Which of the following statements best expresses the relationship between the nervous system and the endocrine system?

An endocrine-disrupting compound was defined by the U.S. Environmental Protection Agency (EPA) as “an exogenous agent that interferes with synthesis, secretion, transport, metabolism, binding action, or elimination of natural blood-borne hormones that are present in the body and are responsible for homeostasis, reproduction, and developmental process.” Our understanding of the mechanisms by which endocrine disruptors exert their effect has grown. Endocrine-disrupting chemicals (EDCs) were originally thought to exert actions primarily through nuclear hormone receptors, including estrogen receptors (ERs), androgen receptors (ARs), progesterone receptors, thyroid receptors (TRs), and retinoid receptors, among others. Today, basic scientific research shows that the mechanisms are much broader than originally recognized. Thus, endocrine disruptors act via nuclear receptors, nonnuclear steroid hormone receptors (e.g., membrane ERs), nonsteroid receptors (e.g., neurotransmitter receptors such as the serotonin receptor, dopamine receptor, norepinephrine receptor), orphan receptors [e.g., aryl hydrocarbon receptor (AhR)—an orphan receptor], enzymatic pathways involved in steroid biosynthesis and/or metabolism, and numerous other mechanisms that converge upon endocrine and reproductive systems. Thus, from a physiological perspective, an endocrine-disrupting substance is a compound, either natural or synthetic, which, through environmental or inappropriate developmental exposures, alters the hormonal and homeostatic systems that enable the organism to communicate with and respond to its environment.

The group of molecules identified as endocrine disruptors is highly heterogeneous and includes synthetic chemicals used as industrial solvents/lubricants and their byproducts [polychlorinated biphenyls (PCBs), polybrominated biphenyls (PBBs), dioxins], plastics [bisphenol A (BPA)], plasticizers (phthalates), pesticides [methoxychlor, chlorpyrifos, dichlorodiphenyltrichloroethane (DDT)], fungicides (vinclozolin), and pharmaceutical agents [diethylstilbestrol (DES)].

Natural chemicals found in human and animal food (e.g., phytoestrogens, including genistein and coumestrol) can also act as endocrine disruptors. These substances, whereas generally thought to have relatively low binding affinity to ERs, are widely consumed and are components of infant formula (1,2). A recent study reported that urinary concentrations of the phytoestrogens genistein and daidzein were about 500-fold higher in infants fed soy formula compared with those fed cow’s milk formula (3). Therefore, the potential for endocrine disruption by phytoestrogens needs to be considered.

A challenge to the field of endocrine disruption is that these substances are diverse and may not appear to share any structural similarity other than usually being small molecular mass (<1000 Daltons) compounds. Thus, it is difficult to predict whether a compound may or may not exert endocrine-disrupting actions. Nevertheless, in very broad terms, EDCs such as dioxins, PCBs, PBBs, and pesticides often contain halogen group substitutions by chlorine and bromine. They often have a phenolic moiety that is thought to mimic natural steroid hormones and enable EDCs to interact with steroid hormone receptors as analogs or antagonists. Even heavy metals and metalloids may have estrogenic activity, suggesting that these compounds are EDCs as well as more generalized toxicants. Several classes of EDCs act as antiandrogens and as thyroid hormone receptor agonists or antagonists, and more recently, androgenic EDCs have been identified.

The sources of exposure to EDCs are diverse and vary widely around the world. The situation is constantly evolving because some EDCs were banned decades ago and others more recently, with significant differences between countries. In this respect, migrating people provide a model to study cessation and/or onset of exposure depending on contamination of the original and new milieus. There are also several historical examples of toxic spills or contamination from PCBs and dioxins that show a direct causal relationship between a chemical and the manifestation of an endocrine or reproductive dysfunction in humans and wildlife. However, these types of single exposures are not representative of more common widespread persistent exposure to a broad mix of indoor and outdoor chemicals and contaminants. Industrialized areas are typically characterized by contamination from a wide range of industrial chemicals that may leach into soil and groundwater. These complex mixtures enter the food chain and accumulate in animals higher up the food chain such as humans, American bald eagles, polar bears, and other predatory animals. Exposure occurs through drinking contaminated water, breathing contaminated air, ingesting food, or contacting contaminated soil. People who work with pesticides, fungicides, and industrial chemicals are at particularly high risk for exposure and thus for developing a reproductive or endocrine abnormality.

Some EDCs were designed to have long half-lives; this was beneficial for their industrial use, but it has turned out to be quite detrimental to wildlife and humans. Because these substances do not decay easily, they may not be metabolized, or they may be metabolized or broken down into more toxic compounds than the parent molecule; even substances that were banned decades ago remain in high levels in the environment, and they can be detected as part of the body burden of virtually every tested individual animal or human (4,5). In fact, some endocrine disruptors are detectable in so-called “pristine” environments at remote distances from the site they were produced, used, or released due to water and air currents and via migratory animals that spend part of their life in a contaminated area, to become incorporated into the food chain in an otherwise uncontaminated region. Others, such as BPA, may not be as persistent [although recent evidence (e.g., Ref. 6) suggests longer half-lives) but are so widespread in their use that there is prevalent human exposure.

A number of issues have proven to be key to a full understanding of mechanisms of action and consequences of exposure to EDCs. These have been reviewed previously in detail (7), and several of them are listed here in brief.

Exposure of an adult to an EDC may have very different consequences from exposure to a developing fetus or infant. In fact, the field of endocrine disruption has embraced the terminology “the fetal basis of adult disease” (8) to describe observations that the environment of a developing organism, which includes the maternal environment (eutherian mammals), the egg (other vertebrates), and the external environment, interacts with the individual’s genes to determine the propensity of that individual to develop a disease or dysfunction later in life. In this Scientific Statement, we extend this concept beyond the fetal period to the early postnatal developmental period when organs continue to undergo substantial development. Thus, we will henceforward use the terminology “the developmental basis of adult disease.”

The developmental basis of adult disease also has implicit in its name the concept that there is a lag between the time of exposure and the manifestation of a disorder. In other words, consequences of developmental exposure may not be immediately apparent early in life but may be manifested in adulthood or during aging.

If individuals and populations are exposed to an EDC, it is likely that other environmental pollutants are involved because contamination of environments is rarely due to a single compound. Furthermore, effects of different classes of EDCs may be additive or even synergistic (9).

There are several properties of EDCs that have caused controversy. First, even infinitesimally low levels of exposure—indeed, any level of exposure at all—may cause endocrine or reproductive abnormalities, particularly if exposure occurs during a critical developmental window (10). Surprisingly, low doses may even exert more potent effects than higher doses. Second, EDCs may exert nontraditional dose-response curves, such as inverted-U or U-shaped curves (11). Both of these concepts have been known for hormone and neurotransmitter actions, but only in the past decade have they begun to be appreciated for EDCs.

EDCs may affect not only the exposed individual but also the children and subsequent generations. Recent evidence suggests that the mechanism of transmission may in some cases involve the germline (12) and may be nongenomic. That is, effects may be transmitted not due to mutation of the DNA sequence, but rather through modifications to factors that regulate gene expression such as DNA methylation and histone acetylation.

The field of endocrine disruption has particular pertinence to endocrinologists. In general, persistent endocrine disruptors have low water solubility and extremely high lipid solubility, leading to their bioaccumulation in adipose tissue. The properties of these substances are particularly well suited for study by endocrinologists because they so often activate or antagonize hormone receptors. There is no endocrine system that is immune to these substances, because of the shared properties of the chemicals and the similarities of the receptors (13) and enzymes involved in the synthesis, release, and degradation of hormones (Fig 1). Therefore, the role of this Scientific Statement is to provide perspectives on representative outcomes of exposures to endocrine disruptors and evidence for their effects in wildlife, laboratory animals, and humans.

Which of the following statements best expresses the relationship between the nervous system and the endocrine system?

Model of the endocrine systems targeted by endocrine-disrupting chemicals as discussed in this article. This figure demonstrates that all hormone-sensitive physiological systems are vulnerable to EDCs, including brain and hypothalamic neuroendocrine systems; pituitary; thyroid; cardiovascular system; mammary gland; adipose tissue; pancreas; ovary and uterus in females; and testes and prostate in males.

Development and function of the female reproductive tract depends on coordinated biological processes that, if altered by endogenous or exogenous factors during critical periods of development or during different life stage, could have significantly adverse effects on women’s health and reproductive function and outcomes. For example, the full complement of cell types in the human ovary depends on successful germ cell migration from the yolk sac during the first trimester and differentiation into oocytes with associated somatic cells to form the functional unit of the primordial follicle by the second to third trimesters of gestation. Factors that interfere with germ cell migration or follicle formation can result in abnormal functioning of this tissue with significant reproductive consequences. Also, the oocyte is arrested in the diplotene stage of late prophase until meitic divisions occur beginning at puberty (meiosis I) and after fertilization (meiosis II), and abnormalities in these processes can have a profound impact on reproductive outcomes, such as aneuploidy, premature ovarian failure (POF), and miscarriage. In addition, whereas Mullerian tract formation begins at 8 wk gestation with fusion of the Mullerian ducts and subsequent differentiation into the uterus (endometrium, myometrium), cervix, and upper vagina, uterine differentiation with regard to formation of luminal epithelium, glandular epithelium, and stromal components is mostly a postnatal event, with functionality of response to steroid hormones beginning at puberty. Interference with these processes can predispose women to infertility, ectopic gestation, poor pregnancy outcomes, and other reproductive disorders that may be programmed during development (e.g., endometriosis, uterine fibroids). Thus, abnormal development or alterations at other times in the life cycle can alter anatomy and functionality of the female reproductive tract and thus can alter the reproductive potential of affected individuals and their offspring.

Most female reproductive disorders are well described with regard to clinical presentation, histological evaluation of involved tissues where applicable, and diagnostic classification. However, whereas few are polygenic inherited traits and some are due to infections, the pathogenesis of the vast majority of female reproductive disorders is not well understood. This has hindered a preventive strategy to their development and/or exacerbation, and in some cases limited the development of effective therapies for symptoms and associated morbidities.

A key question arises as to whether EDCs contribute to the development of female reproductive disorders, particularly those occurring during a critical window of susceptibility: in utero, neonatally, in childhood, during puberty, and during adulthood. There are increasing data from wildlife studies and laboratory studies with rodents, ungulates, and nonhuman primates that support a role of EDCs in the pathogenesis of several female reproductive disorders, including polycystic ovarian syndrome, aneuploidy, POF, reproductive tract anomalies, uterine fibroids, endometriosis, and ectopic gestation (for reviews, see Refs. 29 and 49,50,51,52,53,54; also see Table 4). Many of the mechanisms are understood and, moreover, are conserved between animals and humans. Herein, we describe some of the clinical implications of these associations.

Female reproductive disorders and their possible relationships to EDCs: Some experimental and human data

Female reproductive disorderExperimental dataHuman epidemiological data
Reproductive tract abnormalities/ malignanciesMice prenatally exposed to DES have structural abnormalities of the oviduct, uterus, cervix, and vagina, leiomyoma, infertility-subfertility, immune dysfunction, ovarian cysts, ovarian tumors, vaginal adenocarcinoma (480)In utero exposure to DES: abnormal cervical, uterine, and oviduct anatomy (481), vaginal adenocarcinoma (19), subfertility and infertility, ectopic pregnancy (480)
EndometriosisAdult monkey exposed to TCDD (dioxin): promotion of growth and survival of endometriosis impants (110)↑ plasma concentrations of DEHP in women with endometriosis vs. controls (113); ↑ levels of phthalates (DnBP, BBP, DnOP, DEHP) in Indian women with endometriosis vs. controls (114)
Precocious pubertyImmature female rat exposed to DDT: sexual precocity (27)High levels of the DDT metabolite p,p′-DDE, in plasma from foreign immigrant girls with precocious puberty in Belgium (482)
Female mouse fetuses exposed to BPA: early puberty (474)Breastfed girls exposed to high levels of PBB in utero (≥7 ppm): earlier age at menarche (483)
Premature thelarcheHigher levels of phthalates and its major metabolite mono-(2-ethylhexyl) phthalate in serum of girls from Puerto Rico with premature breast development (26)
Disturbed lactationRodents exposed to atrazine: impaired lactation through prolactin inhibition (484)Negative correlation between DDE (metabolic product of DDT) and duration of lactation (484)
Breast abnormalities/cancerFetal rats exposed to dioxins (TCDD): altered breast development and ↑ susceptibility for mammary cancer (478)Limited and conflicting evidence
Mice exposed to BPA: altered organization of the mammary anlagen, accelerated ductal development, and inhibition of lumen formation in the fetus (128)M2 polymorphism in the cytochrome P450 1A1 gene modify the association between PCB exposure and risk of breast cancer (51)
Mice exposed to BPA: increased number of epithelial structures (145,146)
Rats exposed perinatally to BPA: development of preneoplastic lesions (intraductal hyperplasias) and carcinomas in situ (148)
Rats exposed perinatally to BPA; increased susceptibility to neoplastic development (149)
Rats: lactational exposure to BPA: shortening of the latency period and increased tumor multiplicity after carcinogen challenge (150)
Mice exposed to BPA: development of preneoplastic lesions (intraductal hyperplasias) (147)
PCOSPrenatal exposure to high levels of testosterone results in fetal programming of PCOS traits (60,61)Increased levels of serum AGEs in women with PCOS and positive correlation between AGE proteins and testosterone levels (64)
Rats fed with high vs. low AGE diet: ↑ androgens–↑ ovarian volume and AGE ovarian deposition (461)In polycystic ovaries, increased immunostaining of colocalized AGEs, RAGEs, and activated nuclear factor-κB (211,485)
Fertility and fecundityMice prenatally exposed to DES (480)Isolation of persistent organochlorine chemicals from ovarian follicular fluid of women undergoing IVF (51)
Indications that exposure to pesticides may contribute to female infertility in some occupationally exposed groups (484)

PCOS is a heterogeneous syndrome characterized by persistent anovulation, oligo- or amenorrhea, and hyperandrogenism in the absence of thyroid, pituitary, and/or adrenal disease (55,56,57). At the level of the ovary, there is recruitment and growth of follicles to the small antral stage, without selection of a dominant, preovulatory follicle, leading to accumulation of multiple, small, antral follicles (58). Hyperfunctioning of the theca and relative hypofunctioning of the granulosa cells accompany the acyclicity of the syndrome. Many, but not all women with PCOS have relatively high circulating levels of LH, compared with FSH, believed to be due to insensitivity to steroid hormone feedback. However, this does not fully account for the observed increase in thecal androgen production or the relative quiescence and sometimes frank FSH resistance of the granulosa cells. This complex disorder likely has its origins both within and outside the hypothalamic-pituitary-ovarian axis, and metabolic, neuroendocrine, and other endocrine regulators likely contribute to its manifestation. Obesity and insulin resistance occur in about 50% of women with PCOS, and obese women have a 12% risk of having PCOS (59). PCOS has multiple physiological processes (e.g., neuroendocrine functioning and feedback mechanisms, ovarian steroidogenesis, insulin resistance, and obesity) that are regulated by hormonal and metabolic parameters. Hence, endocrine disruption by environmental chemicals may indeed contribute to the pathogenesis of PCOS.

In sheep and rhesus monkeys, prenatal exposure to high levels of testosterone results in fetal programming of PCOS traits (60). Specifically, high levels of testosterone exposure at gestational d 40–60 and 100–115 result in rhesus monkey females who, in adulthood, have anovulatory infertility, hypersecretion of LH, elevated circulating levels of testosterone, neuroendocrine feedback defects, central adiposity and compensatory insulin resistance, and polycystic ovaries with ovarian hyperandrogenism and follicular arrest in adulthood (60,61). In the sheep model, a similar PCOS phenotype, along with IUGR and compensatory catch-up growth after birth, derives from prenatal exposure to exogenous testosterone (60,62). In rhesus monkey and sheep, unlike rodents, follicular differentiation is completed during fetal life. Thus, it is plausible that in utero exposure of human female fetuses to androgen-like EDCs could result in PCOS in adulthood, along with associated metabolic disorders. Very recent evidence for androgenic properties of personal-care products such as triclocarban (63) add to the possibility of environmental androgens, although a connection to PCOS has not yet been drawn.

There are numerous candidate genes associated with predisposition to developing PCOS in women (57,64), and how and if these interact with prenatal androgen-like factors to promote the PCOS phenotype in women has not been determined. Nonetheless, PCOS is a debilitating disorder in women, occurring in 6.6% of the reproductive-age population (65,66,67); it is a leading cause of subfertility and is associated with increased lifetime risks for cardiovascular disease and type II diabetes (55). In addition to these clinical impacts on patients, the cost to the health care system for PCOS diagnosis and treatment is substantial, totaling in 2004 about $4.4 billion in the United States alone (68). These facts underscore the need to understand potential EDC contributions to the development of PCOS in an effort to minimize such exposures and maximize prevention.

Other pathways may be involved in endocrine disruption of PCOS. Women with PCOS have higher levels of the EDC BPA (69), and increased testosterone in these women is consistent with decreased clearance of BPA (70). Although adult exposures do not necessarily imply earlier exposures in life, especially with EDCs of relatively short half-lives, there are data demonstrating nearly 5-fold higher levels of BPA in amniotic fluid compared with other body fluids, suggesting significant prenatal exposure (71). Although a cause and effect of BPA and PCOS have not been demonstrated definitively, the biological plausibility is interesting and worthy of further consideration.

POF (cessation of proper ovarian function before the age of 40) occurs in about 1% of reproductive-age women (72). Although in some cases the causation is known, for the vast majority of women with POF this is not the case, and there are stages of susceptibility during organogenesis and adult exposures that could contribute to POF.

Because the total ovarian follicle complement is established before birth in humans (73), anything that interferes with this, resulting in a decreased ovarian follicle resting pool, can result in POF. For example, disruption of germ cell migration from the genital ridge into the developing gonad results in ovarian dysgenesis. The resting pool undergoes a baseline level of apoptosis, and TNF-α, Fas ligand, and androgens stimulate this in the resting pool, as well as in the growing pool (74). Also, once a cohort of follicles is recruited during a given cycle in women, survival factors (FSH, estradiol, and growth factors, e.g., IGFs) are important for escape from apoptosis of the dominant follicle. Recent data in the mouse show that selective activation of the K-ras pathway in the oocyte results in rapid follicular development and depletion (75). Interestingly, adult and in utero exposures of mice to BPA have resulted in damage to oocytes (76,77). Specifically, adult exposures result in abnormalities in alignment of chromosomes on the meiotic spindle and aneuploidy, which, while not leading to ovarian senescence, does lead to aneuploid gametes and offspring (76). However, BPA given to pregnant dams during midgestation affects the developing ovary with resulting abnormalities in meiotic prophase, including synaptic defects, and mature animals exposed in utero have an increase in aneuploid oocytes and embryos (77). Such alterations also lead to cell cycle arrest and oocyte death, thus depleting the complement of normal oocytes (77). Currently, there are no data on in utero or adult exposure to BPA and aneuploidy in humans, but the possibility that there are parallels is compelling.

Interestingly, mice exposed in utero to DES, between d 9–16 gestation, have a dose-dependent decrease in reproductive capacity, including decreased numbers of litters and litter size and decreased numbers of oocytes (30%) ovulated in response to gonadotropin stimulation with all oocytes degenerating in the DES-exposed group, as well as numerous reproductive tract anatomic abnormalities (78). In women with in utero exposure to DES, Hatch et al. (79) reported an earlier age of menopause between the 43–55 yr olds, and the average age of menopause was 52.2 yr in unexposed women and 51.5 yr in exposed women. The effect of DES increased with cumulative doses and was highest in a cohort of highest in utero exposure during the 1950s (79). These observations are consistent with a smaller follicle pool and fewer oocytes ovulated, as in DES-exposed mice after ovulation induction (78).

Of interest are human data that demonstrate unequivocally that adult exposure in women to cigarette smoke results in decreased fecundity, decreased success rates in in vitro fertilization (IVF), decreased ovarian reserve (higher basal cycle d 3 FSH and stimulated parameters), earlier menopause by 1–4 yr, and an increased miscarriage rate (80,81). The mechanism appears to be mediated by the AhR-mediated apoptosis of oocytes, with accelerated loss of ovarian follicles. Interestingly, exposure of rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in utero and through the end of reproductive life results in a dose-dependent onset of premature reproductive senescence, likely due to direct effects on ovarian function (82).

Thus, whereas POF may occur in a relatively small percentage of the population, there are several alarming signals that should not be ignored. For example, the age group with the fastest growing rate of involuntary subfertility is 15- to 24-yr-old women (83). Also, the known effects of environmental contaminants on oocyte survival, aneuploidy, decreased ovarian reserve, and infertility described above underscore how much at risk the population may be for reproductive compromise.

With regard to ovarian granulosa steroidogenesis, several EDCs have effects on this process (84). For example, TCDD (10 ppm) decreases FSH-stimulated LH receptor mRNA expression and half-life in cultured granulosa (85). DDE increases vascular endothelial growth factor and IGF-I expression in luteinized granulosa from IVF patients, suggesting a contribution to impaired steroidogenesis and perhaps infertility (86). Recently, Kwintkiewicz and Giudice (87,88) have shown, in preliminary studies, that BPA decreases proliferation and FSH-induced aromatase expression via activation of peroxisome proliferator-activated receptor γ (PPAR-γ) and increases IGF-I and IGF receptor type I in human granulosa-like tumor cells and luteinized human granulosa from IVF subjects. These data suggest that EDCs may have local effects on ovarian function in adult women.

Disruption of female reproductive tract development by the EDC DES is well documented (89). A characteristic T-shaped uterus, abnormal oviductal anatomy and function, and abnormal cervical anatomy are characteristic of this in utero exposure, observed in adulthood (90), as well as in female fetuses and neonates exposed in utero to DES (91). Some of these effects are believed to occur through ERα (92) and abnormal regulation of Hox genes (93,94). Clinically, an increased risk of ectopic pregnancy, preterm delivery, miscarriage, and infertility all point to the devastating effect an endocrine disruptor may have on female fertility and reproductive health (89). It is certainly plausible that other EDCs with similar actions as DES could result in some cases of unexplained infertility, ectopic pregnancies, miscarriages, and premature deliveries. Although another major health consequence of DES exposure in utero was development of rare vaginal cancer in DES daughters, this may be an extreme response to the dosage of DES or specific to pathways activated by DES itself. Other EDCs may not result in these effects, although they may contribute to the fertility and pregnancy compromises cited above. Of utmost importance clinically is the awareness of DES exposure (and perhaps other EDC exposures) and appropriate physical exam, possible colposcopy of the vagina/cervix, cervical and vaginal cytology annually, and careful monitoring for fertility potential and during pregnancy for ectopic gestation and preterm delivery (89,95).

Uterine leiomyomas (fibroids) are benign smooth muscle tumors of the myometrium that can cause morbidity for women, including menorrhagia, abdominal pain, pelvic prolapse, and infertility and miscarriage (96). They are the most common tumor of the reproductive tract in women and comprise the leading cause for hysterectomy and the second leading cause of inpatient surgery in the United States, with health care costs exceeding $2 billion in 2004 (97). The prevalence rate of uterine leiomyomas is approximately 25–50%, with a preponderance occurring in African-American women (97). The greatest risk factor in adult women is prolonged exposure to unopposed estrogen. Whether in utero exposure to DES increases a woman’s lifetime risk of developing uterine fibroids is controversial, as the method to detect fibroids in two different studies influenced the outcome (98,99). Specifically, in a study of 1731 women exposed to DES and 848 matched unexposed controls, no association was found (P = 0.68) when histological confirmation after myomectomy or hysterectomy was used to document uterine fibroids (98). In contrast, when ultrasound was used to determine the presence of fibroids in DES-exposed vs. DES-unexposed women, a significant relationship was found (odds ratio, 2.4; 95% confidence interval, 1.1–5.4) in DES-exposed women and uterine fibroids (99). However, there are strong animal data to support development of uterine fibroids in adulthood after in utero exposure to EDCs, especially DES (for reviews, see Refs. 49, 50, and 52). Newbold et al. (100) reported that CD-1 mice develop uterine leiomyomas if exposed in utero or neonatally to DES, whereas unexposed mice do not. Furthermore, the Eker rat, which has a germ-line mutation in the rat homolog of the tuberus sclerosis complex 2 tumor suppressor gene, spontaneously develops uterine leiomyomas (101). The number, size, and growth rate of the fibroids increase significantly when the rat is exposed to DES on postnatal d 3–5 and 10–12, but not 17–19 (102), an effect that can be diminished with prior oophorectomy (102). These data overall strongly suggest developmental programming and gene-environment interactions for the increased risk of uterine lyomyomas in this rat model (103). In addition to mice, the Eker rat, and some dogs, the Baltic gray seal that has high organochlorine body burden also develops uterine leiomyomas (104). As with most environmental causes of abnormalities in the reproductive tract (and other tissues and organs), direct cause and effect relationships are difficult to establish. However, as in many of the other abnormalities in this Scientific Statement, the likelihood of such a relationship is plausible.

Endometriosis is an estrogen-dependent gynecological disorder associated with pelvic pain and infertility. It occurs in 6–10% of women and up to 50% of women with pelvic pain and infertility. In 2002, the total health care costs estimated in the United States for diagnosis and treatment of endometriosis totaled approximately $22 billion (105). There are suggestive animal data of adult exposure to EDCs and development of or exacerbation of existing disease, and there is evidence that in utero exposure in humans to DES results in an increased relative risk = 1.9 (95% confidence interval, 1.2–2.8) (106). Most striking are the observations of rhesus monkeys administered different doses of TCDD and their subsequent development of endometriosis (107,108). Although this study had low sample size and confounding variables that brought into question the relationship between endometriosis and TCDD (49,52,109), another study revealed that adult exposure of cynomolgus monkey to TCDD promotes growth and survival of endometriosis implants (110), indicating that this EDC is involved in the progression, if not pathogenesis, of this disorder. Similar data were obtained in rodent models of endometriosis in which human endometrium is transplanted into mouse and rat peritoneum, and the established lesions grew larger when animals were exposed to TCDD in utero and as adults (111,112), underscoring the estrogen (and EDC) dependence of this disorder.

There are also correlative findings of phthalate levels in plasma and endometriosis. For example, Cobellis et al. (113) found high plasma concentrations of di-(2-ethylhexyl)-phthalate in women with endometriosis, and an association of phthalate esters with endometriosis was found among Indian women (114). Thus, the evidence is accumulating of correlations between EDCs in the circulation of women with endometriosis, although a cause-and-effect relationship has yet to be established, which is not uncommon in reproductive environmental toxicity.

Endometriosis is believed to be due to retrograde menstruation and transplantation of endometrial fragments and cells into the peritoneal cavity. Because nearly all women have retrograde menstruation but relatively few have endometriosis, the disorder is also believed to involve a dysfunctional immune response, i.e., activated macrophages in the peritoneal cavity with robust secretion of inflammatory cytokines but without clearance of disease. An interesting model of early-life immune insult and developmental immunotoxicity suggests that in utero exposures to specific insults may reprogram the immune system, resulting in disorders such as chronic fatigue syndrome, cancer, and autoimmune disorders. Whether this has any relevance to the development or progression of endometriosis in adult women has not been explored but warrants further evaluation. Interestingly, TCDD and a therapy for endometriosis, danazol, both have effects on the adult immune system, although effects on the developing immune system are not known.

Although the infertility associated with endometriosis for the most part can be treated with advanced reproductive technologies, less success has been achieved with treatment of endometriosis-related pain. Because the pathogenesis of the associated pain is not known with certainty, therapies are empiric and include agents directed to minimize inflammation (nonsteroidal antiinflammatory drugs, danazol), progestins and androgens (to oppose estrogen actions), GnRH analogs (to inhibit gonadotropin secretion and thus ovarian estradiol production), and aromatase inhibitors (to inhibit estradiol synthesis by the ovary and endometriotic lesions), as well as surgical ablation or excision of the disease, when possible. Most of these therapies are effective in up to 50–60% of affected women, with either intolerable side effects (e.g., profound hypoestrogenism) or recurrence of pain (e.g., after surgery) (115). Thus, prevention is key to this disorder, as is understanding the pathogenesis so that therapies for pain can be devised appropriately and administered.

  • Kuiper GG, Lemmen JG, Carlsson B, Corton JC, Safe SH, van der Saag PT, van der Burg B, Gustafsson JA 1998 Interaction of estrogenic chemicals and phytoestrogens with estrogen receptor β. Endocrinology 139:4252–4263 [PubMed] [Google Scholar]
  • Dickerson SM, Gore AC 2007 Estrogenic environmental endocrine-disrupting chemical effects on reproductive neuroendocrine function and dysfunction across the life cycle. Rev Endocr Metab Disord 8:143–159 [PubMed] [Google Scholar]
  • Cao Y, Calafat AM, Doerge DR, Umbach DM, Bernbaum JC, Twaddle NC, Ye X, Rogan WJ 2009 Isoflavones in urine, saliva and blood of infants—data from a pilot study on the estrogenic activity of soy formula. J Expo Sci Environ Epidemiol 19:223–234 [PMC free article] [PubMed] [Google Scholar]
  • Calafat AM, Needham LL 2007 Human exposures and body burdens of endocrine-disrupting chemicals. In: Gore AC, ed. Endocrine-disrupting chemicals: from basic research to clinical practice. Totowa, NJ: Humana Press; 253–268 [Google Scholar]
  • Porte C, Janer G, Lorusso LC, Ortiz-Zarragoitia M, Cajaraville MP, Fossi MC, Canesi L 2006 Endocrine disruptors in marine organisms: approaches and perspectives. Comp Biochem Physiol C Toxicol Pharmacol 143:303–315 [PubMed] [Google Scholar]
  • Stahlhut RW, Welshons WV, Swan SH 2009 Bisphenol A data in NHANES suggest longer than expected half-life, substantial non-food exposure, or both. Environ Health Perspect 117:784–789 [PMC free article] [PubMed] [Google Scholar]
  • Gore AC, Crews D 2009 Environmental endocrine disruption of brain and behavior. In: Pfaff DW, Arnold AP, Etgen A, Fahrbach S, Rubin R, eds. Hormones, Brain and Behavior. San Diego, Academic Press, pp. 1789–1816 [Google Scholar]
  • Barker DJP 2003 The developmental origins of adult disease. Eur J Epidemiol 18:733–736 [PubMed] [Google Scholar]
  • Crews D, Putz O, Thomas P, Hayes T, Howdeshell K 2003 Animal models for the study of the effects of mixtures, low doses, and the embryonic environment on the action of endocrine disrupting chemicals. Pure and Applied Chemistry, SCOPE/IUPAC Project Implications of Endocrine Active Substances for Humans and Wildlife 75:2305–2320 [Google Scholar]
  • Sheehan DM, Willingham EJ, Bergeron JM, Osborn CT, Crews D 1999 No threshold dose for estradiol-induced sex reversal of turtle embryos: how little is too much? Environ Health Perspect 107:155–159 [PMC free article] [PubMed] [Google Scholar]
  • vom Saal FS, Akingbemi BT, Belcher SM, Birnbaum LS, Crain DA, Eriksen M, Farabollini F, Guillette Jr LJ, Hauser R, Heindel JJ, Ho SM, Hunt PA, Iguchi T, Jobling S, Kanno J, Keri RA, Knudsen KE, Laufer H, LeBlanc GA, Marcus M, McLachlan JA, Myers JP, Nadal A, Newbold RR, Olea N, Prins GS, Richter CA, Rubin BS, Sonnenschein C, Soto AM, Talsness CE, Vandenbergh JG, Vandenberg LN, Walser-Kuntz DR, Watson CS, Welshons WV, Wetherill Y, Zoeller RT 2007 Chapel Hill Bisphenol A Expert Panel Consensus Statement: integration of mechanisms, effects in animals and potential to impact human health at current levels of exposure. Reprod Toxicol 24:131–138 [PMC free article] [PubMed] [Google Scholar]
  • Anway MD, Skinner MK 2006 Epigenetic transgenerational actions of endocrine disruptors. Endocrinology 147:S43–S49 [PubMed] [Google Scholar]
  • Thornton JW 2001 Evolution of vertebrate steroid receptors from an ancestral estrogen receptor by ligand exploitation and serial genome expansions. Proc Natl Acad Sci USA 98:5671–5676 [PMC free article] [PubMed] [Google Scholar]
  • Bay K, Asklund C, Skakkebaek NE, Andersson AM 2006 Testicular dysgenesis syndrome: possible role of endocrine disrupters. Best Pract Res Clin Endocrinol Metab 20:77–90 [PubMed] [Google Scholar]
  • Toppari J, Kaleva M, Virtanen HE 2001 Trends in the incidence of cryptorchidism and hypospadias, and methodological limitations of registry-based data. Hum Reprod Update 7:282–286 [PubMed] [Google Scholar]
  • Hemminki K, Li X 2002 Cancer risks in second-generation immigrants to Sweden. Int J Cancer 99:229–237 [PubMed] [Google Scholar]
  • Parent AS, Teilmann G, Juul A, Skakkebaek NE, Toppari J, Bourguignon JP 2003 The timing of normal puberty and the age limits of sexual precocity: variations around the world, secular trends and changes after migration. Endocr Rev 24:668–693 [PubMed] [Google Scholar]
  • Ibáñez L, de Zegher F 2006 Puberty and prenatal growth. Mol Cell Endocrinol 254–255:22–25 [PubMed] [Google Scholar]
  • Herbst AL, Ulfelder H, Poskanzer DC 1971 Adenocarcinoma of vagina. Association of maternal stilbestrol therapy with tumor appearance in young women. N Engl J Med 284:878–881 [PubMed] [Google Scholar]
  • Sharpe RM 2006 Pathways of endocrine disruption during male sexual differentiation and masculinisation. Best Pract Res Clin Endocrinol Metab 20:91–110 [PubMed] [Google Scholar]
  • Skakkebaek NE, Rajpert-De Meyts E, Main KM 2001 Testicular dysgenesis syndrome: an increasingly common developmental disorder with environmental aspects. Hum Reprod 16:972–978 [PubMed] [Google Scholar]
  • Gray Jr LE, Ostby J, Furr J, Price M, Veeramachaneni DN, Parks L 2000 Perinatal exposure to the phthalates DEHP, BBP, and DINP, but not DEP, DMP, or DOTP, alters sexual differentiation of the male rat. Toxicol Sci 58:350–365 [PubMed] [Google Scholar]
  • Swan SH, Main KM, Liu F, Stewart SL, Kruse RL, Calafat AM, Mao CS, Redmon JB, Ternand CL, Sullivan S, Teague JL 2005 Decrease in anogenital distance among male infants with prenatal phthalate exposure. Environ Health Perspect 113:1056–1061 [PMC free article] [PubMed] [Google Scholar]
  • Maffini M, Rubin B, Sonnenschein C, Soto A 2006 Endocrine disruptors and reproductive health: the case of bisphenol-A. Mol Cell Endocrinol 254–255:179–186 [PubMed] [Google Scholar]
  • Buck Louis GM, Gray Jr LE, Marcus M, Ojeda SR, Pescovitz OH, Witchel SF, Sippell W, Abbott DH, Soto A, Tyl RW, Bourguignon JP, Skakkebaek NE, Swan SH, Golub MS, Wabitsch M, Toppari J, Euling SY 2008 Environmental factors and puberty timing: expert panel research needs. Pediatrics 121:S192–S207 [PubMed] [Google Scholar]
  • Colón I, Caro D, Bourdony CJ, Rosario O 2000 Identification of phthalate esters in the serum of young Puerto Rican girls with premature breast development. Environ Health Perspect 108:895–900 [PMC free article] [PubMed] [Google Scholar]
  • Rasier G, Parent AS, Gérard A, Lebrethon MC, Bourguignon JP 2007 Early maturation of gonadotropin-releasing hormone secretion and sexual precocity after exposure of infantile female rats to estradiol or dichlorodiphenyltrichloroethane. Biol Reprod 77:734–742 [PubMed] [Google Scholar]
  • Li S, Hursting SD, Davis BJ, McLachlan JA, Barrett JC 2003 Environmental exposure, DNA methylation and gene regulation. Lessons from diethylstilbestrol-induced cancers. Ann NY Acad Sci 983:161–169 [PubMed] [Google Scholar]
  • McLachlan JA, Simpson E, Martin M 2006 Endocrine disrupters and female reproductive health. Best Pract Res Clin Endocrinol Metab 20:63–75 [PubMed] [Google Scholar]
  • Darbre PD 2006 Environmental oestrogens, cosmetics and breast cancer. Best Pract Res Clin Endocrinol Metab 20:121–143 [PubMed] [Google Scholar]
  • Fenton SE 2006 Endocrine-disrupting compounds and mammary gland development: Early exposure and later life consequences. Endocrinology 147:S18–S24 [PubMed] [Google Scholar]
  • Newbold RR, Jefferson WN, Padilla-Banks E 2007 Long-term adverse effects of neonatal exposure to bisphenol A on the murine female reproductive tract. Reprod Toxicol 24:253–258 [PMC free article] [PubMed] [Google Scholar]
  • Crews D, McLachlan JA 2006 Epigenetics, evolution, endocrine disruption, health, and disease. Endocrinology 147:S4–S10 [PubMed] [Google Scholar]
  • Anway MD, Skinner MK 2008 Transgenerational effects of the endocrine disruptor vinclozolin on the prostate transcriptome and adult onset disease. Prostate 68:517–529 [PMC free article] [PubMed] [Google Scholar]
  • Anway MD, Cupp AS, Uzumcu M, Skinner MK 2005 Epigenetic transgenerational actions of endocrine disruptors and male fertility. Science 308:1466–1469 [PubMed] [Google Scholar]
  • Christiansen S, Scholze M, Axelstad M, Boberg J, Kortenkamp A, Hass U 2008 Combined exposure to anti-androgens causes markedly increased frequencies of hypospadias in the rat. Int J Androl 31:241–248 [PubMed] [Google Scholar]
  • Shono T, Suita S, Kai H, Yamaguchi Y 2004 Short-time exposure to vinclozolin in utero induces testicular maldescent associated with a spinal nucleus alteration of the genitofemoral nerve in rat. J Pediatr Surg 39:217–219 [PubMed] [Google Scholar]
  • Monosson E, Kelce WR, Lambright C, Ostby J, Gray Jr LE 1999 Peripubertal exposure to the antiandrogenic fungicide, vinclozolin, delays puberty, inhibits the development of androgen-dependent tissues, and alters androgen receptor function in the male rat. Toxicol Ind Health 15:65–79 [PubMed] [Google Scholar]
  • Newbold RR, Hanson RB, Jefferson WN, Bullock BC, Haseman J, McLachlan JA 1998 Increased tumors but uncompromised fertility in the female descendants of mice exposed developmentally to diethylstilbestrol. Carcinogenesis 19:1655–1663 [PubMed] [Google Scholar]
  • Kadlubar FF, Berkowitz GS, Delongchamp RR, Wang C, Green BL, Tang G, Lamba J, Schuetz E, Wolff MS 2003 The CYP3A4*1B variant is related to the onset of puberty, a known risk factor for the development of breast cancer. Cancer Epidemiol Biomarkers Prev 12:327–331 [PubMed] [Google Scholar]
  • Rivas A, Fisher JS, McKinnell C, Atanassova N, Sharpe RM 2002 Induction of reproductive tract developmental abnormalities in the male rat by lowering androgen production or action in combination with a low dose of diethylstilbestrol: evidence for importance of the androgen-estrogen balance. Endocrinology 143:4797–4808 [PubMed] [Google Scholar]
  • Ceccarelli I, Della Seta D, Fiorenzani P, Farabollini F, Aloisi AM 2007 Estrogenic chemicals at puberty change ERα in the hypothalamus of male and female rats. Neurotoxicol Teratol 29:108–115 [PubMed] [Google Scholar]
  • Atanassova N, McKinnell C, Williams K, Turner KJ, Fisher JS, Saunders PT, Millar MR, Sharpe RM 2001 Age-, cell- and region- specific immunoexpression of estrogen receptor α (but not estrogen receptor β) during postnatal development of the epididymis and vas deferens of the rat and disruption of this pattern by neonatal treatment with diethylstilbestrol. Endocrinology 142:874–886 [PubMed] [Google Scholar]
  • Khurana S, Ranmal S, Ben-Jonathan N 2000 Exposure of newborn male and female rats to environmental estrogens: delayed and sustained hyperprolactinemia and alterations in estrogen receptor expression. Endocrinology 141:4512–4517 [PubMed] [Google Scholar]
  • Kester MH, Bulduk S, van Toor H, Tibboel D, Meinl W, Glatt H, Falany CN, Coughtrie MW, Schuur AG, Brouwer A, Visser TJ 2002 Potent inhibition of estrogen sulfotransferase by hydroxylated metabolites of polyhalogenated aromatic hydrocarbons reveals alternative mechanism for estrogenic activity of endocrine disrupters. J Clin Endocrinol Metab 87:1142–1150 [PubMed] [Google Scholar]
  • Waring RH, Harris RM 2005 Endocrine disrupters: a human risk? Mol Cell Endocrinol 244:2–9 [PubMed] [Google Scholar]
  • Odum J, Tinwell H, Tobin G, Ashby J 2004 Cumulative dietary energy intake determines the onset of puberty in female rats. Environ Health Perspect 112:1472–1480 [PMC free article] [PubMed] [Google Scholar]
  • Kortenkamp A 2008 Low dose mixture effects of endocrine disrupters: implications for risk assessment and epidemiology. Int J Androl 31:233–240 [PubMed] [Google Scholar]
  • Woodruff TJ, Carlson A, Schwartz JM, Giudice LC 2008 Proceedings of the Summit on Environmental Challenges to Reproductive Health and Fertility: Executive Summary. Fertil Steril 89:281–300 [PMC free article] [PubMed] [Google Scholar]
  • Woodruff TK, Walker CL 2008 Fetal and early postnatal environmental exposures and reproductive health effects in the female. Fertil Steril 89:e47–e51 [PMC free article] [PubMed] [Google Scholar]
  • Caserta D, Maranghi L, Mantovani A, Marci R, Maranghi F, Moscarini M 2008 Impact of endocrine disruptor chemicals in gynaecology. Hum Reprod Update 14:59–72 [PubMed] [Google Scholar]
  • Crain DA, Janssen SJ, Edwards TM, Heindel J, Ho SM, Hunt P, Iguchi T, Juul A, McLachlan JA, Schwartz J, Skakkebaek N, Soto AM, Swan S, Walker C, Woodruff TK, Woodruff TJ, Giudice LC, Guillette Jr LJ 2008 Female reproductive disorders: the roles of endocrine-disrupting compounds and developmental timing. Fertil Steril 90:911–940 [PMC free article] [PubMed] [Google Scholar]
  • Foster WG, Neal MS, Han MS, Dominguez MM 2008 Environmental contaminants and human infertility: hypothesis or cause for concern. J Toxicol Environ Health B Crit Rev 11:162–176 [PubMed] [Google Scholar]
  • Mendola P, Messer LC, Rappazzo K 2008 Science linking environmental contaminant exposures with fertility and reproductive health impacts in the adult female. Fertil Steril 89:e81–e94 [PubMed] [Google Scholar]
  • The Rotterdam ESHRE/ASRM-Sponsored PCOS Consensus Workshop Group 2004 Revised 2003 consensus on diagnostic criteria and long-term health risks related to polycystic ovary syndrome (PCOS). Hum Reprod 19:41–47 [PubMed] [Google Scholar]
  • Azziz R, Carmina E, Dewailly D, Diamanti-Kandarakis E, Escobar-Morreale HF, Futterweit W, Janssen OE, Legro RS, Norman RJ, Taylor AE, Witchel SF 2006 Position statement: criteria for defining polycystic ovary syndrome as a predominantly hyperandrogenic syndrome. An Androgen Excess Society guideline. J Clin Endocrinol Metab 91:4237–4245 [PubMed] [Google Scholar]
  • Legro RS, Azziz R, Giudice L 2006 A twenty-first century research agenda for polycystic ovary syndrome. Best Pract Res Clin Endocrinol Metab 20:331–336 [PubMed] [Google Scholar]
  • Franks S, Mason H, Willis D 2000 Follicular dynamics in the polycystic ovary syndrome. Mol Cell Endocrinol 163:49–52 [PubMed] [Google Scholar]
  • Yildiz BO, Knochenhauer ES, Azziz R 2008 Impact of obesity on the risk for polycystic ovary syndrome. J Clin Endocrinol Metab 93:162–168 [PMC free article] [PubMed] [Google Scholar]
  • Dumesic DA, Abbott DH, Padmanabhan V 2007 Polycystic ovary syndrome and its developmental origins. Rev Endocr Metab Disord 8:127–141 [PMC free article] [PubMed] [Google Scholar]
  • Abbott DH, Barnett DK, Bruns CM, Dumesic DA 2005 Androgen excess fetal programming of female reproduction: a developmental aetiology for polycystic ovary syndrome? Hum Reprod Update 11:357–374 [PubMed] [Google Scholar]
  • West C, Foster DL, Evans NP, Robinson J, Padmanabhan V 2001 Intra-follicular activin availability is altered in prenatally-androgenized lambs. Mol Cell Endocrinol 185: 51–59 [PubMed] [Google Scholar]
  • Chen J, Ahn KC, Gee NA, Ahmed MI, Duleba AJ, Zhao L, Gee SJ, Hammock BD, Lasley BL 2008 Triclocarban enhances testosterone action: a new type of endocrine disruptor? Endocrinology 149:1173–1179 [PMC free article] [PubMed] [Google Scholar]
  • Diamanti-Kandarakis E, Piperi C 2005 Genetics of polycystic ovary syndrome: searching for the way out of the labyrinth. Hum Reprod Update 11:631–643 [PubMed] [Google Scholar]
  • Azziz R, Woods KS, Reyna R, Key TJ, Knochenhauer ES, Yildiz BO 2004 The prevalence and features of the polycystic ovary syndrome in an unselected population. J Clin Endocrinol Metab 89:2745–2749 [PubMed] [Google Scholar]
  • Diamanti-Kandarakis E, Kouli CR, Bergiele AT, Filandra FA, Tsianateli TC, Spina GG, Zapanti ED, Bartzis MI 1999 A survey of the polycystic ovary syndrome in the Greek island of Lesbos: hormonal and metabolic profile. J Clin Endocrinol Metab 84:4006–4011 [PubMed] [Google Scholar]
  • Asunción M, Calvo RM, San Millán JL, Sancho J, Avila S, Escobar-Morreale HF 2000 A prospective study of the prevalence of the polycystic ovary syndrome in unselected Caucasian women from Spain. J Clin Endocrinol Metab 85:2434–2438 [PubMed] [Google Scholar]
  • Azziz R, Marin C, Hoq L, Badamgarav E, Song P 2005 Health care-related economic burden of the polycystic ovary syndrome during the reproductive life span. J Clin Endocrinol Metab 90:4650–4658 [PubMed] [Google Scholar]
  • Takeuchi T, Tsutsumi O, Ikezuki Y, Takai Y, Taketani Y 2004 Positive relationship between androgen and the endocrine disruptor, bisphenol A, in normal women and women with ovarian dysfunction. Endocr J 51:165–169 [PubMed] [Google Scholar]
  • Takeuchi T, Tsutsumi O, Ikezuki Y, Kamei Y, Osuga Y, Fujiwara T, Takai Y, Momoeda M, Yano T, Taketani Y 2006 Elevated serum bisphenol A levels under hyperandrogenic conditions may be caused by decreased UDP-glucuronsyltransferase activity. Endocr J 53:485–491 [PubMed] [Google Scholar]
  • Ikezuki Y, Tsutsumi O, Takai Y, Kamei Y, Taketani Y 2002 Determination of bisphenol A concentrations in human biological fluids reveals significant early prenatal exposure. Hum Reprod 17:2839–2841 [PubMed] [Google Scholar]
  • Sinha P, Kuruba N 2007 Premature ovarian failure. J Obstet Gynaecol 27:16–19 [PubMed] [Google Scholar]
  • McNatty KP, Reader K, Smith P, Heath DA, Juengel JL 2007 Control of ovarian follicular development to the gonadotrophin-dependent phase: a 2006 perspective. Soc Reprod Fertil Suppl 64:55–68 [PubMed] [Google Scholar]
  • Kaipia A, Hsueh AJ 1997 Regulation of ovarian follicle atresia. Annu Rev Physiol 59:349–363 [PubMed] [Google Scholar]
  • Reddy P, Liu L, Adhikari D, Jagarlamudi K, Rajareddy S, Shen Y, Du C, Tang W, HämäläinenT, Peng SL, Lan ZJ, Cooney AJ, Huhtaniemi I, Liu K 2008 Oocyte-specific deletion of Pten causes premature activation of the primordial follicle pool. Science 319:611–613 [PubMed] [Google Scholar]
  • Hunt PA, Koehler KE, Susiarjo M, Hodges CA, Ilagan A, Voigt RC, Thomas S, Thomas BF, Hassold TJ 2003 Bisphenol A exposure causes meiotic aneuploidy in the female mouse. Curr Biol 13:546–553 [PubMed] [Google Scholar]
  • Susiarjo M, Hassold TJ, Freeman E, Hunt PA 2007 Bisphenol A exposure in utero disrupts early oogenesis in the mouse. PLoS Genet 3:e5 [PMC free article] [PubMed] [Google Scholar]
  • McLachlan JA, Newbold RR, Shah HC, Hogan MD, Dixon RL 1982 Reduced fertility in female mice exposed transplacentally to diethylstilbestrol (DES). Fertil Steril 38:364–371 [PubMed] [Google Scholar]
  • Hatch EE, Troisi R, Wise LA, Hyer M, Palmer JR, Titus-Ernstoff L, Strohsnitter W, Kaufman R, Adam E, Noller KL, Herbst AL, Robboy S, Hartge P, Hoover RN 2006 Age at natural menopause in women exposed to diethylstilbestrol in utero. Am J Epidemiol 164:682–688 [PubMed] [Google Scholar]
  • Sharara FI, Seifer DB, Flaws JA 1998 Environmental toxicants and female reproduction. Fertil Steril 70:613–622 [PubMed] [Google Scholar]
  • Genuis SJ 2006 Health issues and the environment—an emerging paradigm for providers of obstetrical and gynaecological health care. Hum Reprod 21:2201–2208 [PubMed] [Google Scholar]
  • Shi Z, Valdez KE, Ting AY, Franczak A, Gum SL, Petroff BK 2007 Ovarian endocrine disruption underlies premature reproductive senescence following environmentally relevant chronic exposure to the aryl hydrocarbon receptor agonist 2,3,7,8-tetrachlorodibenzo-p-dioxin. Biol Reprod 76:198–202 [PubMed] [Google Scholar]
  • Chandra A, Martinez GM, Mosher WD, Abma JC, Jones J 2005 Fertility, family planning, and reproductive health of U.S. women: data from the 2002 National Survey of Family Growth. Vital Health Stat 23:1–160 [PubMed] [Google Scholar]
  • Kwintkiewicz J, Giudice LC 2009 The interplay of insulin-like growth factors, gonadotropins, and endocrine disruptors in ovarian follicular development and function. Semin Reprod Med 27:43–51 [PubMed] [Google Scholar]
  • Minegishi T, Hirakawa T, Abe K, Kishi H, Miyamoto K 2003 Effect of IGF-1 and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on the expression of LH receptors during cell differentiation in cultured granulosa cells. Mol Cell Endocrinol 202:123–131 [PubMed] [Google Scholar]
  • Holloway AC, Petrik JJ, Younglai EV 2007 Influence of dichlorodiphenylchloroethylene on vascular endothelial growth factor and insulin-like growth factor in human and rat ovarian cells. Reprod Toxicol 24:359–364 [PubMed] [Google Scholar]
  • Kwintkiewicz J, Giudice LC, Endocrine disruptor bisphenol A reduces FSH-stimulated cyp19 expression and downstream estradiol production in human granulosa KGN cells. Proc 55th Annual Scientific Meeting of the Society for Gynecologic Investigation, San Diego, CA, 2008 (Abstract) [Google Scholar]
  • Kwintkiewicz J, Giudice LC, Endocrine disruptor bisphenol A induces expression of peroxisome proliferators-activated receptor which contributes to down-regulation of FSH-stimulated aromatase expression and estradiol production in human granulosa KGN cells. Proc 41st Annual Meeting of the Society for the Study of Reproduction, Kona, HI, 2008 (Abstract) [Google Scholar]
  • Swan SH 2000 Intrauterine exposure to diethylstilbestrol: long-term effects in humans. APMIS 108:793–804 [PubMed] [Google Scholar]
  • Jefferies JA, Robboy SJ, O'Brien PC, Bergstralh EJ, Labarthe DR, Barnes AB, Noller KL, Hatab PA, Kaufman RH, Townsend DE 1984 Structural anomalies of the cervix and vagina in women enrolled in the Diethylstilbestrol Adenosis (DESAD) Project. Am J Obstet Gynecol 148:59–66 [PubMed] [Google Scholar]
  • Johnson LD, Driscoll SG, Hertig AT, Cole PT, Nickerson RJ 1979 Vaginal adenosis in stillborns and neonates exposed to diethylstilbestrol and steroidal estrogens and progestins. Obstet Gynecol 53:671–679 [PubMed] [Google Scholar]
  • Couse JF, Dixon D, Yates M, Moore AB, Ma L, Maas R, Korach KS 2001 Estrogen receptor-α knockout mice exhibit resistance to the developmental effects of neonatal diethylstilbestrol exposure on the female reproductive tract. Dev Biol 238:224–238 [PubMed] [Google Scholar]
  • Ma L, Benson GV, Lim H, Dey SK, Maas RL 1998 Abdominal B (AbdB) Hoxa genes: regulation in adult uterus by estrogen and progesterone and repression in müllerian duct by the synthetic estrogen diethylstilbestrol (DES). Dev Biol 197:141–154 [PubMed] [Google Scholar]
  • Block K, Kardana A, Igarashi P, Taylor HS 2000 In utero diethylstilbestrol (DES) exposure alters Hox gene expression in the developing müllerian system. FASEB J 14:1101–1108 [PubMed] [Google Scholar]
  • Schrager S, Potter BE 2004 Diethylstilbestrol exposure. Am Fam Physician 69:2395–2400 [PubMed] [Google Scholar]
  • Buttram Jr VC, Reiter RC 1981 Uterine leiomyomata: etiology, symptomatology, and management. Fertil Steril 36:433–445 [PubMed] [Google Scholar]
  • Agency for Healthcare Research and Quality 2007 Management of uterine fibroids: an update of the evidence. Publication no. 07-E011. Bethesda, MD: National Library of Medicine [Google Scholar]
  • Wise LA, Palmer JR, Rowlings K, Kaufman RH, Herbst AL, Noller KL, Titus-Ernstoff L, Troisi R, Hatch EE, Robboy SJ 2005 Risk of benign gynecologic tumors in relation to prenatal diethylstilbestrol exposure. Obstet Gynecol 105:167–173 [PubMed] [Google Scholar]
  • Baird DD, Newbold R 2005 Prenatal diethylstilbestrol (DES) exposure is associated with uterine leiomyoma development. Reprod Toxicol 20:81–84 [PubMed] [Google Scholar]
  • Newbold RR, Moore AB, Dixon D 2002 Characterization of uterine leiomyomas in CD-1 mice following developmental exposure to diethylstilbestrol (DES). Toxicol Pathol 30:611–616 [PubMed] [Google Scholar]
  • Everitt JI, Wolf DC, Howe SR, Goldsworthy TL, Walker C 1995 Rodent model of reproductive tract leiomyomata. Clinical and pathological features. Am J Pathol 146:1556–1567 [PMC free article] [PubMed] [Google Scholar]
  • Cook JD, Davis BJ, Goewey JA, Berry TD, Walker CL 2007 Identification of a sensitive period for developmental programming that increases risk for uterine leiomyoma in Eker rats. Reprod Sci 14:121–136 [PubMed] [Google Scholar]
  • Cook JD, Davis BJ, Cai SL, Barrett JC, Conti CJ, Walker CL 2005 Interaction between genetic susceptibility and early-life environmental exposure determines tumor-suppressor-gene penetrance. Proc Natl Acad Sci USA 102: 8644–8649 [PMC free article] [PubMed] [Google Scholar]
  • Bäcklin BM, Eriksson L, Olovsson M 2003 Histology of uterine leiomyoma and occurrence in relation to reproductive activity in the Baltic gray seal (Halichoerus grypus). Vet Pathol 40:175–180 [PubMed] [Google Scholar]
  • Simoens S, Hummelshoj L, D'Hooghe T 2007 Endometriosis: cost estimates and methodological perspective. Hum Reprod Update 13:395–404 [PubMed] [Google Scholar]
  • Missmer SA, Hankinson SE, Spiegelman D, Barbieri RL, Michels KB, Hunter DJ 2004 In utero exposures and the incidence of endometriosis. Fertil Steril 82:1501–1508 [PubMed] [Google Scholar]
  • Rier SE, Martin DC, Bowman RE, Dmowski WP, Becker JL 1993 Endometriosis in rhesus monkeys (Macaca mulatta) following chronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Fundam Appl Toxicol 21:433–441 [PubMed] [Google Scholar]
  • Rier SE, Turner WE, Martin DC, Morris R, Lucier GW, Clark GC 2001 Serum levels of TCDD and dioxin-like chemicals in Rhesus monkeys chronically exposed to dioxin: correlation of increased serum PCB levels with endometriosis. Toxicol Sci 59:147–159 [PubMed] [Google Scholar]
  • Guo SW 2004 The link between exposure to dioxin and endometriosis: a critical reappraisal of primate data. Gynecol Obstet Invest 57:157–173 [PubMed] [Google Scholar]
  • Yang JZ, Agarwal SK, Foster WG 2000 Subchronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin modulates the pathophysiology of endometriosis in the cynomolgus monkey. Toxicol Sci 56:374–381 [PubMed] [Google Scholar]
  • Cummings AM, Metcalf JL, Birnbaum L 1996 Promotion of endometriosis by 2,3,7,8-tetrachlorodibenzo-p-dioxin in rats and mice: time-dose dependence and species comparison. Toxicol Appl Pharmacol 138:131–139 [PubMed] [Google Scholar]
  • Nayyar T, Bruner-Tran KL, Piestrzeniewicz-Ulanska D, Osteen KG 2007 Developmental exposure of mice to TCDD elicits a similar uterine phenotype in adult animals as observed in women with endometriosis. Reprod Toxicol 23:326–336 [PMC free article] [PubMed] [Google Scholar]
  • Cobellis L, Latini G, De Felice C, Razzi S, Paris I, Ruggieri F, Mazzeo P, Petraglia F 2003 High plasma concentrations of di-(2-ethylhexyl)-phthalate in women with endometriosis. Hum Reprod 18:1512–1515 [PubMed] [Google Scholar]
  • Reddy BS, Rozati R, Reddy BV, Raman NV 2006 Association of phthalate esters with endometriosis in Indian women. BJOG 113:515–520 [PubMed] [Google Scholar]
  • Burney R, Giudice LC 2008 Pathogenesis of endometriosis. In: Nezhat CR, ed. Operative gynecologic laparoscopy: principles and techniques. Oxford, UK: Cambridge University Press; 253–259 [Google Scholar]
  • Davis DL, Bradlow HL, Wolff M, Woodruff T, Hoel DG, Anton-Culver H 1993 Medical hypothesis: xenoestrogens as preventable causes of breast cancer. Environ Health Perspect 101:372–377 [PMC free article] [PubMed] [Google Scholar]
  • Sharpe RM, Skakkebaek NE 1993 Are oestrogens involved in falling sperm counts and disorders of the male reproductive tract? Lancet 341:1392–1395 [PubMed] [Google Scholar]
  • Williams EM, Jones L, Vessey MP, McPherson K 1990 Short term increase in risk of breast cancer associated with full term pregnancy. BMJ 300:578–579 [PMC free article] [PubMed] [Google Scholar]
  • Lambe M, Hsieh CC, Chan HW, Ekbom A, Trichopoulos D, Adami HO 1996 Parity, age at first and last birth, and risk of breast cancer: a population-based study in Sweden. Breast Cancer Res Treat 38:305–311 [PubMed] [Google Scholar]
  • Trichopoulos D 1990 Is breast cancer initiated in utero? Epidemiology 1:95–96 [PubMed] [Google Scholar]
  • Hahn WC, Weinberg RA 2002 Modelling the molecular circuitry of cancer. Nat Rev Cancer 2:331–341 [PubMed] [Google Scholar]
  • Ho SM, Tang WY, Belmonte de Frausto J, Prins GS 2006 Developmental exposure to estradiol and bisphenol A increases susceptibility to prostate carcinogenesis and epigenetically regulates phosphodiesterase type 4 variant 4. Cancer Res 66:5624–5632 [PMC free article] [PubMed] [Google Scholar]
  • Sonnenschein C, Soto AM 1999 The society of cells: cancer and control of cell proliferation. New York: Springer-Verlag. [Google Scholar]
  • Sonnenschein C, Soto AM 2008 Theories of carcinogenesis: an emerging perspective. Semin Cancer Biol 18:372–377 [PMC free article] [PubMed] [Google Scholar]
  • Soto AM, Sonnenschein C 2004 The somatic mutation theory of cancer: growing problems with the paradigm? Bioessays 26:1097–1107 [PubMed] [Google Scholar]
  • Maffini MV, Calabro JM, Soto AM, Sonnenschein C 2005 Stromal regulation of neoplastic development: age-dependent normalization of neoplastic mammary cells by mammary stroma. Am J Pathol 167:1405–1410 [PMC free article] [PubMed] [Google Scholar]
  • Markey CM, Rubin BS, Soto AM, Sonnenschein C 2002 Endocrine disruptors: from wingspread to environmental developmental biology. J Steroid Biochem Mol Biol 83:235–244 [PubMed] [Google Scholar]
  • Vandenberg LN, Maffini MV, Wadia PR, Sonnenschein C, Rubin BS, Soto AM 2007 Exposure to environmentally relevant doses of the xenoestrogen bisphenol-A alters development of the fetal mouse mammary gland. Endocrinology 148:116–127 [PMC free article] [PubMed] [Google Scholar]
  • Høyer AP, Grandjean P, Jørgensen T, Brock JW, Hartvig HB 1998 Organochlorine exposure and risk of breast cancer. Lancet 352:1816–1820 [PubMed] [Google Scholar]
  • Cohn BA, Wolff MS, Cirillo PM, Sholtz RI 2007 DDT and breast cancer in young women: new data on the significance of age at exposure. Environ Health Perspect 115:1406–1414 [PMC free article] [PubMed] [Google Scholar]
  • Ibarluzea Jm J, Fernández MF, Santa-Marina L, Olea- Serrano MF, Rivas AM, Aurrekoetxea JJ, Expósito J, Lorenzo M, Torné P, Villalobos M, Pedraza V, Sasco AJ, Olea N 2004 Breast cancer risk and the combined effect of environmental estrogens. Cancer Causes Control 15: 591–600 [PubMed] [Google Scholar]
  • Meyer JS 1977 Cell proliferation in normal human breast ducts, fibroadenomas, and other ductal hyperplasias measured by nuclear labeling with tritiated thymidine. Effects of menstrual phase, age, and oral contraceptive hormones. Hum Pathol 8:67–81 [PubMed] [Google Scholar]
  • Maffini MV, Soto AM, Calabro JM, Ucci AA, Sonnenschein C 2004 Rat mammary gland chemical carcinogenesis: the stroma as a crucial target. J Cell Sci 117:1495–1502 [PubMed] [Google Scholar]
  • Palmer JR, Hatch EE, Rosenberg CL, Hartge P, Kaufman RH, Titus-Ernstoff L, Noller KL, Herbst AL, Rao RS, Troisi R, Colton T, Hoover RN 2002 Risk of breast cancer in women exposed to diethylstilbestrol in utero: preliminary results (United States). Cancer Causes Control 13:753–758 [PubMed] [Google Scholar]
  • Palmer JR, Wise LA, Hatch EE, Troisi R, Titus-Ernstoff L, Strohsnitter W, Kaufman R, Herbst AL, Noller KL, Hyer M, Hoover RN 2006 Prenatal diethylstilbestrol exposure and risk of breast cancer. Cancer Epidemiol Biomarkers Prev 15:1509–1514 [PubMed] [Google Scholar]
  • Boylan ES, Calhoon RE 1979 Mammary tumorigenesis in the rat following prenatal exposure to diethylstilbestrol and postnatal treatment with 7,12-dimethylbenz[a]anthracene. J Toxicol Environ Health 5:1059–1071 [PubMed] [Google Scholar]
  • Rothschild TC, Boylan ES, Calhoon RE, Vonderhaar BK 1987 Transplacental effects of diethylstilbestrol on mammary development and tumorigenesis in female ACI rats. Cancer Res 47:4508–4516 [PubMed] [Google Scholar]
  • Halakivi-Clarke L, Cho E, Onojafe I, Liao DJ, Clarke R 2000 Maternal exposure to tamoxifen during pregnancy increases carcinogen-induced mammary tumorigenesis among female rat offspring. Clin Cancer Res 6:305–308 [PubMed] [Google Scholar]
  • Ohtake F, Takeyama K, Matsumoto T, Kitagawa H, Yamamoto Y, Nohara K, Tohyama C, Krust A, Mimura J, Chambon P, Yanagisawa J, Fujii-Kuriyama Y, Kato S 2003 Modulation of oestrogen receptor signalling by association with the activated dioxin receptor. Nature 423:545–550 [PubMed] [Google Scholar]
  • Brown NM, Manzolillo PA, Zhang JX, Wang J, Lamartiniere CA 1998 Prenatal TCDD and predisposition to mammary cancer in the rat. Carcinogenesis 19:1623–1629 [PubMed] [Google Scholar]
  • Vandenberg LN, Hauser R, Marcus M, Olea N, Welshons WV 2007 Human exposure to bisphenol A (BPA). Reprod Toxicol 24:139–177 [PubMed] [Google Scholar]
  • Calafat AM, Ye X, Wong LY, Reidy JA, Needham LL 2008 Exposure of the U.S. population to bisphenol A and 4-tertiary-octylphenol: 2003–2004. Environ Health Perspect 116:39–44 [PMC free article] [PubMed] [Google Scholar]
  • Richter CA, Birnbaum LS, Farabollini F, Newbold RR, Rubin BS, Talsness CE, Vandenbergh JG, Walser-Kuntz DR, vom Saal FS 2007 In vivo effects of bisphenol A in laboratory rodent studies. Reprod Toxicol 24:199–224 [PMC free article] [PubMed] [Google Scholar]
  • Zalko D, Soto AM, Dolo L, Dorio C, Rathahao E, Debrauwer L, Faure R, Cravedi JP 2003 Biotransformations of bisphenol A in a mammalian model: answers and new questions raised by low-dose metabolic fate studies in pregnant CD1 mice. Environ Health Perspect 111:309–319 [PMC free article] [PubMed] [Google Scholar]
  • Muñoz-de-Toro M, Markey CM, Wadia PR, Luque EH, Rubin BS, Sonnenschein C, Soto AM 2005 Perinatal exposure to bisphenol-A alters peripubertal mammary gland development in mice. Endocrinology 146:4138–4147 [PMC free article] [PubMed] [Google Scholar]
  • Markey CM, Luque EH, Munoz De Toro M, Sonnenschein C, Soto AM 2001 In utero exposure to bisphenol A alters the development and tissue organization of the mouse mammary gland. Biol Reprod 65:1215–1223 [PubMed] [Google Scholar]
  • Vandenberg LN, Maffini MV, Schaeberle CM, Ucci AA, Sonnenschein C, Rubin BS, Soto AM 2008 Perinatal exposure to the xenoestrogen bisphenol-A induces mammary intraductal hyperplasias in adult CD-1 mice. Reprod Toxicol 26:210–219 [PMC free article] [PubMed] [Google Scholar]
  • Murray TJ, Maffini MV, Ucci AA, Sonnenschein C, Soto AM 2007 Induction of mammary gland ductal hyperplasia and carcinoma in situ following fetal bisphenol A exposure. Reprod Toxicol 23:383–390 [PMC free article] [PubMed] [Google Scholar]
  • Durando M, Kass L, Piva J, Sonnenschein C, Soto AM, Luque EH, Muñoz-de-Toro M 2007 Prenatal bisphenol A exposure induces preneoplastic lesions in the mammary gland in Wistar rats. Environ Health Perspect 115:80–86 [PMC free article] [PubMed] [Google Scholar]
  • Jenkins S, Raghuraman N, Eltoum I, Carpenter M, Russo J, Lamartiniere CA 7 January 2009 Oral exposure to bisphenol A increases dimethylbenzanthracene-induced mammary cancer in rats. Environ Health Perspect 10.1289/ehp.11751 [PMC free article] [PubMed] [Google Scholar]
  • 2005 Third National Report on Human Exposure to Environmental Chemicals. Atlanta: Center for Disease Control and Prevention [Google Scholar]
  • Słowikowska-Hilczer J, Szarras-Czapnik M, Kula K 2001 Testicular pathology in 46,XY dysgenetic male pseudohermaphroditism: an approach to pathogenesis of testis cancer. J Androl 22:781–792 [PubMed] [Google Scholar]
  • Carlsen E, Giwercman A, Keiding N, Skakkebaek NE 1992 Evidence for decreasing quality of semen during past 50 years. BMJ 305:609–613 [PMC free article] [PubMed] [Google Scholar]
  • Auger J, Kunstmann JM, Czyglik F, Jouannet P 1995 Decline in semen quality among fertile men in Paris during the past 20 years. N Engl J Med 332:281–285 [PubMed] [Google Scholar]
  • Swan SH, Elkin EP, Fenster L 1997 Have sperm densities declined? A reanalysis of global trend data. Environ Health Perspect 105:1228–1232 [PMC free article] [PubMed] [Google Scholar]
  • Bujan L, Mansat A, Pontonnier F, Mieusset R 1996 Time series analysis of sperm concentration in fertile men in Toulouse, France between 1977 and 1992. BMJ 312:471–472 [PMC free article] [PubMed] [Google Scholar]
  • Fisch H, Goluboff ET, Olson JH, Feldshuh J, Broder SJ, Barad DH 1996 Semen analyses in 1,283 men from the United States over a 25-year period: no decline in quality. Fertil Steril 65:1009–1014 [PubMed] [Google Scholar]
  • Paulsen CA, Berman NG, Wang C 1996 Data from men in greater Seattle area reveals no downward trend in semen quality: further evidence that deterioration of semen quality is not geographically uniform. Fertil Steril 65:1015–1020 [PubMed] [Google Scholar]
  • David R, McKee R, Butala J, Barter R, Kaiser M 2001 Esters of aromatic mono-, di-, and tricarboxylic acids, aromatic diacids, and di-, tri-, or polyalcohols. In: Bingham E, Cohrssen B, Powell C, eds. Patty’s toxicology. New York: John Wiley and Sons; 635–932 [Google Scholar]
  • 1997 Toxicological profile for di-n-octyl phthalate (DNOP). Atlanta: Agency for Toxic Substances and Disease Registry [Google Scholar]
  • 2002 Toxicological profile for di-n-octyl phthalate (DNOP). Atlanta: Agency for Toxic Substances and Disease Registry [Google Scholar]
  • 1995 Toxicological profile for di-n-octyl phthalate (DNOP). Atlanta: Agency for Toxic Substances and Disease Registry [Google Scholar]
  • 2001 Toxicological profile for di-n-octyl phthalate (DNOP). Atlanta: Agency for Toxic Substances and Disease Registry [Google Scholar]
  • Adibi JJ, Perera FP, Jedrychowski W, Camann DE, Barr D, Jacek R, Whyatt RM 2003 Prenatal exposures to phthalates among women in New York City and Krakow, Poland. Environ Health Perspect 111:1719–1722 [PMC free article] [PubMed] [Google Scholar]
  • Rudel RA, Camann DE, Spengler JD, Korn LR, Brody JG 2003 Phthalates, alkylphenols, pesticides, polybrominated diphenyl ethers, and other endocrine-disrupting compounds in indoor air and dust. Environ Sci Technol 37:4543–4553 [PubMed] [Google Scholar]
  • Green R, Hauser R, Calafat AM, Weuve J, Schettler T, Ringer S, Huttner K, Hu H 2005 Use of di(2-ethylhexyl) phthalate-containing medical products and urinary levels of mono(2-ethylhexyl) phthalate in neonatal intensive care unit infants. Environ Health Perspect 113:1222–1225 [PMC free article] [PubMed] [Google Scholar]
  • Duty SM, Silva MJ, Barr DB, Brock JW, Ryan L, Chen Z, Herrick RF, Christiani DC, Hauser R 2003 Phthalate exposure and human semen parameters. Epidemiology 14:269–277 [PubMed] [Google Scholar]
  • Hauser R, Meeker JD, Duty S, Silva MJ, Calafat AM 2006 Altered semen quality in relation to urinary concentrations of phthalate monoester and oxidative metabolites. Epidemiology 17:682–691 [PubMed] [Google Scholar]
  • Jönsson BA, Richthoff J, Rylander L, Giwercman A, Hagmar L 2005 Urinary phthalate metabolites and biomarkers of reproductive function in young men. Epidemiology 16:487–493 [PubMed] [Google Scholar]
  • Longnecker MP, Rogan WJ, Lucier G 1997 The human health effects of DDT (dichlorodiphenyltrichloroethane) and PCBS (polychlorinated biphenyls) and an overview of organochlorines in public health. Annu Rev Public Health 18:211–244 [PubMed] [Google Scholar]
  • Brown JF 1994 Determination of PCB metabolic, excretion, and accumulation rates for use as indicators of biological response and relative risk. Environ Sci Technol 28:2295–2305 [PubMed] [Google Scholar]
  • Phillips DL, Smith AB, Burse VW, Steele GK, Needham LL, Hannon WH 1989 Half-life of polychlorinated biphenyls in occupationally exposed workers. Arch Environ Health 44:351–354 [PubMed] [Google Scholar]
  • Kato Y, Haraguchi K, Shibahara T, Masuda Y, Kimura R 1998 Reduction of thyroid hormone levels by methylsulfonyl metabolites of polychlorinated biphenyl congeners in rats. Arch Toxicol 72:541–544 [PubMed] [Google Scholar]
  • Hansen LG 1998 Stepping backward to improve assessment of PCB congener toxicities. Environ Health Perspect 106(Suppl 1):171–189 [PMC free article] [PubMed] [Google Scholar]
  • Dallinga JW, Moonen EJ, Dumoulin JC, Evers JL, Geraedts JP, Kleinjans JC 2002 Decreased human semen quality and organochlorine compounds in blood. Hum Reprod 17:1973–1979 [PubMed] [Google Scholar]
  • Richthoff J, Rylander L, Jönsson BA, Akesson H, Hagmar L, Nilsson-Ehle P, Stridsberg M, Giwercman A 2003 Serum levels of 2,2′,4,4′,5,5′-hexachlorobiphenyl (CB-153) in relation to markers of reproductive function in young males from the general Swedish population. Environ Health Perspect 111:409–413 [PMC free article] [PubMed] [Google Scholar]
  • Hauser R, Chen Z, Pothier L, Ryan L, Altshul L 2003 The relationship between human semen parameters and environmental exposure to polychlorinated biphenyls and p,p’-DDE. Environ Health Perspect 111:1505–1511 [PMC free article] [PubMed] [Google Scholar]
  • Rignell-Hydbom A, Rylander L, Giwercman A, Jönsson BA, Nilsson-Ehle P, Hagmar L 2004 Exposure to CB-153 and p,p’-DDE and male reproductive function. Hum Reprod 19:2066–2075 [PubMed] [Google Scholar]
  • Guo YL, Hsu PC, Hsu CC, Lambert GH 2000 Semen quality after prenatal exposure to polychlorinated biphenyls and dibenzofurans. Lancet 356:1240–1241 [PubMed] [Google Scholar]
  • Hsu PC, Huang W, Yao WJ, Wu MH, Guo YL, Lambert GH 2003 Sperm changes in men exposed to polychlorinated biphenyls and dibenzofurans. JAMA 289:2943–2944 [PubMed] [Google Scholar]
  • Mocarelli P, Gerthoux PM, Patterson Jr DG, Milani S, Limonta G, Bertona M, Signorini S, Tramacere P, Colombo L, Crespi C, Brambilla P, Sarto C, Carreri V, Sampson EJ, Turner WE, Needham LL 2008 Dioxin exposure, from infancy through puberty, produces endocrine disruption and affects human semen quality. Environ Health Perspect 116:70–77 [PMC free article] [PubMed] [Google Scholar]
  • Abell A, Ernst E, Bonde JP 2000 Semen quality and sexual hormones in greenhouse workers. Scand J Work Environ Health 26:492–500 [PubMed] [Google Scholar]
  • Juhler RK, Larsen SB, Meyer O, Jensen ND, Spanò M, Giwercman A, Bonde JP 1999 Human semen quality in relation to dietary pesticide exposure and organic diet. Arch Environ Contam Toxicol 37:415–423 [PubMed] [Google Scholar]
  • Oliva A, Spira A, Multigner L 2001 Contribution of environmental factors to the risk of male infertility. Hum Reprod 16:1768–1776 [PubMed] [Google Scholar]
  • Larsen SB, Giwercman A, Spanò M, Bonde JP 1998 A longitudinal study of semen quality in pesticide spraying Danish farmers. The ASCLEPIOS Study Group. Reprod Toxicol 12:581–589 [PubMed] [Google Scholar]
  • Padungtod C, Savitz DA, Overstreet JW, Christiani DC, Ryan LM, Xu X 2000 Occupational pesticide exposure and semen quality among Chinese workers. J Occup Environ Med 42:982–992 [PubMed] [Google Scholar]
  • Kamijima M, Hibi H, Gotoh M, Taki K, Saito I, Wang H, Itohara S, Yamada T, Ichihara G, Shibata E, Nakajima T, Takeuchi Y 2004 A survey of semen indices in insecticide sprayers. J Occup Health 46:109–118 [PubMed] [Google Scholar]
  • Lifeng T, Shoulin W, Junmin J, Xuezhao S, Yannan L, Qianli W, Longsheng C 2006 Effects of fenvalerate exposure on semen quality among occupational workers. Contraception 73:92–96 [PubMed] [Google Scholar]
  • Whorton MD, Milby TH, Stubbs HA, Avashia BH, Hull EQ 1979 Testicular function among carbaryl-exposed exployees. J Toxicol Environ Health 5:929–941 [PubMed] [Google Scholar]
  • Wyrobek AJ, Watchmaker G, Gordon L, Wong K, Moore 2nd D, Whorton D 1981 Sperm shape abnormalities in carbaryl-exposed employees. Environ Health Perspect 40:255–265 [PMC free article] [PubMed] [Google Scholar]
  • Tan LF, Sun XZ, Li YN, Ji JM, Wang QL, Chen LS, Bian Q, Wang SL 2005 [Effects of carbaryl production exposure on the sperm and semen quality of occupational male workers]. Zhonghua Lao Dong Wei Sheng Zhi Ye Bing Za Zhi 23:87–90 [PubMed] [Google Scholar]
  • Swan SH, Kruse RL, Liu F, Barr DB, Drobnis EZ, Redmon JB, Wang C, Brazil C, Overstreet JW 2003 Semen quality in relation to biomarkers of pesticide exposure. Environ Health Perspect 111:1478–1484 [PMC free article] [PubMed] [Google Scholar]
  • Meeker JD, Ryan L, Barr DB, Herrick RF, Bennett DH, Bravo R, Hauser R 2004 The relationship of urinary metabolites of carbaryl/naphthalene and chlorpyrifos with human semen quality. Environ Health Perspect 112:1665–1670 [PMC free article] [PubMed] [Google Scholar]
  • Boisen KA, Kaleva M, Main KM, Virtanen HE, Haavisto AM, Schmidt IM, Chellakooty M, Damgaard IN, Mau C, Reunanen M, Skakkebaek NE, Toppari J 2004 Difference in prevalence of congenital cryptorchidism in infants between two Nordic countries. Lancet 363:1264–1269 [PubMed] [Google Scholar]
  • Paulozzi LJ 1999 International trends in rates of hypospadias and cryptorchidism. Environ Health Perspect 107: 297–302 [PMC free article] [PubMed] [Google Scholar]
  • Thonneau PF, Gandia P, Mieusset R 2003 Cryptorchidism: incidence, risk factors, and potential role of environment; an update. J Androl 24:155–162 [PubMed] [Google Scholar]
  • Barthold JS, González R 2003 The epidemiology of congenital cryptorchidism, testicular ascent and orchidopexy. J Urol 170:2396–2401 [PubMed] [Google Scholar]
  • Dolk H, Vrijheid M, Scott JE, Addor MC, Botting B, de Vigan C, de Walle H, Garne E, Loane M, Pierini A, Garcia-Minaur S, Physick N, Tenconi R, Wiesel A, Calzolari E, Stone D 2004 Toward the effective surveillance of hypospadias. Environ Health Perspect 112:398–402 [PMC free article] [PubMed] [Google Scholar]
  • Aho M, Koivisto AM, Tammela T, Auvinen A 2000 Is the incidence of hypospadias increasing? Analysis of Finnish hospital discharge data 1970–1994. Environ Health Perspect 108:463–465 [PMC free article] [PubMed] [Google Scholar]
  • Martínez-Frías ML, Prieto D, Prieto L, Bermejo E, Rodríguez-Pinilla E, Cuevas L 2004 Secular decreasing trend of the frequency of hypospadias among newborn male infants in Spain. Birth Defects Res A Clin Mol Teratol 70:75–81 [PubMed] [Google Scholar]
  • Longnecker MP, Klebanoff MA, Brock JW, Zhou H, Gray KA, Needham LL, Wilcox AJ 2002 Maternal serum level of 1,1-dichloro-2,2-bis(p-chlorophenyl)ethylene and risk of cryptorchidism, hypospadias, and polythelia among male offspring. Am J Epidemiol 155:313–322 [PubMed] [Google Scholar]
  • Bhatia R, Shiau R, Petreas M, Weintraub JM, Farhang L, Eskenazi B 2005 Organochlorine pesticides and male genital anomalies in the child health and development studies. Environ Health Perspect 113:220–224 [PMC free article] [PubMed] [Google Scholar]
  • Mol NM, Sørensen N, Weihe P, Andersson AM, Jørgensen N, Skakkebaek NE, Keiding N, Grandjean P 2002 Spermaturia and serum hormone concentrations at the age of puberty in boys prenatally exposed to polychlorinated biphenyls. Eur J Endocrinol 146:357–363 [PubMed] [Google Scholar]
  • Hosie S, Loff S, Witt K, Niessen K, Waag KL 2000 Is there a correlation between organochlorine compounds and undescended testes? Eur J Pediatr Surg 10:304–309 [PubMed] [Google Scholar]
  • Garry VF, Schreinemachers D, Harkins ME, Griffith J 1996 Pesticide appliers, biocides and birth defects in rural Minnesota. Environ Health Perspect 104:394–399 [PMC free article] [PubMed] [Google Scholar]
  • García-Rodríguez J, García-Martín M, Nogueras-Ocaña M, de Dios Luna-del-Castillo J, Espigares García M, Olea N, Lardelli-Claret P 1996 Exposure to pesticides and cryptorchidism: geographical evidence of a possible association. Environ Health Perspect 104:1090–1095 [PMC free article] [PubMed] [Google Scholar]
  • Kristensen P, Irgens LM, Andersen A, Bye AS, Sundheim L 1997 Birth defects among offspring of Norwegian farmers, 1967–1991. Epidemiology 8:537–544 [PubMed] [Google Scholar]
  • Weidner IS, Møller H, Jensen TK, Skakkebaek NE 1998 Cryptorchidism and hypospadias in sons of gardeners and farmers. Environ Health Perspect 106:793–796 [PMC free article] [PubMed] [Google Scholar]
  • Pierik FH, Burdorf A, Deddens JA, Juttmann RE, Weber RF 2004 Maternal and paternal risk factors for cryptorchidism and hypospadias: a case-control study in newborn boys. Environ Health Perspect 112:1570–1576 [PMC free article] [PubMed] [Google Scholar]
  • Carbone P, Giordano F, Nori F, Mantovani A, Taruscio D, Lauria L, Figà-Talamanca I 2006 Cryptorchidism and hypospadias in the Sicilian district of Ragusa and the use of pesticides. Reprod Toxicol 22:8–12 [PubMed] [Google Scholar]
  • Diamanti-Kandarakis E, Katsikis I, Piperi C, Kandaraki E, Piouka A, Papavassiliou AG, Panidis D 2008 Increased serum advanced glycation end-products is a distinct finding in lean women with polycystic ovary syndrome (PCOS). Clin Endocrinol (Oxf) 69:634–641 [PubMed] [Google Scholar]
  • Adami HO, Bergström R, Möhner M, Zatoñski W, Storm H, Ekbom A, Tretli S, Teppo L, Ziegler H, Rahu M 1994 Testicular cancer in nine northern European countries. Int J Cancer 59:33–38 [PubMed] [Google Scholar]
  • 1996 Increase in testicular cancer incidence in six European countries: a birth cohort phenomenon. J Natl Cancer Inst 88:727–733 [PubMed] [Google Scholar]
  • Huyghe E, Matsuda T, Thonneau P 2003 Increasing incidence of testicular cancer worldwide: a review. J Urol 170:5–11 [PubMed] [Google Scholar]
  • Richiardi L, Bellocco R, Adami HO, Torrång A, Barlow L, Hakulinen T, Rahu M, Stengrevics A, Storm H, Tretli S, Kurtinaitis J, Tyczynski JE, Akre O 2004 Testicular cancer incidence in eight northern European countries: secular and recent trends. Cancer Epidemiol Biomarkers Prev 13:2157–2166 [PubMed] [Google Scholar]
  • McGlynn KA, Devesa SS, Sigurdson AJ, Brown LM, Tsao L, Tarone RE 2003 Trends in the incidence of testicular germ cell tumors in the United States. Cancer 97:63–70 [PubMed] [Google Scholar]
  • Hemminki K, Li X 2002 Cancer risks in Nordic immigrants and their offspring in Sweden. Eur J Cancer 38:2428–2434 [PubMed] [Google Scholar]
  • Henderson BE, Benton B, Jing J, Yu MC, Pike MC 1979 Risk factors for cancer of the testis in young men. Int J Cancer 23:598–602 [PubMed] [Google Scholar]
  • Weir HK, Marrett LD, Kreiger N, Darlington GA, Sugar L 2000 Pre-natal and peri-natal exposures and risk of testicular germ-cell cancer. Int J Cancer 87:438–443 [PubMed] [Google Scholar]
  • Hardell L, van Bavel B, Lindström G, Carlberg M, Dreifaldt AC, Wijkström H, Starkhammar H, Eriksson M, Hallquist A, Kolmert T 2003 Increased concentrations of polychlorinated biphenyls, hexachlorobenzene, and chlordanes in mothers of men with testicular cancer. Environ Health Perspect 111:930–934 [PMC free article] [PubMed] [Google Scholar]
  • Hardell L, Bavel B, Lindström G, Eriksson M, Carlberg M 2006 In utero exposure to persistent organic pollutants in relation to testicular cancer risk. Int J Androl 29:228–234 [PubMed] [Google Scholar]
  • Jemal A, Siegel R, Ward E, Murray T, Xu J, Smigal C, Thun MJ 2006 Cancer statistics, 2006. CA Cancer J Clin 56:106–130 [PubMed] [Google Scholar]
  • Moore R 1947 Endocrinology of neoplastic disease. New York: Oxford University Press; 194 [Google Scholar]
  • Huggins C, Hodges CF 1941 Studies on prostatic cancer. I. The effect of castration, of estrogen, and of androgen injection on serum phosphatases in metastatic carcinoma of the prostate. Cancer Res 1:293–297 [PubMed] [Google Scholar]
  • Leav I, Ho SM, Ofner P, Merk FB, Kwan PW, Damassa D 1988 Biochemical alterations in sex hormone-induced hyperplasia and dysplasia of the dorsolateral prostates of Noble rats. J Natl Cancer Inst 80:1045–1053 [PubMed] [Google Scholar]
  • Thomas JA, Keenan EJ 1994 Effects of estrogens on the prostate. J Androl 15:97–99 [PubMed] [Google Scholar]
  • Modugno F, Weissfeld JL, Trump DL, Zmuda JM, Shea P, Cauley JA, Ferrell RE 2001 Allelic variants of aromatase and the androgen and estrogen receptors: toward a multigenic model of prostate cancer risk. Clin Cancer Res 7:3092–3096 [PubMed] [Google Scholar]
  • Raghow S, Hooshdaran MZ, Katiyar S, Steiner MS 2002 Toremifene prevents prostate cancer in the transgenic adenocarcinoma of mouse prostate model. Cancer Res 62:1370–1376 [PubMed] [Google Scholar]
  • Steiner MS, Pound CR 2003 Phase IIA clinical trial to test the efficacy and safety of toremifene in men with high-grade prostatic intraepithelial neoplasia. Clin Prostate Cancer 2:24–31 [PubMed] [Google Scholar]
  • Prins GS, Korach KS 2008 The role of estrogens and estrogen receptors in normal prostate growth and disease. Steroids 73:233–244 [PMC free article] [PubMed] [Google Scholar]
  • Henderson BE, Bernstein L, Ross RK, Depue RH, Judd HL 1988 The early in utero oestrogen and testosterone environment of blacks and whites: potential effects on male offspring. Br J Cancer 57:216–218 [PMC free article] [PubMed] [Google Scholar]
  • Henderson BE, Ross RK, Pike MC 1991 Toward the primary prevention of cancer. Science 254:1131–1138 [PubMed] [Google Scholar]
  • Prins GS, Birch L, Tang WY, Ho SM 2007 Developmental estrogen exposures predispose to prostate carcinogenesis with aging. Reprod Toxicol 23:374–382 [PMC free article] [PubMed] [Google Scholar]
  • Morrison H, Savitz D, Semenciw R, Hulka B, Mao Y, Morison D, Wigle D 1993 Farming and prostate cancer mortality. Am J Epidemiol 137:270–280 [PubMed] [Google Scholar]
  • Meyer TE, Coker AL, Sanderson M, Symanski E 2007 A case-control study of farming and prostate cancer in African-American and Caucasian men. Occup Environ Med 64:155–160 [PMC free article] [PubMed] [Google Scholar]
  • Alavanja MC, Samanic C, Dosemeci M, Lubin J, Tarone R, Lynch CF, Knott C, Thomas K, Hoppin JA, Barker J, Coble J, Sandler DP, Blair A 2003 Use of agricultural pesticides and prostate cancer risk in the Agricultural Health Study cohort. Am J Epidemiol 157:800–814 [PubMed] [Google Scholar]
  • Van Maele-Fabry G, Libotte V, Willems J, Lison D 2006 Review and meta-analysis of risk estimates for prostate cancer in pesticide manufacturing workers. Cancer Causes Control 17:353–373 [PubMed] [Google Scholar]
  • Mahajan R, Bonner MR, Hoppin JA, Alavanja MC 2006 Phorate exposure and incidence of cancer in the agricultural health study. Environ Health Perspect 114: 1205–1209 [PMC free article] [PubMed] [Google Scholar]
  • Usmani KA, Rose RL, Hodgson E 2003 Inhibition and activation of the human liver microsomal and human cytochrome P450 3A4 metabolism of testosterone by deployment-related chemicals. Drug Metab Dispos 31:384–391 [PubMed] [Google Scholar]
  • Usmani KA, Cho TM, Rose RL, Hodgson E 2006 Inhibition of the human liver microsomal and human cytochrome P450 1A2 and 3A4 metabolism of estradiol by deployment-related and other chemicals. Drug Metab Dispos 34:1606–1614 [PubMed] [Google Scholar]
  • Kester MH, Bulduk S, Tibboel D, Meinl W, Glatt H, Falany CN, Coughtrie MW, Bergman A, Safe SH, Kuiper GG, Schuur AG, Brouwer A, Visser TJ 2000 Potent inhibition of estrogen sulfotransferase by hydroxylated PCB metabolites: a novel pathway explaining the estrogenic activity of PCBs. Endocrinology 141:1897–1900 [PubMed] [Google Scholar]
  • Noble RL 1977 The development of prostatic adenocarcinoma in Nb rats following prolonged sex hormone administration. Cancer Res 37:1929–1933 [PubMed] [Google Scholar]
  • Driscoll SG, Taylor SH 1980 Effects of prenatal maternal estrogen on the male urogenital system. Obstet Gynecol 56:537–542 [PubMed] [Google Scholar]
  • Yonemura CY, Cunha GR, Sugimura Y, Mee SL 1995 Temporal and spatial factors in diethylstilbestrol-induced squamous metaplasia in the developing human prostate. II. Persistent changes after removal of diethylstilbestrol. Acta Anat (Basel) 153:1–11 [PubMed] [Google Scholar]
  • Giusti RM, Iwamoto K, Hatch EE 1995 Diethylstilbestrol revisited: a review of the long-term health effects. Ann Intern Med 122:778–788 [PubMed] [Google Scholar]
  • Arai Y, Mori T, Suzuki Y, Bern HA 1983 Long-term effects of perinatal exposure to sex steroids and diethylstilbestrol on the reproductive system of male mammals. Int Rev Cytol 84:235–268 [PubMed] [Google Scholar]
  • Rajfer J, Coffey DS 1978 Sex steroid imprinting of the immature prostate. Long-term effects. Invest Urol 16:186–190 [PubMed] [Google Scholar]
  • Prins GS, Birch L, Habermann H, Chang WY, Tebeau C, Putz O, Bieberich C 2001 Influence of neonatal estrogens on rat prostate development. Reprod Fertil Dev 13:241–252 [PubMed] [Google Scholar]
  • Huang L, Pu Y, Alam S, Birch L, Prins GS 2004 Estrogenic regulation of signaling pathways and homeobox genes during rat prostate development. J Androl 25:330–337 [PubMed] [Google Scholar]
  • Gupta C 2000 The role of estrogen receptor, androgen receptor and growth factors in diethylstilbestrol-induced programming of prostate differentiation. Urol Res 28:223–229 [PubMed] [Google Scholar]
  • Lemmen JG, Arends RJ, van der Saag PT, van der Burg B 2004 In vivo imaging of activated estrogen receptors in utero by estrogens and bisphenol A. Environ Health Perspect 112:1544–1549 [PMC free article] [PubMed] [Google Scholar]
  • Kurosawa T, Hiroi H, Tsutsumi O, Ishikawa T, Osuga Y, Fujiwara T, Inoue S, Muramatsu M, Momoeda M, Taketani Y 2002 The activity of bisphenol A depends on both the estrogen receptor subtype and the cell type. Endocr J 49:465–471 [PubMed] [Google Scholar]
  • Song KH, Lee K, Choi HS 2002 Endocrine disruptor bisphenol A induces orphan nuclear receptor Nur77 gene expression and steroidogenesis in mouse testicular Leydig cells. Endocrinology 143:2208–2215 [PubMed] [Google Scholar]
  • Walsh DE, Dockery P, Doolan CM 2005 Estrogen receptor independent rapid non-genomic effects of environmental estrogens on [Ca2+]i in human breast cancer cells. Mol Cell Endocrinol 230:23–30 [PubMed] [Google Scholar]
  • Keri RA, Ho SM, Hunt PA, Knudsen KE, Soto AM, Prins GS 2007 An evaluation of evidence for the carcinogenic activity of bisphenol A. Reprod Toxicol 24:240–252 [PMC free article] [PubMed] [Google Scholar]
  • Wetherill YB, Fisher NL, Staubach A, Danielsen M, de Vere White RW, Knudsen KE 2005 Xenoestrogen action in prostate cancer: pleiotropic effects dependent on androgen receptor status. Cancer Res 65:54–65 [PubMed] [Google Scholar]
  • Wetherill YB, Hess-Wilson JK, Comstock CE, Shah SA, Buncher CR, Sallans L, Limbach PA, Schwemberger S, Babcock GF, Knudsen KE 2006 Bisphenol A facilitates bypass of androgen ablation therapy in prostate cancer. Mol Cancer Ther 5:3181–3190 [PubMed] [Google Scholar]
  • Prins GS, Tang WY, Belmonte J, Ho SM 2008 Perinatal exposure to oestradiol and bisphenol A alters the prostate epigenome and increases susceptibility to carcinogenesis. Basic Clin Pharmacol Toxicol 102:134–138 [PMC free article] [PubMed] [Google Scholar]
  • Hardell L, Andersson SO, Carlberg M, Bohr L, van Bavel B, Lindström G, Björnfoth H, Ginman C 2006 Adipose tissue concentrations of persistent organic pollutants and the risk of prostate cancer. J Occup Environ Med 48:700–707 [PubMed] [Google Scholar]
  • Prince MM, Ruder AM, Hein MJ, Waters MA, Whelan EA, Nilsen N, Ward EM, Schnorr TM, Laber PA, Davis-King KE 2006 Mortality and exposure response among 14,458 electrical capacitor manufacturing workers exposed to polychlorinated biphenyls (PCBs). Environ Health Perspect 114:1508–1514 [PMC free article] [PubMed] [Google Scholar]
  • Ritchie JM, Vial SL, Fuortes LJ, Guo H, Reedy VE, Smith EM 2003 Organochlorines and risk of prostate cancer. J Occup Environ Med 45:692–702 [PubMed] [Google Scholar]
  • Charles LE, Loomis D, Shy CM, Newman B, Millikan R, Nylander-French LA, Couper D 2003 Electromagnetic fields, polychlorinated biphenyls, and prostate cancer mortality in electric utility workers. Am J Epidemiol 157:683–691 [PubMed] [Google Scholar]
  • Schlumpf M, Schmid P, Durrer S, Conscience M, Maerkel K, Henseler M, Gruetter M, Herzog I, Reolon S, Ceccatelli R, Faass O, Stutz E, Jarry H, Wuttke W, Lichtensteiger W 2004 Endocrine activity and developmental toxicity of cosmetic UV filters–an update. Toxicology 205:113–122 [PubMed] [Google Scholar]
  • Schlumpf M, Jarry H, Wuttke W, Ma R, Lichtensteiger W 2004 Estrogenic activity and estrogen receptor β binding of the UV filter 3-benzylidene camphor. Comparison with 4-methylbenzylidene camphor. Toxicology 199:109–120 [PubMed] [Google Scholar]
  • Hofkamp L, Bradley S, Tresguerres J, Lichtensteiger W, Schlumpf M, Timms B 2008 Region-specific growth effects in the developing rat prostate following fetal exposure to estrogenic ultraviolet filters. Environ Health Perspect 116:867–872 [PMC free article] [PubMed] [Google Scholar]
  • Parent ME, Siemiatycki J 2001 Occupation and prostate cancer. Epidemiol Rev 23:138–143 [PubMed] [Google Scholar]
  • Benbrahim-Tallaa L, Liu J, Webber MM, Waalkes MP 2007 Estrogen signaling and disruption of androgen metabolism in acquired androgen-independence during cadmium carcinogenesis in human prostate epithelial cells. Prostate 67:135–145 [PubMed] [Google Scholar]
  • Waalkes MP 2000 Cadmium carcinogenesis in review. J Inorg Biochem 79:241–244 [PubMed] [Google Scholar]
  • Watson WH, Yager JD 2007 Arsenic: extension of its endocrine disruption potential to interference with estrogen receptor-mediated signaling. Toxicol Sci 98:1–4 [PubMed] [Google Scholar]
  • Davey JC, Bodwell JE, Gosse JA, Hamilton JW 2007 Arsenic as an endocrine disruptor: effects of arsenic on estrogen receptor-mediated gene expression in vivo and in cell culture. Toxicol Sci 98:75–86 [PubMed] [Google Scholar]
  • Benbrahim-Tallaa L, Webber MM, Waalkes MP 2007 Mechanisms of acquired androgen independence during arsenic-induced malignant transformation of human prostate epithelial cells. Environ Health Perspect 115:243–247 [PMC free article] [PubMed] [Google Scholar]
  • Chen CJ, Kuo TL, Wu MM 1988 Arsenic and cancers. Lancet 1:414–415 [PubMed] [Google Scholar]
  • Lewis DR, Southwick JW, Ouellet-Hellstrom R, Rench J, Calderon RL 1999 Drinking water arsenic in Utah: a cohort mortality study. Environ Health Perspect 107:359–365 [PMC free article] [PubMed] [Google Scholar]
  • Kavlock R, Cummings A 2005 Mode of action: inhibition of androgen receptor function—vinclozolin-induced malformations in reproductive development. Crit Rev Toxicol 35:721–726 [PubMed] [Google Scholar]
  • Yu WJ, Lee BJ, Nam SY, Ahn B, Hong JT, Do JC, Kim YC, Lee YS, Yun YW 2004 Reproductive disorders in pubertal and adult phase of the male rats exposed to vinclozolin during puberty. J Vet Med Sci 66:847–853 [PubMed] [Google Scholar]
  • Cowin PA, Foster P, Pedersen J, Hedwards S, McPherson SJ, Risbridger GP 2008 Early-onset endocrine disruptor-induced prostatitis in the rat. Environ Health Perspect 116:923–929 [PMC free article] [PubMed] [Google Scholar]
  • Cocco P, Benichou J 1998 Mortality from cancer of the male reproductive tract and environmental exposure to the anti-androgen p,p’-dichlorodiphenyldichloroethylene in the United States. Oncology 55:334–339 [PubMed] [Google Scholar]
  • Gore A 2007 Endocrine-disrupting chemicals: from basic research to clinical practice. In: Conn PM, ed. Contemporary endocrinology. Totowa, NJ: Humana Press [Google Scholar]
  • Walker DM, Gore AC 2007 Endocrine-disrupting chemicals and the brain. In: Gore AC, ed. Endocrine-disrupting chemicals: from basic research to clinical practice. Totowa, NJ: Humana Press; 63–109 [Google Scholar]
  • Gore AC 2008 Neuroendocrine systems as targets for environmental endocrine-disrupting chemicals. Fertil Steril 89:e101–e102 [PMC free article] [PubMed] [Google Scholar]
  • Gore A 2002 GnRH: the master molecule of reproduction. Norwell, MA: Kluwer Academic Publishers [Google Scholar]
  • Wintermantel TM, Campbell RE, Porteous R, Bock D, Gröne HJ, Todman MG, Korach KS, Greiner E, Pérez CA, Schütz G, Herbison AE 2006 Definition of estrogen receptor pathway critical for estrogen positive feedback to gonadotropin-releasing hormone neurons and fertility. Neuron 52:271–280 [PMC free article] [PubMed] [Google Scholar]
  • Mellon PL, Windle JJ, Goldsmith PC, Padula CA, Roberts JL, Weiner RI 1990 Immortalization of hypothalamic GnRH neurons by genetically targeted tumorigenesis. Neuron 5:1–10 [PubMed] [Google Scholar]
  • Gore AC, Wu TJ, Oung T, Lee JB, Woller MJ 2002 A novel mechanism for endocrine-disrupting effects of polychlorinated biphenyls: direct effects on gonadotropin-releasing hormone neurones. J Neuroendocrinol 14:814–823 [PubMed] [Google Scholar]
  • Gore AC 2002 Organochlorine pesticides directly regulate gonadotropin-releasing hormone gene expression and biosynthesis in the GT1–7 hypothalamic cell line. Mol Cell Endocrinol 192:157–170 [PubMed] [Google Scholar]
  • Cook R, Calabrese EJ 2006 The importance of hormesis to public health. Environ Health Perspect 114:1631–1635 [PMC free article] [PubMed] [Google Scholar]
  • Gore AC, Heindel JJ, Zoeller RT 2006 Endocrine disruption for endocrinologists (and others). Endocrinology 147:S1–S3 [PubMed] [Google Scholar]
  • Rasier G, Parent AS, Gérard A, Denooz R, Lebrethon MC, Charlier C, Bourguignon JP 2008 Mechanisms of interaction of endocrine-disrupting chemicals with glutamate-evoked secretion of gonadotropin-releasing hormone. Toxicol Sci 102:33–41 [PubMed] [Google Scholar]
  • McGarvey C, Cates PA, Brooks A, Swanson IA, Milligan SR, Coen CW, O'Byrne KT 2001 Phytoestrogens and gonadotropin-releasing hormone pulse generator activity and pituitary luteinizing hormone release in the rat. Endocrinology 142:1202–1208 [PubMed] [Google Scholar]
  • Patisaul HB, Fortino AE, Polston EK 2007 Differential disruption of nuclear volume and neuronal phenotype in the preoptic area by neonatal exposure to genistein and bisphenol-A. Neurotoxicology 28:1–12 [PubMed] [Google Scholar]
  • Gore AC 2001 Environmental toxicant effects on neuroendocrine function. Endocrine 14:235–246 [PubMed] [Google Scholar]
  • Bisenius ES, Veeramachaneni DN, Sammonds GE, Tobet S 2006 Sex differences and the development of the rabbit brain: effects of vinclozolin. Biol Reprod 75:469–476 [PubMed] [Google Scholar]
  • Khan IA, Thomas P 2001 Disruption of neuroendocrine control of luteinizing hormone secretion by aroclor 1254 involves inhibition of hypothalamic tryptophan hydroxylase activity. Biol Reprod 64:955–964 [PubMed] [Google Scholar]
  • Gore AC 2008 Developmental programming and endocrine disruptor effects on reproductive neuroendocrinology. Front Neuroendocrinol 29:358–374 [PMC free article] [PubMed] [Google Scholar]
  • George FW, Wilson JD 1978 Conversion of androgen to estrogen by the human fetal ovary. J Clin Endocrinol Metab 47:550–555 [PubMed] [Google Scholar]
  • Steinberg RM, Juenger TE, Gore AC 2007 The effects of prenatal PCBs on adult female paced mating reproductive behaviors in rats. Horm Behav 51:364–372 [PMC free article] [PubMed] [Google Scholar]
  • Chung YW, Clemens LG 1999 Effects of perinatal exposure to polychlorinated biphenyls on development of female sexual behavior. Bull Environ Contam Toxicol 62:664–670 [PubMed] [Google Scholar]
  • Chung YW, Nunez AA, Clemens LG 2001 Effects of neonatal polychlorinated biphenyl exposure on female sexual behavior. Physiol Behav 74:363–370 [PubMed] [Google Scholar]
  • Patisaul HB, Luskin JR, Wilson ME 2004 A soy supplement and tamoxifen inhibit sexual behavior in female rats. Horm Behav 45:270–277 [PubMed] [Google Scholar]
  • Whitten PL, Lewis C, Russell E, Naftolin F 1995 Phytoestrogen influences on the development of behavior and gonadotropin function. Proc Soc Exp Biol Med 208:82–86 [PubMed] [Google Scholar]
  • Kouki T, Okamoto M, Wada S, Kishitake M, Yamanouchi K 2005 Suppressive effect of neonatal treatment with a phytoestrogen, coumestrol, on lordosis and estrous cycle in female rats. Brain Res Bull 64:449–454 [PubMed] [Google Scholar]
  • Crews D, Gore AC, Hsu TS, Dangleben NL, Spinetta M, Schallert T, Anway MD, Skinner MK 2007 Transgenerational epigenetic imprints on mate preference. Proc Natl Acad Sci USA 104:5942–5946 [PMC free article] [PubMed] [Google Scholar]
  • Harvey PW, Everett DJ, Springall CJ 2007 Adrenal toxicology: a strategy for assessment of functional toxicity to the adrenal cortex and steroidogenesis. J Appl Toxicol 27:103–115 [PubMed] [Google Scholar]
  • Zoeller RT, Tan SW, Tyl RW 2007 General background on the hypothalamic-pituitary-thyroid (HPT) axis. Crit Rev Toxicol 37:11–53 [PubMed] [Google Scholar]
  • Khan MA, Hansen LG 2003 Ortho-substituted polychlorinated biphenyl (PCB) congeners (95 or 101) decrease pituitary response to thyrotropin releasing hormone. Toxicol Lett 144:173–182 [PubMed] [Google Scholar]
  • Kuriyama SN, Wanner A, Fidalgo-Neto AA, Talsness CE, Koerner W, Chahoud I 2007 Developmental exposure to low-dose PBDE-99: tissue distribution and thyroid hormone levels. Toxicology 242:80–90 [PubMed] [Google Scholar]
  • Sørmo EG, Jüssi I, Jüssi M, Braathen M, Skaare JU, Jenssen BM 2005 Thyroid hormone status in gray seal (Halichoerus grypus) pups from the Baltic Sea and the Atlantic Ocean in relation to organochlorine pollutants. Environ Toxicol Chem 24:610–616 [PubMed] [Google Scholar]
  • Newbold RR, Padilla-Banks E, Jefferson WN 2006 Adverse effects of the model environmental estrogen diethylstilbestrol are transmitted to subsequent generations. Endocrinology 147:S11–S17 [PubMed] [Google Scholar]
  • Noaksson E, Tjärnlund U, Bosveld AT, Balk L 2001 Evidence for endocrine disruption in perch (Perca fluviatilis) and roach (Rutilus rutilus) in a remote Swedish lake in the vicinity of a public refuse dump. Toxicol Appl Pharmacol 174:160–176 [PubMed] [Google Scholar]
  • Simerly RB, Chang C, Muramatsu M, Swanson LW 1990 Distribution of androgen and estrogen receptor mRNA-containing cells in the rat brain: an in situ hybridization study. J Comp Neurol 294:76–95 [PubMed] [Google Scholar]
  • Quadros PS, Pfau JL, Goldstein AY, De Vries GJ, Wagner CK 2002 Sex differences in progesterone receptor expression: a potential mechanism for estradiol-mediated sexual differentiation. Endocrinology 143:3727–3739 [PubMed] [Google Scholar]
  • Chakraborty TR, Gore AC 2004 Aging-related changes in ovarian hormones, their receptors, and neuroendocrine function. Exp Biol Med (Maywood) 229:977–987 [PubMed] [Google Scholar]
  • Tokumoto T, Tokumoto M, Thomas P 2007 Interactions of diethylstilbestrol (DES) and DES analogs with membrane progestin receptor-α and the correlation with their nongenomic progestin activities. Endocrinology 148: 3459–3467 [PubMed] [Google Scholar]
  • Thomas P, Doughty K 2004 Disruption of rapid, nongenomic steroid actions by environmental chemicals: interference with progestin stimulation of sperm motility in Atlantic croaker. Environ Sci Technol 38:6328–6332 [PubMed] [Google Scholar]
  • Watson CS, Bulayeva NN, Wozniak AL, Alyea RA 2007 Xenoestrogens are potent activators of nongenomic estrogenic responses. Steroids 72:124–134 [PMC free article] [PubMed] [Google Scholar]
  • Brucker-Davis F 1998 Effects of environmental synthetic chemicals on thyroid function. Thyroid 8:827–856 [PubMed] [Google Scholar]
  • Howdeshell KL 2002 A model of the development of the brain as a construct of the thyroid system. Environ Health Perspect 110(Suppl 3):337–348 [PMC free article] [PubMed] [Google Scholar]
  • Delange F 1989 Cassava and the thyroid. In: Gaitan E, ed. Environmental goitrogenesis. Boca Raton, FL: CRC Press, Inc.; 173–194 [Google Scholar]
  • Gaitan E 1989 Environmental goitrogenesis. Boca Raton, FL: CRC Press, Inc. [Google Scholar]
  • Carrasco N 2000 Thyroid iodide transport: the Na+/I− symporter (NIS). In: Braverman LE, Utiger RD, eds. Werner & Ingbar’s The Thyroid: a fundamental and clinical text. 8th ed. Philadelphia: Lippincott, Williams & Wilkins; 52–61 [Google Scholar]
  • Zimmermann MB 2007 The adverse effects of mild-to-moderate iodine deficiency during pregnancy and childhood: a review. Thyroid 17:829–835 [PubMed] [Google Scholar]
  • National Research Council 2005 Health implications of perchlorate ingestion. Washington, DC: National Academies Press [Google Scholar]
  • Murray CW, Egan SK, Kim H, Beru N, Bolger PM 2008 US Food and Drug Administration’s Total Diet Study: dietary intake of perchlorate and iodine. J Expo Sci Environ Epidemiol 18:571–580 [PubMed] [Google Scholar]
  • Blount BC, Valentin-Blasini L, Osterloh JD, Mauldin JP, Pirkle JL 2007 Perchlorate exposure of the US population, 2001–2002. J Expo Sci Environ Epidemiol 17:400–407 [PubMed] [Google Scholar]
  • Greer MA, Goodman G, Pleus RC, Greer SE 2002 Health effects assessment for environmental perchlorate contamination: the dose response for inhibition of thyroidal radioiodine uptake in humans. Environ Health Perspect 110:927–937 [PMC free article] [PubMed] [Google Scholar]
  • Blount BC, Pirkle JL, Osterloh JD, Valentin-Blasini L, Caldwell KL 2006 Urinary perchlorate and thyroid hormone levels in adolescent and adult men and women living in the United States. Environ Health Perspect 114:1865–1871 [PMC free article] [PubMed] [Google Scholar]
  • Steinmaus C, Miller MD, Howd R 2007 Impact of smoking and thiocyanate on perchlorate and thyroid hormone associations in the 2001–2002 national health and nutrition examination survey. Environ Health Perspect 115:1333–1338 [PMC free article] [PubMed] [Google Scholar]
  • Zoeller RT, Rovet J 2004 Timing of thyroid hormone action in the developing brain: clinical observations and experimental findings. J Neuroendocrinol 16:809–818 [PubMed] [Google Scholar]
  • Pearce EN, Leung AM, Blount BC, Bazrafshan HR, He X, Pino S, Valentin-Blasini L, Braverman LE 2007 Breast milk iodine and perchlorate concentrations in lactating Boston-area women. J Clin Endocrinol Metab 92:1673–1677 [PubMed] [Google Scholar]
  • Ginsberg GL, Hattis DB, Zoeller RT, Rice DC 2007 Evaluation of the U.S. EPA/OSWER preliminary remediation goal for perchlorate in groundwater: focus on exposure to nursing infants. Environ Health Perspect 115:361–369 [PMC free article] [PubMed] [Google Scholar]
  • Amitai Y, Winston G, Sack J, Wasser J, Lewis M, Blount BC, Valentin-Blasini L, Fisher N, Israeli A, Leventhal A 2007 Gestational exposure to high perchlorate concentrations in drinking water and neonatal thyroxine levels. Thyroid 17:843–850 [PubMed] [Google Scholar]
  • Wolff J 1998 Perchlorate and the thyroid gland. Pharmacol Rev 50:89–105 [PubMed] [Google Scholar]
  • De Groef B, Decallonne BR, Van der Geyten S, Darras VM, Bouillon R 2006 Perchlorate versus other environmental sodium/iodide symporter inhibitors: potential thyroid-related health effects. Eur J Endocrinol 155:17–25 [PubMed] [Google Scholar]
  • Taurog A 2000 Hormone synthesis: thyroid iodine metabolism. In: Braverman L, Utiger R, eds. Werner & Ingbar’s The Thyroid: a fundamental and clinical text. 8th ed. Philadelphia: Lippincott, Williams & Wilkins; 61–85 [Google Scholar]
  • Taurog A 1964 The biosynthesis of thyroxine. Mayo Clin Proc 39:569–585 [PubMed] [Google Scholar]
  • Cooper DS 2003 Antithyroid drugs in the management of patients with Graves’ disease: an evidence-based approach to therapeutic controversies. J Clin Endocrinol Metab 88:3474–3481 [PubMed] [Google Scholar]
  • Mestman JH 1998 Hyperthyroidism in pregnancy. Endocrinol Metab Clin North Am 27:127–149 [PubMed] [Google Scholar]
  • Engler H, Taurog A, Nakashima T 1982 Mechanism of inactivation of thyroid peroxidase by thioureylene drugs. Biochem Pharmacol 31:3801–3806 [PubMed] [Google Scholar]
  • Doerge DR, Sheehan DM 2002 Goitrogenic and estrogenic activity of soy isoflavones. Environ Health Perspect 110(Suppl 3):349–353 [PMC free article] [PubMed] [Google Scholar]
  • Labib M, Gama R, Wright J, Marks V, Robins D 1989 Dietary maladvice as a cause of hypothyroidism and short stature. Brit Med J 298:232–233 [PMC free article] [PubMed] [Google Scholar]
  • Chorazy PA, Himelhoch S, Hopwood NJ, Greger NG, Postellon DC 1995 Persistent hypothyroidism in an infant receiving a soy formula: case report and review of the literature. Pediatrics 96:148–150 [PubMed] [Google Scholar]
  • Jabbar MA, Larrea J, Shaw RA 1997 Abnormal thyroid function tests in infants with congenital hypothyroidism: the influence of soy-based formula. J Am Coll Nutr 16:280–282 [PubMed] [Google Scholar]
  • Fort P, Moses N, Fasano M, Goldberg T, Lifshitz F 1990 Breast and soy-formula feedings in early infancy and the prevalence of autoimmune thyroid disease in children. J Am Coll Nutr 9:164–167 [PubMed] [Google Scholar]
  • Boker LK, Van der Schouw YT, De Kleijn MJ, Jacques PF, Grobbee DE, Peeters PH 2002 Intake of dietary phytoestrogens by Dutch women. J Nutr 132:1319–1328 [PubMed] [Google Scholar]
  • Schussler GC 2000 The thyroxine-binding proteins. Thyroid 10:141–149 [PubMed] [Google Scholar]
  • Vranckx R, Savu L, Maya M, Nunez EA 1990 Characterization of a major development-regulated serum thyroxine-binding globulin in the euthyroid mouse. Biochem J 271:373–379 [PMC free article] [PubMed] [Google Scholar]
  • Robbins J 2000 Thyroid hormone transport proteins and the physiology of hormone binding. In: Braverman L, Utiger R, eds. Werner & Ingbar’s The Thyroid: a fundamental and clinical text. 8th ed. Philadelphia: Lippincott, Williams and Wilkins; 105–120 [Google Scholar]
  • Refetoff S 1989 Inherited thyroxine-binding globulin abnormalities in man. Endocr Rev 10:275–293 [PubMed] [Google Scholar]
  • Mendel CM, Weisiger RA, Jones AL, Cavalieri RR 1987 Thyroid hormone-binding proteins in plasma facilitate uniform distribution of thyroxine within tissues: a perfused rat liver study. Endocrinology 120:1742–1749 [PubMed] [Google Scholar]
  • Power DM, Elias NP, Richardson SJ, Mendes J, Soares CM, Santos CR 2000 Evolution of the thyroid hormone-binding protein, transthyretin. Gen Comp Endocrinol 119:241–255 [PubMed] [Google Scholar]
  • Zheng W, Lu YM, Lu GY, Zhao Q, Cheung O, Blaner WS 2001 Transthyretin, thyroxine, and retinol-binding protein in human cerebrospinal fluid: effect of lead exposure. Toxicol Sci 61:107–114 [PMC free article] [PubMed] [Google Scholar]
  • Robbins J 2002 Transthyretin from discovery to now. Clin Chem Lab Med 40:1183–1190 [PubMed] [Google Scholar]
  • Palha JA, Nissanov J, Fernandes R, Sousa JC, Bertrand L, Dratman MB, Morreale de Escobar G, Gottesman M, Saraiva MJ 2002 Thyroid hormone distribution in the mouse brain: the role of transthyretin. Neuroscience 113:837–847 [PubMed] [Google Scholar]
  • Palha JA, Fernandes R, de Escobar GM, Episkopou V, Gottesman M, Saraiva MJ 2000 Transthyretin regulates thyroid hormone levels in the choroid plexus, but not in the brain parenchyma: study in a transthyretin-null mouse model. Endocrinology 141:3267–3272 [PubMed] [Google Scholar]
  • Meerts IA, van Zanden JJ, Luijks EA, van Leeuwen-Bol I, Marsh G, Jakobsson E, Bergman A, Brouwer A 2000 Potent competitive interactions of some brominated flame retardants and related compounds with human transthyretin in vitro. Toxicol Sci 56:95–104 [PubMed] [Google Scholar]
  • Chauhan KR, Kodavanti PR, McKinney JD 2000 Assessing the role of ortho-substitution on polychlorinated biphenyl binding to transthyretin, a thyroxine transport protein. Toxicol Appl Pharmacol 162:10–21 [PubMed] [Google Scholar]
  • Lans MC, Klasson-Wehler E, Willemsen M, Meussen E, Safe S, Brouwer A 1993 Structure-dependent, competitive interaction of hydroxy-polychlorobiphenyls, -dibenzo-p-dioxins, and -dibenzofurans with human transthyretin. Chem Biol Interact 88:7–21 [PubMed] [Google Scholar]
  • Brouwer A, Morse DC, Lans MC, Schuur AG, Murk AJ, Klasson-Wehler E, Bergman A, Visser TJ 1998 Interactions of persistent environmental organohalogens with the thyroid hormone system: mechanisms and possible consequences for animal and human health. Toxicol Ind Health 14:59–84 [PubMed] [Google Scholar]
  • Richardson SJ, Lemkine GF, Alfama G, Hassani Z, Demeneix BA 2007 Cell division and apoptosis in the adult neural stem cell niche are differentially affected in transthyretin null mice. Neurosci Lett 421:234–238 [PubMed] [Google Scholar]
  • Meerts IA, Assink Y, Cenijn PH, Van Den Berg JH, Weijers BM, Bergman A, Koeman JH, Brouwer A 2002 Placental transfer of a hydroxylated polychlorinated biphenyl and effects on fetal and maternal thyroid hormone homeostasis in the rat. Toxicol Sci 68:361–371 [PubMed] [Google Scholar]
  • Fisher JW, Campbell J, Muralidhara S, Bruckner JV, Ferguson D, Mumtaz M, Harmon B, Hedge JM, Crofton KM, Kim H, Almekinder TL 2006 Effect of PCB 126 on hepatic metabolism of thyroxine and perturbations in the hypothalamic-pituitary-thyroid axis in the rat. Toxicol Sci 90:87–95 [PubMed] [Google Scholar]
  • Glatt CM, Ouyang M, Welsh W, Green JW, Connor JO, Frame SR, Everds NE, Poindexter G, Snajdr S, Delker DA 2005 Molecular characterization of thyroid toxicity: anchoring gene expression profiles to biochemical and pathologic end points. Environ Health Perspect 113:1354–1361 [PMC free article] [PubMed] [Google Scholar]
  • Tseng LH, Li MH, Tsai SS, Lee CW, Pan MH, Yao WJ, Hsu PC 2008 Developmental exposure to decabromodiphenyl ether (PBDE 209): effects on thyroid hormone and hepatic enzyme activity in male mouse offspring. Chemosphere 70:640–647 [PubMed] [Google Scholar]
  • Kretschmer XC, Baldwin WS 2005 CAR and PXR: xenosensors of endocrine disrupters? Chem Biol Interact 155:111–128 [PubMed] [Google Scholar]
  • Rathore M, Bhatnagar P, Mathur D, Saxena GN 2002 Burden of organochlorine pesticides in blood and its effect on thyroid hormones in women. Sci Total Environ 295:207–215 [PubMed] [Google Scholar]
  • Langer P 2008 Persistent organochlorinated pollutants (PCB, DDE, HCB, dioxins, furans) and the thyroid—review 2008. Endocr Regul 42:79–104 [PubMed] [Google Scholar]
  • Köhrle J 2007 Thyroid hormone transporters in health and disease: advances in thyroid hormone deiodination. Best Pract Res Clin Endocrinol Metab 21:173–191 [PubMed] [Google Scholar]
  • Schwartz CE, May MM, Carpenter NJ, Rogers RC, Martin J, Bialer MG, Ward J, Sanabria J, Marsa S, Lewis JA, Echeverri R, Lubs HA, Voeller K, Simensen RJ, Stevenson RE 2005 Allan-Herndon-Dudley syndrome and the monocarboxylate transporter 8 (MCT8) gene. Am J Hum Genet 77:41–53 [PMC free article] [PubMed] [Google Scholar]
  • St Germain DL, Hernandez A, Schneider MJ, Galton VA 2005 Insights into the role of deiodinases from studies of genetically modified animals. Thyroid 15:905–916 [PubMed] [Google Scholar]
  • Schneider MJ, Fiering SN, Pallud SE, Parlow AF, St Germain DL, Galton VA 2001 Targeted disruption of the type 2 selenodeiodinase gene (DIO2) results in a phenotype of pituitary resistance to T4. Mol Endocrinol 15:2137–2148 [PubMed] [Google Scholar]
  • Kato Y, Ikushiro S, Haraguchi K, Yamazaki T, Ito Y, Suzuki H, Kimura R, Yamada S, Inoue T, Degawa M 2004 A possible mechanism for decrease in serum thyroxine level by polychlorinated biphenyls in Wistar and Gunn rats. Toxicol Sci 81:309–315 [PubMed] [Google Scholar]
  • Morse DC, Wehler EK, Wesseling W, Koeman JH, Brouwer A 1996 Alterations in rat brain thyroid hormone status following pre- and postnatal exposure to polychlorinated biphenyls (Aroclor 1254). Toxicol Appl Pharmacol 136:269–279 [PubMed] [Google Scholar]
  • McKinney JD, Waller CL 1994 Polychlorinated biphenyls as hormonally active structural analogues. Environ Health Perspect 102:290–297 [PMC free article] [PubMed] [Google Scholar]
  • Chana A, Concejero MA, de Frutos M, González MJ, Herradón B 2002 Computational studies on biphenyl derivatives. Analysis of the conformational mobility, molecular electrostatic potential, and dipole moment of chlorinated biphenyl: searching for the rationalization of the selective toxicity of polychlorinated biphenyls (PCBs). Chem Res Toxicol 15:1514–1526 [PubMed] [Google Scholar]
  • Breivik K, Sweetman A, Pacyna JM, Jones KC 2002 Towards a global historical emission inventory for selected PCB congeners—a mass balance approach. 1. Global production and consumption. Sci Total Environ 290:181–198 [PubMed] [Google Scholar]
  • Fisher BE1999 Most unwanted. Environ Health Perspect 107:A18–A23 [PMC free article] [PubMed] [Google Scholar]
  • Huisman M, Koopman-Esseboom C, Lanting CI, van der Paauw CG, Tuinstra LG, Fidler V, Weisglas-Kuperus N, Sauer PJ, Boersma ER, Touwen BC 1995 Neurological condition in 18-month-old children perinatally exposed to polychlorinated biphenyls and dioxins. Early Hum Dev 43:165–176 [PubMed] [Google Scholar]
  • Jackson TA, Richer JK, Bain DL, Takimoto GS, Tung L, Horwitz KB 1997 The partial agonist activity of antagonist-occupied steroid receptors is controlled by a novel hinge domain-binding coactivator L7/SPA and the corepressors N-CoR or SMRT. Mol Endocrinol 11:693–705 [PubMed] [Google Scholar]
  • Osius N, Karmaus W, Kruse H, Witten J 1999 Exposure to polychlorinated biphenyls and levels of thyroid hormones in children. Environ Health Perspect 107:843–849 [PMC free article] [PubMed] [Google Scholar]
  • Ayotte P, Muckle G, Jacobson JL, Jacobson SW, Dewailly E 2003 Assessment of pre- and postnatal exposure to polychlorinated biphenyls: lessons from the Inuit Cohort Study. Environ Health Perspect 111:1253–1258 [PMC free article] [PubMed] [Google Scholar]
  • Walkowiak J, Wiener JA, Fastabend A, Heinzow B, Krämer U, Schmidt E, Steingrüber HJ, Wundram S, Winneke G 2001 Environmental exposure to polychlorinated biphenyls and quality of the home environment: effects on psychodevelopment in early childhood. Lancet 358:1602–1607 [PubMed] [Google Scholar]
  • Zoeller RT, Dowling AL, Vas AA 2000 Developmental exposure to polychlorinated biphenyls exerts thyroid hormone-like effects on the expression of RC3/neurogranin and myelin basic protein messenger ribonucleic acids in the developing rat brain. Endocrinology 141:181–189 [PubMed] [Google Scholar]
  • Goldey ES, Kehn LS, Lau C, Rehnberg GL, Crofton KM 1995 Developmental exposure to polychlorinated biphenyls (Aroclor 1254) reduces circulating thyroid hormone concentrations and causes hearing deficits in rats. Toxicol Appl Pharmacol 135:77–88 [PubMed] [Google Scholar]
  • Bastomsky CH 1977 Enhanced thyroxine metabolism and high uptake goiters in rats after a single dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Endocrinology 101:292–296 [PubMed] [Google Scholar]
  • Crofton KM 2004 Developmental disruption of thyroid hormone: correlations with hearing dysfunction in rats. Risk Anal 24:1665–1671 [PubMed] [Google Scholar]
  • Brouwer A, Longnecker MP, Birnbaum LS, Cogliano J, Kostyniak P, Moore J, Schantz S, Winneke G 1999 Characterization of potential endocrine-related health effects at low-dose levels of exposure to PCBs. Environ Health Perspect 107(Suppl 4):639–649 [PMC free article] [PubMed] [Google Scholar]
  • Schell LM, Gallo MV, Denham M, Ravenscroft J, DeCaprio AP, Carpenter DO 2008 Relationship of thyroid hormone levels to levels of polychlorinated biphenyls, lead, p,p’- DDE, and other toxicants in Akwesasne Mohawk youth. Environ Health Perspect 116:806–813 [PMC free article] [PubMed] [Google Scholar]
  • Herbstman JB, Sjödin A, Apelberg BJ, Witter FR, Halden RU, Patterson DG, Panny SR, Needham LL, Goldman LR 2008 Birth delivery mode modifies the associations between prenatal polychlorinated biphenyl (PCB) and polybrominated diphenyl ether (PBDE) and neonatal thyroid hormone levels. Environ Health Perspect 116:1376–1382 [PMC free article] [PubMed] [Google Scholar]
  • Herbstman J, Apelberg BJ, Witter FR, Panny S, Goldman LR 2008 Maternal, infant, and delivery factors associated with neonatal thyroid hormone status. Thyroid 18:67–76 [PubMed] [Google Scholar]
  • Goldey ES, Crofton KM 1998 Thyroxine replacement attenuates hypothyroxinemia, hearing loss, and motor deficits following developmental exposure to Aroclor 1254 in rats. Toxicol Sci 45:94–105 [PubMed] [Google Scholar]
  • Koibuchi N, Chin WW 2000 Thyroid hormone action and brain development. Trends Endocrinol Metab 11: 123–128 [PubMed] [Google Scholar]
  • Nguon K, Baxter MG, Sajdel-Sulkowska EM 2005 Perinatal exposure to polychlorinated biphenyls differentially affects cerebellar development and motor functions in male and female rat neonates. Cerebellum 4:112–122 [PubMed] [Google Scholar]
  • Granholm AC 1985 Effects of thyroid hormone deficiency on glial constituents in developing cerebellum of the rat. Exp Brain Res 59:451–456 [PubMed] [Google Scholar]
  • Stewart P, Fitzgerald S, Reihman J, Gump B, Lonky E, Darvill T, Pagano J, Hauser P 2003 Prenatal PCB exposure, the corpus callosum, and response inhibition. Environ Health Perspect 111:1670–1677 [PMC free article] [PubMed] [Google Scholar]
  • Berbel P, Guadaño-Ferraz A, Angulo A, Ramón Cerezo J 1994 Role of thyroid hormones in the maturation of interhemispheric connections in rats. Behav Brain Res 64:9–14 [PubMed] [Google Scholar]
  • Gauger KJ, Kato Y, Haraguchi K, Lehmler HJ, Robertson LW, Bansal R, Zoeller RT 2004 Polychlorinated biphenyls (PCBs) exert thyroid hormone-like effects in the fetal rat brain but do not bind to thyroid hormone receptors. Environ Health Perspect 112:516–523 [PMC free article] [PubMed] [Google Scholar]
  • Bansal R, You SH, Herzig CT, Zoeller RT 2005 Maternal thyroid hormone increases HES expression in the fetal rat brain: an effect mimicked by exposure to a mixture of polychlorinated biphenyls (PCBs). Brain Res Dev Brain Res 156:13–22 [PubMed] [Google Scholar]
  • Kitamura S, Jinno N, Suzuki T, Sugihara K, Ohta S, Kuroki H, Fujimoto N 2005 Thyroid hormone-like and estrogenic activity of hydroxylated PCBs in cell culture. Toxicology 208:377–387 [PubMed] [Google Scholar]
  • Fritsche E, Cline JE, Nguyen NH, Scanlan TS, Abel J 2005 Polychlorinated biphenyls disturb differentiation of normal human neural progenitor cells: clue for involvement of thyroid hormone receptors. Environ Health Perspect 113:871–876 [PMC free article] [PubMed] [Google Scholar]
  • Arulmozhiraja S, Morita M 2004 Structure-activity relationships for the toxicity of polychlorinated dibenzofurans: approach through density functional theory-based descriptors. Chem Res Toxicol 17:348–356 [PubMed] [Google Scholar]
  • Kimura-Kuroda J, Nagata I, Kuroda Y 2005 Hydroxylated metabolites of polychlorinated biphenyls inhibit thyroid-hormone-dependent extension of cerebellar Purkinje cell dendrites. Brain Res Dev Brain Res 154:259–263 [PubMed] [Google Scholar]
  • Bogazzi F, Raggi F, Ultimieri F, Russo D, Campomori A, McKinney JD, Pinchera A, Bartalena L, Martino E 2003 Effects of a mixture of polychlorinated biphenyls (Aroclor 1254) on the transcriptional activity of thyroid hormone receptor. J Endocrinol Invest 26:972–978 [PubMed] [Google Scholar]
  • Iwasaki T, Miyazaki W, Takeshita A, Kuroda Y, Koibuchi N 2002 Polychlorinated biphenyls suppress thyroid hormone-induced transactivation. Biochem Biophys Res Commun 299:384–388 [PubMed] [Google Scholar]
  • Miyazaki W, Iwasaki T, Takeshita A, Kuroda Y, Koibuchi N 2004 Polychlorinated biphenyls suppress thyroid hormone receptor-mediated transcription through a novel mechanism. J Biol Chem 279:18195–18202 [PubMed] [Google Scholar]
  • Frumess RD, Larsen PR 1975 Correlation of serum triodothyronine (T3) and thyroxine (T4) with biologic effects of thyroid hormone replacement in propylthiouracil-treated rats. Metabolism 24:547–554 [PubMed] [Google Scholar]
  • Howe S, Borodinsky L, Lyon R 1998 Potential exposure to bisphenol A from food-contact use of epoxy coated cans. J Coat Tech 70:69–74 [Google Scholar]
  • Lewis JB, Rueggeberg FA, Lapp CA, Ergle JW, Schuster GS 1999 Identification and characterization of estrogen-like components in commercial resin-based dental restorative materials. Clin Oral Investig 3:107–113 [PubMed] [Google Scholar]
  • Schönfelder G, Wittfoht W, Hopp H, Talsness CE, Paul M, Chahoud I 2002 Parent bisphenol A accumulation in the human maternal-fetal-placental unit. Environ Health Perspect 110:A703–A707 [PMC free article] [PubMed] [Google Scholar]
  • 1995 Tetrabromobisphenol A and derivatives. Environmental Health Critera no. 172. Geneva: World Health Organization [Google Scholar]
  • 1997 Flame-retardants: a general introduction. Environmental Health Critera no. 192. Geneva: World Health Organization [Google Scholar]
  • Thomsen C, Lundanes E, Becher G 2002 Brominated flame retardants in archived serum samples from Norway: a study on temporal trends and the role of age. Environ Sci Technol 36:1414–1418 [PubMed] [Google Scholar]
  • Moriyama K, Tagami T, Akamizu T, Usui T, Saijo M, Kanamoto N, Hataya Y, Shimatsu A, Kuzuya H, Nakao K 2002 Thyroid hormone action is disrupted by bisphenol A as an antagonist. J Clin Endocrinol Metab 87:5185–5190 [PubMed] [Google Scholar]
  • Staples CA, Dorn PB, Klecka GM, O'Block ST, Harris LR 1998 A review of the environmental fate, effects, and exposures of bisphenol A. Chemosphere 36:2149–2173 [PubMed] [Google Scholar]
  • Krishnan AV, Stathis P, Permuth SF, Tokes L, Feldman D 1993 Bisphenol-A: an estrogenic substance is released from polycarbonate flasks during autoclaving. Endocrinology 132:2279–2286 [PubMed] [Google Scholar]
  • Gaido KW, Leonard LS, Lovell S, Gould JC, Babaï D, Portier CJ, McDonnell DP 1997 Evaluation of chemicals with endocrine modulating activity in a yeast-based steroid hormone receptor gene transcription assay. Toxicol Appl Pharmacol 143:205–212 [PubMed] [Google Scholar]
  • Wetherill YB, Akingbemi BT, Kanno J, McLachlan JA, Nadal A, Sonnenschein C, Watson CS, Zoeller RT, Belcher SM 2007 In vitro molecular mechanisms of bisphenol A action. Reprod Toxicol 24:178–198 [PubMed] [Google Scholar]
  • Kitamura S, Jinno N, Ohta S, Kuroki H, Fujimoto N 2002 Thyroid hormonal activity of the flame retardants tetrabromobisphenol A and tetrachlorobisphenol A. Biochem Biophys Res Commun 293:554–559 [PubMed] [Google Scholar]
  • Zoeller RT, Bansal R, Parris C 2005 Bisphenol-A, an environmental contaminant that acts as a thyroid hormone receptor antagonist in vitro, increases serum thyroxine, and alters RC3/neurogranin expression in the developing rat brain. Endocrinology 146:607–612 [PubMed] [Google Scholar]
  • Cheng SY 2005 Thyroid hormone receptor mutations and disease: beyond thyroid hormone resistance. Trends Endocrinol Metab 16:176–182 [PubMed] [Google Scholar]
  • Seiwa C, Nakahara J, Komiyama T, Katsu Y, Iguchi T, Asou H 2004 Bisphenol A exerts thyroid-hormone-like effects on mouse oligodendrocyte precursor cells. Neuroendocrinology 80:21–30 [PubMed] [Google Scholar]
  • Vermiglio F, Lo Presti VP, Moleti M, Sidoti M, Tortorella G, Scaffidi G, Castagna MG, Mattina F, Violi MA, Crisà A, Artemisia A, Trimarchi F 2004 Attention deficit and hyperactivity disorders in the offspring of mothers exposed to mild-moderate iodine deficiency: a possible novel iodine deficiency disorder in developed countries. J Clin Endocrinol Metab 89:6054–6060 [PubMed] [Google Scholar]
  • Siesser WB, Cheng SY, McDonald MP 2005 Hyperactivity, impaired learning on a vigilance task, and a differential response to methylphenidate in the TRβPV knock-in mouse. Psychopharmacology (Berl) 181:653–663 [PubMed] [Google Scholar]
  • Ishido M, Masuo Y, Kunimoto M, Oka S, Morita M 2004 Bisphenol A causes hyperactivity in the rat concomitantly with impairment of tyrosine hydroxylase immunoreactivity. J Neurosci Res 76:423–433 [PubMed] [Google Scholar]
  • Dowling AL, Martz GU, Leonard JL, Zoeller RT 2000 Acute changes in maternal thyroid hormone induce rapid and transient changes in gene expression in fetal rat brain. J Neurosci 20:2255–2265 [PMC free article] [PubMed] [Google Scholar]
  • Dowling AL, Iannacone EA, Zoeller RT 2001 Maternal hypothyroidism selectively affects the expression of neuroendocrine-specific protein A messenger ribonucleic acid in the proliferative zone of the fetal rat brain cortex. Endocrinology 142:390–399 [PubMed] [Google Scholar]
  • Dowling AL, Zoeller RT 2000 Thyroid hormone of maternal origin regulates the expression of RC3/neurogranin mRNA in the fetal rat brain. Brain Res Mol Brain Res 82:126–132 [PubMed] [Google Scholar]
  • Wu Y, Liu Y, Levine EM, Rao MS 2003 Hes1 but not Hes5 regulates an astrocyte versus oligodendrocyte fate choice in glial restricted precursors. Dev Dyn 226:675–689 [PubMed] [Google Scholar]
  • Schuurmans C, Guillemot F 2002 Molecular mechanisms underlying cell fate specification in the developing telencephalon. Curr Opin Neurobiol 12:26–34 [PubMed] [Google Scholar]
  • Talsness CE 2008 Overview of toxicological aspects of polybrominated diphenyl ethers: a flame-retardant additive in several consumer products. Environ Res 108:158–167 [PubMed] [Google Scholar]
  • Marsh G, Bergman A, Bladh L, Glillner M, Jackobsson E 1998 Synthesis of p-hydroxybromodiphnyl ethers and binding to the thyroid hormone receptor. Organohalogen Compounds (37):305–308 [Google Scholar]
  • Newbold RR, Padilla-Banks E, Snyder RJ, Phillips TM, Jefferson WN 2007 Developmental exposure to endocrine disruptors and the obesity epidemic. Reprod Toxicol 23:290–296 [PMC free article] [PubMed] [Google Scholar]
  • Newbold RR, Padilla-Banks E, Snyder RJ, Jefferson WN 2007 Perinatal exposure to environmental estrogens and the development of obesity. Mol Nutr Food Res 51:912–917 [PubMed] [Google Scholar]
  • De Ferranti SD, Osganian SK 2007 Epidemiology of paediatric metabolic syndrome and type 2 diabetes mellitus. Diabetes Vasc Dis Res 4:285–296 [PubMed] [Google Scholar]
  • Baillie-Hamilton PF 2002 Chemical toxins: a hypothesis to explain the global obesity epidemic. J Altern Complement Med 8:185–192 [PubMed] [Google Scholar]
  • Grün F, Blumberg B 2006 Environmental obesogens: organotins and endocrine disruption via nuclear receptor signaling. Endocrinology 147:S50–S55 [PubMed] [Google Scholar]
  • Tabb MM, Blumberg B 2006 New modes of action for endocrine-disrupting chemicals. Mol Endocrinol 20:475–482 [PubMed] [Google Scholar]
  • Hales CN, Barker DJ 2001 The thrifty phenotype hypothesis. Br Med Bull 60:5–20 [PubMed] [Google Scholar]
  • Buresova M, Zidek V, Musilova A, Simakova M, Fucikova A, Bila V, Kren V, Kazdova L, Di Nicolantonio R, Pravenec M 2006 Genetic relationship between placental and fetal weights and markers of the metabolic syndrome in rat recombinant inbred strains. Physiol Genomics 26:226–231 [PubMed] [Google Scholar]
  • de Moura EG, Passos MC 2005 Neonatal programming of body weight regulation and energetic metabolism. Biosci Rep 25:251–269 [PubMed] [Google Scholar]
  • Cooke PS, Naaz A 2004 Role of estrogens in adipocyte development and function. Exp Biol Med 229:1127–1135 [PubMed] [Google Scholar]
  • Wada K, Sakamoto H, Nishikawa K, Sakuma S, Nakajima A, Fujimoto Y, Kamisaki Y 2007 Life style-related diseases of the digestive system: Endocrine disruptors stimulate lipid accumulation in target cells related to metabolic syndrome. J Pharmacol Sci 105:133–137 [PubMed] [Google Scholar]
  • Knouff C, Auwerx J 2004 Peroxisome proliferator-activated receptor-γ calls for activation in moderation: lessons from genetics and pharmacology. Endocr Rev 25:899–918 [PubMed] [Google Scholar]
  • Kanayama T, Kobayashi N, Mamiya S, Nakanishi T, Nishikawa J 2005 Organotin compounds promote adipocyte differentiation as agonists of the peroxisome proliferator-activated receptor -γ /retinoid X receptor pathway. Mol Pharmacol 67:766–774 [PubMed] [Google Scholar]
  • Grün F, Watanabe H, Zamanian Z, Maeda L, Arima K, Cubacha R, Gardiner DM, Kanno J, Iguchi T, Blumberg B 2006 Endocrine-disrupting organotin compounds are potent inducers of adipogenesis in vertebrates. Mol Endocrinol 20:2141–2155 [PubMed] [Google Scholar]
  • Dang ZC, Audinot V, Papapoulos SE, Boutin JA, Löwik CW 2003 Peroxisome proliferator-activated receptor γ (PPAR-γ) as a molecular target for the soy phytoestrogen genistein. J Biol Chem 278:962–967 [PubMed] [Google Scholar]
  • Penza M, Montani C, Romani A, Vignolini P, Pampaloni B, Tanini A, Brandi ML, Alonso-Magdalena P, Nadal A, Ottobrini L, Parolini O, Bignotti E, Calza S, Maggi A, Grigolato PG, Di Lorenzo D 2006 Genistein affects adipose tissue deposition in a dose-dependent and gender-specific manner. Endocrinology 147:5740–5751 [PubMed] [Google Scholar]
  • Alonso-Magdalena P, Morimoto S, Ripoll C, Fuentes E, Nadal A 2006 The estrogenic effect of bisphenol-A disrupts pancreatic β-cell function in vivo and induces insulin resistance. Environ Health Perspect 114:106–112 [PMC free article] [PubMed] [Google Scholar]
  • Wild S, Roglic G, Green A, Sicree R, King H 2004 Global prevalence of diabetes estimates for the year 2000 and projections for 2030. Diabetes Care 27:1047–1053 [PubMed] [Google Scholar]
  • Rosenbloom AL, Joe JR, Young RS, Winter WE 1999 Emerging epidemic of type 2 diabetes in youth. Diabetes Care 22:345–354 [PubMed] [Google Scholar]
  • Remillard RB, Bunce NJ 2002 Linking dioxins to diabetes: epidemiology and biologic plausibility. Environ Health Perspect 110:853–858 [PMC free article] [PubMed] [Google Scholar]
  • Hugo ER, Brandebourg TD, Woo JG, Loftus J, Alexander JW, Ben-Jonathan N 2008 Bisphenol A at environmentally relevant doses inhibits adiponectin release from human adipose tissue explants and adipocytes. Environ Health Perspect 116:1642–1647 [PMC free article] [PubMed] [Google Scholar]
  • Alonso-Magdalena P, Laribi O, Ropero AB, Fuentes E, Ripoll C, Soria B, Nadal A 2005 Low doses of bisphenol A and diethylstilbestrol impair Ca2+ signals in pancreatic α-cells through a nonclassical membrane estrogen receptor within intact islets of Langerhans. Environ Health Perspect 113:969–977 [PMC free article] [PubMed] [Google Scholar]
  • Poirier P, Després JP 2003 Waist circumference, visceral obesity, and cardiovascular risk. J Cardiopulm Rehabil 23:161–169 [PubMed] [Google Scholar]
  • Gardner JD, Brower GL, Voloshenyuk TG, Janicki JS 2008 Cardioprotection in female rats subjected to chronic volume overload: synergistic interaction of estrogen and phytoestrogens. Am J Physiol Heart Circ Physiol 294:H198–H204 [PubMed] [Google Scholar]
  • Deodato B, Altavilla D, Squadrito G, Campo GM, Arlotta M, Minutoli L, Saitta A, Cucinotta D, Calapai G, Caputi AP, Miano M, Squadrito F 1999 Cardioprotection by the phytoestrogen genistein in experimental myocardial ischaemia-reperfusion injury. Br J Pharmacol 128:1683–1690 [PMC free article] [PubMed] [Google Scholar]
  • Chan YH, Lau KK, Yiu KH, Li SW, Chan HT, Tam S, Shu XO, Lau CP, Tse HF 2007 Isoflavone intake in persons at high risk of cardiovascular events: implications for vascular endothelial function and the carotid atherosclerotic burden. Am J Clin Nutr 86:938–945 [PubMed] [Google Scholar]
  • de Kleijn MJ, van der Schouw YT, Wilson PW, Grobbee DE, Jacques PF 2002 Dietary intake of phytoestrogens is associated with a favorable metabolic cardiovascular risk profile in postmenopausal U.S. women: the Framingham study. J Nutr 132:276–282 [PubMed] [Google Scholar]
  • Cerami C, Founds H, Nicholl I, Mitsuhashi T, Giordano D, Vanpatten S, Lee A, Al-Abed Y, Vlassara H, Bucala R, Cerami A 1997 Tobacco smoke is a source of toxic reactive glycation products. Proc Natl Acad Sci USA 94:13915–13920 [PMC free article] [PubMed] [Google Scholar]
  • Koschinsky T, He CJ, Mitsuhashi T, Bucala R, Liu C, Buenting C, Heitmann K, Vlassara H 1997 Orally absorbed reactive glycation products (glycotoxins): an environmental risk factor in diabetic nephropathy. Proc Natl Acad Sci USA 94:6474–6479 [PMC free article] [PubMed] [Google Scholar]
  • Sandu O, Song K, Cai W, Zheng F, Uribarri J, Vlassara H 2005 Insulin resistance and type 2 diabetes in high-fat-fed mice are linked to high glycotoxin intake. Diabetes 54:2314–2319 [PubMed] [Google Scholar]
  • Diamanti-Kandarakis E, Piperi C, Korkolopoulou P, Kandaraki E, Levidou G, Papalois A, Patsouris E, Papavassiliou AG 2007 Accumulation of dietary glycotoxins in the reproductive system of normal female rats. J Mol Med 85:1413–1420 [PubMed] [Google Scholar]
  • Uribarri J, Stirban A, Sander D, Cai W, Negrean M, Buenting CE, Koschinsky T, Vlassara H 2007 Single oral challenge by advanced glycation end products acutely impairs endothelial function in diabetic and nondiabetic subjects. Diabetes Care 30:2579–2582 [PubMed] [Google Scholar]
  • McLachlan JA 2006 Commentary: prenatal exposure to diethylstilbestrol (DES): a continuing story. Int J Epidemiol 35:868–870 [PubMed] [Google Scholar]
  • Lang IA, Galloway TS, Scarlett A, Henley WE, Depledge M, Wallace RB, Melzer D 2008 Association of urinary bisphenol A concentration with medical disorders and laboratory abnormalities in adults. JAMA 300:1303–1310 [PubMed] [Google Scholar]
  • Emmen JM, McLuskey A, Adham IM, Engel W, Verhoef-Post M, Themmen AP, Grootegoed JA, Brinkmann AO 2000 Involvement of insulin-like factor 3 (Insl3) in diethylstilbestrol-induced cryptorchidism. Endocrinology 141:846–849 [PubMed] [Google Scholar]
  • Ben Rhouma K, Tébourbi O, Krichah R, Sakly M 2001 Reproductive toxicity of DDT in adult male rats. Hum Exp Toxicol 20:393–397 [PubMed] [Google Scholar]
  • Fisher JS, Macpherson S, Marchetti N, Sharpe RM 2003 Human testicular dysgenesis syndrome: a possible model using in-utero exposure of the rat to dibutyl phthalate. Hum Reprod 18:1383–1394 [PubMed] [Google Scholar]
  • Parks LG, Ostby JS, Lambright CR, Abbott BD, Klinefelter GR, Barlow NJ, Gray Jr LE 2000 The plasticizer diethylhexyl phthalate induces malformations by decreasing fetal testosterone synthesis during sexual differentiation in the male rat. Toxicol Sci 58:339–349 [PubMed] [Google Scholar]
  • Gupta C 2000 Reproductive malformation of the male offspring following maternal exposure to estrogenic chemicals. Proc Soc Exp Biol Med 224:61–68 [PubMed] [Google Scholar]
  • Timms BG, Howdeshell KL, Barton L, Bradley S, Richter CA, vom Saal FS 2005 Estrogenic chemicals in plastic and oral contraceptives disrupt development of the fetal mouse prostate and urethra. Proc Natl Acad Sci USA 102:7014–7019 [PMC free article] [PubMed] [Google Scholar]
  • Ramos JG, Varayoud J, Sonnenschein C, Soto AM, MuñozDe Toro M, Luque EH 2001 Prenatal exposure to low doses of bisphenol A alters the periductal stroma and glandular cell function in the rat ventral prostate. Biol Reprod 65:1271–1277 [PubMed] [Google Scholar]
  • Cohn BA, Cirillo PM, Wolff MS, Schwingl PJ, Cohen RD, Sholtz RI, Ferrara A, Christianson RE, van den Berg BJ, Siiteri PK 2003 DDT and DDE exposure in mothers and time to pregnancy in daughters. Lancet 361:2205–2206 [PubMed] [Google Scholar]
  • Markey CM, Wadia PR, Rubin BS, Sonnenschein C, Soto AM 2005 Long-term effects of fetal exposure to low doses of the xenoestrogen bisphenol-A in the female mouse genital tract. Biol Reprod 72:1344–1351 [PubMed] [Google Scholar]
  • Howdeshell KL, Hotchkiss AK, Thayer KA, Vandenbergh JG, vom Saal FS 1999 Exposure to bisphenol A advances puberty. Nature 401:763–764 [PubMed] [Google Scholar]
  • Honma S, Suzuki A, Buchanan DL, Katsu Y, Watanabe H, Iguchi T 2002 Low dose effect of in utero exposure to bisphenol A and diethylstilbestrol on female mouse reproduction. Reprod Toxicol 16:117–122 [PubMed] [Google Scholar]
  • Zsarnovszky A, Le HH, Wang HS, Belcher SM 2005 Ontogeny of rapid estrogen-mediated extracellular signal-regulated kinase signaling in the rat cerebellar cortex: potent nongenomic agonist and endocrine disrupting activity of the xenoestrogen bisphenol A. Endocrinology 146:5388–5396 [PubMed] [Google Scholar]
  • Steinberg RM, Walker DM, Juenger TE, Woller MJ, Gore AC 2008 The effects of perinatal PCBs on adult female rat reproduction: development, reproductive physiology, and second generational effects. Biol Reprod 78:1091–1101 [PMC free article] [PubMed] [Google Scholar]
  • Jenkins S, Rowell C, Wang J, Lamartiniere CA 2007 Prenatal TCDD exposure predisposes for mammary cancer in rats. Reprod Toxicol 23:391–396 [PMC free article] [PubMed] [Google Scholar]
  • Mizuyachi K, Son DS, Rozman KK, Terranova PF 2002 Alteration in ovarian gene expression in response to 2,3,7,8-tetrachlorodibenzo-p-dioxin: reduction of cyclooxygenase-2 in the blockage of ovulation. Reprod Toxicol 16:299–307 [PubMed] [Google Scholar]
  • Newbold R 1995 Cellular and molecular effects of developmental exposure to diethylstilbestrol: implications for other environmental estrogens. Environ Health Perspect 103(Suppl 7):83–87 [PMC free article] [PubMed] [Google Scholar]
  • Kaufman RH 1982 Structural changes of the genital tract associated with in utero exposure to diethylstilbestrol. Obstet Gynecol Annu 11:187–202 [PubMed] [Google Scholar]
  • Krstevska-Konstantinova M, Charlier C, Craen M, Du Caju M, Heinrichs C, de Beaufort C, Plomteux G, Bourguignon JP 2001 Sexual precocity after immigration from developing countries to Belgium: evidence of previous exposure to organochlorine pesticides. Hum Reprod 16:1020–1026 [PubMed] [Google Scholar]
  • Blanck HM, Marcus M, Tolbert PE, Rubin C, Henderson AK, Hertzberg VS, Zhang RH, Cameron L 2000 Age at menarche and Tanner stage in girls exposed in utero and postnatally to polybrominated biphenyl. Epidemiology 11:641–647 [PubMed] [Google Scholar]
  • Martin OV, Lester JN, Voulvoulis N, Boobis AR 2007 Forum: human health and endocrine disruption: a simple multicriteria framework for the qualitative assessment of end point-specific risks in a context of scientific uncertainty. Toxicol Sci 98:332–347 [PubMed] [Google Scholar]
  • Diamanti-Kandarakis E, Piperi C, Patsouris E, Korkolopoulou P, Panidis D, Pawelczyk L, Papavassiliou AG, Duleba AJ 2007 Immunohistochemical localization of advanced glycation end-products (AGEs) and their receptor (RAGE) in polycystic and normal ovaries. Histochem Cell Biol 127:581–589 [PubMed] [Google Scholar]


Page 2

Disorders of the human reproductive system possibly involving EDCs in their pathogenesis: A sexually dimorphic life cycle perspective

Fetal/neonatalPrepubertalPubertalAdult
ProcessesIntrauterine growthAdrenarcheGonadarcheSpermatogenesis
Sexual differentiationOvulation
Hormonal control of prostate, breast, uterus, and lactation
Male disordersIUGR (15)Premature pubarcheSmall testes and high FSH (18)Oligospermia (14,20)a
Cryptorchidism (14,20)aEarly puberty (25)Testicular cancer (14,20)a
Hypospadias (14,20)aDelayed puberty (25)Prostate hyperplasia (24)
Female disordersIUGRPremature thelarche (25)Secondary central precocious puberty (17,27)Vaginal adenocarcinoma (19,28)
Peripheral precocious puberty (17)PCOS (18,25)Disorders of ovulation (29)
Premature pubarche (18)Delayed ovulatory cycles (17,18)Benign breast disease (29,31)
Breast cancer (30,31)
Uterine fibroids (29)
Disturbed lactation (29)