated with schizophrenia in a Scottish population Citation20. A case‐control association study identified significant association of DISC1 with bipolar disorder in women (P = 0.00026) Citation20. Another study investigating 28 SNPs including TRAX, DISC1, and DISC2 in a Finnish cohort of 458 individuals determined a significant undertransmission of the HEP3 haplotype to affected females (P = 0.00024), suggesting that this haplotype may confer a protective effect against the disease in women Citation21. Finally, a recent quantitative trait loci [QTL] analysis of various other complex diseases found that of the 17 QTLs investigated, 11 were significantly sexually dimorphic (P<0.05) Citation22. Among the sexually dimorphic traits were triglycerides, high‐density lipoprotein cholesterol, forced expiratory volume, all anthropometric traits and cortisol levels Citation22.

Table I. Some loci and markers exhibiting gender‐specific evidence for association (A) and linkage (B) in complex diseases and traits. Not all significant findings from each reference are reported. A. Association studies:

B. Linkage studies:

While sex hormones have been the usual ‘culprit’ in explanations of gender effects in complex diseases based on the myriad of data associating hormonal differences with disease states and their critical involvement in human biology Citation5, there are no underlying mechanisms proposed as to how such hormones predispose to or protect from a disease relating to the specific molecular mechanisms of hormone action. The gender‐specific effects in genetic linkage and association studies suggest that chromosomes and individual genes can be the target of sex hormones. While such hormones cannot change DNA sequence, they can be potent modifiers of epigenetic status, which controls gene expression and various other genomic activities. It is known that hormones, including sex hormones, can control gene expression via epigenetic modifications Citation2. Therefore, it can be hypothesized that differential susceptibility to complex disease in males and females is mediated by sex hormone‐ induced differences in epigenetic regulation of genes. In the next section, a brief introduction to epigenetics is provided.

Epigenetics and its relevance to complex disease

Epigenetics refers to regulation of various genomic functions that are brought about by heritable, but potentially reversible changes in DNA modification and chromatin structure Citation23. The epigenetic information is encoded in two types of synergistically acting covalent modifications: DNA methylation, more specifically methylated cytosines (metC), and chromatin proteins (modifications of chromatin proteins such as histone acetylation, methylation, phosphorylation) Citation24. In mammals, DNA methylation occurs most commonly where cytosine is directly followed by guanine, forming what is known as a CpG dinucleotide. Clusters of CpG dinucleotides are referred to as CpG islands Citation25. The epigenetic complexity conferred is through DNA and chromatin modifications that can directly affect the regulation of gene activity Citation26–32, and other important genomic functions Citation33, Citation34 including genetic recombination Citation35 and DNA mutability Citation36. Maintenance of existing DNA methylation and de novo DNA methylation is catalyzed by several types of enzymes known as DNA‐methyltransferases Citation37. The importance of proper maintenance of methylation throughout the genome is highlighted by the observation that a complete loss of maintenance DNA‐methyltranferase function results in death of mice in early embryogenesis Citation38. A large number of genes exhibit an inverse correlation between the degree of DNA methylation and the magnitude of gene expression Citation39, Citation40. Transcription factor binding affinity may be limited by the presence of methylation at the binding sites Citation28, Citation31. In addition to the ‘critical site’ effects of metC, the density of metC in a gene regulatory region also contributes to gene activity. DNA modification acts in concert with histone modifications, another epigenetic mechanism. Alterations in chromatin structure occur through acetylation, methylation and phosphorylation of various histone amino acid residues including lysine, arginine and serine Citation41–45. The information within the patterns of such modifications is sometimes referred to as the ‘histone code’ Citation24. Epigenetic modifications regulate genomic functioning not only in terms of gene expression but also in the suppression of repetitive DNA sequences Citation46 and the formation of architecturally functional chromatin structures such as centromeric regions Citation47. There has been significant progress in understanding the complementary functioning of epigenetic mechanisms, i.e. how changes in DNA modification affect chromatin conformation and vice versa Citation48–50. It is important to note that some studies have demonstrated that genetic differences such as single nucleotide polymorphisms (SNPs) can be predictors of differentially methylated alleles Citation51–53 allowing for a reinterpretation of the potential importance of some seemingly functionally irrelevant SNPs associated with various complex diseases. Sex effects linked to or associated with specific SNPs or haplotypes can be explained by hormone‐induced modifications such that a specific allele or haplotype only becomes a risk factor after some endocrinologically mediated epigenetic modification.

There is increasing evidence that putative epigenetic misregulation of genes may explain numerous epidemiological, clinical, and molecular complexities of complex disease Citation2. It has been argued that the heuristic value of the epigenetic model of complex disease lies in the possibility of integrating a variety of unrelated data into a new theoretical framework, providing the basis for new experimental approaches Citation54–58. Shifting the emphasis from DNA sequence variation and a hazardous environment as the main etiological factors to putative epigenetic misregulation provides a new perspective on the myriad of the above listed non‐Mendelian features (prevalence of sporadic cases, discordance of identical twins, parental origin effects, late age of disease onset and fluctuating course, among others) that cannot be explained by more traditional mechanisms Citation59. In the epigenetic model of complex non‐Mendelian disease, pathology is seen to result from a chain of unfavorable epigenetic events that begin with a primary epigenetic defect, or preepimutation, occurring in the germline during the error‐prone epigenetic‐reprogramming process Citation60, Citation61. Preepimutation increases the risk of developing disease, but unlike the deterministic DNA mutations in Mendelian disorders, a preepimutation does not necessarily indicate that the disease is inevitable. Such preepimutations may not cause any immediate clinical problems, and thus the age of onset may be delayed for a relatively long time. It may take decades until the epigenetic misregulation reaches a critical threshold beyond which the cell is no longer able to function normally. The phenotypic outcome depends on the overall effect of a series of pre‐ and postnatal impacts on such a preepimutation. Preepimutations are subject to further changes by the multidirectional effects of tissue differentiation, stochastic events, some external environmental factors (nutrition, medications, etc.) Citation62–64, and also the hormonal milieu. It is known that various hormones, including sex hormones, have a significant impact on gene expression, and this is achieved by changing chromatin conformation Citation65–68 and/or local pattern of gene methylation Citation69, Citation70. Some examples of how sex hormones can influence epigenetic regulation of genes are provided below.

Epigenetics and sex hormones

A number of earlier studies have demonstrated the direct effects of sex hormone administration on epigenetic states. An example is the effect of estradiol administration on the methylation status of various CpG dinucleotides located in an estradiol‐mediated regulatory region for the avian vitellogenin II gene Citation69, Citation71–73. Two of these methylatable cytosines located within the estradiol‐receptor binding site are actively demethylated in response to estradiol treatment in hormone responsive tissues in a strand‐specific manner Citation69. This DNA demethylation persists after transcription has ended Citation71, Citation73 and has been suggested to result in a ‘memory effect’ resulting in a quicker induction of vitellogenin II mRNA synthesis in response to subsequent estradiol stimulations Citation69. This finding highlights the fact that endocrinological influence can mediate long‐lasting epigenetic change to levels of gene transcription.

One of the mechanisms of sex hormone action is through the molecular epigenetic signatures of particular chromosomal regions that modulate the access of transcription factors to the transcribed sequences. If DNA is tightly packed around the histone proteins, the access of transcription factors to their respective transcription factor binding sites is restricted Citation24, Citation74–76. Such compacted chromatin is supported by high density of metC in the DNA sequence and various types of suppressing modifications of amino‐terminal residues of histone proteins, e.g. deacetylation Citation75. Conversely, low density of metC in the DNA sequence and high levels of histone acetylation unravels the chromatin, allowing for access of DNA binding elements and transcriptional activation Citation24, Citation74–76.

Chromatin modification can be directly affected by members of the nuclear hormone receptor (NHR) superfamily, and the steroid receptor (SR) subset of NHR is of particular interest as it contains the androgen receptor (AR) and estrogen receptor (ER) Citation77. These sex hormone receptors can have activating or repressing transcriptional activity dependent on the presence or absence of ligand, respectively Citation78, Citation79. Steroid hormone mediated transcriptional activation or repression results from the SR recruitment of coactivator and corepressor complexes—protein complexes which associate with various epigenetic modifiers such as histone deacetylases (HDAC), histone acetyltransferases (HAT) and histone methyltransferases (HMT) Citation78, Citation80. It is these ‘coregulatory complexes’ that achieve the epigenetic modification necessary for chromatin remodeling, allowing or restricting access of transcription factors and RNA polymerase II and thus mediating the epigenetic effects of the sex hormones (Figure ).

Figure 2 In the absence of ligand or in the presence of antagonist, steroid receptors(SR) recruit transcriptional corepressor complexes to the DNA, producing an epigenetically silenced state through histone deacetylation or methylation. In the presence of hormone, SR recruit coactivator complexes, histone‐modifying machines or ATP‐dependent chromatin remodeling machines, to acetylate histones and produce a euchromatic chromatin formation suitable for transcriptional activation.

Figure 2 In the absence of ligand or in the presence of antagonist, steroid receptors(SR) recruit transcriptional corepressor complexes to the DNA, producing an epigenetically silenced state through histone deacetylation or methylation. In the presence of hormone, SR recruit coactivator complexes, histone‐modifying machines or ATP‐dependent chromatin remodeling machines, to acetylate histones and produce a euchromatic chromatin formation suitable for transcriptional activation.

There are two primary types of coregulatory complexes, the first being the adenosine triphosphate (ATP)‐dependant chromatin remodeling complexes (Figure ), such as the switch/sucrose nonfermenting (SWI/SNF) complex Citation77. The ATP‐dependent chromatin remodeling complexes may be primarily responsible for opening of chromatin Citation81, Citation82, formation of nucleosome arrays, homologous strand pairing and DNA transcription Citation83. Only recently, the actions of SWI/SNF have been linked to changes in DNA methylation through a mechanism mediated by interactions with methyl CpG binding protein 2 (MeCP2) Citation84. SWI/SNF interacts with both the ER and AR, and mutations in a core subunit have been implicated in the development of ovarian cancer Citation85.

Histone‐modifying complexes constitute the second type of coregulatory complex that interact with the sex hormone receptors. Coactivators include the cyclic adenosine monophosphate (cAMP) response element‐binding protein (CBP)/p300, while the most widely studied corepressor complexes are histone‐modifying machines such as the nuclear receptor corepressor complex (NCoR) and the silencing mediator of retinoic acid and thyroid hormone receptor (SMRT) Citation77 (Figure ). Histone modifications mediated by histone‐modifying machines unravel and compact chromatin through acetylation and deacetylation of lysine residues on histone H4, respectively, and alternatively through methylation of histone lysine and arginine residues Citation74, Citation86. In addition to receptor agonist binding, antagonists to the progesterone receptor (PR) and androgen receptor (AR) elicit an interaction with NCoR and SMRT, demonstrating that antagonistic actions are mediated through corepressor recruitment Citation77.

There are numerous molecular mechanisms through which AR and ER coactivator/corepressor‐binding modulate chromatin structure and influence epigenetic marks; however, the distribution and expression of the genes encoding the coregulatory factors have been shown to vary between the sexes. Gender differences in the expression of steroid receptor co‐activator‐1 (SCR‐1), CBP/p300, NCoR, and SMRT were observed in various tissues in male and female rats Citation87. Furthermore, regulation of transcript levels can be influenced by the differences in circulating hormones between males and females, as estradiol differentially influenced SCR‐1 and SMRT transcript levels in the anterior pituitary of male and female rats Citation87. Therefore, despite being ubiquitously expressed, the gender‐ and tissue‐specific coregulator transcript differences have been suggested to result in variability in the level of response to hormone Citation87, a finding that could explain observations of tissue‐specific differences in ER‐mediated response to ligand.

AR and ER differentially interact with coregulatory complexes through binding to different types of α‐helix LXXLL motifs on the complexes Citation88, Citation89. Through these interactions, recent studies have suggested that different types of ligand binding to the sex hormone receptors or conformational changes over time may influence the ability to recruit various cofactors, as in the case of AR and ER where changes in coactivator‐binding have been linked to prostate and breast cancer development, respectively Citation90, Citation91. In a similar way, differences in AR and ER interactions with coactivator LXXLL motifs Citation88 may result in the recruitment of different epigenetic modifying coregulatory complexes between the sexes.

Finally, various studies have demonstrated that tissue‐specific distribution of the sex hormone receptors varies between the sexes Citation92 as do the target genes for transcriptional regulation, and thus the epigenetic modifications mediated through the sex hormone receptors will likely vary by gene and tissue between the sexes. A study investigating the transcriptional regulation of the prolactin gene tested the effects of estradiol administration in GH4 cells Citation93. Estradiol was found to have a stimulating effect on histone H4 acetylation in the promoter region of the prolactin gene Citation93. In agreement with other studies, such histone modifications are implicated as necessary factors in gene transcription via the formation of a euchromatic state of the promoter region Citation94. Various forms of histone methylation are associated with estradiol binding to the estrogen receptor (ER), affecting the transcription at the target promoter in vitroCitation77, Citation95 and further highlighting that sex hormone‐regulated gene transcription is associated with the formation of an epigenetic mark. In rat, estrogen treatment was shown to increase DNA methylation on the prolactin gene in pituitary and liver tissues with subsequent reduction in mRNA amounts, suggesting a direct effect on transcription in a tissue‐specific manner Citation96. Prolactin administration caused demethylation of DNA in rat liver and kidney in both mature and immature rats Citation97. Prenatal exposure to diethylstilbestrol, a synthetic form of estrogen that has been implicated in development of cancers in humans and mice in later life, permanently alters the DNA methylation pattern of the oncogene, c‐fos, in mice, implicating an epigenetic mechanism for hormone‐induced carcinogenesis Citation98. Testosterone and estradiol administration have been shown to affect amyloid precursor protein (APP) levels in AD model mice Citation99. DNA methylation levels at the APP promoter exhibit a sex difference in mice and are generally higher in females Citation99. Increased levels of estrogens in transgenic breast cancer model mice lead to epigenetic modifications in cancer relevant genes involved in cell cycle, cell proliferation and apoptosis Citation100. Finally, investigation of the effects of estrogen on Japanese medaka fish demonstrated that DNA methylation varied between tissues in a gender‐specific manner Citation101.

Molecular strategies for identification of gender specific epigenetic differences

The epidemiological and molecular evidence of gender effects in various complex diseases warrant dedicated molecular studies that would uncover underlying epigenetic mechanisms and genomic sites that are the primary targets of sex hormone action. As mentioned above, evidence for linkage or association of a particular genomic region could be a marker for differential epigenetic modification of that region, therefore transforming those loci in Table  into candidate targets for gender‐specific epigenetic evaluation. Site‐specific methods for the epigenetic investigation of candidate loci exist, such as in sodium bisulfite modification‐based mapping of methylated cytosines, or chromatin immunoprecipitation assays for histone modification studies. Sodium bisulfite modification is considered the gold standard in DNA methylation measurement and functions through selectively deaminating unmethylated cytosines to uracil while methylated cytosines remain unchanged. After polymerase chain reaction (PCR) amplification, the relative proportions of DNA methylation can be determined at each CpG dinucleotide through a variety of methods including sequencing of clones Citation102, Citation103, pyrosequencing Citation80, MALDI mass spectrometry Citation104, Citation105, and SNaPshot Citation106, amongst others. Chromatin immunoprecipitation is performed with antibodies for modified histone residues, such as H4‐K9 methylation for example, followed by real time PCR amplification of a candidate region in order to quantify levels of histone protein modification Citation107. Use of these methods would allow for the locus‐specific determination of epigenetic differences in populations stratified for sex; however, while these candidate loci may represent targets of epigenetic modification through the actions of sex hormones, the overwhelming majority of sex hormone targets in the genome remain unknown. For this reason, a comprehensive epigenetic interrogation of the entire genome would be optimal for identifying the yet unknown targets of sex hormone‐mediated epigenetic modification that may be associated with sex effects in disease.

The advent of microarray technologies that interrogate a large number of DNA/RNA fragments in a highly parallel fashion has also opened new opportunities for epigenetic studies Citation55, Citation57, Citation58, Citation108. The microarray‐based large‐scale DNA modification analysis is based on interrogation of the unmethylated (or hypermethylated) fraction of genomic DNA on the microarray containing oligonucleotides that represent the genomic regions of interest. Enrichment of the unmethylated fraction of genomic DNA is performed through a number of steps that include digestion with methylation‐sensitive restriction enzymes. Such enzymes cut the unmethylated fraction of genomic DNA into short fragments, whereas the methylated DNA sequences remain undigested. In the next step, the restriction enzyme‐treated DNA is ligated to DNA adaptors and subjected to polymerase chain reaction (PCR). PCR primers complementary to the adaptor sequences preferentially amplify the shorter DNA fragments, namely the unmethylated DNA fraction. Fluorescently labeled amplicons are then hybridized to the microarrays. The lower the degree of DNA methylation at some specific site of the gene of interest, the stronger the hybridization signal seen on the array. Additionally, antibodies specific to methylated cytosines have been developed and applied in accordance with microarray technology to interrogate DNA methylation within the genome Citation109. In addition to DNA methylation, histone residue modification can be investigated using chromatin immunoprecipitation on microarray chips (ChIP‐on‐chip) Citation108.

As an example of the microarray‐based approaches, a recent analysis adds weight to the hypothesis that some molecular causes of gender differences may manifest in different profiles of epigenetic markers between the sexes. As a part of our ongoing study on epigenomics of human brain we performed a pilot analysis of epigenomic profiles of postmortem brain samples stratified for gender. The unmethylated fraction of DNA extracted from the prefrontal cortex of postmortem brains was interrogated on the 12K CpG island microarray Citation58, Citation61. The group of males (n = 20) was compared to the group of females (n = 7). DNA methylation differences in males and females (P<0.05) identified by the microarray between the sexes are shown in Figure . While this pilot data is not corrected for multiple testing, the identified regions might serve as a starting point for the in‐depth epigenetic studies of sex differences, mapping of gender‐dependent epigenetic effects across the entire genome, and analysis of the relevance of such sites to various complex diseases.

Figure 3 Loci identified as significantly different in DNA methylation between male and female postmortem brain samples. DNA methylation profiles were measured by human CpG island microarrays in a common reference design. Each microarray interrogates 12,912 loci whose locations on the chromosomes are shown in white in the plot. Data were logarithmically transformed and then normalized, by the print‐tip loess method, to remove nonbiological effects. Vertically: human chromosomes 1 to 22. The t‐statistic of the male group versus female group was calculated for each locus and its P‐value was also obtained. Loci with P‐values below 0.05 are painted black in the plot.

Figure 3 Loci identified as significantly different in DNA methylation between male and female postmortem brain samples. DNA methylation profiles were measured by human CpG island microarrays in a common reference design. Each microarray interrogates 12,912 loci whose locations on the chromosomes are shown in white in the plot. Data were logarithmically transformed and then normalized, by the print‐tip loess method, to remove nonbiological effects. Vertically: human chromosomes 1 to 22. The t‐statistic of the male group versus female group was calculated for each locus and its P‐value was also obtained. Loci with P‐values below 0.05 are painted black in the plot.

Experimental techniques applied to investigate longitudinal changes in gender‐specific epigenetic patterns will be of particular interest. Also, studies on epigenetic effects of hormones may allow for the identification of critical preclinical epigenetic states that could be used in early diagnostics as well as in designing new epigenetic therapies Citation110. It is our hope that through an acceleration of molecular biology technologies of epigenetic investigation and a combination of animal and human studies, the role of epigenetic factors in complex disease—both mediated through sex hormones and other mechanisms—will soon be elucidated.

Acknowledgements

This research has been supported by the Ontario Mental Health Foundation, Canadian Institutes for Health and Research, National Institute of Mental Health (R01 MH074127‐01), as well as NARSAD and the Stanley Foundation. Zachary Kaminsky is a CIHR Doctoral Fellow. Special thanks to Sigrid Ziegler for her critical reading of the manuscript.

References

  • Petronis A. Epigenetics and twins: three variations on the theme. Trends Genet 2006; 22: 347–50
  • Petronis A. Human morbid genetics revisited: relevance of epigenetics. Trends Genet 2001; 17: 142–6
  • Ostrer H. Sex‐based differences in gene transmission and gene expression. Lupus 1999; 8: 365–9
  • Seeman M. V. Psychopathology in women and men: focus on female hormones. Am J Psychiatry 1997; 154: 1641–7
  • Committee on Understanding the Biology of Sex and Gender Differences, Board on Health Science Policy, Institute of Medicine;. Exploring the Biological Contributions to Human Health, Does Sex Matter? 2001, Theresa M. Wizemann, Mary‐Lou Pardue, eds. Washington D.C. National Academy Press
  • Simerly R. B. Wired on hormones: endocrine regulation of hypothalamic development. Curr Opin Neurobiol 2005; 15: 81–5
  • Simerly R. B. Wired for reproduction: organization and development of sexually dimorphic circuits in the mammalian forebrain. Annu Rev Neurosci 2002; 25: 507–36
  • Lustig R. H. Sex hormone modulation of neural development in vitro. Horm Behav 1994; 28: 383–95
  • Kudwa A. E., Michopoulos V., Gatewood J. D., Rissman E. F. Roles of estrogen receptors alpha and beta in differentiation of mouse sexual behavior. Neuroscience 2006; 138: 921–8
  • Bolduc C., Larose M., Yoshioka M., Ye P., Belleau P., Labrie C., et al. Effects of dihydrotestosterone on adipose tissue measured by serial analysis of gene expression. J Mol Endocrinol 2004; 33: 429–44
  • Mayes J. S., Watson G. H. Direct effects of sex steroid hormones on adipose tissues and obesity. Obes Rev 2004; 5: 197–216
  • Tan Z. S., Seshadri S., Beiser A., Zhang Y., Felson D., Hannan M. T., et al. Bone mineral density and the risk of Alzheimer disease. Arch Neurol 2005; 62: 107–11
  • Rosario E. R., Chang L., Stanczyk F. Z., Pike C. J. Age‐related testosterone depletion and the development of Alzheimer disease. JAMA 2004; 292: 1431–2
  • Ivakine E. A., Fox C. J., Paterson A. D., Mortin‐Toth S. M., Canty A., Walton D. S., et al. Sex‐specific effect of insulin‐dependent diabetes 4 on regulation of diabetes pathogenesis in the nonobese diabetic mouse. J Immunol 2005; 174: 7129–40
  • Hawkins T., Gala R. R., Dunbar J. C. The effect of neonatal sex hormone manipulation on the incidence of diabetes in nonobese diabetic mice. Proc Soc Exp Biol Med 1993; 202: 201–5
  • Stone J. L., Merriman B., Cantor R. M., Yonan A. L., Gilliam T. C., Geschwind D. H., et al. Evidence for sex‐specific risk alleles in autism spectrum disorder. Am J Hum Genet 2004; 75: 1117–23
  • Lamb J. A., Barnby G., Bonora E., Sykes N., Bacchelli E., Blasi F., et al. Analysis of IMGSAC autism susceptibility loci: evidence for sex limited and parent of origin specific effects. J Med Genet 2005; 42: 132–7
  • Abkevich V., Camp N. J., Hensel C. H., Neff C. D., Russell D. L., Hughes D. C., et al. Predisposition locus for major depression at chromosome 12q22‐12q23.2. Am J Hum Genet 2003; 73: 1271–81
  • Zubenko G. S., Maher B., Hughes H. B 3rd., Zubenko W. N., Stiffler J. S., Kaplan B. B., et al. Genome‐wide linkage survey for genetic loci that influence the development of depressive disorders in families with recurrent, early‐onset, major depression. Am J Med Genet B Neuropsychiatr Genet 2003; 123B: 1–18
  • Thomson P. A., Wray N. R., Millar J. K., Evans K. L., Hellard S. L., Condie A., et al. Association between the TRAX/DISC locus and both bipolar disorder and schizophrenia in the Scottish population. Mol Psychiatry 2005; 10: 657–68; 616
  • Hennah W., Varilo T., Kestila M., Paunio T., Arajarvi R., Haukka J., et al. Haplotype transmission analysis provides evidence of association for DISC1 to schizophrenia and suggests sex‐dependent effects. Hum Mol Genet 2003; 12: 3151–9
  • Weiss L. A., Pan L., Abney M., Ober C. The sex‐specific genetic architecture of quantitative traits in humans. Nat Genet 2006; 38: 218–22
  • Henikoff S., Matzke M. A. Exploring and explaining epigenetic effects. Trends Genet 1997; 13: 293–5
  • Jenuwein T., Allis C. D. Translating the histone code. Science 2001; 293: 1074–80
  • Takai D., Jones P. A. Comprehensive analysis of CpG islands in human chromosomes 21 and 22. Proc Natl Acad Sci U S A 2002; 99: 3740–5
  • Siegfried Z., Eden S., Mendelsohn M., Feng X., Tsuberi B. Z., Cedar H. DNA methylation represses transcription in vivo. Nat Genet 1999; 22: 203–6
  • Razin A., Shemer R. Epigenetic control of gene expression. Results Probl Cell Differ 1999; 25: 189–204
  • Ehrlich M., Ehrlich K. Effect of DNA methylation and the binding of vertebrate and plant proteins to DNA. DNA Methylation: Molecular Biology and Biological Significance, J Jost, P Saluz. Birkhauser Verlag, Basel, Switzerland 1993; 145–68
  • Jones P. L., Veenstra G. J., Wade P. A., Vermaak D., Kass S. U., Landsberger N., et al. Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat Genet 1998; 19: 187–91
  • Nan X., Ng H. H., Johnson C. A., Laherty C. D., Turner B. M., Eisenman R. N., et al. Transcriptional repression by the methyl‐CpG‐binding protein MeCP2 involves a histone deacetylase complex. Nature 1998; 393: 386–9
  • Riggs A. D., Xiong Z., Wang L., LeBon J. M. Methylation dynamics, epigenetic fidelity and X chromosome structure. Epigenetics 1998; 214–27, In: Wolffe A, ed. Chistester. John Wiley & Sons
  • Constancia M., Pickard B., Kelsey G., Reik W. Imprinting mechanisms. Genome Res 1998; 8: 881–900
  • Bestor T., Chandler V. L., Feinberg A. P. Epigenetic effects in eukaryotic gene expression. Develop Genet 1994; 15: 458
  • Riggs A., Porter T. Overview of epigenetic mehanisms. Epigenetic mechanisms of gene regulation, V. E. A Russo, R. A Martienssen, A. D Riggs. Cold Spring Harbor Laboratory Press, Plainview, N.Y., 29–45
  • Petronis A. Genomic imprinting in unstable DNA diseases. Bioessays 1996; 18: 587–90
  • Yang A. S., Jones P. A., Shibata A. The mutational burden of 5‐methylcytosine. Epigenetic mechanisms of gene regulation 1996; 77–94, In: Russo V. E. A, Martienssen R. A, Riggs A. D, eds. Plainview, N.Y. Cold Spring Harbor Laboratory Press
  • Bestor T. H. The DNA methyltransferases of mammals. Hum Mol Genet 2000; 9: 2395–402
  • Li E., Bestor T. H., Jaenisch R. Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 1992; 69: 915–26
  • Yeivin A., Razin A. Gene methylation patterns and expression. DNA Methylation: Molecular Biology and Biological Significance, J Jost, H Saluz. Birkhauser Verlag, Basel 1993; 523–68
  • Holliday R., Ho T., Paulin R. Gene silencing in mammalian cells. Epigenetic mechanisms of gene regulation, V. E. A Russo, R. A Martienssen, A. D Riggs. Cold Spring Harbor Laboratory Press, Plainview, N.Y. 1996; 47–59
  • Schotta G., Lachner M., Peters A. H., Jenuwein T. The indexing potential of histone lysine methylation. Novartis Found Symp 2004; 259: 22–37; discussion 37–47, 163–9
  • Wang Y., Fischle W., Cheung W., Jacobs S., Khorasanizadeh S., Allis C. D. Beyond the double helix: writing and reading the histone code. Novartis Found Symp 2004; 259: 3–17; discussion 17‐21, 163–9
  • Robertson K. D. DNA methylation and chromatin—unraveling the tangled web. Oncogene 2002; 21: 5361–79
  • Geiman T. M., Robertson K. D. Chromatin remodeling, histone modifications, and DNA methylation—how does it all fit together?. J Cell Biochem 2002; 87: 117–25
  • Li E. Chromatin modification and epigenetic reprogramming in mammalian development. Nat Rev Genet 2002; 3: 662–73
  • Druker R., Whitelaw E. Retrotransposon‐derived elements in the mammalian genome: a potential source of disease. J Inherit Metab Dis 2004; 27: 319–30
  • Ekwall K. The roles of histone modifications and small RNA in centromere function. Chromosome Res 2004; 12: 535–42
  • El‐Osta A., Wolffe A. P. DNA methylation and histone deacetylation in the control of gene expression: basic biochemistry to human development and disease. Gene Expr 2000; 9: 63–75
  • Dobosy J. R., Selker E. U. Emerging connections between DNA methylation and histone acetylation. Cell Mol Life Sci 2001; 58: 721–7
  • Baylin S. B., Esteller M., Rountree M. R., Bachman K. E., Schuebel K., Herman J. G. Aberrant patterns of DNA methylation, chromatin formation and gene expression in cancer. Hum Mol Genet 2001; 10: 687–92
  • Yamada Y., Watanabe H., Miura F., Soejima H., Uchiyama M., Iwasaka T., et al. A comprehensive analysis of allelic methylation status of CpG islands on human chromosome 21q. Genome Res 2004; 14: 247–66
  • Polesskaya O. O., Aston C., Sokolov B. P. Allele C‐specific methylation of the 5‐HT2A receptor gene: evidence for correlation with its expression and expression of DNA methylase DNMT1. J Neurosci Res 2006; 83: 362–73
  • Murrell A., Heeson S., Cooper W. N., Douglas E., Apostolidou S., Moore G. E., et al. An association between variants in the IGF2 gene and Beckwith‐Wiedemann syndrome: interaction between genotype and epigenotype. Hum Mol Genet 2004; 13: 247–55
  • Huang T. H., Perry M. R., Laux D. E. Methylation profiling of CpG islands in human breast cancer cells. Hum Mol Genet 1999; 8: 459–70
  • Shi H., Wei S. H., Leu Y. W., Rahmatpanah F., Liu J. C., Yan P. S., et al. Triple analysis of the cancer epigenome: an integrated microarray system for assessing gene expression, DNA methylation, and histone acetylation. Cancer Res 2003; 63: 2164–71
  • Yan P. S., Efferth T., Chen H. L., Lin J., Rodel F., Fuzesi L., et al. Use of CpG island microarrays to identify colorectal tumors with a high degree of concurrent methylation. Methods 2002; 27: 162–9
  • Yan P. S., Chen C. M., Shi H., Rahmatpanah F., Wei S. H., Huang T. H. Applications of CpG island microarrays for high‐throughput analysis of DNA methylation. J Nutr 2002; 132: 2430S–4S
  • Schumacher A., Kapranov P., Kaminsky Z., Flanagan J., Assadzadeh A., Yau P., et al. Microarray‐based DNA methylation profiling: technology and applications. Nucleic Acids Res 2006; 34: 528–42
  • Schumacher A., Petronis A. Epigenetics of Complex Disease: from the General Theory to Laboratory Experiment. Curr Top Microbiol Immunol 2006; 310: 81–115
  • Jablonka E., Lamb M. Epigenetic Inheritance and Evolution. Oxford University Press, Oxford; New York 1995
  • Flanagan J. M., Popendikyte V., Pozdniakovaite N., Sobolev M., Assadzadeh A., Schumacher A., et al. Intra‐ and Interindividual Epigenetic Variation in Human Germ Cells. Am J Hum Genet 2006; 79: 67–84
  • Sutherland J. E., Costa M. Epigenetics and the environment. Ann N Y Acad Sci 2003; 983: 151–60
  • Jaenisch R., Bird A. Epigenetic regulation of gene expression: how the genome integrates intrinsic and environmental signals. Nat Genet 2003; 33
  • Petronis A. The origin of schizophrenia: genetic thesis, epigenetic antithesis, and resolving synthesis. Biol Psychiatry 2004; 55: 965–70
  • Jantzen K., Fritton H. P., Igo‐Kemenes T., Espel E., Janich S., Cato A. C., et al. Partial overlapping of binding sequences for steroid hormone receptors and DNaseI hypersensitive sites in the rabbit uteroglobin gene region. Nucleic Acids Res 1987; 15: 4535–52
  • Truss M., Chalepakis G., Pina B., Barettino D., Bruggemeier U., Kalff M., et al. Transcriptional control by steroid hormones. J Steroid Biochem Mol Biol 1992; 41: 241–8
  • Csordas A., Puschendorf B., Grunicke H. Increased acetylation of histones at an early stage of oestradiol‐mediated gene activation in the liver of immature chicks. J Steroid Biochem 1986; 24: 437–42
  • Pasqualini J. R., Mercat P., Giambiagi N. Histone acetylation decreased by estradiol in the MCF‐7 human mammary cancer cell line. Breast Cancer Res Treat 1989; 14: 101–5
  • Saluz H. P., Jiricny J., Jost J. P. Genomic sequencing reveals a positive correlation between the kinetics of strand‐specific DNA demethylation of the overlapping estradiol/glucocorticoid‐receptor binding sites and the rate of avian vitellogenin mRNA synthesis. Proc Natl Acad Sci U S A 1986; 83: 7167–71
  • Yokomori N., Moore R., Negishi M. Sexually dimorphic DNA demethylation in the promoter of the Slp (sex‐limited protein) gene in mouse liver. Proc Natl Acad Sci U S A 1995; 92: 1302–6
  • Wilks A., Seldran M., Jost J. P. An estrogen‐dependent demethylation at the 5' end of the chicken vitellogenin gene is independent of DNA synthesis. Nucleic Acids Res 1984; 12: 1163–77
  • Meijlink F. C., Philipsen J. N., Gruber M., Ab G. Methylation of the chicken vitellogenin gene: influence of estradiol administration. Nucleic Acids Res 1983; 11: 1361–73
  • Wilks A. F., Cozens P. J., Mattaj I. W., Jost J. P. Estrogen induces a demethylation at the 5' end region of the chicken vitellogenin gene. Proc Natl Acad Sci U S A 1982; 79: 4252–5
  • Strahl B. D., Allis C. D. The language of covalent histone modifications. Nature 2000; 403: 41–5
  • Rice J. C., Allis C. D. Code of silence. Nature 2001; 414: 258–61
  • Nemeth A., Langst G. Chromatin higher order structure: opening up chromatin for transcription. Brief Funct Genomic Proteomic 2004; 2: 334–43
  • Kinyamu H. K., Archer T. K. Modifying chromatin to permit steroid hormone receptor‐dependent transcription. Biochim Biophys Acta 2004; 1677: 30–45
  • Fu M., Wang C., Zhang X., Pestell R. G. Acetylation of nuclear receptors in cellular growth and apoptosis. Biochem Pharmacol 2004; 68: 1199–208
  • Xu L., Glass C. K., Rosenfeld M. G. Coactivator and corepressor complexes in nuclear receptor function. Curr Opin Genet Dev 1999; 9: 140–7
  • Fu M., Rao M., Wang C., Sakamaki T., Wang J., Di Vizio D., et al. Acetylation of androgen receptor enhances coactivator binding and promotes prostate cancer cell growth. Mol Cell Biol 2003; 23: 8563–75
  • Gross D. S., Garrard W. T. Nuclease hypersensitive sites in chromatin. Annu Rev Biochem 1988; 57: 159–97
  • Cartwright I. L., Hertzberg R. P., Dervan P. B., Elgin S. C. Cleavage of chromatin with methidiumpropyl‐EDTA. iron(II). Proc Natl Acad Sci U S A 1983; 80: 3213–7
  • Lusser A., Kadonaga J. T. Chromatin remodeling by ATP‐dependent molecular machines. Bioessays 2003; 25: 1192–200
  • Harikrishnan K. N., Chow M. Z., Baker E. K., Pal S., Bassal S., Brasacchio D., et al. Brahma links the SWI/SNF chromatin‐remodeling complex with MeCP2‐dependent transcriptional silencing. Nat Genet 2005; 37: 254–64
  • Kiskinis E., Garcia‐Pedrero J. M., Villaronga M. A., Parker M. G., Belandia B. Identification of BAF57 mutations in human breast cancer cell lines. Breast Cancer Res Treat 2006; 98: 191–8
  • Kraus W. L., Wong J. Nuclear receptor‐dependent transcription with chromatin. Is it all about enzymes?. Eur J Biochem 2002; 269: 2275–83
  • Misiti S., Schomburg L., Yen P. M., Chin W. W. Expression and hormonal regulation of coactivator and corepressor genes. Endocrinology 1998; 139: 2493–500
  • Dubbink H. J., Hersmus R., Verma C. S., van der Korput H. A., Berrevoets C. A., van Tol J., et al. Distinct recognition modes of FXXLF and LXXLL motifs by the androgen receptor. Mol Endocrinol 2004; 18: 2132–50
  • Needham M., Raines S., McPheat J., Stacey C., Ellston J., Hoare S., et al. Differential interaction of steroid hormone receptors with LXXLL motifs in SRC‐1a depends on residues flanking the motif. J Steroid Biochem Mol Biol 2000; 72: 35–46
  • Baek S. H., Ohgi K. A., Nelson C. A., Welsbie D., Chen C., Sawyers C. L., et al. Ligand‐specific allosteric regulation of coactivator functions of androgen receptor in prostate cancer cells. Proc Natl Acad Sci U S A 2006; 103: 3100–5
  • Ko Y. J., Balk S. P. Targeting steroid hormone receptor pathways in the treatment of hormone dependent cancers. Curr Pharm Biotechnol 2004; 5: 459–70
  • Azzi L., El‐Alfy M., Labrie F. Gender differences and effects of sex steroids and dehydroepiandrosterone on androgen and oestrogen alpha receptors in mouse sebaceous glands. Br J Dermatol 2006; 154: 21–7
  • Liu J. C., Baker R. E., Chow W., Sun C. K., Elsholtz H. P. Epigenetic mechanisms in the dopamine D2 receptor‐dependent inhibition of the prolactin gene. Mol Endocrinol 2005; 19: 1904–17
  • Chakrabarti S. K., Francis J., Ziesmann S. M., Garmey J. C., Mirmira R. G. Covalent histone modifications underlie the developmental regulation of insulin gene transcription in pancreatic beta cells. J Biol Chem 2003; 278: 23617–23
  • Metivier R., Penot G., Hubner M. R., Reid G., Brand H., Kos M., et al. Estrogen receptor‐alpha directs ordered, cyclical, and combinatorial recruitment of cofactors on a natural target promoter. Cell 2003; 115: 751–63
  • Kulig E., Landefeld T. D., Lloyd R. V. The effects of estrogen on prolactin gene methylation in normal and neoplastic rat pituitary tissues. Am J Pathol 1992; 140: 207–14
  • Reddy P. M., Reddy P. R. Effect of prolactin on DNA methylation in the liver and kidney of rat. Mol Cell Biochem 1990; 95: 43–7
  • Li S., Hansman R., Newbold R., Davis B., McLachlan J. A., Barrett J. C. Neonatal diethylstilbestrol exposure induces persistent elevation of c‐fos expression and hypomethylation in its exon‐4 in mouse uterus. Mol Carcinog 2003; 38: 78–84
  • Mani S. T., Thakur M. K. In the cerebral cortex of female and male mice, amyloid precursor protein (APP) promoter methylation is higher in females and differentially regulated by sex steroids. Brain Res 2006; 1067: 43–7
  • Tekmal R. R., Kirma N., Gill K., Fowler K. Aromatase overexpression and breast hyperplasia, an in vivo model—continued overexpression of aromatase is sufficient to maintain hyperplasia without circulating estrogens, and aromatase inhibitors abrogate these preneoplastic changes in mammary glands. Endocr Relat Cancer 1999; 6: 307–14
  • Contractor R. G., Foran C. M., Li S., Willett K. L. Evidence of gender‐and tissue‐specific promoter methylation and the potential for ethinylestradiol‐induced changes in Japanese medaka (Oryzias latipes) estrogen receptor and aromatase genes. J Toxicol Environ Health A 2004; 67: 1–22
  • Petronis A., Gottesman I. I., Kan P., Kennedy J. L., Basile V. S., Paterson A. D., Popendikyte V. Monozygotic twins exhibit numerous epigenetic differences: clues to twin discordance?. Schizophr Bull 2003; 29: 169–78
  • Dahl C., Guldberg P. DNA methylation analysis techniques. Biogerontology 2003; 4: 233–50
  • Tost J., Schatz P., Schuster M., Berlin K., Gut I. G. Analysis and accurate quantification of CpG methylation by MALDI mass spectrometry. Nucleic Acids Res 2003; 31: e50
  • Sauer S., Lechner D., Berlin K., Plancon C., Heuermann A., Lehrach H., et al. Full flexibility genotyping of single nucleotide polymorphisms by the GOOD assay. Nucleic Acids Res 2000; 28: e100
  • Kaminsky Z. A., Assadzadeh A., Flanagan J., Petronis A. Single nucleotide extension technology for quantitative site‐specific evaluation of metC/C in GC‐rich regions. Nucleic Acids Res 2005; 33: e95
  • Puppo F., Musso M., Pirulli D., Griseri P., Bachetti T., Crovella S., et al. Comparative genomic sequence analysis coupled to chromatin immunoprecipitation: a screening procedure applied to search for regulatory elements at the RET locus. Physiol Genomics 2005; 23: 269–74
  • van Steensel B., Henikoff S. Epigenomic profiling using microarrays. Biotechniques 2003; 35: 346–50; 52–4; 56–7
  • Weber M., Davies J. J., Wittig D., Oakeley E. J., Haase M., Lam W. L., et al. Chromosome‐wide and promoter‐specific analyses identify sites of differential DNA methylation in normal and transformed human cells. Nat Genet 2005; 37: 853–62
  • Flanagan J., Petronis A. Pharmacoepigenetics: from the basic mechanisms to therapeutic applications. Pharmacogenomics, W Kalow, U. A Meyer, R Tyndale. Marcel Dekker, Inc., New York 2006; 461–90
  • Harrison's Principles of Internal Medicine. 15th ed, E Braunwald. McGraw‐Hill Medical Publishing, New York 2001
  • Wodarz N., Bobbe G., Eichhammer P., Weijers H. G., Wiesbeck G. A., Johann M. The candidate gene approach in alcoholism: are there gender‐specific differences?. Arch Women Ment Health 2003; 6: 225–30
  • Gunther C., von Hadeln K., Muller‐Thomsen T., Alberici A., Binetti G., Hock C., et al. Possible association of mitochondrial transcription factor A (TFAM) genotype with sporadic Alzheimer disease. Neurosci Lett 2004; 369: 219–23
  • Sweet R. A., Devlin B., Pollock B. G., Sukonick D. L., Kastango K. B., Bacanu S. A., et al. Catechol‐O‐methyltransferase haplotypes are associated with psychosis in Alzheimer disease. Mol Psychiatry 2005; 10: 1026–36
  • Szczeklik W., Sanak M., Szczeklik A. Functional effects and gender association of COX‐2 gene polymorphism G‐765C in bronchial asthma. J Allergy Clin Immunol 2004; 114: 248–53
  • Vaskuring A., Izakovicova Holla L., Vaskuring V. V., Tschoplova S., Stejskalova A. Polymorphisms in angiotensinogen gene (M235T and G(‐6)A) in multifactorial diseases. Pathophysiology 2001; 8: 113–8
  • Sutcliffe J. S., Delahanty R. J., Prasad H. C., McCauley J. L., Han Q., Jiang L., et al. Allelic heterogeneity at the serotonin transporter locus (SLC6A4) confers susceptibility to autism and rigid‐compulsive behaviors. Am J Hum Genet 2005; 77: 265–79
  • Chen Y., Vaughan R. W., Kondeatis E., Fortune F., Graham E. M., Stanford M. R., et al. Chemokine gene polymorphisms associate with gender in patients with uveitis. Tissue Antigens 2004; 63: 41–5
  • Underwood S. L., Christoforou A., Thomson P. A., Wray N. R., Tenesa A., Whittaker J., et al. Association analysis of the chromosome 4p‐located G protein‐coupled receptor 78 (GPR78) gene in bipolar affective disorder and schizophrenia. Mol Psychiatry 2006; 11: 384–94
  • Mukherjee M., Joshi S., Bagadi S., Dalvi M., Rao A., Shetty K. R. A low prevalence of the C677T mutation in the methylenetetrahydrofolate reductase gene in Asian Indians. Clin Genet 2002; 61: 155–9
  • de Andrade F. M., Silveira F. R., Arsand M., Antunes A. L., Torres M. R., Zago A. J., et al. Association between ‐250G/A polymorphism of the hepatic lipase gene promoter and coronary artery disease and HDL‐C levels in a Southern Brazilian population. Clin Genet 2004; 65: 390–5
  • McCarthy J. J., Meyer J., Moliterno D. J., Newby L. K., Rogers W. J., Topol E. J. Evidence for substantial effect modification by gender in a large‐scale genetic association study of the metabolic syndrome among coronary heart disease patients. Hum Genet 2003; 114: 87–98
  • Kantarci O. H., Goris A., Hebrink D. D., Heggarty S., Cunningham S., Alloza I., et al. IFNG polymorphisms are associated with gender differences in susceptibility to multiple sclerosis. Genes Immun 2005; 6: 153–61
  • Denys D., Van Nieuwerburgh F., Deforce D., Westenberg H. Association between the dopamine D(2) receptor TaqI A2 allele and low activity COMT allele with obsessive‐compulsive disorder in males. Eur Neuropsychopharmacol 2006; 16: 446–50
  • Villadsen M. M., Bunger M. H., Carstens M., Stenkjaer L., Langdahl B. L. Methylenetetrahydrofolate reductase (MTHFR) C677T polymorphism is associated with osteoporotic vertebral fractures, but is a weak predictor of BMD. Osteoporos Int 2005; 16: 411–6
  • Long J. R., Liu P. Y., Liu Y. J., Lu Y., Shen H., Zhao L. J., et al. APOE haplotypes influence bone mineral density in Caucasian males but not females. Calcif Tissue Int 2004; 75: 299–304
  • Maraganore D. M., Wilkes K., Lesnick T. G., Strain K. J., de Andrade M., Rocca W. A., et al. A limited role for DJ1 in Parkinson disease susceptibility. Neurology 2004; 63: 550–3
  • Qin W., Gao J., Xing Q., Yang J., Qian X., Li X., et al. A family‐based association study of PLP1 and schizophrenia. Neurosci Lett 2005; 375: 207–10
  • Gombos Z., Hermann R., Veijola R., Knip M., Simell O., Pollanen P., et al. Androgen receptor gene exon 1 CAG repeat polymorphism in Finnish patients with childhood‐onset type 1 diabetes. Eur J Endocrinol 2003; 149: 597–600
  • Galinsky D., Tysoe C., Brayne C. E., Easton D. F., Huppert F. A., Dening T. R., et al. Analysis of the apo E/apo C‐I, angiotensin converting enzyme and methylenetetrahydrofolate reductase genes as candidates affecting human longevity. Atherosclerosis 1997; 129: 177–83
  • Butt C., Zheng H., Randell E., Robb D., Parfrey P., Xie Y. G. Combined carrier status of prothrombin 20210A and factor XIII‐A Leu34 alleles as a strong risk factor for myocardial infarction: evidence of a gene‐gene interaction. Blood 2003; 101: 3037–41
  • Shearman A. M., Cupples L. A., Demissie S., Peter I., Schmid C. H., Karas R. H., et al. Association between estrogen receptor alpha gene variation and cardiovascular disease. JAMA 2003; 290: 2263–70
  • Petrkova J., Cermakova Z., Drabek J., Lukl J., Petrek M. CC chemokine receptor (CCR)2 polymorphism in Czech patients with myocardial infarction. Immunol Lett 2003; 88: 53–5
  • Weiss L. A., Abney M., Cook EH J. r., Ober C. Sex‐specific genetic architecture of whole blood serotonin levels. Am J Hum Genet 2005; 76: 33–41
  • Camp N. J., Lowry M. R., Richards R. L., Plenk A. M., Carter C., Hensel C. H., et al. Genome‐wide linkage analyses of extended Utah pedigrees identifies loci that influence recurrent, early‐onset major depression and anxiety disorders. Am J Med Genet B Neuropsychiatr Genet 2005; 135: 85–93
  • McGuffin P., Knight J., Breen G., Brewster S., Boyd P. R., Craddock N., et al. Whole genome linkage scan of recurrent depressive disorder from the depression network study. Hum Mol Genet 2005; 14: 3337–45
  • Ralston S. H., Galwey N., MacKay I., Albagha O. M., Cardon L., Compston J. E., et al. Loci for regulation of bone mineral density in men and women identified by genome wide linkage scan: the FAMOS study. Hum Mol Genet 2005; 14: 943–51
  • Peacock M., Koller D. L., Lai D., Hui S., Foroud T., Econs M. J. Sex‐specific quantitative trait loci contribute to normal variation in bone structure at the proximal femur in men. Bone 2005; 37: 467–73
  • Wilson S. G., Reed P. W., Bansal A., Chiano M., Lindersson M., Langdown M., et al. Comparison of genome screens for two independent cohorts provides replication of suggestive linkage of bone mineral density to 3p21 and 1p36. Am J Hum Genet 2003; 72: 144–55
  • Kammerer C. M., Schneider J. L., Cole S. A., Hixson J. E., Samollow P. B., O'Connell J. R., et al. Quantitative trait loci on chromosomes 2p, 4p, and 13q influence bone mineral density of the forearm and hip in Mexican Americans. J Bone Miner Res 2003; 18: 2245–52

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.