Bao Ting Zhu and Allan H.Conney1


Laboratory for Cancer Research, Department of Chemical Biology, College of Pharmacy, Rutgers-The State University of New Jersey, Piscataway, NJ 08854-8020, USA.


* Abbreviations: NADPH, β-nicotinamide adenine dinucleotide phosphate (reduced form); TCDD, 2,3,7,8-tetrachlorodibenzo-p-dioxin; COMT, catechol-methyltransferase; 17β-HSD, 17β-hydroxysteroid dehydrogenase.


1:To whom correspondence should be addressed; E-mail: aconney@rci.rutgers.edu.


Cytochrome P450 enzymes that metabolize estrogens are expressed in the mammary gland, uterus, brain and other target tissues for estrogen action, and this results in the formation of hydroxylated estrogens in these tissues. Estra-diol metabolites formed in target tissues at or near estrogen receptors may either be inactive or have important bio-logical effects, and changes in the activities of estrogen-etabolizing enzymes in target tissues may profoundly influence estrogen action. Although some active estrogen metabolites exert hormonal effects in target tissues by interaction with the classical estrogen receptor, other meta-bolites appear to elicit unique biological responses that are not associated with activation of this receptor. Therefore, some of the many actions of estradiol may not be caused by estradiol per se, but may result from the formation of active estrogen metabolite(s) which function as local mediators or may activate their own unique receptors or effectors. This is an important area in need of more research. The present paper represents a review of the literature and perspectives by the authors on the functional role of estrogen metabolism in target tissues.



Introduction


Estrogens exert diverse biological effects in animals and humans, and many of these effects result from a direct interaction of the estrogen with an intracellular receptor that activates the expression of genes encoding proteins with important biological functions (1-4). One of the most important and notable effects of estrogens is a superpotent mitogenic action in hormone sensitive tissues such as the uterus (5,6) and breast (7-9). Prolonged exposure of target tissues or cells to excessive mitogenic stimulation by natural or synthetic estrogens has long been considered an important etiological factor for the induction of estrogen-associated cancers in experimental animals (10,11) and humans (10,12-18).


Estrogenic hormones are eliminated from the body by metabolic conversion to hormonally inactive (or less active) water-soluble metabolites that are excreted in the urine and/or feces. The metabolic disposition of estrogens includes oxidative metabolism (largely hydroxylations; reviewed in ref. 19) and conjugative metabolism by glucuronidation (20-22), sulfonation (23-27) and/or D-methylation (28-30). Members of the cytochrome P450 family are the major enzymes catalyzing β-nicotinamide adenine dinucleotide phosphate (reduced form) (NADPH*)-dependent oxidative metabolism of estrogens to multiple hydroxylated metabolites (summarized in Figure 1 and Table I). Although most of the oxidative metabolism of estrogens takes place in liver, some estrogen-metabolizing isoforms of the cytochromes P450 that are usually expressed at low or undetectable levels in liver are selectively expressed in certain extrahepatic tissues.



Figure 1. . Sites of oxidative metabolism of estradiol by NADPH-dependent cytochrome P450 enzymes. Major hepatic pathways of estradiol metabolism (2-hydroxylation, 16α-hydroxylation, and estrone formation) are indicated by the solid arrows. Additional information about oxidative estradiol metabolism is summarized in Table I.


Table 1. Hydroxylated and keto metabolites of 17β-estradiol (estradiol) and estrone. All estrogen metabolites listed were found in biological samples(e.g. tissues, blood, bile, urine) or were formed during in vitro incubations of estrogens with enzyme preparations from animals or humans.



An early suggestion that an estrogen may exert some of its effects because of metabolism in a target tissue came from studies by Fishman and Norton (31). These investigators demonstrated the 2-hydroxylation of estradiol by the rat brain. This work, coupled with earlier studies showing an inhibitory effect of 2-hydroxyestradiol on the inactivation of centrally active catecholamines by catechol-D-methyltransferase (32-34), indicated that a locally formed estrogen metabolite may exert a biological effect important for the action of the parent hormone. Although research on the metabolism of estrogens by target tissues has been pursued during the past 20 years, and reviews on certain aspects of this topic have appeared (30,35-40), the functional role or importance of the NADPH-ependent hydroxylations of estradiol and estrone by multiple cytochrome P450 enzymes in target tissues or cells is largely unknown. In this paper, we have reviewed some of our data as well as data by others which collectively suggest a general concept that certain hydroxylated estrogen metabolites formed by specific enzymes in target cells may possess important and unique biological functions that are not directly associated the estrogen. Based on a review of data scattered in the literature, we suggest that some of the effects exerted by active estrogen metabolites may be mediated by specific intracellular receptors or effectors which are different from the classical estrogen receptor. In this context, it is noteworthy that a novel estrogen receptor with unknown function(s) has recently been identified in rat prostate and ovary (41) and it is likely that additional isoforms of the classical estrogen receptor also exist (42-45). Studies on possible interactions of the multiple estrogen metabolites with these novel receptors are needed.


In addition to the NADPH-dependent hydroxylation of estrogens catalyzed mainly by cytochrome P450 enzymes, there are also several other pathways of estrogen metabolism in target cells (such as sulfonation, desulfonation, and interconversion between estradiol and estrone) which have been quite well studied in the past. It is known that desulfonation of estrogen sulfates by sulfatase or aromatization of androgens by aromatase in target cells contributes substantially to the formation of parent estrogen in these cells and thereby enhances the hormonal stimulation of the classical estrogen receptor. In contrast, metabolism of estrogens by conjugative enzymes (e.g. sulfotransferase) present in target cells may decrease the hormonal activity of estrogens by lowering the intracellular concentration of the parent hormone (25,35,36,39).


Although a major focus of this paper is on NADPH-dependent hydroxylation of estrogens in extrahepatic target tissues, the importance of several other metabolic pathways for the meta-bolism of estrogens in these target tissues has also been discussed.



Multiple pathways of NADPH-dependent estrogen hydroxylation


Endogenous estrogens (estradiol and estrone) can be hydroxylated at multiple positions (labeled with arrows in Figure 1) by NADPH-dependent cytochrome P450 enzymes. Table I summarizes the presence of hydroxylated and keto metabolites of estradiol and/or estrone in biological samples (e.g. tissues, blood, urine) or the formation of these metabolites during in vitro incubations of estrogens with enzyme prepara-tions from animals or humans. For instance, incubation of [4-14С]estradiol and NADPH with liver microsomes (a crude preparation containing many cytochrome P450 isozymes) from adult male rats resulted in the formation of up to 20 detectable estrogen metabolites (46). Similar formation of multiple estrogen metabolites was observed after incubating [4-14С]-strone and NADPH with female hamster liver or kidney microsomes (47) or after incubating [4-14С]estrone or [4-14С]estradiol and NADPH with female mouse liver micro-somes (48). Since catechol estrogens can undergo metabolic O-methylation (30,49), several additional methoxyestrogen metabolites would have been formed if a methylating enzyme system had been included during the incubations.


Several extrahepatic target tissues or cultured cells from target tissues express estrogen-hydroxylating enzyme activities (31,47,50-82). At least nine different isoforms of cytochrome P450 (some are known to metabolize estradiol) have been detected in the mammary gland of the female rat (83), and several isoforms in rat and human breast are subject to developmental and endocrine regulation (83,84). Recent studies showed the presence of high levels of estradiol 4-hydroxylase activity (mediated by cytochrome P450 1B1) in human uterine myoma (80), human breast cancer tissue (81) and a human breast cancer cell line (70,75,85,86). In addition, a unique estradiol 4-hydroxylase activity is expressed in the male Syrian hamster kidney (76,77,79,82), a target organ for estrogen-nduced carcinogenesis (87,88). Because of these observations and the strong carcinogenic activity of 4-hydroxyestradiol in a hamster kidney tumor model (88,89), this estrogen metabolite is suspected of playing a role in the development of estrogen-ssociated cancers in target organs of animals and possibly humans. In contrast to the carcinogenic potential of 4-hydroxy-stradiol (88,89), 2-hydroxyestradiol has little or no carcinogenic activity (88,89), and 2-methoxyestradiol (an O-ethylated product of 2-hydroxyestradiol) exerts potent cytostatic (90-95) and antiangiogenic effects (93,95) which may inhibit the development of estrogen-induced cancers.


Multiple cytochromes P450 are present and are selectively expressed in certain areas of the brain (96-103), but their role for metabolism of estrogens is largely unknown. Enzymes that catalyze the formation of catechol estrogens have been detected in the central nervous system of rats (31,51-53,55-57,66) and the levels of these enzymes in certain regions of the brain are markedly increased (up to 3.5-fold) during lactation (57). It has been postulated that changes in the formation of catechol estrogens in the central nervous system may play a role in gonadotropin release (54,104), ovulation (105,106), and the function of catecholamines (32-34). In addition to the presence of catechol estrogen-forming enzymes in the brain, studies with a partially purified cytochrome P450 fraction from rat brain revealed high catalytic activity for the 6α-hydroxylation of estradiol, and the partially purified enzyme preparation also catalyzed to a lesser extent the 2-, 4-, 6β-, 15α-nd 16α-ydroxylations of estradiol (97). The functional role of these hydroxylated estrogen metabolites formed in the brain is not known. It is noteworthy that treatment of rats with ethanol induces the levels of certain cytochromes P450 in the brain (100), but the physiological significance of this effect for the metabolism and action of estrogens in the brain is not known.


Possibly because of very low levels of conjugating enzymes in certain extrahepatic target organs, in situ metabolism of estrogens may result in accumulation of significant amounts of unconjugated estrogen metabolites in these target tissues. Previous studies showed that very high levels of catechol estrogens are present in the pituitary, hypothalamus, and cerebral cortex (30). High concentrations of 6α-hydroxyestra-iol, 16α-hydroxyestradiol (estriol) and 2-methoxyestradiol were present in human follicular fluid specimens, and several additional metabolites were also observed (107).


The observation of multiple pathways for estrogen hydroxylation in liver and, in particular, in estrogen target organs raises the important question of why so many estrogen metabolites are formed? As discussed below, estrogen-etabolizing enzymes in liver and extrahepatic target tissues/ cells are under regulatory control by endogenous factors such as sex hormones (83, 108) and by environmental substances such as drugs, pesticides, polycyclic aromatic hydrocarbons, and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) (70,75,108-111). We believe that many of the multiple estrogen metabolites that are formed in liver or estrogen target tissues may have important but unrecognized biological effects that are necessary for some of the actions of estrogens.


2-Hydroxylation

2-Hydroxylation of estradiol or estrone to a catechol is a major metabolic pathway in the liver whereas 4-hydroxylation to a different catechol represents a quantitatively minor path-way (usually < 15% of 2-hydroxylation) in this organ (46,77,108,112). Many different isoforms of cytochrome P450 contribute to the 2-hydroxylation of estradiol in the liver. In rats, hepatic 2-hydroxylation of estradiol is catalyzed by cytochromes P450 1A2, 2B1/2B2, 2C6, 2C11, C-M/F (one or more members of the 2D family) and the 3A family (108,111,113-117). In humans, cytochrome P450 1A2 and the 3A family are major enzymes for hepatic estrogen 2-ydroxylation (112,118-121). It is of considerable interest that there are large interindividual differences in the 2-ydroxylation of estradiol or estrone by human liver samples (112,118,119) and these inter-individual differences may be reflected by person-to-person differences in estrogen action in different individuals.


NADPH-dependent 2-hydroxylation of estradiol and/or estrone has been observed with microsomes prepared from various extrahepatic tissues such as uterus (64,72,73,80), breast (74,81,122), placenta (50,63,80), kidney (76,79,82), brain (31,51,53,55-57,59) and pituitary (53,66). Catechol estrogen formation by microsomal monooxygenase(s) from human term placenta is predominantly 2-hydroxylation and the 4-ydroxylation is a very minor metabolic pathway in this tissue (50,63,80). Based on studies with selective inhibitors and on studies using selectively expressed human aromatase (cytochrome P450 19), the 2-hydroxylation of estradiol in human placenta is likely catalyzed by aromatase (123). In contrast to these observations, the 2-hydroxylation of estradiol by MCF-7 human breast cancer cells treated with TCDD appears to be predominantly catalyzed by cytochrome P450 1A1/1A2 (70,75). Since cytochrome P45o 3A4 (which has high estradiol 2-hydroxylase activity) is present in several extrahepatic tissues (83,124,125), it is believed that this cytochrome P450 isoform may contribute substantially to estradiol 2-hydroxylation in these tissues.


In addition to the NADPH-dependent 2-hydroxylation of estradiol and estrone, the organic hydroperoxide-dependent peroxidatic pathway may also contribute significantly to the 2-nd 4-hydroxylation of these estrogens in liver (126) and estrogen target tissues (63,76,127). This peroxidatic pathway of catechol estrogen formation appears to be catalyzed largely by cytochrome P450 enzymes (40,126). It will be of interest to identify the cytochrome P450 isoforms that have high activity for the peroxidatic formation of catechol estrogens in liver or estrogen target tissues, and to ascertain the physiological significance of this peroxidatic pathway in vivo.


Although estradiol and estrone are extensively 2-ydroxylated in liver and extrahepatic tissues, the concentra-tions of unconjugated 2-hydroxyestradiol and 2-hydroxyestrone metabolites are very low in the systemic circulation (128-130) and in several tissues (131,132), which is probably due to rapid conjugative metabolism (O-methylation, glucuronidation, sulfonation, etc.) followed by urinary excretion


Several important consequences of locally-formed 2-hydro-xyestradiol and 2-hydroxyestrone have been suggested: (i) Both catechol estrogens can bind to the classical estrogen receptor, but with a markedly reduced binding affinity (17,30,133,134), and these metabolites possess much weaker hormonal potency as compared with the parent hormone, estradiol (30,135-138). 2-Hydroxyestrone was reported to partially antagonize the growth-stimulatory effect of estradiol in cultured human MCF-7 breast cancer cells (139,140). This growth-inhibitory effect of 2-hydroxyestrogens (at high concentrations) may be due to the interaction of these com-pounds with the estrogen receptor (139) and/or due to their metabolic redox cycling to generate reactive estrogen quinones and free radicals which are highly cytotoxic (discussed in ref. 141). (ii) By serving as a co-oxidant, 2-hydroxyestradiol strongly stimulates the metabolic cooxidation of arachidonic acid to prostaglandins in the uterus during certain periods of pregnancy (142-145) and thereby may modulate the physiol-ogic effects of arachidonic acid and prostaglandins during pregnancy (146). (iii) 2-Hydroxyestradiol inhibits catechol-O-ethyltransferase-catalyzed O-methylation of catecholamines (32-34), which may exert a modulatory effect on the neurophy-iological/pharmacological effects of catecholamines in the central nervous system (147). In addition, catechol estrogens may also modulate intracellular signal transduction (148). It is possible that these modulatory effects of catechol estrogens in the brain may play a role in the feeling of well-being that results from administration of estrogen to ovariectomized or post-menopausal women. (iv) Administration of 2-hydroxyestradiol alters the secretion of prolactin (149,150) and the secretion of luteinizing hormone/follicle-stimulating hormone (LH/FSH) by the anterior pituitary (135). Consistent with a role of locally-formed catechol estrogens in neuroendocrine regulation, estradiol 2-nd 4-hydroxylase activity was markedly increased in the anterior pituitary and hypothalamus of female rats during lactation (57). Additional studies suggest that 2-hydroxyestradiol and 2-hydroxyestrone modulate the interaction of the neurotransmitter dopamine with its receptor (151,152), and this may be a mechanism responsible for regulating prolactin secretion (149,150) and other neuroendo-crine effects of dopamine. Reports have appeared suggesting the presence of a specific membrane binding site for 2-ydroxyestradiol in the anterior pituitary and endometrial cells, but these observations need to be considered in more detail (153-156). The membrane estrogen binding site in cultured GH3 pituitary tumor cells is suspected of playing a role in causing the rapid release of prolactin (156). (v) 2-Hydroxyestra-iol is present in the ovarian follicular fluid of humans (107) and horses (157). An autocrine/paracrine regulatory role for this estrogen metabolite in follicular function has been proposed (158). (vi) 2-Hydroxyestradiol and 2-hydroxyestrone (like 4-ydroxyestradiol) can undergo metabolic redox cycling to generate free radicals such as superoxide and the chemically-eactive estrogen semiquinone/quinone intermediates (159161) which may damage DNA and other cellular constituents (162-166). Despite their potential for undergoing metabolic redox cycling and generating free radicals, 2-hydroxyestradiol and 2-hydroxyestrone (but not 4-hydroxyestradiol) have little or no tumorigenic activity towards the male Syrian hamster kidney (88,89). Moreover, some studies indicate that treatment of rodents with certain inducers of estradiol 2-hydroxylation may decrease spontaneous tumorigenesis in estrogen-sensitive tissues and this is discussed later.


The lack of carcinogenic activity of 2-hydroxyestradiol and 2-hydroxyestrone (but not 4-hydroxyestradiol) is possibly because the 2-hydroxylated estrogen metabolites, when com-pared with 4-hydroxyestradiol, have a faster rate of metabolism by catechol-O-methyltransferase-catalyzed O-methylation (167,168), a more rapid clearance in vivo (130,169), and possess weaker hormonal potency in estrogen target tissues (30,134,136-140,170). Moreover, 2-methoxyestradiol (a prod-uct of subsequent enzymatic O-methylation of 2-hydroxyestra-iol) is a very potent inhibitor of tumor cell proliferation (9095) and angiogenesis (93,95), which may be an important reason for the lack of carcinogenicity of 2-hydroxyestradiol in vivo.


4-Hydroxylation

Although 2-hydroxylation of estradiol and estrone is the dominant pathway for catechol estrogen formation in liver microsomes, small amounts of 4-hydroxylated estradiol and estrone are also formed (46,77,82,108,111,112). In rat liver microsomes, 4-hydroxylation of estradiol is catalyzed by cytochromes P450 1A2, 2B1/2, and the 3A family (108,111). In human liver microsomes, the cytochrome P450 3A family is believed to play a major role in the 4-hydroxylation of estradiol (112,121).


In contrast to the above observations indicating that 4-ydroxylation of estrogens is a minor pathway for catechol estrogen formation in liver, recent studies showed that 4-ydroxylation of estradiol is a dominant pathway for catechol estrogen formation in several extrahepatic target tissues. For instance, an estradiol 4-hydroxylase activity is expressed in rat pituitary which has little or no activity for 2-hydroxyestradiol formation (66), and an estradiol 4-hydroxylase activity with a low Km (2-5 μM ) was observed in human uterine myometrial and myoma tissues (80). Recent studies indicate that cytochrome P450 1B1 is an important enzyme for the 4-ydroxylation of estradiol in human breast and uterus (80,86). Expression of the human P450 1B1 gene in Saccharomyces cerevisiae produced an enzyme that catalyzed the 4-nd 2-ydroxylation of estradiol with Km values of 0.7-0.8 μM, and turnover numbers of 1.39 and 0.27 nmol product/min/nmol P450, respectively (86). In addition to the NADPH-dependent pathway of catechol estrogen formation, an organic hydroper-oxide-dependent peroxidatic pathway may also contribute significantly to the 4-nd 2-hydroxylation of estradiol and estrone in liver (126) and estrogen target organs (63,76,127). The different pathways for metabolic formation of 4-hydroxy-stradiol by several estrogen target tissues was reviewed earlier by Weisz (38,40).


Selective expression of estradiol 4-hydroxylase activity in target cells does not inactivate the parent estrogen, but may be a mechanism for maintaining strong hormonal activity in these cells or for exerting other unknown biological effects that are not shared with estradiol. As summarized below, several important functions of 4-hydroxyestradiol are already known or have been suggested: (i) 4-Hydroxyestradiol is similar to estradiol in its ability to bind to and activate the classical estrogen receptor (30,133,134,170). Interestingly, the interaction of this estrogen metabolite with the estrogen recep-tor appears to occur with a reduced dissociation rate compared with estradiol (134,171), suggesting that the association of 4-ydroxyestradiol with the estrogen receptor may last longer than that for its parent hormone, estradiol. (ii) 4-Hydroxyestra-iol is hormonally active for stimulating uterine growth when injected into animals (30,170,172), but its uterotropic potency is slightly weaker than that of estradiol (172), possibly due to a faster metabolic clearance than estradiol. It is of considerable interest that addition of 4-hydroxyestradiol to surviving rat or mouse uterine segments in vitro was much more effective than estradiol in stimulating the incorporation of [14С]formate into protein (173). Similarly, another study reported that 4-ydroxyestradiol had stronger activity than estradiol in inducing progesterone receptor formation in the rat pituitary under experimental conditions where 4-hydroxyestradiol and estradiol induced a similar level of nuclear estrogen receptor (134). (iii) Administration of 4-hydroxyestradiol (or 2-hydroxyestra-iol) alters the secretion of luteinizing hormone/follicle-stimu-ating hormone (LH/FSH) by the anterior pituitary (135). (iv) 4-Hydroxyestradiol (like estradiol) supports embryo implantation in mice (105,174,175). An increased selective expression of estradiol 4-hydroxylase activity occurs in the pig blastocyst during the preimplantation period (61,65,67,176) and in the uteri of rabbits and mice during embryo implantation (72,73). These observations suggest a possible role of 4-hydroxyestra-iol during embryo implantation. (v) Like 2-hydroxyestradiol, 4-hydroxyestradiol also serves as a cooxidant and strongly stimulates the metabolic cooxidation of arachidonic acid to prostaglandins in the uterus during certain periods of pregnancy (142-145). This may modulate the physiological effects of prostaglandins during pregnancy (146). (vi) 4-Hydroxyestra-iol inhibits techol-O-methyltransferase-catalyzed O-methyl-tion of catecholamines (32-34), which may exert a modulatory effect on the neuropsychologic/pharmacologic effects of cate-cholamines in the central nervous system (147). (vii) 4-ydroxyestradiol undergoes metabolic redox cycling (159161) to generate free radicals such as superoxide and the chemically-reactive estrogen semiquinone/quinone inter-mediates (shown in Figure 2). These metabolic intermediates may damage DNA and other cellular constituents (162-165), induce cell transformation (166) and initiate tumorigenesis (141,177,178). (viii) 4-Hydroxyestradiol is a strong carcinogen towards the hamster kidney (~100% tumor incidence) under conditions where 2-hydroxyestradiol is not carcinogenic (88,89). The strong carcinogenicity of 4-hydroxyestradiol may be due to its potential genotoxicity (redox cycling plus reactive semiquinone intermediate) as mentioned above and its potent growth stimulatory effect as recently demonstrated with cul-tured kidney proximal tubule cells (179-181). Although direct injection of estrone-3,4-quinone (derived from 4-hydroxy-strone) into the rat mammary gland did not induce the formation of mammary tumors (182), injection of this reactive estrogen metabolite into 12-day-old male B6/C3/F1 mice four times daily for 4 days did induce hepatomas (178). The ability of 4-hydroxyestradiol or estrone-3,4-quinone to cause kidney or liver tumors in certain animal models is noteworthy, and implicates 4-hydroxylated estrogens as carcinogenic meta-bolites.



Figure 2. Metabolic redox cycling of 4-hydroxyestradiol catalyzed by cytochrome P450 enzymes. Although we have only shown the structure of 4-hydroxyestradiol, 2-hydroxyestradiol can also undergo a similar metabolic redox cycling.


Recent studies suggest that high levels of estradiol 4-ydroxylase activity in estrogen target tissues may play important roles in the development of estradiol-induced tumorigenesis (reviewed in refs 141 and 177). A high level of estradiol 4-ydroxylase activity is expressed in the kidney of male Syrian hamsters (76,77,79,82), the uterus of CD-1 mice (64,73), and the pituitary of rats (66), which are all target organs susceptible to estrogen-induced carcinogenesis (11,87,88,183,184). Interestingly, each of the above three target organs contains very high concentrations of endogenous catecholamines (185) which may provide significant inhibition of catechol-O-methyltrans-erase-catalyzed O-methylation of 4-nd 2-hydroxyestradiol in vivo. Moreover, catechol-O-methyltransferase-catalyzed O-ethylation of 4-hydroxyestradiol is inhibited by 2-hydroxyes-radiol, whereas the O-methylation of 2-hydroxyestradiol is not inhibited by 4-hydroxyestradiol (168). Therefore, it is likely that 4-hydroxyestradiol will accumulate in these target organs because of inhibition of its O-methylation and also because of its rapid formation. 4-Hydroxyestradiol can mediate damage to cellular DNA and other macromolecules according to the mechanism depicted in Figure 2. Moreover, the potent mitogenic effects of estradiol and its hormonally active 4-ydroxylated metabolite formed locally in target cells may stimulate the growth of transformed cells, which is believed to be a necessary component for the full development of estrogen-associated cancers (10,17,179-181).


It is of great interest that human uterine myoma expresses estradiol 4-hydroxylase activity to a greater extent than 2-ydroxylase activity, and the former activity was much higher in tumors than in the ‘normal-appearing’ surrounding myome-trium (80). Similarly, elevated estradiol 4-hydroxylase activity has also been observed in human breast cancer tissue compared to normal breast tissue (81). In line with this observation, a previous study reported that 4-hydroxyestradiol appears to be the most abundant estrogen metabolite (4.16 nmol/g tissue) in an extract from a human breast cancer specimen, and several additional metabolites were also observed (186). More studies are needed to confirm the presence of high tissue levels of 4-hydroxyestradiol in human breast cancers. These above interesting results (described in refs 80,81,186), together with studies indicating strong tumorigenic activity for 4-hydroxyes-radiol in animals, are consistent with a possible role of metabolically formed 4-hydroxyestradiol in the genesis of estrogen-associated cancers.


Several recent studies have attempted to characterize cytochrome P450 isoform(s) with high estradiol 4-hydroxylase activity in estrogen target tissues, as well as their distribution and regulation. In microsomes from MCF-7 human breast cancer cells treated with TCDD or from human uterine myoma, estradiol 4-hydroxylation is catalyzed predominantly by a newly identified member of the cytochrome P450 1 family, designated as cytochrome P450 1B1 (80,85,86,187). As indicated above, a recent study showed that human cytochrome P450 1B1 isolated from a yeast expression system catalyzes both 4- and 2-hydroxylation of estradiol with low Km values (,1 µM), but the Vmax for the 4-hydroxylation of estradiol is ~4-fold higher than the Vmax for the 2-hydroxylation of estradiol (86).


In mouse or human tissues examined, the mRNA for cytochrome P450 1B1 is expressed in many different tissues or cells (liver, kidney, brain, placenta, breast, uterus, prostate and lymphocytes), and it is also expressed in various steroidogenic tissues such as adrenal gland and ovary (188-190). Cytochrome P450 1B1 mRNA appears to be one of the major cytochrome P450 mRNAs in uterine endometrium and mammary gland (expressed mainly in stromal cells) (187, 188, 190, 191). In contrast, the mRNA for cytochrome P450 1A1 (which catalyzes the 2-hydroxylation of estradiol to a much greater extent than its 4-hydroxylation) is expressed mainly in epithelial cells of mouse uterine endometrium and mammary gland (188,190). The expression of cytochrome P450 1B1 mRNA in several cultured mammary cell lines is not correlated with the presence of the classical estrogen receptor (191).


The expression of cytochrome P450 1B1 mRNA is regulated by multiple endogenous and exogenous factors. In cultured mouse embryo fibroblasts, the expression of cytochrome P450 1B1 mRNA can be regulated by cyclic AMP, adrenocorticotropic hormone (ACTH), TCDD, and polycyclic aromatic hydrocarbons (187,191). The expression of cytochrome P450 1B1-dependent estradiol 4-hydroxylase activity in MCF-7 human breast cancer cells is stimulated by treatment with TCDD (70,85,86), or indolo[3,2-b]carbazole (86), a dietary-erived Ah-receptor ligand (192). However, the expression of cytochrome P450 1B1 mRNA in treatment with 1 pM to 1 μM of TCDD although the expression of P450 1A1 mRNA was markedly induced (189). It will be of considerable interest to characterize the selective expression and differential regulation of cytochrome P450 1B1, and other cytochrome P450 isoforms with estradiol 2-nd 4-hydroxylase activity in different cell types in human breast. Such studies may help reveal the functional roles of 2-nd 4-hydroxylated estrogens in the physiology and patho-physiology of the human breast.


6 and 7-hydroxylation

The NADPH-dependent 6-hydroxylation of estradiol and estrone by liver from animals and humans has been known for 40 years (193). Recent studies indicate that cytochrome P450 2B1/2B2 in rat liver microsomes is a major enzyme responsible for the 6α-nd 6-4-hydroxylation of estradiol and that cytochrome P450 1A1/1A2 also catalyzes the hepatic 6α-nd 7α-hydroxylation reactions (117). Additional studies showed that treatment of female rats with phenobarbital or dexamethasone markedly stimulated the hepatic 6α-nd 6 b-ydroxylation of estradiol and that treatment of rats with 3-methylcholanthrene stimulated the hepatic 6α-nd 7α-ydroxylation of estradiol (111). It should be noted that cytochromes P450 1A1 and P450 2B1/2 in human liver are very low unless the animals or human subjects are treated with inducers.


The presence of large amounts of 6α-hydroxyestradiol (13.2 ± 0.9 ng/ml, from 11 follicular fluid specimens), in addition to several other estrogen metabolites, was observed in human ovarian follicular fluid (107). 6α-nd 7α-hydroxylated metabolites are major biotransformation products of estrone and estradiol in pig uterus (58,78). Pig endometrial cells effectively catalyze the 6α-nd 7α-hydroxylation of estrone with low Km values ( <1 μM), and these estrogen-metabolizing enzyme activities were inhibited by cytochrome P450 inhibitors such as ketoconazole (78). 6α-Hydroxylation of estradiol by microsomes from MCF-7 human breast cancer cells is markedly stimulated by exposure of MCF-7 cells to TCDD (70,75), and the addition of anti-rat cytochrome P450 1A1/1A2 IgG to the incubation mixture inhibited by ~70% the 6α-hydroxylation as well as the 2-nd 15α-hydroxylation of estradiol (75). In additional studies, a partially purified cytochrome P450 enzyme from rat brain was shown to catalyze the 6α-nd 6β-ydroxylation of estradiol, and this enzyme preparation also catalyzed the formation of 2-hydroxyestradiol, 4-hydroxy-stradiol, 6-ketoestradiol, 15α-hydroxyestradiol and 16α-ydroxyestradiol (97). 6- or 7-hydroxylated metabolites of estrone or estradiol do not have appreciable binding affinities for the classical estrogen receptor, and their uterotropic activity is very weak (194,195).


It is noteworthy that a preliminary study by Takagi et al. with two chemically synthesized analogs of estrogen-6-sulfates (pyridinium 3-methoxyestra-1,3,5(10)-triene-6α-yl and -6β-yl sulfates) suggest that a highly reactive benzylic carbocation intermediate (at the C-6 position) can be generated in aqueous solution (196), a process which is similar to the formation of chemically reactive intermediates from sulfonated benzylic hydroxyl groups on safrole, estragole or 7,12-dimethyl-enz[a]anthracene (197-199). Although 6α-nd 6β-hydroxyl-tions of estrogens are quantitatively significant pathways in rodents and humans, it is not known whether sulfonation (at the C-6 position) of 6α-r 6β-hydroxylated estradiol or estrone metabolites takes place in vivo. If estradiol-r estrone-6-ulfate are formed, they may be highly reactive and genotoxic. Therefore, it will be of interest to determine the relative rate of metabolic formation of 6-sulfonated estrogen metabolites in the body (particularly in estrogen target tissues or cells) and to evaluate the potential genotoxicity or mutagenicity of 6-sulfonated estrogens in these cells.


15α-Hydroxylation

A substantial amount of estrogen 15α-hydroxylase activity is present in certain microorganisms (200), in the human adrenal gland (201,202) and in human fetal liver (203). However, little or no 15α-hydroxylated metabolite was detected during incubations of estradiol with microsomes from non-fetal human liver. In male rats, 15α-hydroxylation is a significant metabolic pathway (46,204) which has been attributed to the male-pecific cytochrome P450 2C13 (115,204). It has been reported that male (but not female) rats metabolize estradiol to 15 a-ydroxyestrone and 15α-hydroxyestradiol that are excreted into the bile in vivo (204). A recent study from our laboratory showed that treatment of immature or adult female rats with 3-methylcholanthrene increased hepatic microsomal estradiol 15α-hydroxylase activity by 5-to 16-fold and cytochrome P450 1A1 was the predominant isoform responsible for the increased 15α-hydroxylation reaction (111).


15α-Hydroxyestrone is present in large amounts in human feces during pregnancy (205). Several earlier studies suggested that the fetal-placental unit is a principal producer of 15α-ydroxylated estrogens (206-209). In the late stages of human pregnancy, urinary 15α-hydroxyestriol (15α,16α-dihydroxyes-radiol; estetrol) is excreted in amounts that exceed those of all other estrogens except 16α-hydroxyestradiol (estriol) and 16α-hydroxyestrone (210-214). Large amounts of estetrol have been reported to originate from fetal liver at the late stages of pregnancy, and high levels of this metabolite are found in the serum of infants shortly after birth (215). Although the physiological role of estetrol is unknown, some studies suggested that the amount of estetrol excreted in the urine by the expectant mother can be used as an indicator of fetal wellbeing (211,213,214,216,217).


On days 12 and 13 of pregnancy in pigs, peri-implantation blastocysts selectively expressed estradiol 15α-hydroxylase activity at 15-o 70-fold higher levels than were expressed during other days of pregnancy (218). The biological signific-ance of this remarkable and short-lived increase of estradiol 15α-hydroxylase activity is not known. Since the 15α-ydroxylation of estradiol results in metabolites that have <1% of estradiol’s binding affinity for the classical estrogen receptor and little or no estrogenic activity (170,219), perhaps the rapid metabolic conversion of estradiol to 15α-hydroxyes-radiol and estetrol protects certain target cells of the embryo and developing infant from exposure to very high concentrations of endogenous estrogenic hormones. Alternatively, these 15α-hydroxylated estrogens may exert other important yet still unrecognized effects. Further research is needed to determine the physiological function(s) of 15α-hydroxyestradiol and estetrol.


16 a-Hydroxylation

In male and female rats, the constitutively expressed hepatic cytochrome P450C-M/F (thought to be a member of the 2D family; see ref. 116) has high activity for the 2-nd 16α-ydroxylation of estradiol and weak activity for 6α-, 6β-nd 15α-hydroxylation of estradiol and for the conversion of estradiol to estrone (114). Cytochrome P450 2C11, a male specific isoform in rats, also catalyzes estrogen 16α-ydroxylation (115) and liver microsomes from male rats catalyze the 16α-hydroxylation of estradiol to a much greater extent than liver microsomes from female rats (46). These results suggest that the male specific cytochrome P450 2C11 may have an important role in the 16α-hydroxylation of estrogens. Studies with prototype liver microsomal enzyme inducers indicate little or no induction of the 16α-hydroxylation of estradiol in female rats treated with phenobarbital, 3-ethylcholanthrene, dexamethasone, isoniazid or clofibrate (111). A preliminary report by Blume et al. suggests that a new member of the cytochrome P450 2D family may have an important role in the 16α-hydroxylation of estrogens in mouse mammary cells (220). In humans, a recent study showed that cytochrome P450 3A4 has strong catalytic activity for estrone 16α-hydroxylation (119).


16α-Hydroxylated estrogen metabolites were found to pos-sess some unique properties: (i) 16α-hydroxyestrone and 16α-hydroxyestradiol, like 4-hydroxyestradiol, retain potent hormonal activity by activating the classical estrogen receptor (221). (ii) A covalent reaction of 16α-hydroxyestrone with the estrogen receptor has been reported (222), and there is a preliminary study suggesting that 16α-hydroxyestrone may activate classical estrogen receptor-mediated oncogene expres-sion and growth stimulation for a prolonged period (223). Mechanistically, a Schiff base is formed from 16α-hydroxyes-rone by reacting with amino groups in proteins (illustrated in Figure 3; ref. 224). In principle, 16α-hydroxyestrone may also react covalently with other amino-containing macromolecules (e.g. DNA), but this has not been described.



Figure 3. . Formation of stable adducts of 16α-hydroxyestrone with amino-containing macromolecules. The reversible reaction between 16α-hydroxyestrone and the amino group results in the formation of a Schiff base, which will further form a stable 16-keto-17ß-amino estrogen adduct via a Heyns rearrangement.


Bradlow, Fishman and their colleagues (reviewed in refs 225-227) have suggested that increased formation of 16α-ydroxyestrogen metabolites may be associated with an increased risk for developing mammary cancer in mice and humans based on the following observations. (i) By using [16α-3H]estradiol and measuring 3H release into body water, these investigators reported that whole-body 16α-hydroxylation of estradiol was ~50% greater in post-meno-pausal patients with breast cancer than in healthy control subjects and enhanced 16α-hydroxylation was also detected in healthy women at high risk for breast cancer (from cancer-rone families) (228,229). (ii) Estrogen 16α-hydroxylation was higher in terminal duct lobular units in cancerous or noncancerous tissues from women with breast cancer compared with breast tissue (reduction mammoplasties) from women without cancer (74). Treatment of cultured human mammary tissue with 7,12-dimethylbenz[a]anthracene (DMBA) coordinately increased the ras protooncogene expression and estradiol 16α-hydroxylation in terminal duct lobular units (230). (iii) 16α-Hydroxyestrone increased unscheduled DNA synthesis and anchorage-independent growth of mouse mammary epithelial cells in culture, which suggests possible genotoxicity for this estrogen metabolite (231). In addition, 16α-hydroxyestrone (but not 2-hydroxyestrone) enhanced the carcinogen-initiated growth stimulation of cultured mouse mammary epithelial cells (231,232). (iv) Animal studies showed that in several different strains of mice with varying incidence of spontaneous mammary tumors, the extent of 16α-hydroxylation of estradiol was positively correlated with their mammary tumor incidence (233). (v) Treatment of C3H/OuJ mice with indole-3-carbinol, a compound derived from vegetables of the Brassica family (such as broccoli and brussel sprouts) and a precursor of a potent Ah-receptor ligand, indolo[3,2-b]carbazole (192), increased the hepatic 2-hydroxylation but not the 16α-ydroxylation of estradiol (234-236). This treatment of mice or rats with indole-3-carbinol was associated with a decreased incidence of spontaneous mammary or endometrial cancer (237,238). (vi) Certain pesticides and polychlorinated biphenyls, which have been postulated to increase the risk of human breast cancer (239,240), were reported to increase the rate of 16α-hydroxylation of estrone and decrease the rate of 2-hydroxylation of estrone (241) in MCF-7 cells, a human breast cancer cell line.


Although the data described above suggest that increased formation of 16α-hydroxylated estrogen metabolites (relative to the formation of 2-hydroxylated estrogen metabolites) may be associated with an elevated risk of breast cancer, it should be noted that these studies have not yet received sufficient confirmation by other investigators using different experimental settings. One concern is the questionable validity of the 3H release assay used in the above studies which determines the release of tritium from 2-3H-r 16α-3H-labeled estrogens to measure 2-nd 16α-hydroxylation in vivo and, in particular, to measure the very low levels of estrogen 2-nd 16α-ydroxylase activity in uninduced mammary explants or cells in vitro (usually <1 pmol/mg microsomal protein/min). It of tritium from [2-3H]estradiol has been noted earlier in the tritium release assay for estradiol 2-hydroxylation (55,56,59). A recent study has suggested that there indeed exists substantial differences in the absolute rate of liver microsomal 2-hydroxylation of estradiol when measured by the tritium release assay vs. a gas chromatography/mass spectrometry (GC/MS)-based product isolation assay (242). In addition, if enolization at the C-16 and C-17 positions is a necessary step for estrogen 16α-hydroxylation as suggested earlier (243), this may also make tritium at the C-16 position labile.


In contrast to the conclusions of Bradlow and his colleagues, a study by Lemon et al. suggested that increased urinary excretion of catechol estrogens but not 16α-hydroxylated estrogens is correlated with an increased risk for non-familial breast cancer (244). More recent studies by Aldercreutz and his colleagues showed that a Finnish population (with high risk for developing breast cancer) had an increased urinary excretion of catechol estrogens relative to 16α-hydroxylated estrogens when compared with an Oriental population at a lower risk (245,246). These investigators suggested that the main risk factor for the Finnish women may be related to high estrogen levels and greater estrogen production than occurred in the Oriental women.


Finally, several additional questions concerning the possible etiological role of 16α-hydroxylated estrogens in hormonal cancer still need to be addressed: (i) 16α-hydroxyestrone and 16α-hydroxyestradiol (estriol) are only very weak carcinogens in the estrogen-induced hamster kidney tumor model under experimental conditions that produced a 100% tumor incidence in animals treated with estradiol or 4-hydroxyestradiol (88,89,180). The question of whether the lack of substantial carcinogenicity for 16α-hydroxyestrone and 16α-hydroxy-stradiol (estriol) observed in this tumor model also occurs in other animal tumor models or humans is not known. (ii) Although pregnant women produce very large amounts of 16α-hydroxyestradiol (estriol) and 16α-hydroxyestrone during normal pregnancy (247,248), full term pregnancy does not increase their breast cancer risk, but actually decreases their risk of breast cancer (249-251). (iii) Although 16α-hydroxyes-rone is capable of interacting covalently with amino groups in proteins (as depicted in Figure 3), there is not sufficient evidence to indicate in vivo genotoxicity or strong carcinogenicity for the 16α-hydroxylated estrogens. In view of these concerns, we conclude that the evidence is not sufficiently strong to support the hypothesis of an etiological role of 16α-ydroxylated estrogens in cancer development.


In summary, we have described the 2-, 4-, 6α-, 6β-, 7α-, 15α-nd 16α-hydroxylation of estradiol or estrone by cytochrome P450 enzymes present in liver and, in particular, in estrogen target tissues or cells. Many other hydroxylated metabolites of estradiol and estrone (not discussed here, but summarized in Table I) have been identified during in vitro incubation of an estrogen with enzyme preparations from animals or humans and/or have been identified in urine, blood, or tissue samples obtained during in vivo studies. It is not known whether these metabolites are formed in extrahepatic target cells and very little is known about their physiological activities. This is an area in need of more research.


As indicated above, many different isoforms of cytochrome P450 are expressed in estrogen target tissues (e.g. breast, uterus and brain), but the role of these cytochrome P450 enzymes for the metabolism and action of estrogens in estrogen target organs has not received adequate attention. We believe that metabolic conversion of estradiol or estrone to multiple hydroxylated metabolites by cytochrome P450 enzymes in target organs is not only for the inactivation of the hormone, but may also diversify the action of estrogens and provide a mechanism that enables estradiol and estrone to exert unique effects in a tissue/cell-specific manner. It will be of considerable interest to examine each of the hydroxylated estrogen metabolites listed in Table I for their estrogenic activity (mediated by the classical estrogen receptor), as well as for other potentially unique biological effects. A recent study showed that 17-piestriol (a quantitatively minor metabolite of estradiol) is a potent stimulator of the promoter of the human TGF-P3 gene transfected into cultured cells whereas estradiol displayed little or no effect (252).


It is noteworthy that several recent studies have suggested a role for estrogens in the prevention of Alzhiemer’s disease in humans (253-256) possibly by enhancing the plasticity and interactions of brain neurons (257). Multiple cytochromes P450 are present in the brain (96-103), and recent studies have shown that alcohol and other inducing agents stimulate the expression of certain cytochromes P450 in the brain (100). These effects of environmental chemicals may modulate estrogen metabolism and action in the central nervous system. It will be of interest to determine whether the reported beneficial effects of estrogen in the central nervous system depend on or are modulated by the metabolism of the hormone by specific isoforms of cytochrome P450 in the brain.


Modulation of estrogen metabolism and action by inducers of estrogen hydroxylation


In earlier studies, we pointed out that steroid hormones are metabolized by the same monooxygenases (cytochrome P450 enzymes) that metabolize drugs and other xenobiotics (258,259). Accordingly, factors that influence the metabolism of xenobiotics also influence the metabolism of steroid hor-mones (258,259). Cytochrome P450-dependent xenobiotic-nd steroid-metabolizing enzymes in liver are regulated by genetic and environmental factors (108-110, 260-267). In studies started over 30 years ago, we found that treatment of rats with phenobarbital or other drugs or with certain halogenated hydrocarbon insecticides increased liver microsomal monooxygenase activity for the metabolism of estradiol, estrone, progesterone, and testosterone (reviewed in refs 109,110,260). The stimulatory effect of liver microsomal enzyme inducers on steroid metabolism was paralleled in vivo by a decreased action of the steroids. Pretreatment of rats with phenobarbital or several other enzyme inducers decreased (i) the uterotropic effects of estradiol, estrone and certain oral contraceptive steroids (268-271), (ii) the anesthetic effects of progesterone and deoxycorticosterone (272,273), and (iii) the growth-promoting effects of testosterone on the seminal vesicles (274,275). In contrast to these results, inhibition of hepatic cytochrome P450 enzymes by treatment of rats with carbon tetrachloride enhanced the action of estradiol on the uterus (276).


Additional studies from our laboratory showed that treatment of adult female rats with phenobarbital stimulated the liver microsomal metabolism of estradiol to 2-, 4-, 6α-, 6β-nd 14a-ydroxyestradiol whereas treatment with 3-methylcholanthrene stimulated the 6α-, 7α-nd 15α-hydroxylation of estradiol (111). Treatment of female rats with dexamethasone stimulated the hepatic microsomal enzyme activities for the 2-, 4-, 6β-, 7α-nd 14a-hydroxylation of estradiol and the formation of several nonpolar unidentified metabolites of estradiol (111). Altogether, at least a dozen estradiol metabolites were formed by liver microsomes from female rats treated with different inducing agents (111).


Studies in humans showed that cigarette smoking (exposure to polycyclic aromatic hydrocarbons and other inducers) or treatment of people with phenobarbital and other anticonvulsant drugs stimulated the metabolism of xenobiotics and steroid hormones (reviewed in refs 110 and 260). Female smokers have enhanced 2-hydroxylation of estradiol (277,278), and lower serum and urinary levels of estradiol and estrone (279281) which helps explain why these individuals have a higher risk for osteoporosis (282-284) and a lower risk for endometrial cancer (282,284,285). Women taking phenobarbital and other anticonvulsants have an increased risk of osteoporosis which was thought to result from enhanced metabolic inactivation of 25-hydroxyvitamin D3 (286-288), but enhanced metabolism of estrogen to inactive metabolites may also play a role in the increased risk of osteoporosis seen in women treated with anticonvulsants. More research is needed to evaluate this possibility. It will also be of interest to determine whether women taking phenobarbital and other anticonvulsant drugs that enhance estrogen metabolism have a decreased risk of breast or endometrial cancer.


Studies by Bradlow and his colleagues showed that chronic administration of indole-3-carbinol stimulates the 2-ydroxylation of estradiol (234-237) and inhibits mammary preneoplasia and spontaneous mammary tumors in female C3H/OuJ mice (234,237). An inhibitory effect of indole-3-arbinol on the formation of spontaneous endometrial cancer was also observed in rats (238). In addition, Bradlow and his colleagues reported a stimulatory effect of indole-3-carbinol on the 2-hydroxylation of estradiol in humans (289,290) and suggested that indole-3-carbinol may be an effective chemopreventive agent for breast cancer in women (226,229). However, the use of indole-3-carbinol and other Ah receptor agonists as chemopreventive agents against estrogen-dependent human cancers through induction of estrogen 2-hydroxylation should receive more careful evaluation because: (i) Some studies suggest that total catechol estrogen production is positively associated with an increased risk of breast cancer in women (244-246). (ii) Indole-3-carbinol, which is converted to a potent Ah-receptor agonist, indolo[3,2-b]carbazole, in the acidic stomach (192), and other Ah-receptor agonists (such as TCDD) have been reported to increase the rates of estrogen 4- and 2-hydroxylation in liver and estrogen target cells (70,75,86,235). For instance, treatment of rats with indole-3-arbinol resulted in increased rates of hepatic 4-nd 2-ydroxylation of estradiol by 4-nd 2-fold, respectively (235), and treatment of cultured MCF-7 breast cancer cells with indolo-3-carbazole or TCDD increased the rates of 4-nd 2-hydroxylation of estradiol by > 10-fold (70,75,86). The stimulatory effect of indole-3-carbinol on estradiol 4-ydroxylation has not been carefully considered in mammary cancer chemoprevention studies in animal models and humans (226,289,290). A chemopreventive strategy for human breast cancer by employing Ah receptor agonists that also enhance the 4-hydroxylation of estrogens to potential carcinogens should receive more attention prior to extensive clinical trials, although indole-3-carbinol treatment may turn out to be an effective modality for inhibition of estrogen-related carcinogenesis and studies in this area should be encouraged.


Recent studies from our laboratory indicated that chronic administration of sodium phenobarbital (0.05% in the drinking water) for 16 months very strongly inhibited the formation of spontaneous mammary tumors in female C3H/OuJ mice. This effect was accompanied by a several-fold increase in the liver microsomal 2-hydroxylation of estradiol with little or no change in estradiol 4-nd 16α-hydroxylation (B.T.Zhu and A.H.Conney, unpublished data). It was observed in this study, however, that sodium phenobarbital administration increased the formation of liver tumors. It will be of interest to do additional dose-response studies to determine whether lower doses of sodium phenobarbital will inhibit mammary carcino-genesis without enhancing liver tumor formation. It is important to note that epidemiology studies have not found an enhanced risk of liver cancer in people treated with phenobarbital and other anticonvulsants.


Increased hepatic estradiol metabolism should result in decreased circulating levels of estradiol as well as decreased hormonal activity (269,291), but selective modulation of estrogen metabolism in target tissues or cells should not be expected to affect the blood and urinary levels of estrogens and their metabolites because of the very low activity of estrogen-metabolizing enzymes in target tissues relative to the liver. Previous studies showed that treatment of animals with TCDD (an environmental chemical with potent monooxygen-ase inducing activity) resulted in a marked antiestrogenic effect (292-294), but the blood level of estradiol was not significantly influenced (294). Additional studies revealed a marked stimulatory effect of TCDD on the 2-, 4-, 6α-nd 15α-hydroxylation of estradiol in cultured human MCF-7 breast cancer cells (70). The stimulatory effect of TCDD on oxidative estradiol metabolism in cultured MCF-7 cells was associated with an inhibitory effect on estradiol-induced formation of transformed foci and plasminogen activator activity (295,296). Other studies indicated that administration of TCDD to mice injected with MCF-7 human breast cancer cells inhibited the growth of these explants (296), presumably by enhancing estrogen metabolism in the breast cancer cell explant and/or in host tissues. These results indicate that Ah receptor agonists such as TCDD can exert multiple effects on estrogen action in target cells due to altered estrogen metabolism. These effects include decreased estrogenic hormonal stimulation due to lower levels of estradiol, increased exposure to 2-hydroxyestradiol or 2-methoxyes-radiol (potentially protective metabolites), and an increased exposure to 4-hydroxyestradiol (a genotoxic/carcinogenic metabolite of estradiol) (70). The net biological effect of TCDD in different target tissues/cells may differ depending on the relative amounts of the different metabolites formed.


It is of considerable interest that although prolonged administration of TCDD to female rats enhanced the incidence of liver tumors, this treatment inhibited the formation of spontaneous tumors in the uterus, mammary gland, and pituitary (297), which are all target tissues for estrogen action. Whether the inhibitory effect of TCDD on spontaneous tumor formation in estrogen target tissues is attributable to its strong stimulatory effect on estrogen metabolic inactivation in these target tissues is worthy of further investigation. Careful dose-response studies with TCDD are needed to determine whether the beneficial or toxic effects predominate at low dose levels of TCDD.


Enzyme-catalyzed interconversion between estrone and estradiol


17β-Hydroxysteroid dehydrogenase (17β-HSD) is a group of intracellular isozymes catalyzing interconversions between estradiol and estrone (298). These enzymes are widely distrib-uted in human tissues, not only in classical steroidogenic tissues, such as placenta (299-301) and ovary (302), but also in a large number of peripheral intracrine tissues (303), including adipose tissue (304), skin and vaginal mucosa (305), endometrium (306), breast and cultured mammary cancer cells (307-313), red blood cells (314), and liver (315). Interconversion between estradiol and the less active hormone, estrone, by 17β-HSD in target cells has long been recognized as an important regulatory mechanism for modulation of estrogen action in these cells (84,298).


Although 17β-HSD activity has been detected in several target tissues, many early studies focused on the functional role of this enzyme in the uterine endometrium. 17β-HSD activity is localized in the secretory glandular epithelium and proliferative endometrium of the uterus (316), and this enzyme activity is increased > 10-fold in the uterus during the secretory phase of the menstrual cycle compared to that during the proliferative phase (317). In humans as well as in all other mammalian species examined, this increase of uterine 17β-SD during the secretory phase is mediated by increased concentrations of progesterone (310,317,318). Previous studies showed that uterine 17β-HSD predominantly catalyzes the conversion of estradiol to estrone using NAD (and to a lesser extent, NADP) as a cofactor, resulting in significant loss of estrogenic hormonal stimulation in the uterus (298,315). It is noteworthy that progesterone is included in many estrogen preparations (e.g. post-menopausal estrogen supplements and birth control pills) since it reduces the risk of uterine endomet-rial cancer in women taking estrogens (18). A major mechanism for the beneficial effects of progesterone in preventing uterine endometrial carcinoma may be attributed to its induction of uterine 17β-HSD, as well as estrogen sulfotransferase (315,317-319).


It is of considerable interest that 17β-HSD in breast tumors predominantly catalyzes the reductive conversion of estrone to estradiol (308,320), suggesting that breast tumors have different isozymes of 17β-HSD than the uterus which catalyzes the oxidative metabolism of estradiol to estrone (298). Moreover, the reductive activity of 17β-HSD is increased in mammary tumors compared with normal breast that does not contain malignant cells (307,309). This finding suggests that breast tumors may provide themselves with a favorable estrogenic environment for growth through increasing the metabolic conversion of estrone to estradiol. The 17β-HSD reductive activity appears to be under multifactorial control. Several growth factors and cytokines (IL-1 b and IL-6) are among the multiple factors capable of stimulating the expression of this enzyme activity in human breast cancer cells (311,313,321323).


At least five different isozymes of human 17β-HSD (types I-V) have been isolated and sequenced (320,323-327). They have different substrate and cofactor specificities, tissue distributions, subcellular localizations and catalyze opposite reactions in vivo. The diversity of the multiple 17β-HSD isozymes suggests that they have different physiological functions in different target cells (328). For instance, reductive activity in human mammary cells is largely associated with the Type I 17β-HSD that uses NADPH as a cofactor (320), whereas the oxidative activity in uterus is associated with Type II 17β-SD that mainly uses NAD as a cofactor (298,320). It should be emphasized that selective regulation of the oxidative and/ or reductive 17β-HSD activity in estrogen target tissues by chemical intervention or dietary modulation may lead to new approaches for the prevention of estrogen-dependent uterine and mammary cancers in humans.


Finally, it is noteworthy that certain isoforms of cytochrome P450 can also catalyze the NADPH-dependent oxidation of estradiol to estrone (117). Hence, factors that influence the levels of cytochrome P450 enzymes may modulate the action of estradiol both by altering its ring hydroxylation and its conversion to estrone.

Aromatization of androgens

Aromatase (estrogen synthetase; cytochrome P450 19) cata-lyzes the aromatization of androgens to estrogens and is the rate-limiting enzyme in the biosynthesis of endogenous estrogens (329,330). In addition to endocrine tissues such as ovary and placenta (329,331,332), aromatase activity is found in several non-endocrine tissues such as brain (333-335), adipose tissue (336,337), liver (338), fibroblasts (339,340) and mammary glandular cells (341-343). In adipose tissue, much higher enzyme activity and mRNA levels for aromatase are present in stromal cells than in other cell types (344,345).


The extent of conversion of plasma 4-androstene-3,17-dione to estrone is positively correlated with increased obesity and increased age, suggesting adipose tissue as a significant site of estrogen biosynthesis and as a major source of estrogen in post-menopausal women and in elderly men (346). Unlike estrogen synthesis in the ovary, estrogen production in adipose tissue is not cyclic, but continuous. Although the physiological role of extragonadal estrogen production by the aromatase pathway is not fully understood, several investigators (347349) have discussed the possible relationship between estrogen biosynthesis in adipose tissue and several disease states, such as endometrial and breast cancer, as well as osteoporosis and chronic amenorrhea in obese women and gynecomastia in obese men.


It is of interest that immunohistochemical staining of human breast tumors for aromatase indicated the presence of aromatase in focal areas of stromal spindle cells but only very low levels were present in tumor epithelial cells, inflammatory cells and normal breast cells (350). Moreover, the aromatase activity is more strongly expressed in adipose stromal cells surrounding mammary carcinoma cells than in stromal cells distal to the tumor (344,345), suggesting enzymatic aromatization as an important source of estrogen for mammary tumors. Several inhibitors of aromatase have been developed and used in patients with estrogen-dependent tumors with varying degrees of success (351-353).


Recent studies have characterized the human aromatase gene, which consists of 10 exons and spans at least 75 kb (354-358). The aromatase gene contains multiple copies of exon 1, and each copy of exon 1 contains its own promoter for transcriptional activation (356-359). Tissue-specific expression is regulated by alternative use of these multiple promoters (356-360). The expression of aromatase in breast and abdom-inal adipose tissue is regulated by a promoter that flanks exon 1b whereas exon 1c (with its own separate promoter) is specific for the expression of aromatase in the ovary (356-359). Molecular and epidemiological analyses of tissue-specific util-ization of multiple exons 1 and their corresponding promoters revealed a switching from the use of adipose-specific exon 1b to exon 1c in adipose tissue adjacent to carcinomas in most breast cancer patients, whereas aromatase mRNA in adipose tissue distal to the tumor of the same patient was normally transcribed from exon 1b as occurred in the breasts of healthy controls (356-359). A switching from exon 1b to exon 1c was more often observed in breast cancer patients with metastasis than in those without metastasis (361). It is speculated that the switching from the adipose-specific exon 1b to exon 1c may modify the strictly-regulated tissue-specific expression of aromatase thereby leading to over-expression of aromatase in adipose stromal cells adjacent to the tumor.


A variety of factors have been shown to regulate the activity and expression of aromatase in cultured adipose stromal cells. Examples of these regulatory substances include glucocorticoids, cAMP analogs, phorbol esters and a variety of growth factors such as epidermal growth factor, fibroblast growth factor, platelet-derived growth factor, transforming growth factors a and p, and tumor necrosis factor (360). It will be of interest to identify factors that govern the switching from the use of adipose-specific exon 1b to the use of ovary-specific exon 1c in transcriptional activation of aromatase synthesis in adipose stromal cells near mammary tumors.


Formation and hydrolysis of estrogen sulfates and glucuronides

It has been well documented that metabolic conjugation of estradiol and estrone to glucuronides and sulfates by conjugat-ing enzymes in liver and target cells decreases their hormonal activity by facilitating their excretion (25,35,36,39). However, it has also been known for years that dehydroepiandrosterone (DHEA)-sulfate, a sulfonated steroid with a very slow meta-bolic clearance rate (<1% of DHEA; ref. 362), is an important intermediate precursor for the biosynthesis of several steroidal hormones in many different types of cells. The concept that conjugated estrogens (formed mainly in liver) may also be important precursors for metabolic formation of hormonally active estrogens in target cells has received ever-increasing attention and support (363-370).


In post-menopausal women, the plasma level of estradiol is very low, but its concentration in uterine endometrium and breast cancer tissues can be 10-o 50-fold higher than the circulating estradiol concentration (371,372). The concentra-tions of estradiol in the uterine endometrium and breast cancer tissue obtained from post-menopausal women are similar to those observed in premenopausal women (365,369,371,372). It is believed that estrogen biosynthesis in extragonadal tissues and, in particular, in estrogen target organs may contribute significantly to the unexpectedly high tissue levels of unmetab-olized estrogens in post-menopausal women. Since estrone-3-ulfate is a major circulating estrogen metabolite with an overall low clearance rate and a long half-life in humans (~9 h) (373,374), estrone-3-sulfate is thought to be an important precursor of active estrogen peripherally in post-menopausal women (37,367,370,375).


Sulfonated estrogens themselves have almost no estrogen receptor binding affinity and their estrogenic activity results from the release of unconjugated estrogens by enzymatic deconjugation in liver as well as in target tissues or cells (e.g. uterine endometrial cells and breast cells) (25, 363-369,376). A recent study in rodents from our laboratory showed that estrone-3-sulfatase in the uterus (a classical target organ for estrogen action) may play a more important role than hepatic sulfatase in mediating the hormonal action of estrone-3-sulfate in this target organ (377). Since estrone-3-sulfate is transported into human breast cancer cells (378) which contain high levels of estrogen sulfatase activity (25,367,368,378), enzymatic hydrolysis of estrone-3-sulfate (together with estrogen synthesis by aromatase in breast tumors) is likely an important source of parent estrogens in breast cancer cells and these pathways of estrogen synthesis are believed to contribute very significantly to the higher concentration of estradiol in breast tumor tissues than in the systemic circulation (37,369,370,372). Several studies have indicated that estrogen formation in human breast tumors in post-menopausal women through desulfonation of estrone-3-sulfate may be more important than the aromatase pathway of estrogen formation by the breast (367,370,375,378). It is noteworthy that a previous study showed that treatment of rats with estrone-3-sulfate signific-antly stimulated the growth of A-methyl-A-nitrosourea-induced mammary tumors (379), which is in support of the concept that estrogen sulfates can be effective precursors of estrogenic hormones required for mammary tumor growth in vivo.


High levels of estrogen β-glucuronidase, which converts estradiol and estrone glucuronides to their corresponding parent hormones, are expressed in hamster kidney (a target organ for carcinogenesis), and this enzyme activity is stimulated in the kidney following chronic administration of estradiol (380). The high estrogen β-glucuronidase activity in hamster kidney (approximately twice the activity in liver) and the increase of this enzyme activity during chronic estrogen treatment may provide additional amounts of parent estradiol and estrone as well as reactive estrogen metabolites to this target organ and thus may facilitate the development of estradiol-induced kidney tumors (380).


It is of considerable interest that a large amount of β-lucuronidase was found in mammary glandular tissue of female Sprague-Dawley rats at puberty (381), and treatment of these animals with dietary D-glucaro-1,4-lactone (a potent natural inhibitor of β-glucuronidase) strongly inhibited the growth of mammary glandular cells in vivo (381). Additional studies also showed that treatment of female Sprague-Dawley rats with β-glucuronidase inhibitors suppressed mammary tumor promotion (an estrogen-dependent process) after initi-ation with chemical carcinogens (382-384).


It should be noted that estrogen sulfates or glucuronides can also be hydroxylated at multiple positions by NADPH-ependent cytochrome P450 enzymes in liver as well as in extrahepatic target cells (117,385,386). The hydroxylation of conjugated estradiol and estrone in the liver and/or in target cells and further deconjugation (enzymatic hydrolysis) of these metabolites may be an additional source of hydroxylated estrogen metabolites in target tissues or cells.


More research is needed to understand the factors that regulate the expression of estrogen-specific sulfotransferase, UDP-glucuronosyltransferase, sulfatase and β-glucuronidase in estrogen target tissues, as well as to identify genetic and environmental factors that modulate these enzyme activities. The formation of estradiol, estrone and their hydroxylated metabolites by metabolic hydrolysis of estrogen sulfates and glucuronides is a potential target for chemotherapy or chemo-revention of estrogen-associated tumors (e.g. breast and uterine cancers). Developing potent and selective inhibitors of estrogen sulfatase(s) is already an actively-pursued strategy for chemoprevention against mammary tumorigenesis (387389). We believe that developing potent β-glucuronidase inhibitors with low toxicity and high bioavailability in humans will also be of great potential value for the prevention and treatment of estrogen-associated cancers.



Formation and catabolism of methoxyestrogens and estrogen fatty acid esters


Metabolic hydroxylation and/or conjugation of estrogens make(s) them more water soluble and, at the same time, often markedly reduces or completely eliminates their classical estrogenic activity. However, enzymatic O-methylation or esterification with fatty acids leads to the formation of much more lipophilic estrogen metabolites. These metabolites have very long half-lives (390,391) and do not bind to the classical estrogen receptor (392). Recent studies suggest that at least some of these lipophilic estrogen metabolites may have unique biological activities (not associated with the classical estrogen receptor) or they can be further converted to biologically active molecules at or near estrogen receptors.


Methoxyestrogen

The O-methylation of catechol estrogens is catalyzed by catechol-O-methyltransferase (COMT), an enzyme that also catalyzes the O-methylation of physiologically important cate-cholamines and many other catechols (49,393-395). Catechol-methyltransferase activity is present in large amounts in liver and kidney, and it also exists in significant amounts in red blood cells, uterine endometrium, the mammary gland and many other tissues (396-399). In all tissues examined thus far, catechol-O-methyltransferase activity is found almost exclus-ively in the cytosol, but some activity is also found in a membrane-bound form (394,400). Because of the rapid enzymatic O-methylation of catechol estrogens, 2-methoxyestrone was previously shown to be one of the most abundant estrogen metabolites in human plasma and urine (30,401,402). In pregnant women, the mean plasma concentration of unconjugated 2-methoxyestrone is ~4000 pg/ml (403). Interestingly, 2-methoxyestrone and 2-methoxyestradiol have higher binding affinities for sex hormone-binding globulin than estradiol and 2-hydroxyestradiol (404), and the high binding affinities of these two methoxyestrogens may contribute to their high plasma levels.


The monomethylated estrogen metabolites (structures shown in Figure 4) have little or no estrogen receptor binding affinities (<1%) when compared to estradiol (30,405), and they lack estrogenic effects on the uterus [e.g. lack of effect on uterinewt (221) or peroxidase activity (406)]. Previous studies on the chemical reactivity and potential genotoxicity of catechol estrogens (159-165,178) have led to the suggestion that enzymatic O-methylation was primarily a detoxification path-way for these catechol intermediates. However, there are several studies indicating that 2-methoxyestradiol exerts unique biological effects that are not associated with estradiol, 2-ydroxyestradiol or other methoxy derivatives of estradiol (90-94,407-411). For instance, treatment of rats with 2-ethoxyestradiol decreases cholesterol and triglyceride levels in the blood and this effect is not associated with activation of the classical estrogen receptor (407-409). In addition, 2-ethoxyestradiol inhibits the growth of certain human breast cancer cell lines in vitro and in vivo, and it is a potential inhibitor of estrogen-dependent carcinogenesis. These studies are discussed below.



Figure 4. Structures of monomethylated catechol estradiol metabolites. Estradiol is hydroxylated to form catechol estradiol metabolites (2- and 4-hydroxyestradiol). They can be further O-methylated by catechol-O-methyltransferase (COMT) to form monomethoxy estradiol metabolites. 2-Methoxyestradiol and 4-methoxyestradiol are the major monomethylated isomers formed from their respective catechol precursors. Although the structures for estrone and its catechol metabolites are not shown, the same metabolic pathways as shown also occur for estrone or its metabolites (a 17-keto group in place of the 17β-hydroxy group of estradiol). Since catechol-O-methyltransferase catalyzes the metabolism of a wide spectrum of endogenous and exogenous catechols, the O-methylation of catechol estradiol or estrone metabolites is thus subjected to regulation by these catechols (hydroxylated flavonoids, catecholamines, etc.). Abbreviations used: 2-OH-E2, 2-hydroxyestradiol; 4-OH-E2, 4-hydroxyestradiol; 2-MeO-E2, 2-methoxyestradiol; 2-OH-3-MeO-E2, 2-hydroxy-3-methoxyestradiol; 4-MeO-E2, 4-methoxyestradiol; 4-OH-3-MeO-E2, 4-hydroxy-3-methoxyestradiol.


2-Methoxyestradiol inhibits the proliferation of several can-cer cell lines invitro (90,91,93,95,412), and human breast cancer cell lines (estrogen receptor positive or negative) were particularly sensitive to a cytotoxic effect of 2-methoxyestra-diol (412). Additional studies indicated that 2-methoxyestradiol disrupted microtubule function (92,94) and was a potent inhibitor of angiogenesis (93,95). Administration of 2-ethoxyestradiol inhibited the growth of transplanted meth-A sarcoma and B16 melanoma in C3H mice (93), and oral administration of 2-methoxyestradiol also inhibited the growth of a human breast carcinoma cell line (estrogen receptor negative) in immunodeficient mice (95). It is noteworthy that 2-methoxyestradiol is among the most potent endogenous inhibitors of angiogenesis known, and its antiangiogenic effect as tested in vitro is highly specific and is not shared by several closely related structural analogs (93). The effects of 2-ethoxyestradiol to disrupt microtubule function, to inhibit angiogenesis and to inhibit the proliferation of breast cancer cells in vitro and in vivo suggest that factors enhancing the 2-ydroxylation of estradiol and the subsequent formation of 2-ethoxyestradiol may inhibit estrogen-induced breast cancer. It is possible that the stimulatory effect of indole-3-carbinol and phenobarbital on the 2-hydroxylation of estradiol and the subsequent formation of an increased amount of 2-methoxyestradiol can explain the inhibitory effects of indole-3-carbinol and phenobarbital on spontaneous breast carcinogenesis in C3H/OuJ mice.


Recent studies have shown that chronic administration of quercetin (a substrate and inhibitor of catechol-O-methyltrans-erase) to male Syrian hamsters significantly increased the severity of estradiol-induced kidney tumors (413), which was correlated with inhibition of enzymatic O-methylation of catechol estrogens during quercetin administration (414). It is of considerable interest that male Syrian hamster kidney,CD-1 mouse uterus and rat pituitary (tissues that develop estradiol-induced tumors) have very high levels of endogenous catecholamines (up to 50-fold higher than in several non-target tissues in the same animals or in other strains or species) (185). High concentrations of catecholamines in target tissues may inhibit catechol-O-methyltransferase-catalyzed O-methyl-tion of catechol estrogens (185), which will result in decreased formation of 2-methoxyestradiol and increased tissue concen-trations of 2-nd 4-hydroxylated estrogens. This effect may enhance tumor formation in these target tissues. Earlier studies in animals showed that stressed mice had an increased incidence of spontaneous breast cancer (415). Interestingly, epidemiological studies have also suggested that sustained stress (associated with an increase in endogenous catecholamine levels) is a risk factor for human breast cancer (416-419), which is consistent with decreased formation of 2-methoxyestradiol together with the possible accumulation of 4-hydroxylated catechol estrogens in mammary tissue.


Since human and rodent breast cancers express catechol estrogen-forming enzyme activity (81,122,127,420,421) and elevated levels of catechol-O-methyltransferase (393-395), it is likely that significant amounts of 2-methoxyestradiol are formed in breast cancer cells. Recent studies showed that there are marked person-to-person variations in catechol-O-ethyltransferase activity in red blood cells (422-424) and liver samples (425), and the distribution of catechol-O-methyl-ransferase activity in the American population appears to follow a polymorphic bimodal pattern (423,425,426). It will be of great interest to compare the risk of breast cancer in people with high or low catechol-O-methyltransferase activity. It is expected that individuals with significantly lower COMT activity may have a higher risk of estrogen-associated breast cancer due to decreased formation of antitumorigenic 2-ethoxyestradiol and retarded inactivation of catechol estrogen intermediates (particularly 4-hydroxyestradiol which is hormonally active and potentially genotoxic).


The mechanisms by which 2-methoxyestradiol exerts anti-angiogenic activity, anti-tubulin activity, and antiproliferative effects on tumor cells lacking the classical estrogen receptor are not known. The high degree of selectivity of 2-methoxyes-radiol’s antiangiogenic activity (many closely related analogs have poor activity) suggests the possibility of a specific yet unidentified receptor for this action of 2-methoxyestradiol. It will be of great interest to determine the molecular mechanisms of action of 2-methoxyestradiol, to determine whether endogenously formed 2-methoxyestradiol has antitumor effects in vivo, and to identify endogenous as well as exogenous factors that regulate the formation and metabolic disposition of 2-ethoxyestradiol in target cells. Such studies will help our understanding of the physiological roles of 2-methoxyestradiol, which is a nonpolar estradiol metabolite formed in substantial amounts in many estrogen target cells. These studies may also provide novel mechanism(s) and strategies for inhibiting the formation and growth of hormonal cancers.


Finally, it should be noted that although 4-methoxyestradiol showed little activity in test systems where 2-methoxyestradiol is highly active, whether this methoxyestrogen metabolite has other unique biological functions other than serving as a substrate for 4-hydroxyestradiol formation by demethylation (380) is not known at present.


Estrogen fatty acid esters

In the presence of fatty acid acyl-co-enzyme A, estradiol (at the C-17 position only) can be converted to very lipophilic estrogen fatty acid esters by enzymes present in liver as well as in estrogen target organs such as the uterus, breast and placenta (427-432). In MCF-7 human breast cancer cells, enzyme activity for the formation of estrogen fatty acid esters is highest in the microsomal fraction, but some enzyme activity is also present in the nuclear and mitochondrial fractions (431). Both saturated and unsaturated fatty acid acyl-co-enzyme A derivatives can serve as donors for the esterification of estradiol. It should be noted that although most previous studies have only studied fatty acid ester formation with estradiol as substrate, esterification of 16α-hydroxyestradiol (estriol) has also been shown to occur at the 16α-hydroxy group (432,434).


Estradiol fatty acid esters themselves have little or no estrogen receptor binding affinity (392), and their hormonal activity results from the slow release of the parent estrogen by metabolic cleavage of the fatty acid esters by esterases (430-435). Because of the high lipophilicity of the fatty acid esters of estradiol, these endogenous metabolites are present at very low concentrations in the blood (436-438), but at relatively high concentrations in fatty tissues (432,438). These esters have long half-lives and are extremely potent mammalian estrogens because they serve as a reservoir and direct precursor for estradiol formation (390,391,432,439). Because the mammary gland contains large amounts of fat which can serve as a storage site for fatty acid esters of estradiol and possibly also for fatty acid esters of hydroxylated estrogen metabolites, we suggest that estrogen-fatty acid esters may be particularly important estrogens for the breast (‘mammary selective estrogen metabolites’).


It will be of considerable interest to advance our knowledge on the formation, storage and, in particular, subsequent hydro-lysis of estrogen-fatty acid esters by enzymes selectively expressed in the mammary gland which may be an important pathway for supplying hormonally-active and/or potentially genotoxic estrogens for this target organ. It will also be of interest to study the factors (endogenous or exogenous) that regulate the synthesis and cleavage of estrogen fatty acid esters since factors influencing the formation and hydrolysis of estrogen fatty acid esters in liver and in extrahepatic target cells may profoundly affect the intensity and duration of estrogen action in the body. Recent studies in our laboratory indicated that treatment of rats with clofibrate (a peroxisome-roliferator) stimulates by several-fold the liver microsomal formation of estradiol-fatty acid esters (440). The significance of this effect on estrogen’s hormonal action remains to be elucidated.


Estrogen metabolites formed in target cells may have unique functions not associated with classical estrogen receptors: a general hypothesis


Estradiol activates the classical estrogen receptor to exert its major effects and so do some estrogen metabolites that still retain potent estrogenic activity (e.g. 4-r 16α-hydroxy-stradiol, and 4-r 16α-hydroxyestrone). Although these four estrogen metabolites as well as some other estrogen metabolites have uterotropic activity in rodents, they are generally less potent than estradiol.


Some estrogen metabolites elicit unique physiological responses with higher potency and activity than the parent hormone estradiol, and some of these unique activities as well as some activities of the parent estrogen cannot be blocked by estrogen receptor antagonists. These results suggest that certain effects of estradiol and its metabolites are not associated with activation of the classical estrogen receptor It is becoming increasingly better recognized that certain estrogen metabolites formed in target tissues or cells may serve as local modulators or mediators of specific biochemical events taking place in these target cells. We have indicated earlier in this review that (i) catechol estrogens may modulate the function of dopamine and other catecholamines by inhibiting their enzymatic O-ethylation and by modulating their interaction with specific receptors; (ii) 4-hydroxyestradiol is carcinogenic and this compound may damage cellular macromolecules such as DNA, proteins and lipids because of metabolic redox cycling to generate reactive intermediates and free radicals; (iii) 2-ethoxyestradiol inhibits angiogenesis, the growth of cultured cancer cells and the growth of human mammary cancer cells in immunodeficient mice; and (iv) 17-epiestriol stimulates the promoter of the human TGF-β3 gene transfected into cultured cells whereas estradiol has little or no effect (252). Studies with 17-epiestriol suggest that the interaction of this estradiol metabolite with the classical estrogen receptor is different from the interaction of estradiol with the receptor, and this interaction of 17-epiestriol with the estrogen receptor elicits a new response not shared by estradiol (252). We believe that future studies will identify unique effects of additional estrogen metabolites.


We hypothesize that several active metabolites of estradiol may exert unique effects through interaction with their own specific intracellular receptors or effectors that are refractory to the parent estrogen. Although the functions and structures of these hypothetical intracellular receptors are unknown at present, there are several candidates worthy of consideration. (i) Previous studies showed the presence of multiple variants of the classical estrogen receptor as a result of aberrant splicing of its mRNA (441-444). It will be of interest to determine whether certain active estrogen metabolites may be endogenous ligands for some of these variants. (ii) The nuclear type II estrogen binding site has a low apparent binding affinity for estradiol (445,446) and functionally mediates a growth-nhibitory effect in estrogen-sensitive cells (447). Although previous studies reported methyl -4-hydroxyphenyllactate as an endogenous ligand for the type II estrogen binding site (448), it is not known whether physiologically active estrogen meta-bolites are also ligands for the type II binding site. (iii) A novel estrogen receptor expressed in rat prostate and ovary was recently described (41). Estradiol is bound to this receptor with high affinity, and estrone and 16α-hydroxyestradiol were only moderately effective for inhibiting estradiol binding (ligand competition experiments). The binding of other estrogen metabolites to this receptor either as agonists or antagonists have not been determined. It is likely that additional novel isoforms of the classical estrogen receptor may also exist (42-45). (iv) An estrogen binding site in the plasma membrane has been reported (153-156) and this binding site is suspected of playing a role in the rapid release of prolactin in cultured GH3 pituitary tumor cells (156). Although an earlier study suggested that 2-hydroxyestradiol can bind to a membrane receptor site (153), the binding affinity of many other estrogen metabolites towards this membrane receptor is not known. (v) Several DNA sequences that encode proteins with a high degree of structural similarity to members of the steroid receptor superfamily have been identified in animals and humans (449), but their functions and endogenous ligands are not known. These proteins have been called ‘orphan receptors’. It will be of considerable interest to determine whether estrogen metabolites with unique functions are endo-genous ligands for some of the orphan receptors. (vi) In addition to the candidate receptors mentioned above, other unidentified receptors or specific intracellular effectors (such as transcriptional factors or enzymes) may also be targets for active estrogen metabolites. Whatever the structural identity of intracellular receptors or effectors for active estrogen metabolites, we believe that the interaction of active estrogen metabolites with these specific receptors or effectors in target cells may lead to the expression of new responses or alterations of on-going biochemical processes. This is an area in need of more research.


There are a few examples of biologically active metabolites of certain hormones that have their own receptors. These examples may help our understanding of the hypothesis that certain unique effects of the parent estrogen are mediated by locally-formed metabolites that activate their own specific receptors or effectors in target cells. The pituitary takes up thyroxine (T4) from plasma and converts it locally into triiodothyronine (T3), which is the active receptor-binding form of thyroid hormone (450). In addition, there is enhanced conversion of T4 to T3 in the cerebral cortex and cerebellum of animals with hypothyroidism (451). The prostate converts testosterone to dihydrotestosterone locally and it is the latter compound that binds to a nuclear receptor with a much higher affinity than testosterone (452). The kidney and gut epithelium (two major target tissues for vitamin D action) selectively express 1α-hydroxylase activity, which converts 25-hydroxyvi-amin D3 (a poorly active prehormone that circulates in blood) to the active hormone, 1α,25-dihydroxyvitamin D3, in target sites for hormone action (453). These examples suggest that metabolism of hormones in target cells may be a general mechanism for markedly increasing the original hormone’s activity and cellular specificity. Moreover, the enhanced enzymatic conversion of T4 to T3 in the pituitary of rats with hypothyroidism also exemplifies a metabolic control mechan-ism by which the body maintains homeostasis by regulating the metabolism of certain hormones in target cells (451). The enzymatic conversion of androgen to estrogen by aromatase is a mechanism that modifies and diversifies the original hormone’s action. For instance, the hypothalamic region of the brain in immature rodents converts androgens to estrogens locally by aromatase as a requisite step for male differentiation (334,335). The in situ conversion of testosterone to estradiol by aromatase markedly alters the action of testosterone and provides a source of estrogen in these aromatase-containing cells. In this case, metabolically formed estradiol interacts with its own intracellular receptor which is not responsive to its androgen precursor.


By analogy with the above examples, active estrogen meta-bolites do not need to be present in the systemic circulation in significant quantities to exert their biological effects. The parent hormone estradiol may exert unique functions through metabolic formation of active estrogen metabolite(s) catalyzed by specific metabolizing enzymes that are selectively expressed in estrogen target cells. This is a mechanism that may explain how estradiol can exert diverse, highly selective effects in different target cells.


Concluding remarks and future perspectives


Estradiol is synthesized from testosterone by a cytochrome P450 aromatase present in the ovary and placenta and incertain estrogen target tissues such as brain, prostate, uterus and mammary gland. Once formed, estradiol can be metabolized to multiple hydroxylated products by enzymes of the cytochrome P450 family, and 2-nd 4-hydroxylated catechol estrogen metabolites can be O-methylated by catechol-O-methyltrans-erase. Estradiol and its hydroxylated metabolites can also be esterified to fatty acid esters or conjugated by glucuronidation or sulfonation which in turn may undergo de-esterification by esterase or deconjugation by β-glucuronidase or sulfatase to release active estrogens either in the liver or directly in target cells. This complexity of estrogen synthesis and metabolism (illustrated in Figure 5) provides many potential sites for the regulation of estrogen action.



Figure 5. Complexities of estrogen metabolism. Abbreviations used: ST (sulfotransferase), GT (glucuronosyltransferase), EAT (estrogen acyltransferase; for fatty acid ester formation), 17β-HSD (17β- hydroxysteroid dehydrogenase), COMT (catechol-O-methyltransferase) and P450 (cytochrome P450).


Earlier studies demonstrated marked person-to-person differences in the metabolism of xenobiotics by cytochrome P450 enzymes (reviewed in 260,266,267,454), and these differences are caused by genetic (261-267) and environmental (109,110,260) factors such as drug administration, tobacco smoking, alcohol ingestion, pesticide exposure and dietary composition. Recent studies in humans revealed a large number of mutations in cytochrome P450 genes that impair xenobiotic metabolism (261,262,264-267,455,456), as well as the duplication of cytochrome P450 genes that results in super-fast metabolism of certain xenobiotics (263). Since steroid hormones are metabolized by many of the same cytochrome P450 enzymes that metabolize xenobiotics, changes in xenobiotic metabolism should also be reflected by an altered metabolism of steroid hormones. There is considerable variability in the levels of cytochrome P450 3A4 and 1α2 (cytochromes P450 prominent in catalyzing 2-hydroxylation of estradiol) in different human liver samples (up to 20-fold), and these differences provide an explanation for large interindividual variability in the hydroxylation of estradiol at C-2 and at other positions (112,118). Modulation of oxidative xenobiotic metabolism by cigarette smoking or by changing the ratio of dietary protein to carbohydrate in humans (110,260) also modulates the 2-ydroxylation of estradiol (277,457). Marked person-to-person variations have also been observed for the enzyme systems that catalyze conjugative estrogen metabolism (such as sulfonation; see refs 458,459) and for the O-methylation of catechol estrogens (see refs 422-426). Studies that identify genetic and environmental factors influencing estrogen meta-bolism at or near estrogen receptors in target cells may be of considerable importance since these factors could profoundly modify the biological effects of estrogens in complex manners depending on the pathways of metabolism that are affected and the biological activities of the metabolites that are formed. Such effects need not be associated with an altered profile of estrogen metabolites in the blood or urine.


Despite the widespread belief that increased exposure to estrogens contributes to the development of human mammary cancer, previous studies have only demonstrated a weak association of breast cancer with high estrogen intake or with a high circulating or urinary level of estrogen in women (251) although an elevated exposure to estrogens or high plasma level of estrogens is associated with a substantial increase in the risk for endometrial cancer (12-14,17,18). We believe that measurement of unmetabolized estrogens in plasma may be too crude a parameter to use successfully for relating exposure of target tissues to estrogens with the risk of cancer develop-ment. Several concepts should be stressed which may help our understanding of the underlying complexity of estrogen action: (i) The levels of unmetabolized estradiol and estrone in target tissues or in target cells within these tissues could be very different from estrogen blood levels. Previous studies reported that high concentrations of both albumin and estradiol are present in human mammary tumor cells (460,461). The pres-ence of high levels of intracellular albumin (which has high binding affinity for estrogens) may significantly increase the intracellular levels of estrogens in mammary tumors. Moreover, an active transport of estrogen in addition to passive diffusion may also play a role in regulating the intracellular concentrations of estrogen (462). (ii) The hormonal activity of estrogens can be influenced by the activity of estrogen-metabolizing enzymes at or near estrogen receptors. In addition, metabolic formation of estrogen by aromatization of androgens and by deconjugation of estrogen conjugates in mammary tumor cells, as well as in the surrounding normal glandular and adipose cells may contribute to a high estrogen concentration in tumors. (iii) The carcinogenic activity of estrogens may be mediated or modulated by the formation of hormonally inactive or active metabolite(s) by specific estrogen-metabolizing enzymes in target cells. (iv) It is possible that unique biological effects of estrogen metabolites are mediated by specific receptors in target cells. In such cases, the formation of specific metabolites and the amounts of specific receptors in target cells may both be critical determinants for estrogen action.


It is of great interest that 2-methoxyestradiol was recently identified as an estradiol metabolite with strong activity as an inhibitor of cultured breast cancer cell proliferation in vitro and the growth of human mammary cancer cells in vivo. Further studies are needed to determine if physiologically formed 2-hydroxyestradiol and 2-methoxyestradiol have an inhibitory effect on estrogen-induced carcinogenesis.


In the present review, we suggest that certain estrogen metabolites may function as chemical mediators or as second-ary hormones with unique functions. We point out that estrogen metabolism via multiple pathways can occur in target tissues at or near estrogen receptors. Estrogen metabolism in target tissues not only alters the intensity of estrogen action but may also alter its profile of physiological effects. Many factors modulate the levels of estrogen-metabolizing enzymes in liver and in target tissues, and the biological effects of an estrogen will depend on the profile of multiple metabolites formed and the biological activities of each of the metabolites. We believe that research on factors that influence the synthesis and metabolism of estradiol and its metabolites by diverse pathways at or near receptors in target cells is an underexplored area in need of more attention. Research in this area may lead to an enhanced understanding of estrogen action and to the discovery of pharmacologically active agents that are useful for the prevention and treatment of estrogen-induced cancers.




References


Kumar,V., Green,S., Stack,G., Berry,M., Jin,J.R. and Chambon,P. (1987) Functional domains of the human estrogen receptor. Cell, 51, 941-951. 

Evans,R.M. (1988) The steroid and thyroid hormone receptor superfamily. Science, 240, 889-895. 

Weisz,A., Cicatiello,L., Persicot,E., Scalona,M. and Bresciani,F. (1990) Estrogen stimulates transcription of c-jun protooncogene. Mol. Endocrinol., 4, 1041-1050. 

Parker,M.G. (ed.) (1991) Nuclear Hormone Receptors. Academic Press, London. Martin,L., Finn,C.A. and Trinder,G. (1973) Hypertrophy and hyperplasia in the mouse uterus after estrogen treatment: an autoradiographic study. J. Endocrinol., 56, 133-141. 

Quarmby,V.E. and Korach,K.S. (1984) Influence of 17β-estradiol on the pattern of cell division in the uterus. Endocrinology, 114, 694-702. 

Silver,M. (1953) A quantitative analysis of the role of oestrogen in mammary development in the rat. J. Endocrinol., 10, 17-34. 

Banerjee,M.R. (1976) Responses of mammary cells to hormones. Int. Rev. Cytol. , 4, 1-97. 

Holland,M.B. and Roy,D. (1995) Estrogen-induced cell proliferation and differentiation in the mammary gland of the female Noble rat. 

Carcinogenesis, 16, 1955-1961. 

Nandi,S., Guzma,R.C. and Yang,J. (1995) Hormones and mammary carcinogenesis in mice, rats, and humans: a unifying hypothesis. Proc. Natl Acad. Sci. USA, 92, 3650-3657. 

Li,J.J. and Li,S.A. (1996) The effects of hormones on tumor induction. In Arcos,J.G., Argus,M.F. and Woo,Y.T. (eds) Chemical Induction of Cancer. Birkhauser, Boston, pp. 396-449. 

Ziel,H.K. and Finkle,W.D. (1975) Increased risk of endometrial carcinoma among users of conjugated estrogens. N. Engl. J. Med., 293, 1167-1170. 

Mack,T.M., Pike,M.C., Henderson,B.E., Pfeffer,R.I., Gerkins,V.R., Arthur,M. and Brown,S.E. (1976) Estrogens and endometrial cancer in a retirement community. N. Engl. J. Med., 294, 1262-1267. 

McDonald,T.W., Annegers,J.F., O’Fallow,W.M., Dockerty,M.B., Malkasian, G.D.,Jr and Karland,L.T. (1977) Exogenous estrogens and endometrial carcinoma: case control and incidence study. Am. J. Obstet. Gynecol., 127, 572-580. 

Jick,H., Walker,A.M., Watkins,R.N., D’ewart,D.C., Hunter,J.R., Danford,A., Madsen,S., Dinan,B.J. and Rothman,K.J. (1980) Replacement estrogens and breast cancer. Am. J. Epidemiol., 112, 586-594. 

Pike,M.C., Spicer,D.V., Dahmoush,L. and Press,M.F. (1993) Estrogens, progestogens, normal breast cell proliferation, and breast cancer risk. Epidemiol. Rev., 15, 17-35. 

Feigelson,H.S. and Henderson,B.E. (1996) Estrogens and breast cancer. Carcinogenesis, 17, 2279-2284. 

Grady,D. and Ernster,V.L. (1996) Endometrial cancer. In Schottenfeld,D. and Fraumeni,J.F.,Jr (eds) Cancer Epidemiology and Prevention, 2nd edn. Oxford University Press, Oxford, pp. 1058-1089. 

Martucci,C.P. and Fishman,J. (1993) P450 enzymes of estrogen metabolism. Pharmacol. Ther., 57, 237-257. 

Musey,P.I., Kright,K., Preedy,J.R.K. and Collins,D.C. (1997) Formation and metabolism of steroid conjugates: effect of conjugation on excretion and tissue distribution. In Hobkirk,R. (ed.) Steroid Biochemistry, Vol. II. CRC Press, Boca Raton, FL, pp. 81-132. 

Ritter,J.K., Sheen,Y.Y. and Owens,I.S. (1990) Cloning and expression of human liver UDP-glucuronosyltransferase in COS-1 cells. 3,4-catechol estrogens and estriol as primary substrates. J. Biol. Chem., 265, 7900-7906. 

Zhu,B.T., Suchar,L.S., Huang,M.-T. and Conney,A.H. (1996) Similarities and differences in the glucuronidation of estradiol and estrone by UDP-lucuronosyltransferase in liver microsomes from male and female rats. Biochem. Pharmacol., 51, 1195-1202. 

Brooks,S.C. and Horn,L. (1971) Hepatic sulfonation of estrogen metabolites. Biochim. Biophys. Acta, 231, 233-241. 

Payne,A.H. and Singer,S.S. (1979) The role of steroid sulfatase and sulfotransferase enzymes in the metabolism of C21 and C19 steroids. In Hobkirk,R. (ed.) Steroid Biochemistry, vol I. CRC Press, Boca Raton, FL, pp. 111-145. 

Hobbirk,R. (1985) Steroid sulfotransferases and steroid sulfate sulfatases: Characteristics and biological roles. Can. J. Biochem. Cell Biol., 63, 1127-1144. 

Watanabe,K., Takanashi,K. and Yoshizawa,I. (1988) Determination of estradiol-17-sulfate in human urine by a direct radioimmunoassay: urinary levels throughout the menstrual cycle. Steroids, 52, 123-136. 

Hernandez,J.S., Watson,R.W.G., Wood,T.C. and Weinshilboum,R.M. (1992) Sulfation of estrone and 17β-estradiol in human liver. Catalysis by thermostable phenol sulfotransferase and by dehydroepiandrosterone sulfotransferase. Drug Metab. Dispos., 20, 413-422. 

Axelrod,J. and Tomchick,R. (1958) Enzymatic G-methylation of epinephrine and other catechols. J. Biol. Chem., 233, 702-705. 

Guldberg,H.C. and Marsden,C.A. (1975) Catechol-G-methyltransferase: pharmacological aspects and physiological role. Pharmacol. Rev., 27, 135-206. 

Ball,P. and Knuppen,R. (1980) Catecholoestrogens (2-nd 4-ydroxyoestrogens): chemistry, biogenesis, metabolism, occurrence and physiological significance. Acta Endocrinol., 232, 1-127. 

Fishman,J. and Norton,B. (1975) Catechol estrogen formation in the central nervous system of the rat. Endocrinology, 96, 1054-059. 

Ball,P, Knuppen,R., Haupt,M. and Breuer,H. (1972) Interactions between estrogens and catechol amines. III. Studies on the methylation of catechol estrogens, catechol amines and other catechols by the catechol G-ethyltransferase of human liver. J. Clin. Endocrinol. Metab., 34, 736-746. 

Breuer,H. and Koster,G. (1974) Interaction between estrogens and neurotransmitters at the hypophysial-hypothalamic level. J. Steroid Biochem., 5, 961-967. 

Ghraf,R. and Hiemke,C.H. (1983) Interaction of catechol estrogens with catecholamine synthesis and metabolism. In Merriam,G.R. and Lipsett,M.B. (eds) Catechol Estrogens. Raven Press, New York, pp. 177-187. 

Gurpide,E. (1978) Enzymatic modulation of hormonal action at the target tissue. J. Toxicol. Environ. Hlth, 4, 249-268. 

Vermeulen,A. (1986) Human mammary cancer as a site of sex steroid metabolism. Cancer Surveys, 5, 587-595. 

Santen,R.J. (1986) Determinants of tissue oestradiol levels in human breast cancer. Cancer Surveys, 5, 598-616. 

Weisz,J. (1991) Metabolism of estrogens by target cells: Diversification and amplification of hormone action and the catechol estrogen hypothesis. In Hochberg R.B. and Naftolin, F. (eds), New Biology of Steroid Hormones. Raven Press, New York, pp. 101-112 

Roy,A.K. (1992) Regulation of steroid hormone action in target cells by specific hormone-inactivating enzymes. Proc. Soc. Exp. Biol. Med., 199, 265-272. 

Weisz,J. (1994) Biogenesis of catecholestrogens: metabolic activation of estrogens by Phase I enzymes. Polycyclic Aromatic Comp. , 6, 241-251. 

Kuiper,G.G.J.M., Enmark,E., Pelto-Huikko,M., Nilsson,S. and Gustafsson,J.-Â. (1996) Cloning of a novel estrogen receptor expressed in rat prostate and ovary. Proc. Natl Acad. Sci. USA, 93, 5925-5930. 

Karas,R.H., Baur,W.E., van Eickles,M. and Mendelsohn,M.E. (1995) Human vascular smooth muscle cells express an estrogen receptor isoform. FEBSLett., 377, 103-108. 

Hirata,S., Koh,T., Yamada-Mouri,N. and Kato,J. (1996) The novel untranslated first exon ‘exon ON' of the rat estrogen receptor gene. Biochem. Biophys. Res. Commun., 225, 849-854. 

Mosselman,S., Polman,J. and Dijkema,R. (1996) ER beta: identification and characterization of a novel human estrogen receptor. FEBS Lett. , 392, 49-53. 45. Inoue,S., Hoshino,S., Miyoshi,H., Akishita,M., Hosoi,T., Orimo,H. and Ouchi,Y. (1996) Identification of a novel isoform of estrogen receptor, a potential inhibitor of estrogen action, in vascular smooth muscle cells. 

Biochem. Biophys. Res. Commun., 219, 766-772. 

Suchar,L.A., Chang,R.L., Rosen,R.T, Lech,J. and Conney,A.H. (1995) High performance liquid chromatography separation of hydroxylated estradiol metabolites: formation of estradiol metabolites by liver microsomes from male and female rats. J. Pharmacol. Exp. Ther., 272, 197-206. 

Haaf,H., Metzler,M. and Li,J.J. (1992) Metabolism of [4-14С]estrone in hamster and rat hepatic and renal microsomes: species-, sex-nd age-pecific differences. J. SteroidBiochem. Mol. Biol., 42, 389-397. 

Zhu,B.T., Lech,J., Rosen,R.T. and Conney,A.H. (1997) Effect of dietary 2(3)-tert-butyl-4-hydroxyanisole on the metabolism and action ofestradiol and estrone in female CD-1 mice. Cancer Res., 57, 2419-2427. 

Ball,P., Knuppen,R., Haupt,M. and Breuer,H. (1972) Interactions between estrogens and catechol amines. III. Studies on the methylation of catechol estrogens, catechol amines and other catechols by the catechol-O-ethyltransferase of human liver. J. Clin. Endocrinol., 34, 736-746. 

Fishman,J. and Dixon,D. (1967) 2-Hydroxylation of estradiol by human placental microsomes. Biochemistry, 6, 1683-1687. 

Paul,S.M., Axelrod,J. and Diliberto,E.J. (1977) Catechol estrogen-forming enzyme of brain: demonstration of a cytochrome P450 monooxygenase. Endocrinology, 101, 1604-1610. 

Ball,P., Haupt,M. and Knuppen,R. (1978) Comparative studies on the metabolism of oestriol in the brain, the pituitary and the liver of the rat. 

Acta Endocrinol., 87, 1-11. 

Ball,P. and Knuppen,R. (1978) Formation of 2-nd 4-hydroxyestrogens by brain, pituitary, and liver of the human fetus. J. Clin. Endocrinol. Metab., 47, 732-737. 

Fishman,J., Norton,B.I. and Krey,L. (1980) 2-Hydroxylation of estrogens in the brain participates in the initiation of the preovulatory LH surge in the rat. Biochem. Biophys. Res. Commun., 93, 471-477. 

Hersey,R.M., Gunsalus,P, Lloyd,T. and Weisz,J. (1981) Catechol estrogen formation by brain tissue: a comparison of the release of tritium from [2-3H]2-hydroxyestradiol formation from [6,7-3H]estradiol by rabbit hypothalami in vitro. Endocrinology, 109, 1902-1911. 

Hersey,R.M., Williams,K.I.H. and Weisz,J. (1981) Catechol estrogen formation by brain tissue: characterization of a direct product isolation assay for estrogen-2-nd 4-hydroxyestradiol formation by rabbit hypothalamus. Endocrinology, 109, 1912-1920. 

Hersey,R.M. and Weisz,J. (1983) A product isolation assay for estrogen-/4-hydroxylase activity. InMerriam,G.R. and Lipsett,M.B. (eds) Catechol Estrogens. Raven Press, New York, pp. 37-48. 

Maschler,I., Ball,P., Bayerkohler,G., Gaues,J. and Knuppen,R. (1983) Identification of 6α-nd 7α-hydroxyestrone as major metabolites of estrone and estradiol in porcine uterus. Steroids, 41, 597-607. 

Jellinck,P.H., Hahn,E.F., Norton,B.I. and Fishman,J. (1985) Catechol estrogen formation and metabolism in brain tissue: comparison of tritium release from different portions of ring A of the steroids. Endocrinology, 115, 1850-1856. 

Li,S.A., Klicka,J.K. and Li,J.J. (1985) Estrogen 2-nd 4-hydroxylase activity, catechol estrogen formation, and implications for estrogen carcinogenesis in the hamster kidney. Cancer Res., 45, 181-185. 

Mondschein,J.S., Hersey,R.M., Dey,S.K., Davis,S.K. and Weisz,J. (1985) Catecholestrogen formation by the pig blastocyst during the preimplantation period: biochemical characterization of estrogen-2/4-ydroxylase and correlation with aromatase activity. Endocrinology, 117, 2339-2346. 

Hammond,J.M., Hersey,R.M., Walega,M.A. and Weisz,J. (1986) Catecholestrogen production by porcine ovarian cells. Endocrinology, 118, 2292-2299. 

Bui,Q. and Weisz,J. (1988) Identification of microsomal, organic hydroperoxide-dependent catechol estrogen formation: Comparison with NADPH-dependent mechanism. Pharmacology, 36, 356-364. 

Bunyagidj,C. and McLachlan,J.A. (1988) Catechol estrogen formation in mouse uterus. J. Steroid Biochem., 31, 795-801. 

Chakraborty,C., Davis,D.L. and Dey,S.K. (1988) Characteristics of estrogen-2/4-hydroxylase in pig blastocysts. Inhibition by steroidal and non-steroidal agents. J. Steroid Biochem., 31, 231-235. 

Bui,Q.D. and Weisz,J. (1988) Monooxygenase mediating catecholestrogen formation by rat anterior pituitary is an estrogen-4-hydroxylase. Endocrinology, 124, 1085-1087. 

Chakraborty,C., Dey,S.K. and Davis,D.L. (1989) Pattern and tissue distribution of catechol estrogen forming activity by pig conceptuses during the peri-implantation period. J. Anim. Sci., 67, 991-998. 

Telang,N.T., Bradlow,H.L., Kurihara,H. and Osborne,M.P. (1989) In vitro biotransformation of estradiol by explant cultures of murine mammary tissues. Breast Cancer Res. Treat. , 13, 173-181. 

Brueggemeier,R.W., Katlic,N.E, Palmer,C.W., Jr and Stevens,J.M. (1989) Catechol estrogen formation in MCF-7 cell culture and effects of bromoestrogen inhibitors. Mol. Cell. Endocrinol., 64, 161-167. 

Spink,D.C., Lincoln,D.W., Dickerman,H.W. and Gierthy,J.F. (1990) 2,3,7,8-Tetrachlorodibenzo-p-dioxin causes an extensive alteration of 17β-stradiol metabolism in MCF-7 breast tumor cells. Proc. Natl Acad. Sci. USA, 87, 6917-6921. 

Brueggemeier,R.W., Tseng,K., Katlic,N.E., Beleh,M.A. and Lin,Y.C. Estrogen metabolism in primary kidney cell cultures from Syrian hamsters. J. Steroid Biochem., 36, 325-331. 

Chakaborty,C., Huet-Hudson,Y.M. and Dey,S.K. (1990) Catechol-estrogen synthesis and metabolism in the rabbit uterus during the periimplantation period. J. Steroid Biocehm., 35, 39-46. 

Paria,B.C., Chakraborty,C. and Dey,S.K. (1990) Catechol estrogen formation in the mouse uterus and its role in implantation. Mol. Cell. Endocrinol., 69, 25-32. 

Telang,N.T., Axelrod,D.M., Wong,G.Y., Bradlow,H.L. and Osborne,M.P. Biotransformation of estradiol by explant culture of human mammary tissue. Steroids, 56, 37-43. 

Spink,D.C., Eugster,H.P., Lincoln,D.W., Schuetz,J.D., Schuetz,E.G., Johnson,J.H., Kaminsky,L.S. and Gierthy,J.F. (1992) 17β-Estradiol hydroxylation catalyzed by human cytochrome P450 1A1: a comparison of the activities induced by 2,3,7,8-tetrachlorodibenzo-p-dioxin in MCF-7 cells with those from heterologous expression of the cDNA. Arch. Biochem. Biophys., 293, 342-248. 

Weisz,J., Bui,Q.D., Roy,D. and Liehr,J.G. (1993) Elevated4-hydroxylation of estradiol by hamster kidney microsomes: a potential pathway of metabolic activation of estrogens. Endocrinology, 131, 655-661. 

Zhu,B.T., Roy,D. and Liehr,J.G. (1993) The carcinogenic activity of ethinyl estrogens is determined by both their hormonal characteristics and their conversion to catechol metabolites. Endocrinology, 132, 577-583. 

Adamski,J., Hohls,E. and Jungblut,PW. (1994) Characterization of estrone hydroxylase activities in porcine endometrial cells. Exp. Clin. Endocrinol. , 102, 388-393. 

Zhu,B.T., Bui,Q.D., Weisz,J. and Liehr,J.G. (1994) Conversion of estrone to 2-nd 4-hydroxyestrone by hamster kidney and liver microsomes: implications for the mechanism of estrogen-induced carcinogenesis. Endocrinology, 135, 1772-1779. 

Liehr,J.G., Ricci,M.J., Jefcoate,C.R., Hannigan,E.V., Hokanson,J.A. and Zhu,B.T. (1995) 4-Hydroxylation of estradiol by human uterine myometrium and myoma microsomes: Implications for the mechanism of uterine tumorigenesis. Proc. Natl Acad. Sci. USA, 92, 9220-9224. 

Liehr,J.G. and Ricci,M.J. (1996) 4-Hydroxylation of estrogens as marker of human mammary tumors. Proc. Natl Acad. Sci. USA, 93, 3294-3296. 

Hammond,D.K., Zhu,B.T., Wang,M.Y., Ricci,M.J. and Liehr,J.G. (1997) Cytochrome P450 metabolism of estradiol in hamster liver and kidney. Toxicol. Appl. Pharmacol., 145, 54-60. 

Hellmold,H., Lamb,J.G., Wyss,A., Gustafsson,J.-Â. and Warner,M. (1995) Developmental and endocrine regulation of P-450 isoforms in rat breast. Mol. Pharmacol., 48, 630-638. 

Smith,G., Harrison,D.J., East,N., Rae,F., Wolf,H. and Wolf,C.R. (1993) Regulation of cytochrome P450 gene expression in human and breast tumor xenografts. Br. J. Cancer, 68, 57-63. 

Spink,D.C., Hayes,C.L., Young,N.R., Christou,M., Sutter,T.R., Jefcoate,C.R. and Gierthy,J.F. (1994) The effects of 2,3,7,8-tetra-hlorodibenzo-p-dioxin on estrogen metabolism in MCF-7 breast cancer cells: Evidence for induction of a novel 17β-estradiol 4-hydroxylase. J. Steroid Biochem. Mol. Biol., 51, 251-258. 

Hayes,C.L., Spink,D.C., Spink,B.C., Cao,J.Q., Walker,N.J. and Sutter,T.R. (1996) 17β-Estradiol hydroxylation catalyzed by human cytochrome P450 1B1. Proc. Natl Acad. Sci. USA, 93, 9776-9781. 

Kirkman,H. (1959) Estrogen-induced tumors of the kidney in Syrian hamster. III. Growth characteristics in the Syrian hamster. NCI Monogr., 1, 1-57. Li,J.J. and Li,S.A. (1987) Estrogen carcinogenesis in Syrian hamster tissues: role of metabolism. Fed. Proc., 46, 1858-1863. 

Liehr,J.G., Fang,W.F., Sirbasku,D.A. and Ulubelen,A.A. (1986) Carcinogenicity of catechol estrogens in Syrian hamsters. J. Steroid. Biochem., 24, 353-356. 

Seegers,J.C., Aveling,M.-L., van Aswegen,C.H., Cross,M., Koch,F. and Joubert,W.S. (1989) The cytotoxic effects of estradiol-17β, catecholestradiols and methoxyestradiols on dividing MCF-7 and Hela cells. J. Steroid Biochem., 32, 797-809. 

Lottering,M.L., Haag,M. and Seegers,J.C. (1992) Effects of 17β-estradiol metabolites on cell cycle events in MCF-7 cells. Cancer Res., 52, 5926-5932. 

D’Amato,R.J., Lin,C.M., Flynn,E., Folkman,J. and Hamel,E. (1994) 2-ethoxyestradiol, an endogenous mammalian metabolite, inhibits tubulin polymerization by interacting at the colchicine site. Proc. Natl Acad. Sci. USA, 91, 3964-3968. 

Fotsis,T., Zhang,Y., Pepper,M.S., Adlercreutz,H., Montesano,R., 

Nawroth,P.P. and Schwelgerer,L. (1994) The endogenous oestrogen metabolite 2-methoxyoestradiol inhibits angiogenesis and suppresses tumor growth. Nature, 368, 237-239. 

Hamel,E., Lin,C.M., Flynn,E. and D’Amato,R.J. (1996) Interaction of 2-ethoxyestradiol, an endogenous mammalian metabolite, with unpolymerized tubulin and tubulin polymers. Biochemistry, 35, 1304-1310. 

Klauber,N., Parangi,S., Flynn,E., Hamel,E. and D’Amato,R.J. (1997) Inhibition of angiogenesis and breast cancer in mice by the microtubule inhibitors 2-methoxyestradiol and taxol. Cancer Res., 57, 81-6. 

Kapitulnik,J, Gelboin,H.V., Guengerich,F.P. and Jacobowitz,D.M. (1987) Immunohistochemical localization of cytochrome P-450 in rat brain. Neuroscience, 20, 829-33. 

Sugita, O., Miyairi, S., Sassa, S. and Kappas, A. (1987) Partial purification of cytochrome P-450 from rat brain and demonstration of estradiol hydroxylation. Biochem. Biophys. Res. Commun., 147, 1245-250. 

Warner,M., Kohler,C., Hansson,T. and Gustafsson,J.-Â. (1988) Regional distribution of cytochrome P450 in the rat brain: spectral quantitation and contribution of P450b, e and P450c, d. J. Neurochem., 50, 1057-065. 

Yamada,K., Harada,N., Tamaru,M. and Takagi,Y. (1993) Effects of changes in gonadal hormones on the amount of aromatase RNA in mouse brain diencephalon. Biochem. Biophys. Res. Commun., 195, 462-468. 

Warner,M. and Gustafsson,J-Â. (1994) Effect of ethanol on cytochrome P450 in the rat brain. Proc. Natl Acad. Sci. USA, 91, 1019-023. 

Stromstedt,M., Warner,M. and Gustafsson,J.-Â. (1994) Cytochrome P450s of the 4A subfamily in the brain. J. Neurobiochem., 63, 671-76. 

Ravindranath,V and Boyd,M.R. (1995) Xenobiotic metabolism in brain. Drug Metab. Rev., 27, 419-48. 

Wyss,A., Gustafsson,J.-Â. and Warner,M. (1995) Cytochrome P450s of the 2D subfamily in the rat brain. Mol. Pharmacol., 47, 1148-155. 

Ball,P, Emons,G., Klingebiel,T., Griihn,K.-M. and Knuppen,R. (1981) Effects of catecholestrogens on luteinizing hormone levels in long term ovariectomized adult rats. Endocrinology, 109, 1037-039. 

Ball,P, Franks,S., Naftolin,F. and Ruf,K.B. (1979) Influence of 2-nd 4-ydroxyestrogens (catecholoestrogens) on the PMGS-induced premature ovulation in the immature rat. Acta Endocrinol., 25 (Suppl.), 102. 

Emons,G. and Ball,P. (1982) Induction of ovulation in immature female rats by a single injection of 4-hydroxyoestradiol-dibenzoate. Acta Endocrinol., 100, 105-08. 

Dehennin,L., Blacker,C., Reiffsteck,A, and Scholler,R. (1984) Estrogen 2-, 4-, 6-, or 16-hydroxylation by human follicles shown by gas chromatography-mass spectrometry associated with stable isotope dilution. J. Steroid Biochem., 20, 465-71. 

Dannan,G.A., Porubek,D.J., Nelson,S.D., Waxman,D.J. and Guengerich,F.P (1986) 17β-Estradiol 2-nd 4-hydroxylation catalyzed by rat hepatic cytochrome P-450: roles of individual forms, inductive effects, developmental patterns, and alterations by gonadectomy and hormone replacement. Endocrinology, 118, 1952-960. 

Conney,A.H., Levin,W., Jacobson,M. and Kuntzman,R. (1973) Effects of drugs and environmental chemicals on steroid metabolism. Clin. Pharmacol. Ther., 14, 727-41. 

Conney,A.H. (1986) The induction of microsomal cytochrome P-450 enzymes: The first Bernard B.Brodie lecture at Pennsylvania State University. Life Sci., 39, 2493-518. 

Suchar,L.A., Chang,R.L., Thomas,PE., Rosen,R.T., Lech,J. and Conney,A.H. (1996) Effect of phenobarbital, dexamethasone and 3-ethylcholanthrene administration on the metabolism of 17β-estradiol by liver microsomes from female Long-vans rats. Endocrinology, 137, 663-76. 

Kerlan,V., Dreano,Y., Bercovici,J.P, Beaune,PH., Floch,H.H. and Berthou,F. (1992) Nature of cytochrome P450 involved in the 2-/ 4-hydroxylations of estradiol in human liver microsomes. Biochem. Pharmacol., 44, 1745-756. 

Ryan,D.E., Iida,S., Wood,A.W., Thomas,P.E., Lieber,C.S. and Levin,W. (1984) Characterization of three highly purified cytochromes P-450 from hepatic microsomes of adult male rats. J. Biol. Chem., 259, 1239-250. 

Sugita,O., Sassa,S., Miyairi,S., Fishman,J., Kubota,I., Noguchi,T. and Kappas,A. (1988) Cytochrome P-450C-/F, a new constitutive form of microsomal cytochrome P450 in male and female rat liver with estrogen 2-nd 16α-hydroxylase activity. Biochemistry, 27, 678-86. 

Cheng,K.C. and Schenkman,J.B. (1984) Metabolism of progesterone and estradiol by microsomes and purified cytochrome P-450 RLM3 and RLM5. Drug Metab. Dispos., 12, 222-34. 

Ishida,N., Tawaragi,Y., Inuzuka,C., Sugita,O., Kubota,I., Nakazoto,H., Noguchi,T. and Sassa,S. (1988) Four species of cDNAs for cytochrome P450 isozymes immunorelated to P450-M/F encode for members of P450 IID subfamily, increasing the number of members within the subfamily. Biochem. Biophys. Res. Commun., 156, 681-88. 

Watanabe,K., Takanashi,K., Imaoka,S., Funae,Y, Kawano,S., Inoue,K., Kamataki,T., Takagi,H. and Yoshizawa,I. (1991) Comparison of cytochrome P-450 species which catalyze the hydroxylations of the aromatic ring of estradiol and estradiol 17-sulfate. J. Steroid Biochem. Biol., 38, 737-43. 

Ball,S.E., Forrester,L.M., Wolf,C.R. and Back,D.J. (1990) Differences in the cytochrome P-450 isoenzymes involved in the 2-hydroxylation of oestradiol and 17α-ethinyloestradiol. Biochem. J., 261, 221-26. 

Shou,M., Korzekwa,K.R., Brooks,E.N., Krausz,K.W., Gonzalez,FJ. and Gelboin,H.V. (1997) Role of human hepatic cytochrome P450 1A2 and 3A4 in the metabolic activation of estrone. Carcinogenesis, 18, 207-14. 

Guengerich,F.P. (1988) Oxidation of 17α-ethynylestradiol by human liver cytochrome P-450. Mol. Pharmacol., 33, 500-508. 

Aoyama,T., Korzekwa,K., Gillette,!, Gelboin,H.V. and Gonzalez,F.J. (1990) Estradiol metabolism by complementary deoxyribonucleic acid-expressed human cytochrome P450s. Endocrinology, 126, 3101-106. 

Hoffman,A.R., Paul,S.M. and Axelrod,J. (1979) Catecholestrogen synthesis and metabolism by human breast tumors in vitro. Cancer Res., 39, 4584-4587. 

Osawa,Y., Higashiyama,T., Shimizu,Y. and Yarborough,C. (1993) Multiple functions of aromatase and the active site structure: aromatase is the placental estrogen 2-hydroxylase. J. Steroid Biochem. Mol. Biol. , 44, 469-480. 

de Waziers,I., Cugnenc,PH., Yang,C.S., Leroux,J.P. and Beaune,P.H. (1990) Cytochrome P450 isoenzymes, epoxide hydrolase and glutathione transferases in rat and human hepatic and extrahepatic tissues. J. Pharmacol. Exp. Ther., 253, 387-94. 

Jayyosi,Z., Cooper,K.O. and Thomas,PE. (1992) Brain cytochrome P450 and testosterone metabolism by rat brain subcellular fractions: presence of cytochrome P450 3A immunoreactive protein in rat brain mitochondria. Arch. Biochem. Biophys., 298, 265-70. 

Bui,Q.D., Weisz,J. and Wrighton,S.A. (1990) Hepatic catecholestrogen synthases: Differential effect of sex, inducers of cytochromes P-450 and antibody to glucocorticoid inducible cytochrome P-450 on NADPH-ependent estrogen-2-hydroxylase and on organic hydroperoxide-ependent estrogen-2/4-hydroxylase activity of rat hepatic microsomes. J. Steroid Biochem., 37, 285-93. 

Levin,M., Weisz,J., Bui,Q.D. and Santen,R.J. (1987) Peroxidatic catecholestrogen production by human breast cancer tissue in vitro. J. Steroid Biochem., 28, 513-20. 

Ball,P, Emons,G., Haupt,O., Hoppen,H.-O. and Knuppen,R. (1978) Radioimmunoassay of 2-hydroxy-estrone. Steroids, 31, 249-58. 

Emons,G., Ball,P. and Knuppen,R. (1983) Radioimmunoassays of catecholestrogens. In Merriam,G.R. and Lipsett,M.B. (eds) Catecholestrogens. Raven Press, New York, pp. 71-1. 

Emons,G., Merriam,R., Pfeiffer,D., Loriaux,D.L., Ball,P and Knuppen,R. (1987) Metabolism of exogenous 4-nd 2-hydroxyestradiol in human male. J. Steroid Biochem., 28, 499-04. 

Fishman,J. and Martucci,C. (1979) Absence of measurable catechol estrogens in the rat brain-vidence for rapid turnover. J. Clin. Endocrinol. Metab., 49, 940-42. 

Zhu,B.T. and Liehr,J.G. (1996) Inhibition of catechol O-methyltransferase-atalyzed O-methylation of 2-nd 4-hydroxyestradiol by quercetin. 

J. Biol. Chem., 271, 1357-363. 

van Aswegen,C.H., Purdy,R.H. and Wittliff,J.L. (1989) Binding of 2-ydroxyestradiol and 4-hydroxyestrdiol to estrogen receptor human breast cancers. J. Steroid Biochem., 32, 485-92. 

MacLusky,N.J., Barnea,E.R., Clark,C.R. and Naftolin,F. (1983) Catechol estrogens and estrogen receptors. In Merriam,G.R. and Lipsett,M.B. (eds) Catechol Estrogens. Raven Press, New York, pp. 151-65. 

Martucci,C. and Fishman,J. (1979) Impact of continuously administered catechol estrogens on uterine growth and LH secretion. Endocrinology, 105, 1288-292. 

Fishman,J. (1981) Biological action of catecholestrogens. J. Endocr., 85, 59P-5P. 

Schutze,N., Vollmer,G., Tiemann,I., Geiger,M. and Knuppen,R. (1993) Catecholestrogens are MCF-7 cell estrogen agonists. J. Steroid Biochem. Mol. Biol., 46, 781-89. 

Schutze,N., Vollmer,G. and Knuppen,R. (1994) Catecholestrogens are agonists of estrogen receptor-dependent gene expression in MCF-7 cells. J. SteroidBiochem. Mol. Biol., 48, 453-461. 

Schneider,J., Huh,M.M., Bradlow,H.L. and Fishman,J. (1984) Antiestrogen action of 2-hydroxyestrone on MCF-7 human breast cancer cells. J. Biol. Chem., 259, 4840-4845. 

Vandewalle,B. and Lefebvre,J. (1989) Opposite effects of estrogen and catecholestrogen on hormone-sensitive breast cancer cell growth and differentiation. Mol. Cell. Endocrinol., 61, 239-246. 

Yager,J.D. and Liehr,J.G. (1996) Molecular mechanisms of estrogen carcinogenesis. Ann. Rev. Pharmacol. Toxicol., 36, 203-232. 

Castracane,V.D. and Jordan,V.C. (1975) The effect of estrogen and progesterone on uterine prostaglandin biosynthesis in the ovariectomized rat. Biol. Reprod., 13, 587-596. 

Castracane,V.D. and Jordan,V.C. (1976) Considerations into the mechanism of estrogen-stimulated uterine prostaglandin synthesis. Prostaglandins, 12, 243-251. 

Kelly,R.W. and Abel,M.H. (1981) A comparison of the effects of 4-atechol oestrogens and 2-pyrogallol oestrogens on prostaglandin synthesis by the rat and human uterus. J. Steroid Biochem., 14, 787-791. 

Pakrasi,P.L. and Dey,S.K. (1983) Catechol estrogens stimulate synthesis of prostaglandins in the preimplantation rabbit blastocyst and endometrium. Biol. Reprod., 29, 347-354. 

Biswas,A., Chaudhury,A., Chattoraj,S.C. and Dale,S.L. (1991) Do catechol estrogens participate in the initiation of labor? Am. J. Obstet. Gyn., 165, 984-987. 

Parvizi,N. and Ellendorff,F. (1983) Catechol estrogens in the brain: neuroendocrine integration. J. Steroid Biochem., 19, 615-618. 

Braun,T. (1990) Inhibition of the soluble form of testis adenylate cyclase by catechol estrogens and other catechols. Proc. Soc. Exp. Biol. Med. , 194, 58-63. 

Fishman,J. and Tulchinsky,D. (1980) Suppression of prolactin secretion in normal young women by 2-hydroxyestrone. Science, 210, 73-74. 

Linton,E.A., White,N., de Tineo,O.L. and Jeffcoate,S.L. (1981) 2-ydroxyestradiol inhibits prolactin release from the superfused rat pituitary gland. J. Endocrinol., 90, 315-322. 

Paden,C.M., McEwen,B.S., Fishman,J., Snyder,L. and DeGroff,V. (1982) Comparison by estrogens for catecholamine receptor binding in vitro. J. Neurochem., 39, 512-520. 

Schaeffer,J. and Hsueh,A. (1979) 2-Hydroxyestradiol interactions with dopamine receptor binding in rat anterior pituitary. J. Biol. Chem., 254, 5606-5608. 

Schaefer,J., Stevens,S., Smith,R. and Hsueh,A. (1980) Binding of 2-ydroxyestradiol to rat anterior pituitary cell membrane. J. Biol. Chem. , 255, 9838-9843. 

Pietras,R.J. and Szego,C.M. (1977) Specific binding sites for oestrogen at the outer surfaces of isolated endometrial cells. Nature, 265, 69-72. 

Bression,D., Michard,M., Le Dafniet,M., Pagesy,P and Peillon,F. (1986) Evidence for a specific estradiol binding site on the pituitary membrane. Endocrinology, 119, 1048-1051. 

Watson,C.S., Pappas,T.C. and Gametchu,B. (1995) The other estrogen receptor in the plasma membrane: implications for the actions of environmental estrogens. Environ. Hlth Perspect., 103 (Suppl. 7), 41-50. 

Silberzahn,P, Almahbobi,G., Dehennin,L. and Merouane,A. (1985) Estrogen metabolites in equine ovarian follicles: gas chromatographic-ass spectrometric determinations in relation to follicular ultrastructure and progestin content. J. Steroid Biochem., 22, 501-505. 

Spicer,L.J. and Hammond,J.M. (1989) Regulation of ovarian function by catecholestrogens: current concepts. J. Steroid Biochem., 33, 489-501. 

Liehr,J.G., Ulubelen,A.A. and Strobel,H.W. (1986) Cytochrome P-450-ediated redox cycling of estrogens. J. Biol. Chem., 261, 16865-16870. 

Liehr,J.G. and Roy,D. (1990) Free radical generation by redox cycling of estrogens. Free Radical Biol. Med., 8, 415-423. 

Liehr,J.G. (1990) Genotoxic effects of estrogens. Mutat. Res., 238, 269-276. 

Nutter,L.M., Ngo,E.O. and Abul-Hajj,Y.J. (1991) Characterization of DNA damage induced by 3,4-estrone-o-quinone in human cells. J. Biol. Chem., 266, 16380-16386. 

Han,X. and Liehr,J.G. (1994) DNA single-strand breaks in kidneys of Syrian hamsters treated with steroidal estrogens: hormone-induced free radical damage preceding renal malignancy. Carcinogenesis, 15, 9971000. 

Han,X. and Liehr,J.G. (1994) 8-Hydroxylation of guanine bases in kidney and liver DNA of hamsters treated with estradiol: Role of free radicals in estrogen-induced carcinogenesis. Cancer Res., 54, 5515-5517. 

Cavalieri,E.L., Stack,D.E., Devanesan,P.D., Todorovic,R., Dwiredy,I., Higginbotham,S., Johansson,S.L., Patil,K.D., Gross,M.L., Gooden,J.K., Ramanathan,R., Cerny,L. and Rogan,E.G. (1997) Molecular origin of cancer: catechol estrogen-3,4-quinones as endogenous tumor initiators. Proc. Natl Acad. Sci. USA, 94, 10937-10942. 

Hayashi,N., Hasegawa,K., Komine,A.,Tanaka,Y., McLachlan,J.A., Barrett,J.C. and Tsutsui,T. (1996) Estrogen-induced cell transformation and DNA-adduct formation in cultured Syrian hamster embryo cells. Mol. Carcinogenesis, 16, 149-156. 

Li,S.A., Purdy,R.H. and Li,J.J. (1989) Variations in catechol-O-ethyltransferase activity in rodent tissues: possible role in estrogen carcinogenicity. Carcinogenesis, 10, 63-67. 

Roy,D., Weisz,J. and Liehr,J.G. (1990) The O-methylation of 4-ydroxyestradiol is inhibited by 2-hydroxyestradiol: implications for estrogen-induced carcinogenesis. Carcinogenesis, 11, 459-462. 

Lipsett,M.B., Merriam,G.R., Kono,S., Brandon,D.D., Pfeiffer,D.G. and Loriaux,D.L. (1983) Metabolic clearance of catechol estrogens. In Merriam,G.R. and Lipsett,M.B. (eds) Catechol Estrogens. Raven Press, New York, pp. 105-114 

Martucci,C. and Fishman,J. (1976) Uterine estrogen receptor binding of catecholestrogens and of estetrol (1,3,5(10)estratriene-3,15α,16α,17β-etrol). Steroids, 27, 325-333. 

Barnea,E.R., MacLusky,N.J. and Naftolin,F. (1983) Kinetics of catechol estrogen-estrogen receptor dissociation: a possible factor underlying differences in catechol estrogen biological activity. Steroids, 41, 643. 

Franks,S., MacLusky,N.J. and Naftolin,F. (1982) Comparative pharmacology of oestrogens and catecholestrogens: actions on the immature uterus in vivo and in vitro. J. Endocrinol., 94, 91-98. 

Mueller,G.C. (1995) In vitro stimulation of incorporation of formate-14С in surviving uterine segments by hydroxylated estradiols. Nature, 176, 127-128. 

Hoversland,R.C., Dey,S.K. and Johnson,D.C. (1982) Catechol estradiol induced implantation in the mouse. Life Sci., 30, 1801-1804. 

Kantor,B.S., Dey,S.K. and Johnson,D.C. (1985) Catecholoestrogen induced initiation of implantation in the delayed implanting rat. Acta Endocr., 109, 418-422. 

Dey,S.K., Davis,D.L., Hersey,R.M., Weisz,J., Johnson,D.C. and Pakrasi,P.L. (1984) Physiological aspects ofblastocyst uterine interaction. 

J. Biosci., 6 (Suppl.), 23-31. 

Liehr,J.G. (1994) Mechanism of metabolic activation and inactivation of catechol estrogens: a basis of genotoxicity. Polycyclic Aromatic Compounds, 6, 229-239. 

Cavalieri,E.L. (1994) Minisymposium on endogenous carcinogens: the catechol estrogen pathway. Polycyclic Aromatic Compounds, 6, 223-228. 

Li,J.J., Gonzalez,S., Banerjee,S.K. and Li,S.A. (1993) Estrogen carcinogenesis in the hamster kidney: role of cytotoxicity and cell proliferation. Environ. Hlth Perspect., 101 (Suppl. 5), 259-264. 

Li,J.J., Li,S.A., Oberley,T.D. and Parsons,J.A. (1995) Carcinogenic activities of various steroidal and nonsteroidal estrogens in the hamster kidney: Relation to hormonal activity and cell proliferation. Cancer Res. , 55, 4347-4351. 

Li,J.J. and Li,S.A. (1995) Estrogen carcinogenesis in the hamster kidney: a hormone-driven multi-step process. In Huff,J. Boyd,J. and Barrett,J.C. (eds) Cellular and Molecular Mechanisms of Hormonal Carcinogenesis: Environmental Influences. Wiley-Liss Inc., Philadelphia, pp. 247-259. 

El-Bayoumy,K., Ji,B.-Y., Upadhyaya,P, Chae,Y.-H., Kurtzke.,C., Rivenson,A., Reddy,B.S., Amin,S. and Hecht,S.S. (1996) Lack of tumorigenicity of cholesterol epoxides and estrone-3,4-quinone in the rat mammary gland. Cancer Res., 56, 1970-1973. 

Newbold,R.R., Bullock,B.C. and MacLachlan,J.A. (1990) Uterine adenocarcinoma in mice following developmental treatment with estrogen. Cancer Res., 50, 7677-7681. 

Wiklund,J., Wertz,N. and Gorski,J. (1981) Genetic differences in estrogen-nduced deoxyribonucleic acid synthesis in the rat pituitary: correlation with pituitary tumor susceptibility. Endocrinology, 111, 1140-1149. 

Zhu,B.T. and Liehr,J.G. (1993) Inhibition of the catechol-O-ethyltransferase-catalyzed O-methylation of 2-nd 4-hydroxyestradiol by catecholamines: Implications for the mechanism of estrogen-induced carcinogenesis. Arch. Biochem. Biophys., 304, 248-256. 

Castagnetta,L.A., Granata,O.M., Arcuri,F.P, Polito,L.M., Rosati,F. and Cartoni,G.P (1992) Gas chromatography/mass spectrometry of catechol estrogens. Steroids, 57, 437-443. 

Savas,U., Christou,M. and Jefcoate,C.R. (1993) Mouse endometrium stromal cells express a polycyclic aromatic hydrocarbon-inducible cytochrome P450 in mouse embryo fibroblasts (P450EF). Carcinogenesis, 14, 2013-2018. 

Savas,U., Bhattacharyya,K.K., Christou,M., Alexander,D.L. and Jefcoate,C.R. (1994) Mouse cytochrome P-450EF, representative of a new 1B subfamily of cytochrome P-450s. J. Biol. Chem., 269, 14905-14911. 

Hakkola,J., Pasanen,M., Pelkonen,O., et al. (1997) Expression of CYP1B1 in human adult and fetal tissues and differential inducibility of CYP1B1 and CYP1A1 by Ah receptor ligand in human placenta and cultured cells. Carcinogenesis, 18, 391-397. 

Shimada,T., Hayes,C.L., Yamazaki,H., Amin,S., Hecht,S.S., Guengerich,F.P. and Sutter,T.R. (1996) Activation of chemically diverse procarcinogens by human cytochrome P-450 1B1. Cancer Res., 56, 2979-2984. 

Brake,PB., Christsou,M., Savas,Ü., Bhattacharyya,K., Alexander,D., Larsen,M. and Jefcoate,R. (1995) Selective expression and hormonal regulation of cytochrome P450 1B1 in mammary cells. Proc. Am. Ass. Cancer Res., 36, 257. 

Bjeldanes,L.F., Kim,J.-Y., Grose,K.R., Bartholomew,J.C. and Bradfield,C.A. (1991) Aromatic hydrocarbon responsiveness-receptor agonists generated from indole-3-carbinol in vitro and in vivo: comparisons with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Proc. Natl Acad. Sci. USA, 88, 9543-9547. 

Mueller,G.C. and Rumney,G. (1957) Formation of 6p-hydroxy and 6-eto derivatives of estradiol-16-C14 by mouse liver microsomes. J. Am. Chem. Soc., 79, 1004-1005. 

Huggins,C. and Jensen,E.V. (1955) The depression of estrone-induced uterine growth by phenolic estrogens with oxygenated functions at positions 6 or 16: the impeded estrogens. J. Exp. Med., 102, 335-346. 

Muller,R.E. and Wotiz,H.H. (1977) Post-coital contraceptive activity and estrogen receptor binding affinity of phenolic steroids. Endocrinology, 100, 513-519. 

Takagi,H., Komatsu,K.-I. and Yoshizuwa,I. (1991) Synthesis and mechanism of hydrolysis of estrogen 6-sulfates: model compounds for demonstrating the carcinogenesis of estrogen. Steroids, 56, 173-179. 

Phillips,D.H., Miller,J.A., Miller,E.C. and Adams,B. (1981) Structures of the DNA adducts formed in mouse liver after administration of the proximate hepatocarcinogen 1’-hydroxyestragole. Cancer Res., 41, 176-186. 

Watabe,T., Ishizuka,T., Isobe,M. and Ozawa,N. (1982) A 7-hydroxymethyl sulfate ester as an active metabolite of 7,12-dimethylbenz[a]anthracene. Science, 215, 403-405. 

Bonerg,E.W., Miller,E.C., Miller,J.A., Poland,A. and Liem,A. (1983) Strong evidence from studies with brachymorphic mice and pentachlorophenol that 1’-sulfooxysafrole is the major ultimate electrophilic and carcinogenic metabolite of 1’-hydroxysafrole in mouse liver. Cancer Res., 43, 5163-5173. 

Laskin,A.I., Grabowich,P, Junta,B., de Lisle Meyers,C. and Fried,J. (1964) Microbial hydroxylation of estrone and estradiol in the 6p-, 7α, and 15 a-ositions. J. Org. Chem. , 29, 1333-1336. 

Knuppen,R., Haupt,M. and Breuer,H. (1965) Formation of 15α-ydroxyestradiol-17β and 1a-hydroxyestrone by human adrenal tissue. J. Endocrinol., 33, 529-530. 

Younglai,E.V. and Solomon,S. (1968) Formation of estra-1,3,5(10)-triene-,15α,16α, 17β-tetrol (estetrol) and estra-1,3,5(10)-triene-3,15α,17β-triol from neutral precursors. J. Clin. Endocrinol. Metab., 28, 1611-1617. 

Cantineau,R., Kremers,P., De Graeve,J., Gielen,J.E. and Lambotte,R. (1985) 15-nd 16-Hydroxylations of androgens and estrogens in the human fetal liver: a critical step in estetrol biosynthesis. J. Steroid Biochem., 22, 195-201. 

Maggs,J.L., Morgan,P and Park,B.K. (1992) The sexually differentiated metabolism of [6,7-3H]17β-oestradiol in rats: male-specific 15α-nd male-selective 16α-hydroxylation and female-selective catechol formation. J. SteroidBiochem. Mol. Biol., 42, 65-76. 

Adlercreutz,H. and Martin,F. (1976) Oestrogens in human pregnancy faeces. Acta Endocrinol., 83, 410-419. 

Schwers,J., Eriksson,G. and Diczfalusy,E. (1965) Metabolism of oestrone and oestradiol in the human foeto-placental unit at midpregnancy. Acta Endocrinol., 49, 65-82. 

Gurpide,E., Schwers,J., Welch,M.T., Vande Wiele,R.L. and Lieberman,S. (1966) Fetal and maternal metabolism of estradiol during pregnancy. Clin. Endocrinol. Metab., 26, 1355-1365. 

Schwers,J., Gurpide,E., Vande Wiele,R.L. and Lieberman,S. (1967) Ürinary metabolites of estradiol and estriol administered intra-mniotically. J. Clin. Endocrinol. Metab., 27, 1403-1408. 

Benagiano,G., Mancuso,S., de la Torre,B. and Diczfalusy,E. (1970) Metabolism of 17β-oestradiol-17α-3H by the previable human fetus at midterm. Acta Endocrinol., 63, 39-49. 

Fishman,J., Schut,H. and Solomon,S. (1972) Metabolism, production and excretion rates of 15α-hydroxyestriol in late pregnancy. J. Clin. 

Endocrinol. Metab., 35, 339-344. 

Tulchinsky,D., Frigoletto,F.D., Jr., Ryan,K.J. and Fishman,J. (1975) Plasma estetrol as an index of fetal well-being. J. Clin. Endocrinol. Metab., 40, 560-567. 

Kundu,N. and Grant,M. (1976) Radioimmunoassay of 15α-hydroxyestriol (estetrol) in pregnancy serum. Steroids, 27, 785-796. 

Notation,A.D. and Tagatz,G.E. (1977) Unconjugated estriol and 15α-ydroxyestriol in complicated pregnancies. Am. J. Obstet. Gynecol., 128, 747-756. 

Taylor,N.F. and Schackleton,C.H.L. (1978) 15α-Hydroxyestriol and other polar oestrogens in pregnancy monitoring- review. Ann. Clin. Biochem., 15, 1-11. 

Zucconi,,G., Lisboa,B.P, Simonitsch,E., Roth,L., Hagen,A.A. and Diczfalusy,E. (1967) Isolation of 15-hydroxy-oestriol from pregnancy urine and from the urine of newborn infants. Acta Endocrinol., 56, 413-423. 

Klopper,A., Jandial,V. and Wilson,G. (1975) Plasma steroid assay in the assessment of fetoplacental function. J. Steroid Biochem., 6, 651-656. 

Kunda,N. Wachs,M., Iverson,G. and Petersen,L.P (1981) Comparison of serum unconjugated estriol and estetrol in normal and complicated pregnancies. Obstet. Gynec., 58, 276-281. 

Chakraborty,C., Davis,D.L. and Dey,S.K. (1990) Estradiol-15α-ydroxylation: A new avenue of estrogen metabolism in peri-implantation pig blastocysts. J. Steroid Biochem., 35, 209-218. 

Fishman,J. and Martucci,C. (1978) Differential biological activity of estradiol metabolism. Pediatrics, 62, 1128-1133. 

Blume,N., Xu,Z.J., Arvind,P and Leonard,J. (1996) Identification of a novel CYP2D gene expressed in the mammary gland and its potential role in the 16α-hydroxylation of estrogens. The American Endocrine Society Annual Meeting, Abstract No. P2-806. 

Fishman,J. and Martucci,C.P (1980) Biological properties of 16α-ydroxyestrone: implications in estrogen physiology and pathophysiology. J. Clin. Endocrinol. Metab., 51, 611-615. 

Swaneck,G.E. and Fishman,J. (1988) Covalent binding ofthe endogenous estrogen 16α-hydroxyestrone to estradiol receptor in human breast cancer cells: characterization and intranuclear localization. Proc. Natl Acad. Sci. USA, 85, 7831-7835. 

Hsu,C.-J., Kirkman,B.R. and Fishman,J. (1991) Differential expression of oncogenes c-fos, c-myc and neu/Her-2 induced by estradiol and 16α-ydroxyestrone in human cancer cell line. 73rd Annual Endocrine Society Meeting, Washington D.C., abstract # 586. 

Miyairi,S., Ichikawa,T. and Nambara,T (1991) Structure of the adduct of 16α-hydroxyestrone with a primary amine: evidence for the Heyns rearrangement of steroidal D-ring a-hydroxyamines. Steroids, 56, 361-366. 

Bradlow,H.L., Herschcopf,R.J. and Fishman,J. (1986) Oestradiol 16α-ydroxylase: a risk marker for breast cancer. Cancer Surv., 5, 573-583. 

Bradlow,H.L., Sepkovic,D.W., Telang,N.T. and Osborne,M.P. (1995) Indole-3-carbinol: A novel approach to breast cancer prevention. Ann. NYAcad. Sci., 768, 180-200. 

Fishman,J., Osborne,M.P. and Telang,N.T. (1995) The role of estrogen in mammary carcinogenesis. Ann. NY Acad. Sci., 768, 91-100. 

Fishman,J., Schneider,J., Hershcopf,R.J. and Bradlow,H.L. (1984) Increased estrogen 16α-hydroxylase activity in women with breast and endometrial cancer. J. Steroid. Biochem. Mol. Biol., 20, 1077-1081. 

Bradlow,H.L. and Michnovicz,J.J. (1989) A new approach to the prevention of breast cancer. Proc. Roy. Soc. Edinburgh, 95B, 77-86. 

Telang,N.T., Basu,A., Modak,M.J., Bradlow,H.L. and Osborne,M.P. (1988) Parallel enhancement of ras protooncogene expression and of estradiol-16α-hydroxylation in human mammary terminal duct-lobular units (TDLÜ) by a carcinogen. Breast Cancer Res. Treat., 12, 138. 

Telang,N.T., Sato,A., Wong,G.Y., Osborne,M.P and Bradlow,H.L. (1992) Induction by estrogen metabolite 16α-hydroxyestrone of genotoxic damage and aberrant proliferation in mouse mammary epithelial cells. J. Natl Cancer Inst., 84, 634-638. 

Suto,A., Bradlow,H.L., Wong,G.Y., Osborne,M.P and Telang,N.T. (1993) Experimental down-regulation of intermediate biomarkers of carcinogenesis in mouse mammary epithelial cells. Breast Cancer Res. Treat., 27, 193-202. 

Bradlow,H.L., Hershcopf,R.J., Martucci,C.P. and Fishman,J. (1985) Estradiol 16α-hydroxylation in the mouse correlates with mammary tumor incidence and presence of mammary tumor virus: A possible model for the hormonal etiology of breast cancer in humans. Proc. Natl Acad. Sci. USA, 82, 6295-6299. 

Osborne,M.P., Telang,N.T., Kaur,S. and Bradlow,H.L. (1990) Influence of chemopreventive agents on estradiol metabolism and mammary preneoplasia in the C3H mouse. Steroids, 55, 114-119. 

Jellinck,P.H., Michnovicz,J.J. and Bradlow,H.L. (1991) Influence of indole-3-carbinol on the hepatic microsomal formation of catechol estrogens. Steroids, 56, 446-450. 

Jellinck,P.H., Forkert,P.G., Riddick,D.S., Okey,A.B., Michnovicz,J.J. and Bradlow,H.L. (1993) Ah receptor binding properties of indole carbinols and induction of hepatic estradiol hydroxylation. Biochem. Pharmacol., 45, 1129-1136. 

Bradlow,H.L., Michnovicz,J.J., Telang,N.T. and Osborne,M.P. (1991) Effects of dietary indole-3-carbinol on estradiol metabolism and spontaneous mammary tumors in mice. Carcinogenesis, 12, 1571-1574. 

Kojima,T., Tanaka,T. and Mori,H. (1994) Chemoprevention of spontaneous endometrial cancer in female Donryu rats by dietary indole--carbinol. Cancer Res., 54, 1446-1449. 

Davis,D.L., Bradlow,H.L., Wolff,M., Woodruff,T., Hoel,D.G. and Anton-ulver,H. (1993) Medical hypothesis: Xenoestrogens as preventable causes of breast cancer. Environ. Hlth Perspect., 101, 372-377. 

Davis,D.L. and Bradlow,H.L. (1995) Can environmental estrogens cause breast cancer? Scientific American, 273, 167-172. 

Bradlow,H.L., Davis,D.L., Lin,G., Sepkovic,D. and Tiwari,R. (1995) Effects of pesticides on the ratio of 16α/2-hydroxyestrone: A biological marker of breast cancer risk. Environ. Hlth Perspect., 103 (Suppl. 7), 147-150. 

Sepkovic,D.W., Bradlow,H.L., Michnovicz,J., Murtezani,S., Levy,I. and Osborne,M.P. (1994) Catechol estrogen production in rat microsomes after treatment with indole-3-carbinol, ascorbigen, or p-naphthaflavone: a comparison of stable isotope dilution gas chromatography-mass spectrometry and radiometric methods. Steroids, 59, 318-323. 

Fishman,J. (1966) Stereochemistry of enolization of 17-keto steroids. J. Org. Chem., 31, 520-523. 

Lemon,H.M., Heidel,J.W. and Rodriguez-Sierra,J.F. (1992) Increased catechol estrogen metabolism as a risk factor for nonfamilial breast cancer. Cancer, 69, 457-465. 

Adlercreutz,H., Gorbach,S.L., Goldin,B.R., Woods,M.N., Dwyer,J.T. and Hamalainen,E. (1994) Estrogen metabolism and excretion in Oriental and Caucasian women. J. Natl Cancer Inst., 86, 1076-1082. 

Adlercreutz,H., Gorbach,S.L., Goldin,B.R., Woods,M.N., Dwyer,J.T., Hockerstedt,K., Wahala,K., Hase,T., Hamalainen,E. and Fotsis,T. (1994) Diet and urinary estrogen profile in various populations. A preliminary report. Polycyclic Aromatic Compounds, 6, 261-273. 

Merrill,R.C. (1958) Estriol: a review. Physiol. Rev., 38, 463-480. 

Fotsis,T. (1987) The multicomponent analysis of estrogens in urine by ion exchange chromatography and GC-MS-II. Fractionation and quantitation of the main groups of estrogen conjugates. J. Steroid Biochem., 28, 215-226. 

MacMahon,B., Cole,P., Lin,T.M., Lowe,C.R., Mirra,A.P., Ravnihar,B., Salber,E.J., Valaoras,V.G. and Yuasa,S. (1970) Age at first birth and breast cancer risk. Bull. WldHlth Org., 43, 209-221. 

Yuan,J.M., Yu,M.C., Ross,R.K., Gao,Y.T. and Henderson,B.E. (1988) Risk factors for breast cancer in Chinese women in Shanghai. Cancer Res., 48, 1949-1953. 

Henderson,B.E., Pike,M.C., Bernstein,L. and Ross,R.K. (1996) Breast cancer. In Schottenfeld,D. and Fraumeni,J.F.,Jr (eds) Cancer Epidemiology and Prevention, 2nd edn. Oxford University Press, New York, Oxford, pp. 1022-1039. 

Yang,N.N., Venugopalan,M., Hardikar,S. and Glasebrook,A. (1996) Identification of an estrogen response element activated by metabolites of 17β-estradiol and raloxifene. Science, 273, 1222-1225. 

Simpkins,J.W., Singh,M. and Bishop,J. (1994) The potential role for estrogen replacement therapy in the treatment of the cognitive decline and neurodegeneration associated with Alzheimer’s disease. Neurobiol. Aging, 15, s195-s197. 

Brenner,D.E., Kukull,W.A., Stergachis,A., van Belle,G., Bowen,J.D., McCormick,W.C., Teri,L. and Larson,E.B. (1994) Post-menopausal estrogen replacement therapy and the risk of Alzheimer’s disease: a population-based case-control study. Am. J. Epidemiol., 140, 262-267. 

Barrett-Connor,E. and Kritz-Silverstein,D. (1993) Estrogen replacement therapy and cognitive function in older women. JAMA, 269, 2637-2641. 

Tang,M.-X., Jacobs,D., Stern,Y., Marder,K., Schofield,P., Gurland,B., Andrews,H. and Mayeux,R. (1996) Effect of estrogen during menopause on risk and age at onset of Alzheimer’s disease. Lancet, 348, 429-432. 

Naftolin,F., Leranth,C., Perez,J. and Garcia-Seyura,L.M. (1993) Estrogen induces synaptic plasticity in adult primate neuron. Neuroendocrinology, 57, 935-939. 

Kuntzman,R., Jacobson,M., Schneidman,K. and Conney,A.H. (1964) Similarities between oxidative drug-metabolizing enzymes and steroid hydroxylases in liver microsomes. J. Pharmacol. Exp. Ther., 146, 280-285. 

Conney,A.H. (1967) Pharmacological implications of microsomal enzyme induction. Pharmacol. Rev, 19, 317-366. 

Conney,A.H. (1982) Induction of microsomal enzymes by foreign chemicals and carcinogenesis by polycyclic aromatic hydrocarbons: G.H.A.Clowes Memorial Lecture. Cancer Res., 42, 4875-4917. 

Guengerich,F.P (1989) Polymorphism of cytochrome P-450 in humans. TIPS, 10, 107-109. 

Kagimoto,M., Heim,M., Kagimoto,K., Zeugin,T. and Meyer,U.A. (1990) Multiple mutations of the human cytochrome P450IID6 gene (CYP2D6) in poor metabolizers of debrisoquine. J. Biol. Chem., 265, 17209-17214. 

Johansson,I., Lundqvist,E., Bertilsson,L., Dahl,M.-L., Sjoqvist,F. and Ingelman-Sundberg,M. (1993) Inherited amplification of an active gene in the cytochrome P-450 CYP2D locus as a cause ofultrarapid metabolism of debrisoquine. Proc. Natl Acad. Sci. USA, 90, 11825-11829. 

De Morais,S.M.F., Wilkinson,G.R., Blaisdell,J., Nakamura,K., Meyer,U.A. and Goldstein,J.A. (1994) Identification of a new genetic defect responsible for the polymorphism of (S)-mephenytoin metabolism in humans. J. Biol. Chem., 269, 15419-15422. 

Gonzalez,F.J. and Gelboin,H.V. (1994) Role of human cytochromes P450 in the metabolic activation of chemical carcinogens and toxins. Drug Metab. Rev., 26, 165-183. 

Guengerich,F.P (1995) Human cytochrome P450 enzymes. In Ortiz de Montellano,P.R. (ed.) Cytochrome P450: Structure, Mechanism, and Biochemistry. Plenum Press, New York and London, 2nd edn. pp. 473-535. 

Gonzalez,F.J. (1995) Genetic polymorphism and cancer susceptibility: fourteenth Sapporo cancer seminar. Cancer Res., 55, 710-715. 

Welch,R.M., Levin,W. and Conney,A.H. (1967) Insecticide inhibition and stimulation of steroid hydroxylases in rat liver. J. Pharmacol. Exp. Ther. , 155, 167-173. 

Levin,W., Welch,R.M. and Conney,A.H. (1968) Effect of phenobarbital and other drugs on the metabolism and uterotropic action of estradiol-7p and estrone. J. Pharmacol. Exp. Ther, 159, 362-371. 

Levin,W., Welch,R.M. and Conney,A.H. (1968) Decreased uterotropic potency of oral contraceptives in rats pretreated with phenobarbital. Endocrinology, 83, 149-156. 

Welch,R.M., Levin,W., Kuntzman,R., Jacobson,M. and Conney,A.H. (1971) Effect of halogenated hydrocarbon insecticides on the metabolism and uterotropic action of estrogens in rats and mice. Toxicol. Appl. Pharmacol., 19, 234-246. 

Kuntzman,R., Sansur,M. and Conney,A.H. (1965) Effect of drugs and insecticides on the anesthetic action of steroids. Endocrinology, 77, 952-954. 

Conney,A.H., Jacobson,M., Levin,W., Schneidman,K. and Kuntzman,R. (1966) Decreased central depression effect of progesterone and other steroids in rats pretreated with drugs and insecticides. J. Pharmacol. Exp. Ther., 154, 310-318. 

Levin,W., Welch,R.M. and Conney,A.H. (1969) Inhibitory effect of phenobarbital or chlordane pretreatment on the androgen-induced increase in seminal vesicle weight in the rat. Steroids, 13, 155-161. 

Levin,W., Welch,R.M. and Conney,A.H. (1974) Increased liver microsomal androgen metabolism by phenobarbital: correlation with decreased androgen action on the seminal vesicles ofthe rat. J. Pharmacol. Exp. Ther., 188, 287-292. 

Levin,W., Welch,R.M. and Conney,A.H. (1970) Effect of carbon tetrachloride and other inhibitors of drug metabolism on the metabolism and action of estradiol-17β and estrone in the rat. J. Pharmacol. Exp. Ther., 173, 247-255. 

Michnovicz,J.J., Hershcopf,R.J., Naganuma,H., Bradlow,H.L. and Fishman,J. (1986) Increased 2-hydroxylation of estradiol as a possible mechanism for the anti-estrogenic effect of cigarette smoking. N. Eng. J. Med., 315, 1305-1309. 

Michnovicz,J.J., Naganuma,H., Hershcopf,R.J., Bradlow,H.L. and Fishman,J. (1988) Increased urinary catechol estrogen excretion in female smokers. Steroids, 52, 69-83. 

MacMahon,B., Trichopoulos,D., Cole,P. and Brown,J. (1982) Cigarette smoking and urinary estrogens. N. Engl. J. Med., 307, 1062-1065. 

Jensen,J., Christiansen,C. and Rodbro,P (1985) Cigarette smoking, serum estrogens, and bone loss during hormone-replacement therapy early after menopause. N. Engl. J. Med., 313, 973-975. 

Friedman,A.J., Ravnikar,V.A. and Barbieri,R.L. (1987) Serum steroid hormone profiles in post-menopausal smokers and nonsmokers. Fertil. Steril., 47, 398-401. 

Baron,J.A. (1984) Smoking and estrogen-related disease. Am. J. Epidemiol., 119, 9-22. 

Jensen,J. and Christiansen,C. (1988) Effect of cigarette smoking on serum lipoproteins and bone mineral content during post-menopausal hormone replacement therapy. Am. J. Obstet. Gynecol., 159, 820-825. 

Baron,J.A., Vecchia,C.L. and Levi,F. (1990) The antiestrogenic effect of cigarette smoking in women. Am. J. Obstet. Gynecol., 162, 502-514. 

Lesko,S.M., Rosenberg,L., Kaufman,D.W., etal. (1985) Cigarette smoking and the risk of endometrial cancer. N. Engl. J. Med., 313, 593-596. 

Dent,C.E., Richens,A., Rowe,D.J.F. and Stamp,T.C.B. (1970) Osteomalacia with long-term anticonvulsant therapy in epilepsy. Br. Med. J., 4, 69-72. 

Hunter,J., Maxwell,J.D., Stewart,D.A., Parsons,V. and Williams,R. (1971) Altered calcium metabolism in epileptic children on anticonvulsants. Br. Med. J., 4, 202-204. 

Hahn,TJ. and Halstead,L.R. (1983) Sequential changes in mineral metabolism and serum vitamin D metabolite concentrations produced by phenobarbital administration in the rat. Calcified Tissue Int., 35, 376-382. 

Michnovicz,J.J. and Bradlow,H.L. (1991) Altered estrogen metabolism and excretion in humans following consumption of indole-3-carbinol. Nutr. Cancer, 16, 59-66. 

Bradlow,H.L., Michnovicz,J.J., Wong,G.Y., Halper,M.P., Miller,D. and Osborne,M.P. (1994) Long-term responses of women to indole-3-carbinol or a high fiber diet. Cancer Epidemiol. Biomarkers Prevent., 3, 591-595. 

Welch,R.M., Levin,W. and Conney A.H. (1968) Stimulatory effect of phenobarbital on the metabolism in vivo of estradiol-17β and estrone in the rat. J. Pharmacol. Exp. Ther., 160, 171-178. 

Romkes,M. and Safe,S. (1988) Comparative activities of 2,3,7,8-etrachlorodibenzo-p-dioxin and progesterone as antiestrogens in female rat uterus. Toxicol. Appl. Pharmacol., 92, 368-380. 

Umbreit,T.H., Hesse,E.J., Macdonald,G.J. and Gallo,M.A. (1988) Effects of TCDD-estradiol interactions in three strains of mice. Toxicol. Lett. , 40, 1-9. 

DeVito,M.J., Thomas,T., Martin,E., Umbreit,T.H. and Gallo,M.A. (1992) Antiestrogenic action of 2,3,7,8-tetrachlorodibenzo-p-dioxin: tissue specific regulation of estrogen receptor in CD-1 mice. Toxicol. Appl. Pharmacol., 113, 284-292. 

Gierthy,J.F., Lincoln,D.W., Gillespie,M.B., Seeger,J.I., Martinez,H.L., Dickerman,H.W. and Kumar,S.A. (1987) Suppression of estrogen-egulated extracellular tissue plasminogen activator activity of MCF-7 cells by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Cancer Res., 47, 6198-6203. 

Gierthy,J.F., Bennett,J.A., Bradley,L.M. and Cutler,D.S. (1993) Correlation of in vitro and in vivo growth suppression of MCF-7 human breast cancer by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Cancer Res., 53, 3149-3153. 

Kociba,R.J., Keyes,D.G., Beyer,J.E., etal. (1978) Results of a two-year chronic toxicity and oncogenicity study of 2,3,7,8-tetrachlorodibenzo-p-ioxin in rats. Toxicol. Appl. Pharmacol., 46, 279-303. 

Williamson,D.G. (1979) The biochemistry of the 17-hydroxysteroid dehydrogenases. In Hobkirk, R. (ed.) Steroid Biochemistry, vol I. CRC Press, Boca Raton, FL, pp. 83-110. 

Karavolas,H.J., Baedecker,M.L. and Engel,L.L. (1970) Human placental 17β-estradiol dehydrogenase. V. Purification and partial characterization of the diphosphopyridine nucleotide (triphosphonucleotide) linked enzyme. J. Biol. Chem., 245, 4948-4952. 

Blomquist,C.H., Lindemann,N.J. and Hakanson,E.Y. (1987) Steroid modulation of 17β-hydroxysteroid oxidoreductase activities in human placental villi in vitro. J. Clin. Endocrinol. Metab., 65, 647-652. 

Luu-The,V., Labrie,C., Simard,J., Lachance,Y., Zhao,H.F., Couet,J., Leblanc,G. and Labrie,F. (1990) Structure of two in tandem human 17β-ydroxysteroid dehydrogenase genes. Mol. Endocrinol., 4, 268-275. 

Pittaway,D.E., Andersen,R.N., Coleman,S.A. Jr., Givens,J.R. and Wiser,W.L. (1983) Human ovarian 17β-hydroxysteroid oxidoreductase activity. A comparison of normal and polycystic ovarian tissues. J. Clin. Endocrinol. Metab., 56, 715-719. 

Martel,C., Rheaume,E., Takahashi,M., Trudel,C., Couet,J., Luu-The,V., Simard,J. and Labrie,F. (1992) Distribution of 17β-hydroxysteroid dehydrogenase gene expression and activity in rat and in human tissues. J. SteroidBiochem. Mol. Biol., 41, 597-603. 

Bleau,G., Roberts,K.D. and Chapdelaine,A. (1974) The in vitro and in vivo uptake and metabolism of steroid in human adipose tissue. J. Clin. Endocrinol. Metab., 39, 236-246. 

Weisntein,G.D., Frost,P. and Hsia,S.L. (1968) In vitro interconversion of estrone and 17β-estradiol in human skin and vaginal mucosa. J. Invest. Dermat., 5, 4-10. 

Tseng,L., Stolee,A. and Gurpide,E. (1972) Quantitative studies on the uptake and metabolism of estrogens and progesterone by human endometrium. Endocrinology, 90, 390-404. 

Bonney,R.C., Reed,M.J., Davidson,K., Beranek,PA. and James,V.H.T. (1983) The relationship between 17β-hydroxysteroid dehydrogenase activity and oestrogen concentrations in human breast tumours and in normal breast tissue. Clin. Endocrinol., 19, 727-739. 

McNeill,J.M., Reed,M.J., Beranek,P.A., Bonney,R.C., Ghilchik,M.W., Robinson,D.J. and James,V.H. (1986) A comparison of the in vivo uptake and metabolism of 3H-oestrone and 3H-oestradiol by normal breast and breast tumour tissue in post-menopausal women. Int. J. Cancer, 38, 193-196. 

Vermeulen,A., Deslypere,J.P., Pavidaens,R., Leclevq,G., Roy,F. and Henson,J.C. (1986) Aromatase, 17β-hydroxysteroid dehydrogenase and intratissular sex hormone concentrations in cancerous and normal glandular breast tissue in post-menopausal women. Eur. J. Cancer Clin. Oncol, 32, 515-522. 

Poutanen,M., Isomaa,V., Kainulainen,K. and Vihko,R. (1990) Progestin induction of 17β-hydroxysteroid dehydrogenase enzyme protein in the T-7D human breast cancer cell line. Int. J. Cancer, 46, 897-901. 

Adams,E.F., Rafferty,B. and White,M.C. (1991) Interleukin 6 is secreted by breast fibroblasts and stimulates 17β-oestradiol oxidoreductase activity in MCF-7 cells: possible paracrine regulation of breast oestradiol levels. Int. J. Cancer, 49, 118-121. 

Couture,P., Theriault,C., Simard,J. and Labrie,F. (1993) Androgen receptor-mediated stimulation of 17β-hydroxysteroid dehydrogenase activity by dihydrotestosterone and medroxyprogesterone acetate in ZR-5-1 human breast cancer cells. Endocrinology, 132, 179-185. 

Duncan,L.J., Coldham,N.G. and Reed,M.J. (1994) The interaction of cytokines in regulating oestradiol 17β-hydroxysteroid dehydrogenase activity in MCF-7 cells. J. Steroid Biochem. Mol. Biol., 49, 63-68. 

Jacobson,G.M. and Hochberg,R.B. (1968) 17β-Hydroxysteroid dehydrogenase from human red blood cells. J. Biol. Chem., 243, 2985-2994. 

Kaufmann,M., Carstensen,J., Husen,B. and Adamski,J. (1995) The tissue distribution of porcine 17β-estradiol dehydrogenase and its induction by progesterone. J. Steroid Biochem. Mol. Biol., 55, 535-539. 

Tseng,L. and Gurpide,E. (1972) Changes in the in vitro metabolism of estradiol during the menstrual cycle. Am. J. Obstet. Gynecol. , 114, 1002-1008. 

Tseng,L. and Gurpide,E. (1975) Induction of human endometrial estradiol dehydrogenase by progestins. Endocrinology, 97, 825-833. 

Casey,M.L., MacDonald,P.C. and Andersson,S. (1994) 17β-Hydroxy-teroid dehydrogenase type 2: Chromosome assignment and progestin regulation of gene expression in human endometrium. J. Clin. Invest. , 94, 2135-2141. 

King,R.J.B., Townsend,P.T. and Whitehead,M.I. (1981) The role of estradiol dehydrogenase in mediating progestin effects on endometrium from post-menopausal women receiving estrogens and progestins. J. Steroid Biochem., 14, 235-238. 

Luu-The,V., Zhang,Y., Poirier,D. and Labrie,F. (1995) Characteristics of human types 1, 2 and 3 17β-hydroxysteroid dehydrogenase activities: Oxidation/reduction and inhibition. J. Steroid Biochem. Mol. Biol., 55, 581-587. 

Singh,A. and Reed,M.J. (1991) Insulin-like growth factor I and insulin-like growth factor II stimualte oestradiol 17β-hydroxysteroid dehydrogenase activity in breast cancer cells. J. Endocrinol., 129, R5-R8. 

Singh,A., Blench,I., Morris,H.R., Savoy,L.-A. and Reed,M.J. (1992) Synergistic interaction of growth factors and albumin in regulating estradiol synthesis in breast cancer cells. Mol. Cell. Endocrinol., 85, 165-173. 

Peltoketo,H., Isomaa,V., Maentausta,O. and Vihko,R. (1988) Completeamino acid sequence of human placental 17β-hydroxysteroid dehydrogenase deduced from cDNA. FEBS Lett., 239, 73-77. 

Luu-The,V., Labrie,C., Zhao,H.F., Couet,J., Lachance,Y., Simard,J., LeBlanc,G., Gote,J., Berube,G., Gagne,R. and Labrie,F. (1989) Characterization of cDNAs for human estradiol 17β-dehydrogenase and assignment of the gene to chromosome 17: evidence of two mRNA species with distinct 5' termini in human placenta. Mol. Endocrinol., 3, 1301-1309. 

Wu,L., Einstein,M., Geissler,W.M., Chan,H.K., Elliston,K.O. and Andersson,S. (1993) Expression cloning and characterization of human 17β-hydroxysteroid dehydrogenase type 2, a microsomal enzyme possessing 20a-hydroxysteroid dehydrogenase activity. J. Biol. Chem., 268, 12964-2969. 

Geissler,W.M., Davis,D.L., Wu,L., Bradshaw,K.D., Patel,S., Mendonca,B.B., Elliston,K.O., Wilson,J.D., Russell,D.W. and Andersson,S. (1994) Male pseudohermaphroditism caused by mutations of testicular 17β-hydroxysteroid dehydrogenase 3. Nature Genet., 7, 34-39. 

Leenders,F., Adamski,J., Husen,B., Thole,H.H. and Jungblut,PW. (1994) Molecular cloning and amino acid sequence of the porcine 17 betaestradiol dehydrogenase. Eur. J. Biochem., 222, 221-227. 

Labrie,F. (1991) Intracrinology. Mol. Cell. Endocrinol., 78, C113-C118. 

Ryan,K.J. (1959) Biological aromatization of steroids. J. Biol. Chem., 234, 268-272. 

Simpson,E.R., Mahendroo,M.S., Means,G.D., Kilgore,M.W., Hinshelwood,M.M., Graham-Lorence,S., Amarneh,B., Ito,Y., Fisher,C.R., Michael,M.D., Mendelson,C.R. and Bulun,S.E. (1994) Aromatase cytochrome P450, the isozyme responsible for estrogen metabolism. Endocr. Rev., 15, 342-355. 

McNatty,K.P., Baird,D.T., Bolton,A., Chambers,P., Corker,C.S. and MacLean,H. (1976) Conversion of oestrogens and androgens in human ovarian venous plasma and follicular fluid throughout the menstrual cycle. 

J. Endocrinol., 71, 77-85. Tan,L. and Muto,N. (1986) Purification and reconstitution properties of human placental aromatase. A cytochrome P-450-type monooxygenase. Eur. J. Biochem., 156, 243-250. 

Naftolin,F., Ryan,K.J., Davis,I.J., Reddy,V.V., Flores,F., Petro,Z. and Huhn,M. (1975) The formation of estrogen by central neuroendocrine tissues. Recent Prog. Horm. Res. , 31, 295-315. 

Selmanoff,M.K., Brodkin,L.D., Weiner,L.D. and Siiteri,P.K. (1977) Aromatization and 5a-reduction of androgens in discrete hypothalamic and limbic regions of the male and female rat. Endocrinology, 101, 841-848. 

Naftolin,F. and MacLusky,N.J. (1982) Aromatase in the central nervous system. Cancer Res., 42 (Suppl.), 3274s-3276s. 

Schindler,A.E., Ebert,A. and Frederick,E. (1972) Conversion of androstenedione to estrone by human fat tissue. J. Clin. Endocrinol. Metab., 35, 627-630. 

Nimrod,A. and Ryan,K.J. (1975) Aromatization of androgens by human abdominal and breast fat tissue. J. Clin. Endocrinol. Metab., 40, 367-379. 

Smuk,M. and Schwers,J. (1977) Aromatization of androstenedione by human adult liver in vitro. J. Clin. Endocrinol. Metab., 45, 1009-1012. 

Schweikert,H.U., Milewick,L. and Wilson,J.D. (1976) Aromatization of androstenedione by cultured human fibroblasts. J. Clin. Endocrinol. Metab., 43, 785-795. 

Berkovitz,G.D., Fujimoto,M., Brown,T.R., Brodie,A.M. and Migeon,C.J. (1984) Aromatase activity in cultured human genital skin fibroblasts. J. Clin. Endocrinol. Metab., 59, 665-671. 

Miller,W.R. and Forrest,A.P.M. (1974) Oestradiol synthesis by human breast carcinoma. Lancet, 2, 866-868. 

Lipton,A., Santner,S.J., Santen,R.J., Harvey,H.A., Feil,P.D., White-ershey,D., Bartholomew,M.J. and Antle,C.E. (1987) Aromatase activity in primary and metastatic human breast cancer. Cancer, 59, 779-782. 

Esteban,J.M., Warsi,Z., Haniu,M., Hall,P., Shively,J.E. andChen,S. (1992) Detection of intratumoral aromatase in breast carcinoma: an immunohistochemical study with clinicopathologic correlation. Am. J. Pathol., 140, 337-343. 

O’Neill,J.S., Elton,R.A. and Miller,W.R. (1988) Aromatase activity in adipose tissue from breast quadrants: a link with tumor site. Br. Med. J., 296, 741-743. 

Santen,R.J., Martel,J., Hoagland,M., Naftolin,F., Rao,L., Harada,N., Hafer,L., Zaino,R. and Santner,S.J. (1994) Stromal spindle cells contain aromatase in human breast tumors. J. Clin. Endocrinol. Metab. , 79, 627-632. 

Zumoff,B. (1982) Relationship of obesity to blood estrogens. Cancer Res., 42, 3289s-3294s. 

Siiteri,P.K., Williams,J.E. and Takaki,N.K. (1976) Steroid abnormalities in endometrial and breast carcinoma: a unifying hypothesis. J. Steroid Biochem. , 7, 897-903. 

Siiteri,P.K. (1981) Extraglandular estrogen formation and serum binding of estradiol: relationship to cancer. J. Endocrinol., 89 (Suppl.), 119P-129P. 

Siiteri,P.K. (1982) Review of studies on estrogen biosynthesis in the human. Cancer Res., 42 (suppl.), 3269s-3273s. 

Santen,R.J., Martel,J., Hoagland,M., Naftolin,F., Roa,L., Harada,N., Hafer,L., Zaino,R. and Santener,S.T. (1994) Stromal spindle cell distribution of aromatase in human breast tumors. J. Clin. Endocrinol. Metab., 79, 627-632. 

Miller,W.R. (1989) Aromatase inhibitors in the treatment of advanced breast cancer. Cancer Treat. Rev. , 16, 83-93. 

Demers,L.M. (1994) Effects of fadrozole (CGS 16949) and letrozole (CGS 20267) on the inhibition of aromatase activity in breast cancer patients. Breast Cancer Res. Treat., 30, 95-102. 

Yue,W., Wang,J., Savinov,A. and Brodie,A. (1995) Effect of aromatase inhibitors on growth of mammary tumors in a nude mouse model. Cancer Res., 55, 3073-3077. 

Means,G.D., Mahendroo,M.S., Corbin,C.J., Mathis,M.J., Powell,F.E., Mendelson,C.R. and Simpson,E.R. (1989) Structural analysis of the gene encoding human aromatase cytochrome P-450, the enzyme responsible for estrogen biosynthesis. J. Biol. Chem., 264, 19385-19391. 

Toda,K., Terashima,M., Kawamoto,T., et al. (1990) Structural and functional characterization of human aromatase P450 gene. Eur. J. Biochem., 193, 559-565. 

Kilgore,M.W., Means,G.D., Mendelson,C.R. and Simpson,E.R. (1992) Alternative promotion of aromatase P-450 expression in the human placenta. Mol. Cell. Endocrinol., 83, R9-R16. 

Harada,N., Utsumi,T. and Takagi,Y. (1993) Tissue-specific expression of the human aromatase cytochrome P-450 gene by alternative use of multiple exons 1 and promoters, and switching of tissue-specific exons 1 in carcinogenesis. Proc. Natl Acad. Sci. USA, 90, 11312-11316. 

Simpson,E.R., Michael,M.D., Agarwal,VR., Hinshelwood,M.M., Bulun,S.E. and Zhao,Y. (1997) Expression of the CYP19 (aromatase) gene: an unusual case of alternative promoter usage. FASEB J., 11, 29-36. 

Harada,N., Utsumi,T. and Takagi,Y. (1995) Molecular and epidemiological analyses of abnormal expression of aromatase in breast cancer. Pharmacogenetics, 5, S59-S64. 

Mahendroo,M.S., Mendelson,C.R. and Simpson,E.R. (1993) Tissue-pecific and hormonally controlled alternative promoters regulate aromatase cytochrome P-450 gene expression in human adipose tissue. J. Biol. Chem., 268, 19463-19470. 

Lipton,A., Santner,S.J., Santen,R.J., Harvey,H.A., Feil,P.D., White-ersey,D., Bartholomew,M.J. and Antle,C.E. (1987) Aromatase activity in primary and metastatic human breast cancer. Cancer, 59, 779-782. 

Longcope,C. (1996) Dehydroepiandrosterone metabolism. J. Endocrinol., 150, S125-S127. 

Dao,T.L., Hayes,C. and Libby,P.R. (1974) Steroid sulfatase activities in human breast tumors. Proc. Soc. Exp. Biol. Med., 146, 381-384. 

Vignon,F., Terqui,M., Westley,B., Derocq,D. and Rochefort,H. (1980) Effects of plasma estrogen sulfates in mammary cancer cells. Endocrinology, 106, 1079-1086. 

Carlstrom,K., von Uexkull,A.-K., Einhorn,N., Fredricsson,B., Lunell,N.-O. and Sundelin,P. (1983) Metabolism of estrone sulfate in human endometrium. Acta Obstet. Gynecol. Scand., 62, 519-524. 

Tseng,L., Mazella,J., Lee,L.Y. and Stone,M.L. (1983) Estrogen sulfatase and estrogen sulfotransferase in human primary mammary carcinoma. J. Steroid Biochem. , 19, 1413-1417. 

Santner,S.J., Feil,P.D. and Santen,R.J. (1984) In situ estrogen production via the estrone sulfatase pathway in breast tumors: relative importance versus the aromatase pathway. J. Clin. Endocrinol. Metab. , 59, 29-33. 

MacIndoe,J.H., Woods,G., Jeffries,L. and Hinkhouse,M. (1988) The hydrolysis of estrone sulfate and dehydroepiandrosterone sulfate by MCF- human breast cancer cells. Endocrinology, 123, 1281-1287. 

Carlstrom,K., Bergqvist,A. and Ljungberg,O. (1988) Metabolism of estrone sulfate in endometriotic tissue and in uterine endometrium in proliferative and secretory phase. Fertil. Steril. , 49, 229-233. 

Santner,S.J., Leszczynski,D., Wright,C., Manni,A., Feil,P.D. and Santen,R.J. (1986) Estrone sulfate: a potential source of estradiol in human breast cancer tissues. Breast Cancer Res. Treat., 7, 35-44. 

Millington,D.S. (1975) Determination of hormonal steroid concentrations in biological extracts by high resolution mass fragmentography. J. Steroid Biochem., 6, 239-243. 

Trévoux,R., De Brux,J., Castanier,M., Nahoul,K., Soule,J.-P. and Scholler,R. (1986) Endometrium and plasma hormone profile in the perimenopause and post-menopause. Maturitas, 8, 309-326. 

Twombly,G.H. and Levitz,M. (1960) Metabolism of estrone-C14-16 sulfate in women. Am. J. Obstet. Gynecol. , 80, 889-326. 

Ruder,H.J., Loriaux,L. and Lipsett,M.B. (1972) Estrone sulfate: Production rate and metabolism in man. J. Clin. In-vest., 51, 1020-1033. 

Santen,R.J., Leszczynski,D., Tilson-Mallet,N., Feil,P.D., Wright,C., Manni,A. and Santner,S.J. (1986) Enzymatic control of estrogen production in human breast cancer: relative significance of aromatase versus sulfatase pathways. Ann. N.Y. Acad. Sci., 464, 126-137. 

Milewich,L., Garcia,R.L. and Gerrity,L.W. (1984) Steroid sulfatase and 17β-hydroxysteroid oxidoreductase activities in mouse tissues. J. Steroid Biochem., 21, 529-538. 

Zhu,B.T. and Fu,J.-H. (1997) Uterine estrogen sulfatase plays a more important role than hepatic sulfatase in determining the sensitivity of the uterus to estrone-3-sulfate. Endocrine, 7, 191-198. 

Pasqualini,J.R., Gelly,C., Nguyen,B.-L. and Vella,C. (1989) Importance of estrogen sulfates in breast cancer. J. Steroid Biochem., 34, 155-163. 

Santner,S.J., Levin,M.C. and Santen,R.J. (1990) Estrone sulfate stimulates growth of nitrosomethylurea-induced breast carcinoma in vivo in the rat. Int. J. Cancer, 46, 73-78. 

Zhu,B.T., Evaristus,E.N., Antoniak,S.K., Sarabia,S.F., Ricci,M.J. and Liehr,J.G. (1996) Metabolic deglucuronidation and demethylation of estrogen conjugates as a source of parent estrogens and catechol estrogen metabolites in Syrian hamster kidney, a target organ of estrogen-induced tumorigenesis. Toxicol. Appl. Pharmacol., 136, 186-193. 

Walaszek,Z., Hanusek,M., Sherman,U. and Adams,A.K. (1990) Antiproliferative effect of dietary glucarate on the Sprague-Dawley rat mammary gland. Cancer Lett., 49, 51-57. 

Walaszek,Z., Hanausek-Walaszek,M. and Webb,T.E. (1984) Inhibition of 7,12-dimethylbenz[a]anthracene-induced rat mammary tumorigenesis by 2,5-di-O-acetyl-D-glucaro-1,4: 6,3-dilactone, an in vivo p-glucuronidase inhibitor. Carcinogenesis, 5, 767-772. 

Walaszek,Z., Hanausek-Walaszek,M., Minton,J.P. and Webb,T.E. (1986) Dietary glucarate as anti-promoter of 7,12-dimethylbenz[a]anthracene-nduced mammary tumorigenesis. Carcinogenesis, 7, 1463-1466. 

Walaszek,Z. (1990) Potential use of D-glucaric acid derivatives in cancer prevention. Cancer Lett., 54, 1-8. 

Watanabe,K. and Yoshizawa,I. (1987) Tissue distribution of estradiol 17-ulfate 2-nd 4-hydoxylation enzymes in the rat. J. Pharmacobio-Dyn., 10, 302-308. 

Ingelman-Sundberg,M., Rane,A. and Gustafasson,J.-Â. (1975) Properties of hydroxylase systems in the human fetal liver active on free and sulfoconjugated steroids. Biochemistry, 14, 429-437. 

Duncan,L., Purohit,A., Howarth,N.M., Potter,B.V.L. and Reed,M.J. (1993) Inhibition of estrogen sulfatase activity by estrone-3-methyl-hiophosphonate: A potential therapeutic agent in breast cancer. Cancer Res., 53, 298-303. 

Howarth,N.M., Purohit,A., Reed,M.J. and Potter,B.V.L. (1994) Estrone sulfamates: Potent inhibitors ofestrone sulfatase with therapeutic potential. J. Med. Chem., 37, 219-221. 

Reed,M.J., Purohit,A., Woo,L.W.L. and Potter,B.V.L. (1996) The development of steroid sulfatase inhibitors. Endocr. Rel. Cancer, 3, 9-23. 

Larner,J.M., Maclusky,N.J. and Hochberg,R.B. (1985) The naturally occurring C-17 fatty acid esters of estradiol are long-acting estrogens. J. Steroid Biochem., 22, 407-413. 

Vazquez-Alcantara,M.A., Menjivar,M., Garcia,G.A., Diaz-Zagoya,J.C. and Garza-Flores,J. (1989) Long-acting estrogenic responses of estradiol fatty acid esters. J. Steroid Biochem., 33, 1111-1118. 

Janocko,L., Larner,J.M. and Hochberg,R.B. (1984) The interaction of C-7 esters of estradiol with the estrogen receptor. Endocrinolgoy, 114, 1180-1186. 

Axelrod,J. and Tomchick,R. (1958) Enzymatic O-methylation of epinephrine and other catechols. J. Biol. Chem., 233, 702-705. 

Guldberg,H.C. and Marsden,C.A. (1975) Catechol-O-methyltransferase: Pharmacological aspects and physiological role. Pharmacol. Rev. , 27, 135-206. 

Thakker,D.R. and Creveling,C.R. (1990) O-Methylation. In Mulder,G.J. (ed.) Conjugation Reactions in Drug Metabolism. Taylor and Francis Ltd, London, pp. 193-232. 

Axelrod,J., Albers,W. and Clemente,C.D. (1959) Distribution of catechol--methyltransferase in the nervous system and other tissues. 

J. Neurochem., 5, 68-71. 

Inoue,K., Tice,L.W. and Creveling,C.R. (1977) Immunocytochemical localization of catechol-O-methyltransferase. In Usdin,E, Weiner,N. and Youdim,M.B.H. (eds) Structure and Function of Monoamine Enzymes. Marcel Dekker, New York, pp. 835-859. 

Amin,A.M., Creveling,C.R. and Lowe,M.C. (1983) Immunohistochemical localization of catechol methyltransferase in normal and cancerous breast tissues of mice and rats. J. Natl Cancer Inst., 70, 337-342. 

Assicot,M., Contesso,G. and Bohuon,C. (1977) Catechol-O-ethyltransferase in human breast cancers. Eur. J. Cancer, 13, 961-966. 

Bertocci,B., Miggiano,V, Da Prada,M., Dembic,Z., Lahm,H.W. and Malherbe,P. (1991) Human catechol-O-methyltransferase: cloning and expression of the membrane-associated form. Proc. Natl Acad. Sci. USA, 88, 1416-1420. 

Kraychy,S. and Gallagher,T.F. (1957) 2-Methoxyestrone, anew metabolite of estradiol-17β in man. J. Biol. Chem., 229, 519-526. 

Ball,P. Reu,G., Schwab,J. and Knuppen,R. (1979) Radioimmunoassay of 2-hydroxyestrone and 2-methoxyestrone in human urine. Steroids, 33, 563-576. 

Ball,P. and Knuppen,R. (1990) Formation, metabolism, and physiologic importance of catecholestrogens. Am. J. Obstet. Gynecol., 163,2163-2170. 

Dunn,J.F. (1983) Transport of estrogens in human plasma. In Merriam,G.R. and Lipsett,M.B. (eds) Catechol Estrogens. Raven Press, New York. pp. 167-176. 

Merriam,G.R., MacLusky,N.J., Pickard,M.K. and Naftolin,F. (1980) Comparative properties of the catechol estrogens. I: Methylation by catechol-O-methyltransferase and binding to cytosol estrogen receptors. Steroids, 369, 1-11. 

Jellinck,PH. and Newcombe,A.M. (1980) Effect of catechol estrogens and estriol on the induction of uterine peroxidase. J. Steroid Biochem., 13, 681-683. 

Gordon,S, Cantrall,E.W., Cekleniak,W.P., Albers,H.J., Mauer,S., Stolar,S.M. and Bernstein,S. (1964) Steroid and lipid metabolism. The hypocholesteremic effect of estrogen metabolites. Steroids, 4, 787-791. 

Patt,V., Mollering,M. and Breuer,H. (1971) The effect of 2-r 4-ubstituted derivatives of oestrogens on blood lipids in rats. ArzneimittelForsch., 21, 308-310. 

Ferreri,L.F. and Naito,H.K. (1978) Effect of estrogen on rat serum cholesterol concentrations: consideration of dose, type of estrogen, and treatment duration. Endocrinology, 102, 1621-1627. 

Martucci,C. (1982) Induction of rat plasma renin substrate by 2-ethoxyestrone. The Endocrine Society 64th Annual Meeting, Abstract No. 889. Katayama,S. and Fishman,J. (1982) 2-Hydroxyestrone suppresses and 2-ethoxyestrone augments the preovulatory prolactin surge in the cycling rat. Endocrinology, 110, 1448-1450. 

Cushman,M., He,H.M., Katzenellenbogan,J.A., Lin,C.M. and Hamel,E. (1995) Synthesis, antitubulin and antimitotic activity, and cytotoxicity of analogs of 2-methoxyestradiol, an endogenous mammalian metabolite of estradiol that inhibits tubulin polymerization by binding to the colchicine binding site. J. Med. Chem., 38, 2041-2049. 

Zhu,B.T. and Liehr,J.G. (1994) Quercetin increases the severity of estradiol-induced tumorigenesis in hamster kidney. Toxicol. Appl. Pharmacol, 125, 149-158. 

Zhu,B.T. and Liehr,J.G. (1996) Inhibition of catechol O-methyltransferase-atalyzed O-methylation of 2-nd 4-hydroxyestadiol by quercetin: possible role in estradiol-induced tumorigenesis. J. Biol. Chem., 271, 1357-1363. 

Riley,V. (1975) Mouse mammary tumors: alteration of incidence as apparent function of stress. Science, 189, 465-467. 

Cooper,C.L. (ed.) (1988) Stress and Breast Cancer. John Wiley and Sons, Chichester. Cooper,C.L., Cooper,R. and Faragher,E.B. (1989) Incidence and perception of psychosocial stress: the relationship with breast cancer.Psychol. Med., 19, 415-422. 

Faragher, E.B. and Cooper, C.L. (1990) Type A stress prone behavior and breast cancer. Psychol. Med., 20, 663-670. 

Forsen,A. (1991) Psychosocial stress as a risk for breast cancer. Psychother. Psychosom., 55, 176-185. 

MacLusky,N.J., Naftolin,F., Krey,L.C. and Franks,S. (1981) The catechol estrogen. J. Steroid Biochem., 15, 111-124. 

Levin,M., Weisz,J., Bui,Q.D. and Santen,R.J. (1987) Peroxidatic catecholestrogen production by human breast cancer tissue in vitro. J. Steroid Biochem., 28, 513-520. 

Cohn,C.K., Dunner,D.L. and Axelrod,J. (1970) Reduced catechol-O-ethyltransferase activity in red blood cells of women with primary effective disorder. Science, 170, 1323-1324. 

Weinshilboum,R.M., Raymond,F.A., Elveback,L.R. and Weidman,W.H. (1974) Correlation of erythrocyte catechol-O-methyltransferase activity between siblings. Nature, 252, 490-491. 

Weinshilboum,R.M. and Raymond,F.A. (1977) Inheritance of low erythrocyte catechol-O-methyltransferase activity in man. Am. J. Hum. Genet., 29, 125-135. 

Weinshilboum,R. (1989) Methyltransferase pharmacogenetics. Pharmacol. Ther., 43, 77-90. 

Boudikova,B., Szumlanski,C., Maidak,B. and Weinshilboum,R. (1990) Human liver catechol-O-methyltransferase pharmacogenetics. Clin. Pharmacol. Ther., 48, 381-389. 

Schatz,F. and Hochberg,R.B. (1981) Lipoidal derivative of estradiol: The biosynthesis of a nonpolar estrogen metabolite. Endocrinology, 109, 697-703. 

Mellon-Nussbaum,S.H., Ponticorvo,L., Schatz F. and Hochberg,R.B. (1982) Estradiol fatty acid esters: The isolation and identification of the lipoidal derivative of estradiol synthesized in the bovine uterus. J. Biol. Chem., 257, 5678-5684. 

Abul-Hajj,Y.J. (1982) Formation of estradiol-17β fatty acyl 17-esters in mammary tumors. Steroids, 40, 149-155. 

Adams,J,B., Hall,R.T. and Nott,S. (1986) Esterification-deesterification of estradiol by human mammary cancer cells in culture. J. Steroid Biochem., 24, 1159-1162. 

Martyn,P., Smith,D.L. and Adams,J.B. (1987) Selective turnover of the essential fatty acid ester components of estradiol-17β lipoidal derivatives formed by human mammary cancer cells in culture. J. Steroid Biochem., 28, 393-398. 

Hochberg,R.B., Zielinski,J.E., Larner,J.M. and Pahuja,S.L. (1991) Steroid esterification and hormone action. In Hochberg,R.B. and Naftolin,F. (eds) TheNew Biology ofSteroid Hormones. Raven Press, New York, pp. 31-43. 

Martyn,P., Smith,D.L. and Adams,J.B. (1988) Properties of fatty acyl-coenzyme A: estradiol-17β acyltransferase in bovine placenta microsomes. Mol. Cell. Endocrinol., 60, 7-13. 

Pahuga,S.L., Zielinski,J.E., Giordano,G., McMurray,W.J. and Hochberg,R.B. (1991) The biosynthesis of D-ring fatty acid esters of estriol. J. Biol. Chem., 266, 7410-7416. 

Katz,J., Finlay,T.H., Banerjee,S. and Levitz,M. (1987) An estrogen-ependent esterase activity in MCF-7 cells. J. Steroid Biochem., 26, 687-692. 

Janocko,L. and Hochberg,R.B. (1983) Estradiol fatty acid esters occur naturally in human blood. Science, 222, 1334-336. 

Hershcopf,R.J., Bradlow,H.L., Fishman,J., Swaneck,G.E., Larner,J.M. and Hochberg,R.B. (1985) Metabolism of estradiol fatty acid esters in man. J. Clin. Endocrinol. Metab., 61, 1071-1075. 

Larner,J.M., Shackleton,C.H.L., Roitman,E., Schwartz,P.E. and Hochberg,R.B. (1992) Measurement of estradiol-17-fatty acid esters in human tissues. Endocrinology, 75, 195-200. 

Maclusky,N.J., Larner,J.M. and Hochberg,R.B. (1989) Actions of an estradiol-17-fatty acid ester in estrogen target tissues of the rat: Comparison with other C-17 metabolites and a pharmacological C-17 ester. Endocrinology, 124, 318-324. 

Xu,S., Zhu,B.T. and Conney,A.H. (1997) Stimulatory effect of peroxisome proliferators on the formation of estradiol fatty acid esters by rat liver microsomes. Proc. Am. Ass. Cancer Res., 38, 451. 

Fuqua,A., Fitzgerald,S.D., Chamness,G.C., Tandon,A.K., McDonnell,D.P., Nawaz,Z, O’Malley,B.W. and McGuire,W.L. (1991) Variant human breast tumor estrogen receptor with constitutive transcriptional activity. Cancer Res., 51, 105-109. 

Pfeffer,U., Fecarotta,E. and Vidali,G. (1995) Coexpression of multiple estrogen receptor variant messenger RNAs in normal and neoplastic breast tissues and MCF-7 cells. Cancer Res., 55, 2158-2165. 

Pink,J.J. Wu,S.Q., Wolf,D.M., Bilimoria,M.M. and Jordan,V.C. (1996) A novel 80 kDa human estrogen receptor containing a duplication of exons 6 and 7. Nucleic Acids Res. , 24, 962-969. 

Grandien,K. (1996) Determination oftranscription start sites in the human estrogen receptor gene and identification of a novel, tissue-specific, estrogen receptor-mRNA isoform. Mol. Cell. Endocrinol. , 116, 207-212. 

Markaverich,B.M. and Clark,J.H. (1979) Two binding sites for estradiol in rat uterine nuclei: relationship to uterotropic response. Endocrinology, 105, 1458-1462. 

Markaverich,B.M., Williams,M., Upchurch,S. and ClarkJ.H. (1987) Heterogeneity of nuclear estrogen-binding sites in the rat uterus: a simple method for the quantitation of type I and type II sites by [3H]estradiol exchange. Endocrinology, 109, 62-69. 

Xu,X. and Thomas,M.L. (1995) Biophasic actions of estrogen on colon cancer cell growth: possible mediation by high-nd low-affinity binding sites. Endocrine, 3, 661-665. 

Markaverich,B.M., Gregory,R.R., Alejandro,M.A., Clark,J.H., Johnson,G.A. and Middleditch,B.S. (1988) Methyl p-hydroxy-henyllactate: an inhibitor of cell proliferation and an endogenous ligand for nuclear type II binding sites. J. Biol. Chem., 263, 7203-7210. 

Laudet,V., Hanni,C., Coll,J., Catzeflis,F. and Stehelin,D. (1992) Evolution of the nuclear receptor gene superfamily. EMBO J., 11, 1003-1012. 

Larsen,P.R. (1982) Thyroid-pituitary interaction. Feed back regulation of thyrotropin secretion by thyroid hormones. N. Engl. J. Med., 306, 23-32. 

Crantz,F.R., Silva,J.E. and Larson,P.R. (1982) An analysis of the sources and quantity of 3,5,3,-triiodothyronine specifically bound to nuclear receptors in rat cerebral cortex and cerebellum. Endocrinology, 110, 367-375. 

Wilson,J.D. (1996) Role of dihydrotestosterone in androgen action. Prostate, 6 (Suppl.), 88-92. 

Henry,H.L. (1992) Vitamin D hydroxylases. J. Cell. Biochem., 49, 4-9. 

Conney,A.H. and Kappas,A. (1985) Interindividual differences in the metabolism of xenobiotics. In Huberman,E. and Barr,S.H. (eds) The Role of Chemicals and Radiation in the Etiology of Cancer. Raven Press, New York, pp. 147-166. 

Meyer,U.A. (1994) Molecular genetics of human xenobiotic metabolizing cytochromes P450. Proceedings of the 10th International Symposium on Microsomes and Drug Oxidations, Toronto, Canada, pp. 2-4. 

De Morais,S.M.F., Wilkinson,G.R., Blaisdell,J., Meyer,U.A., 

Nakamura,K. and Goldstein,J.A. (1994) Identification of a new genetic defect responsible for the polymorphism of (S)-mephenytoin metabolism in Japanese. Mol. Pharmacol., 46, 594-98. 

Anderson,K.E., Kappas,A., Conney,A.H., Bradlow,H.L. and Fishman,J. (1984) The influence of dietary protein and carbohydrate on the principal oxidative biotransformation of estradiol in normal subjects. J. Clin. Endocrinol. Metab., 59, 103-107. 

van Loon,J. and Weinshilboum,R. (1984) Human platelet phenol sulfotransferase: familial variations in the thermal stability of the TS form. Biochem. Genet., 22, 997-1014. 

Weinshilboum,R. (1990) Sulfotransferase pharmacogenetics. Pharmacol. Ther., 45, 93-107. 

Lea,O.A., Thorsen,T. and Kvinnsland,S. (1987) Extracellular proteins in breast tumor cytosol. Anticancer Res., 7, 113-118. 

S0reide,J.A., Lea,O.A. and Kvinnsland,S. (1991) Cytosol albumin content in operable breast cancer. Acta Oncol., 30, 797-802. 

Rao,G.S. (1981) Mode of entry of steroid and thyroid hormones into cells. Mol. Cell. Endocrinol., 21, 97-108. 

Fishman,J., Cox,R.I. and Gallagher,T.F. (1960) 2-Hydroxyestrone: Anew metabolite of estradiol in man. Arch. Biochem. Biophys., 90, 318-319. 

Jellinck,P.H. and Lucieer,J. (1965) Sex differences in the metabolism of oestrogens by rat liver microsomes. J. Endocrinol., 32, 91-98. 

King, R.J.B. (1961) Oestriol metabolism by rat-nd rabbit-liver slices. Isolation of 2-methoxyoestriol and 2-hydroxyoestriol. Biochem. J., 79, 355-361. 

Willians,J.G., Longcope,C. and Williams,K.I.H. (1974) 4-Hydroxyestrone: a new metabolite of estradiol from humans. Steroids, 24, 687-701. 

Emons,G., Hoppen,H.-O., Ball,P. and Knuppen,R. (1980) 4-Hydroxy-strone, isolation and identification in human urine. Steroids, 36, 73-79. 

Breuer,H., Knuppen,R. and Pangels,G. (1962) Configuration of a fourth Kober-chromogen from the urine of pregnant women. Biochim. Biophys. Acta, 65, 1-12. 

Breuer,H., Knuppen,R. and Haupt,M. (1966) Metabolism of oestrone and estradiol-17β in human liver in vitro. Nature, 212, 76. 

Zucconi,G. Lisboa,B.P, Simonitsch,E., Roth,L., Hagen,A.A. and Diczfalusy,E. (1967) Isolation of 15α-hydroxy-oestriol from pregnancy urine and from the urine of new born infants. Acta Endocrinol., 56, 413-423. 

Hoppen,H.O., Bohlig,H.J., Werner,M. and Knuppen,R. (1976) Metabolism of oestrone and oestradiol in the isolated perfused rat liver. Hoppe-Seyler’s Z. Physiol. Chem., 357, S67-S74. 

Knuppen,R. and Breuer,H. (1962) Biogenesis of 11p-hydroxyoestrone and 16α-hydroxyoestrone by adrenal tissue. Biochim. Biophys. Acta, 58, 147-148. 

Adlercreutz,H., Ervast,H.-S., Tenhunen,A. and Tikkanen,M.J. (1973) Gas chromatographic and mass spectrometric studies on oestrogens in bile: Part I. Pregnant women. Acta Endocrinol., 73, 543-554. 

Loke,K.H. and Chong-Yong,G. (1968) 14a-Hydroxylation of oestrone by porcine adrenals. Steroids, 11, 863-875. 

Knuppen,R., Haupt,O. and Breuer,H. (1967) 14a-Hydroxyestrone, anew estrogen metabolite. Biochem. J., 105, 971. 

Adlercreutz,H. and Martin,F. (1976) Oestrogen in human pregnancy faeces. Acta Endocrinologica, 83, 410-19. 

Knuppen,R., Haupt,M. and Breuer,H. (1965) Formation of 15α-ydroxyoestradiol-17β and 18-hydroxyoestrone by human adrenal tissue. J. Endocrinol., 33, 529-530. 

Marriam,G.F., Watson,E.J.D. and Panattoni,M. (1957) The isolation of a ketonic dihydroxy Kober chromogen from the urine of pregnant women. Biochem. J., 65, 12-18. 

Marrian,G.F., Loke,K.H., Watson,E.J.D. and Panattoni,M. (1957) 16α-ydroxyoestrone in the urine of pregnant women. Biochem. J., 66, 60-65. 

Levitz,M., Spitzer,J.R. and Twombly,G.H. (1956) The conversion of estradiol-17β-16-C14 to radioactive 16-ketoestradiol-17β in man. J. Biol. Chem., 222, 981-990. 

Layne,D.S. and Marrian,G.F. (1958) Isolation of 16p-hydroxyoestrone from the urine of pregnant women. Nature, 182, 50-51. 

Gallagher,T.F., Kraychy,S., Fishman,J., Brown,J.B. and Marrian,G.F. (1958) A comparison of methods for the analysis of estrone, estradiol, and estriol in extracts of human urine. J. Biol. Chem., 233, 1093-1096. 

Doisy,E.A., Thayer,S.A., Levin,L. and Curtis,J.M. (1930) A new triatomic alcohol from the urine of pregnant women. Proc. Soc. Biol. Med., 28, 88-89. 

Migeon,C.J., Wall,PE. and Bertrand,J. (1959) Some aspects of the metabolism of 16-C14-estrone in normal individuals. J. Clin. Invest., 38, 619-629. 

Breuer,H., Knuppen,R. and Pangels,G. (1959) Stoffwechsel von 16-etooestron in menschlichen geweben. Acta Endocrinol., 30, 247-258.  

Marrian,G.F. (1930) Chemistry of estrin. IV. The chemical nature of crystalline preparation. Biochem. J., 24, 1021-1030. 

Marrian,G.F. and Bauld,W.S. (1955) The isolation of 16-epiestriol from the urine of pregnant women. Biochem. J., 59, 136-141. 

Adlercreutz,H., Diczfalusy,E. and Engstrom,L. (1960) Isolation and identification of bile estrogens in pregnant women. Endocrinology, 66, 80-88. 

Engel,L.L., Cameron,C.B., Stoffyn,A., Alexander,J.A., Klein,O. and Trofimow,N.D. (1961) The estimation of urinary metabolites of administered estradiol-17β-16-C14. Anal. Biochem., 2, 114-125. 

Adlercreutz,H., Ervast,H.S., Tenhunen,A. and Tikkanen,M.J. (1973) Gas chromatographic and mass spectrometric studies on oestrogens in bile: Part I. Pregnant women. Acta Endocrinol., 73, 543-554. 

Layne,D.S. and Marrian,G.F. (1958) The isolation of 16α-hydroxyestrone and 16-oxo-oestradiol-17β from the urine of pregnant women. Biochem. J., 70, 244-248. 

Breuer,H. and Pangels,G. (1960) Isolierung von 17-epi-Ostriol und 16,17-pi-Ostriol aus dem Urin schwangerer Frauen. Hoppe-Seyler’s Z. Physiol. Chem., 322, 177-183. 

Heard,R.D.H., Bauld,W.S. and Hoffman,M.M. (1941) Steroids. V. a-stradiol and progesterone metabolism. J. Biol. Chem., 141, 709-710. 

Adlercreutz,H. and Luukkainen,T. (1965) Gas-chromatographic studies on oestrogens in the bile of pregnant women. Biochim. Biophys. Acta, 97, 134-141. 

Loke,K.H., Marrian,G.F., Johnson,W.S., Meyer,W.L. and Cameron,D.D. (1958) Isolation and identification of 18-hydroxyoestrone from the urine of pregnant women. Biochim. Biophys. Acta, 28, 214. 

Loke,K.H., Marrian,G.F. and Watson,W.J.D. (1959) The isolation of a sixth Kober chromogen from the urine of pregnant women and its identification as 18-hydroxyoestrone. Biochem. J., 71, 43-48.