当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 内分泌进展 > 2005年 > 第2期 > 正文
编号:11168586
Steroid Sulfatase: Molecular Biology, Regulation, and Inhibition
http://www.100md.com 内分泌进展 2005年第2期
     Endocrinology and Metabolic Medicine and Sterix Ltd. (M.J.R., A.P., S.P.N.), Faculty of Medicine, Imperial College, St. Mary’s Hospital, London, W2 1NY, United Kingdom

    Medicinal Chemistry and Sterix Ltd. (L.W.L.W., B.V.L.P.), Department of Pharmacy and Pharmacology, University of Bath, Bath, BA2 7AY, United Kingdom

    Abstract

    Steroid sulfatase (STS) is responsible for the hydrolysis of aryl and alkyl steroid sulfates and therefore has a pivotal role in regulating the formation of biologically active steroids. The enzyme is widely distributed throughout the body, and its action is implicated in physiological processes and pathological conditions. The crystal structure of the enzyme has been resolved, but relatively little is known about what regulates its expression or activity. Research into the control and inhibition of this enzyme has been stimulated by its important role in supporting the growth of hormone-dependent tumors of the breast and prostate. STS is responsible for the hydrolysis of estrone sulfate and dehydroepiandrosterone sulfate to estrone and dehydroepiandrosterone, respectively, both of which can be converted to steroids with estrogenic properties (i.e., estradiol and androstenediol) that can stimulate tumor growth. STS expression is increased in breast tumors and has prognostic significance. The role of STS in supporting tumor growth prompted the development of potent STS inhibitors. Several steroidal and nonsteroidal STS inhibitors are now available, with the irreversible type of inhibitor having a phenol sulfamate ester as its active pharmacophore. One such inhibitor, 667 COUMATE, has now entered a phase I trial in postmenopausal women with breast cancer. The skin is also an important site of STS activity, and deficiency of this enzyme is associated with X-linked ichthyosis. STS may also be involved in regulating part of the immune response and some aspects of cognitive function. The development of potent STS inhibitors will allow investigation of the role of this enzyme in physiological and pathological processes.

    I. Introduction

    II. Molecular Biology of STS

    III. Localization of STS

    A. Immunocytochemical localization of STS

    B. Biochemical localization of STS

    IV. Regulation of STS Activity

    A. Cytokines and growth factors

    B. Steroids

    V. Biological Roles of STS

    A. In hormone-dependent breast cancer

    B. STS in skin

    C. STS in the immune system

    D. STS, neurofunction, and memory

    E. STS in reproductive tract tissues

    F. STS activity in osteoblast cells

    G. STS activity in leukocytes and thrombocytes

    VI. Tissue Availability of Steroid Sulfates

    VII. STS Inhibitors

    A. Alternative substrates

    B. Reversible inhibitors

    C. Irreversible inhibitors

    VIII. Active Pharmacophore Required for Potent Inhibition

    IX. In Vivo Activity of STS Inhibitors and Efficacy in Tumor Models

    X. Dual-Function Inhibitors

    A. Dual aromatase-sulfatase inhibitors

    B. STS and antiangiogenic microtubule disruptors

    C. STS and CA

    XI. Mechanism of Steroid Sulfate Hydrolysis and STS Inhibition

    XII. Future Perspectives

    I. Introduction

    AFTER MORE THAN a decade’s research to develop potent steroid sulfatase (STS) inhibitors, at least one inhibitor has now entered clinical trials to test its efficacy in postmenopausal women with breast cancer. It is therefore timely to review the role that this enzyme has in physiological and pathological conditions and to examine the rapid progress that has recently been made in developing potent STS inhibitors. STS (EC 3.1.6.2, aryl sulfatase C) is the enzyme responsible for the hydrolysis of alkyl [e.g., dehydroepiandrosterone (DHEA) sulfate (DHEAS)] and aryl steroid sulfates [e.g., estrone sulfate (E1S)] to their unconjugated forms. E1S was one of the first steroid conjugates to be isolated from the urine of pregnant mares in 1938 (1). In the past, steroid sulfates were generally considered to be end products of metabolism with their water solubility aiding excretion. However, during the last decade there has been a resurgence of interest in the roles that steroid sulfates, such as DHEAS and E1S, may have as precursors for the formation of biologically active hormones.

    A major impetus to the development of STS inhibitors was to identify new drugs for use in the treatment of hormone-dependent breast cancer. These tumors in postmenopausal women are initially treated with endocrine therapy, such as antiestrogens or, more recently, aromatase inhibitors. Many breast tumors will either fail to respond to such therapies or progress after a relatively short period of time, making it necessary to continue the search for new effective therapeutic agents. While the search for STS inhibitors was in progress, it became apparent that they may also have therapeutic applications in a number of other, nononcological conditions, including regulation of part of the immune response, dermatology, and cognitive function. In this paper, we review the recent advances that have been made in understanding the molecular biology and structure of the STS enzyme. The roles that STS may have in regulating the formation of biologically active hormones are also considered. The research leading to the development of potent STS inhibitors is discussed together with the potential therapeutic importance of this new class of drug.

    II. Molecular Biology of STS

    STS is a member of a superfamily of 12 different mammalian sulfatases (2, 3). The gene for the human STS is located on the distal short arm of the X-chromosome and maps to Xp22.3-Xpter; the gene is pseudoautosomal and escapes X-inactivation. On the Y-chromosome, there is a pseudogene for STS, which is transcriptionally inactive as the promoter, and several exons have been deleted. Sequence divergence has produced numerous stop codons in this pseudogene, and there are several large insertions. The extent of sequence similarity between the two genes suggests that they have been diverging for approximately 40 million years (4).

    The locus for the human STS gene on the X-chromo-some has been cloned, characterized, and sequenced (4) (GenBank accession no. M23945; Ensembl accession no. ENSG00000101846). The structure of the gene is shown in Fig. 1. The gene consists of 10 exons and spans 146 kb, with the intron sizes ranging from 102 bp up to 35 kb. Variable mRNA transcripts, detected by Northern blotting, are due to the use of alternative polyadenylation sites within exon 10 and are not thought to be caused by splice variants (3). The cDNA for STS has been cloned and sequenced (5, 6) (GenBank accession no. M16505 and J04964). It encodes a protein of 583 amino acids, with a signal peptide of 21–23 peptides and four potential glycosylation sites of which at least two are used, at asparagine residues 47 and 259.

    Inactivation of the STS gene results in X-linked ichthyosis (X-LI), one of the most prevalent human inborn errors of metabolism (7). In 80–90% of cases, the X-LI is due to complete deletions of the 146-kb STS gene and substantial flanking regions from the distal short arm of the X-chromosome. However, some patients have been identified who have normal hybridization patterns for genomic DNA and mRNA when probed with STS cDNA (7). Further investigation of these patients led to the identification of six point mutations within the coding sequence of the STS gene (8, 9). These mutations lead to the production of catalytically inactive STS. The loss of activity does not appear to be caused by incorrect localization or posttranslational modifications but may be due to a shortened half-life and/or loss of the substrate binding site (9). A further six mutations have now been identified, all of which lead to catalytically inactive STS (10, 11). To date, all the point mutations reported are located in the carboxyl region of the STS enzyme, which is thought to be important for substrate binding (12).

    So far, there has only been a very limited study of the molecular regulation of STS. The cytokines TNF and IL-6 both up-regulate STS enzyme activity in MCF-7 breast cancer cells. However, upon further investigation, this up-regulation appeared to be posttranslationally mediated rather than occurring via any changes in gene transcription or mRNA stability (13). The promoter region of the STS gene has been characterized, and some potential tissue-specific regulatory elements have been identified (14). The promoter is unusual, because it resembles neither a housekeeping gene nor a tightly regulated gene. It lacks a TATA box, is not GC rich, and lacks binding sites for Sp1 and other known transcription factors. The transcription start sites were mapped by primer extension and S1 nuclease protection assays. The major start site is at –221 with respect to the A nucleotide of the initiating methionine, with other minor transcription start sites mapped to –197, –206, and –241. The basal promoter region was identified as a 110-bp region from –192 to –302 using transient transfection reporter gene assays. Four other upstream regulatory elements (UREs) were identified: URE1 –305 to –572, URE2 –870 to –1086, and URE3 –1087 to –1253, which all act as enhancers. The three enhancer regions are counterbalanced by the presence of a negative regulatory region at –1253 to –1458. The basic promoter and URE activities could be detected only in the human choriocarcinoma JEG-3 cells, which are of placental origin and have high STS activity. Transfections of the basic promoter and UREs into COS-1, HeLa, and B82 cells gave no activity, suggesting that tissue-specific factors are required for activity of the STS promoter.

    III. Localization of STS

    STS activity was first demonstrated in rat liver microsomes by Dodgson et al. (15). Since then, it has been found in testis, ovary, adrenal glands, placenta, prostate, skin, brain, fetal lung, viscera, endometrium, peripheral blood lymphocytes, aorta, kidney, and bone. It is believed to be virtually ubiquitous in small quantities. The organ and tissue distribution varies considerably between different mammals. It is reported to be absent in the guinea pig and some marsupial livers and is undetectable in erythrocytes. The richest source of STS is the placenta. STS has been detected in various tissues by 1) immunohistochemistry, 2) biochemical analysis of hydrolytic products of various sulfated substrates (by colorimetric, fluorimetric, or radiometric methods), and 3) more recently in combination with mRNA expression levels using RT-PCR.

    A. Immunocytochemical localization of STS

    Using an azo-coupling histochemical method, Partanen (16) was unable to demonstrate the presence of STS in the epithelium of ducts and lobules of the normal breast, although activity was detectable in some samples of benign and malignant breast tissues. With the availability of purified preparations of STS (particularly human placental STS), anti-STS (polyclonal and monoclonal) antibodies were obtained, and specific immunohistochemical methods were developed to examine the subcellular localization of STS. In cultured human skin fibroblasts, STS was localized on the rough endoplasmic reticulum, Golgi cisternal, trans-Golgi reticulum, and, to a lesser extent, in plasma membranes and components of the endocytic pathway (i.e., coated pits, endosomes, and multivesticular endosomes). No STS immunostaining was detected in lysosomes (17). Immunohistochemistry with a monoclonal antibody to placental STS, combined with electron microscopy, also localized STS to the membranes of the endoplasmic reticulum, the nuclear envelope in rat hepatocytes, the proximal tubules in the kidney, and in the pineal gland, choroid plexus, and adenohypophysis of the rat brain (18). More recently, immunohistochemical evidence for the presence of STS has been obtained in the cytoplasm of ovarian clear cell adenocarcinomas (19), in glandular epithelial cells of the basilar layer of the endometrium but not the myometrium (20), and in vascular smooth muscle cells from the aorta (21). Immunohistochemistry has also been combined with RT-PCR to examine the localization and expression of STS in human fallopian tubes (22). STS was found to be localized in the secretory cells of fallopian tubes, and a higher number of positive cells were found in tissues obtained during the early luteal phase than in tissues collected during the follicular phase of the menstrual cycle. In agreement with these findings, abundant expression of STS mRNA was found in tissues from the early luteal phase (22). In another study, STS mRNA, enzyme activity, and immunoreactivity were assessed in normal human adult and fetal tissues (23). Amplified STS mRNA transcripts were weakly expressed in adult lung, aorta, liver, thyroid, testis, uterus, and all fetal tissues examined. Relatively high levels of STS activity were found in adult liver and the adrenal gland. The highest activity was detected in the placenta but, in keeping with the lower sensitivity of this technique, STS immunoreactivity was detected only in placental syncytiotrophoblasts. The same researchers detected STS immunoreactivity in breast carcinoma cells in 74% of cases, and this was significantly associated with its mRNA level and enzyme activity (24). An affinity-purified monoclonal antibody (KW 1049), raised against STS purified from human placenta that did not cross-react with arylsulfatases A or B, was used for the investigations in normal and malignant human tissues (23, 24).

    B. Biochemical localization of STS

    Historically, STS has been detected in microsomes or whole-tissue homogenates using biochemical or radiometric assays of substrate hydrolysis. For specific measurements of STS activity, [6,7-3H]E1S or [7-3H]DHEAS are used in buffer at pH 7.4 [based on Burstein and Dorfman (25)]. Phosphate buffer is preferred because it completely inhibits arylsulfatases A and B. The activities of both of these enzymes are relatively low at pH 7.4, which further improves the specificity of the assay. These assays have been used to identify and characterize STS activities in human leukocytes (26), brain (27, 28), osteoblast cell lines (29, 30), ovarian granulosa cells (31, 32), and rat testis (33).

    The central role of placental STS for the formation of estriol in the fetoplacental unit, its abundance in the placenta, and the virtual absence of detectable activity in cases of the inherited disorder of placental STS deficiency and recessive X-LI have led to the enzyme from human placenta being extensively investigated. Human placental STS has been purified to homogeneity and has been well characterized. Depending on the extent of glycosylation, the purified STS has a molecular mass of approximately 65 kDa (6). Whereas evidence from early investigations suggested that aryl sulfatase C and STS may have been different enzymes, biochemical and genetic analyses have confirmed that there is only one enzyme. Chromatography of placental microsomal extracts has revealed that both activities colocalize in the same fractions (34, 35, 36). Purified STS hydrolyzes aryl sulfates (e.g., p-nitrophenyl-sulfate, E1S) as well as alkyl sulfates (DHEAS, pregnenolone sulfate, deoxycorticosterone sulfate, cholesterol sulfate), and, to a lesser extent, iodothyronine sulfates (37, 38, 39). In addition, since the first observation by Jobsis et al. (40), that sons of women with sulfatase-deficient placentas develop X-LI, the link between the deficiency of microsomal STS and X-LI has been confirmed by several groups. In these subjects, enzyme activity toward both aryl- and alkyl-steroid sulfates was lacking in all tissues examined. In keeping with the lack of STS activity, plasma concentrations of all steroid sulfates are elevated. In subjects with STS deficiency, the activities of aryl sulfatases A and B are normal (41, 42, 43). Furthermore, when the cDNA for human placental STS was transfected into COS-1 cells, the expressed protein hydrolyzed aryl (E1S) and alkyl (DHEAS) steroid sulfates, with the hydrolysis of both substrates being blocked by a single inhibitor (44). Although there is only one gene for STS, some evidence has emerged that different isoforms of the enzyme may exist in rodents and humans. After the observation of Nelson et al. (45) that two isoforms may exist in mice, two isoforms (microsomal and nuclear) were shown to exist in rat liver and human placenta (46, 47). In humans, two isoforms (slow and fast) were identified in fibroblasts (48, 49, 50). It is possible that these isoforms are the result of posttranslational modifications. Hence, it is apparent from biochemical and immunohistochemical localization studies that STS is found mainly in target tissues of the reproductive tract (i.e., endometrium, ovarian, prostate, testis, placenta), the breast, skin, brain, bone, and blood. The biological role of STS in these tissues/organs is discussed in Section V.

    IV. Regulation of STS Activity

    The action of STS makes a major contribution to in situ estrogen production in hormone-dependent malignant tissues. Although expression of STS mRNA and STS activity is increased in malignant breast (and endometrial) tissues compared with nonmalignant tissues, little is known about the regulation of its expression or activity. Because the expression of other enzymes of steroidogenesis (such as aromatase) is known to be regulated by cytokines, growth factors, steroids, and prostaglandin E2, some of these factors have also been tested to assess whether they would induce STS.

    A. Cytokines and growth factors

    The cytokines IL-6 and TNF act synergistically to increase STS activity in breast cancer cells (51, 52). Furthermore, these cytokines increase STS activity without the use of promoter/enhancer elements, suggesting that the control of STS activity is via posttranslational modification of cysteine to formyl glycine in the active site or indirectly via changes in membrane fluidity or organic anion transporters, allowing increased uptake of the hydrophilic substrate (13). In contrast, the inflammatory cytokine IL-1? decreases the activity and expression of STS mRNA in human endometrial stromal cells, in a dose-dependent manner, and this effect is antagonized by the IL-1 receptor antagonist (53). IL-1? also suppressed STS activity and mRNA expression in vascular smooth muscle cells derived from human aortas (21). The presence of these cytokines in breast cyst fluid may explain the differential regulation of STS in breast cancer cell lines by breast cyst fluid (54). In a separate study, both basic fibroblast growth factor and IGF-I were found to increase STS activity in a dose- and time-dependent manner in MCF-7 and MDA-MB-231 breast cancer cells. This induction was inhibited by cycloheximide, indicating the requirement for new protein synthesis (55). These growth factors, which are thought to be secreted by breast tumors, may therefore increase local production of estrogens.

    B. Steroids

    Schneider et al. (56) first reported that in utero androgen exposure is required for induction of androgen-responsive hepatic STS in male rats. Lam and Polani (57) used exogenous testosterone treatment and concluded that STS induction is, in part, controlled by the male hormones in the mouse. Moutaouakkil et al. (58) observed that STS was highest in the uteri of pregnant guinea pigs compared with that in the uteri of fetal, castrated, or mature females, suggesting estrogenic regulation. The possibility of substrate induction of in vivo STS activity in liver and white blood cells in ovariectomized rats was confirmed by administration of exogenous E1S to ovariectomized rats (59). In contrast, a decrease in STS mRNA levels was found when MCF-7 breast cancer cells were treated with the progestagen Promegestone (R-5020) (60). However, exposure of MCF-7 and MDA-MB-231 breast cancer cells to the progestagen, medroxyprogesterone acetate, stimulated STS activity in these cells (61). Because medroxyprogesterone acetate is known to affect membrane fluidity, the enhanced STS activity might be explained by increased substrate availability from the medium. In addition, the availability of sulfated substrates may be increased by the induction of specific high-affinity transporters. It has also been reported that progesterone increased the uptake of inorganic sulfate in endometrial epithelial cells through induction of a high-affinity transport system (62). Whether progesterone or other steroids induce specific transporters for sulfated steroids in endometrial and/or other tissues remains to be explored. Recently, retinoids and 1,25-dihydroxy vitamin D3 have been reported to induce STS activity and expression in HL-60 promyelocytic cells (63). However, the molecular mechanisms underlying cytokine or steroid induction of STS activity and/or expression remains to be explored. Furthermore, factors governing the extent of posttranslational modification of cysteine-formyl glycine, glycosylation, and translocation to the endoplasmic reticulum are all likely to influence the activity of STS.

    V. Biological Roles of STS

    A. In hormone-dependent breast cancer

    1. Hydrolysis of E1S.

    Estrogens have a major role in supporting the development and growth of tumors in hormone-dependent tissues such as the breast and endometrium (64, 65). The highest incidence of breast cancer occurs in postmenopausal women after cessation of ovarian production of estrogens. However, estrogens continue to be produced in postmenopausal women by the peripheral conversion of androstenedione (Adione) to estrone (E1), a reaction mediated by the aromatase enzyme complex (66, 67). In postmenopausal women, the production rates for E1 and estradiol (E2) are approximately 40 μg/24 h and 6 μg/24 h, respectively (68). Much of the estrogens that are formed can be converted to estrogen sulfates by the actions of estrone sulfotransferase and phenol sulfotransferase (69, 70, 71, 72). Sulfation of estrogens changes them from being hydrophobic to hydrophilic molecules. In addition, because estrogen sulfates are unable to bind to the estrogen receptor (ER), they are biologically inactive. Circulating concentrations of E1S are much higher than that of the unconjugated estrogens (73, 74). Estrogen sulfates bind to albumin and have a prolonged half-life in blood (up to 9 h) compared with the much shorter half-lives of E1 and E2 (75). The high circulating concentrations of E1S together with its prolonged half-life have given rise to the view that E1S may act as a reservoir for the formation of biologically active estrogens via the action of STS (76, 77, 78, 79).

    In contrast to the low circulating levels of E1 and E2 in postmenopausal women, there is now general agreement that their concentrations are much higher in normal and malignant breast tissues (80, 81). Concentrations of E1 and E2 in malignant breast tissues can be up to 10-fold higher than those found in plasma. There is also evidence for high levels of E1S and estradiol sulfate (E2S) in breast tumors (74). Surprisingly, although plasma estrogen concentrations in postmenopausal women are much lower than in premenopausal women, breast tumor estrogen levels are similar in both groups of women (82, 83). The origin of estrogens in breast tumors has been the subject of intensive research during the last decade. There are two possible mechanisms that could account for this: 1) uptake from the circulation and binding with high affinity to ERs, or 2) in situ synthesis from estrogen precursors. Although uptake and binding to ERs may make an important contribution to tissue estrogen concentrations, the finding that levels are similar in ER-positive (ER+) and ER-negative (ER–) tumors suggests that local synthesis makes a major contribution to breast tumor estrogen concentrations (84, 85).

    Three enzyme systems are required for the formation of E2 from androgen precursors in breast tissues and include the aromatase, which converts Adione to E1 and 17?-hydroxysteroid dehydrogenase (17?HSD) type 1, which reduces E1 to E2, the biologically active estrogen that interacts with the ER. In addition, STS can act on E1S, formed as a result of sulfotransferase activity, to form E1, which can subsequently be converted to E2 (Fig. 2). All of these enzymes have been identified in malignant breast and endometrial tissues (86, 87). However, whereas aromatase activity is detected in only 40–60% of breast tumors, STS activity is present in most breast tumors (86, 88). Furthermore, the activity of STS is considerably higher than that of the aromatase enzyme in breast tumors (86). Using the appropriate substrate concentrations, it was found that as much as 10-fold more E1 could originate from E1S, via the sulfatase pathway, than from Adione by the aromatase route (89).

    Recently, real time RT-PCR techniques have been used to examine STS mRNA expression in breast tissues and to relate expression to a number of clinicopathological variables. Using this technique, it was shown that the level of STS mRNA expression in malignant breast tissue (1458 amol/mg RNA) was significantly higher than in normal tissue (536 amol/mg RNA) (90). This finding is consistent with the higher STS enzymatic activity that has been detected in malignant breast tissue (86, 91).

    STS mRNA expression was found to be an independent prognostic indicator in predicting relapse-free survival, with high levels of expression being associated with a poor prognosis (92). One possible explanation for this finding was suggested, i.e., in breast tissues expressing high levels of STS mRNA, tumor cells that escape surgical removal may grow very fast, and therefore patients may relapse earlier. Whereas previous investigations found no link between time to relapse and STS activity in breast tumors (93, 94), the original findings of Utsumi et al. (92) have now been confirmed in two further investigations (24, 95). In one study, it was found that the association between STS mRNA expression and prognosis applied only to ER+ tumors. Interestingly, high STS mRNA expression was associated with a poor prognosis in both pre- and postmenopausal women. This finding led to the suggestion that even in premenopausal women, intratumoral estrogen synthesis may play an important role in the growth of breast tumors. The role of aromatase mRNA expression analysis as a prognostic marker was also examined in view of the pivotal role that the enzyme is considered to have in regulating tumor estrogen synthesis. Aromatase mRNA expression was found to have no prognostic value, a finding consistent with previous studies that examined aromatase activity as a prognostic indicator (96, 97). The lack of prognostic value of aromatase mRNA determination led the authors to speculate that the sulfatase pathway may be more important than the aromatase route for intratumoral estrogen synthesis. STS mRNA expression was also found to correlate with tumor size and to be significantly higher in tumors with lymph node metastasis than in those without lymph node metastasis (24, 95). An examination of the intratumoral expression of genes from the estradiol metabolic pathway has provided further confirmation of the high expression and prognostic significance of STS mRNA expression (98).

    Immunohistochemistry and STS mRNA expression of laser-captured microdissected samples were also used to examine the location of STS within breast tumors (24). STS immunoreactivity was detected in the cytoplasm of cancer cells (Fig. 3) with STS mRNA expression being detected in microdissected carcinoma cells but not in stromal cells. This contrasts with reports as to the localization of the aromatase enzyme. Biochemical studies have consistently revealed higher aromatase activity in the stromal rather than the epithelial component of breast tumors (99). Immunohistochemical studies, however, have provided evidence for both an epithelial and stromal location for the aromatase enzyme complex (100, 101, 102).

    As previously discussed, when estrogen sulfates were first isolated it was thought that they represented the end products of metabolism, with sulfation rendering them water soluble. Many in vitro studies have now shown convincingly that E1S can be hydrolyzed by breast cancer cells, induce the production of estrogen-sensitive proteins such as pS2 and cathepsin D, and induce cell proliferation (103, 104). To examine whether E1S could support the growth of tumors in vivo, it was infused into rats bearing nitrosomethyl-urea (NMU)-induced mammary tumors (105). This animal model has been widely used to examine the effects of hormones on tumor growth. The tumors are hormone dependent and regress after ovariectomy but can be stimulated to regrow with estrogens. These tumors contain high levels of STS activity but are devoid of aromatase activity. Infusions of E1S at 300 pmol/h inhibited ovariectomy-induced tumor regression, whereas 3000 pmol/h stimulated tumor growth.

    Although this infusion study clearly demonstrated that NMU-induced mammary tumors in rats can be stimulated to grow by E1S, it did not differentiate between hydrolysis of E1S occurring in peripheral tissues, such as liver, and that occurring within the tumor. Two elegant studies have addressed this question, using the NMU-induced mammary tumor model or inoculation of MCF-7 breast cancer cells transfected with the STS cDNA. In the NMU model, a double-isotope infusion technique was used to determine the extent of in situ E1 formation from E1S in the tumor (106), based on a method that had previously been employed to measure the extent of formation of E1 from Adione in human breast tumors (107). For this, 14C-labeled E1 was infused into animals over a 3-d period to ensure that an isotopic steady state had been achieved. By measuring [14C]E1 levels in tumor tissue and blood, an index of the uptake of unconjugated E1 into the tumor can be calculated. By simultaneously infusing [3H]E1S, it is possible to calculate how much E1 is being formed within the tumor. Some of the infused [3H]E1S will be hydrolyzed in peripheral tissues, with some of the released [3H]E1 being taken up by the tumor. As uptake from the circulation can be calculated from the infusion of [14C]E1, any [3H]E1 in the tumor above that expected to be present due to uptake is considered to be formed by in situ synthesis. Using this technique, it was found that as much as 50% of the E1 formed within the tumor could originate from E1S.

    As an alternative approach to investigate the importance of in situ formation of unconjugated estrogen from estrogen sulfates, MCF-7 cells transfected with either a vector (MCF-7v) or vector containing the STS cDNA (MCF-7STS) were inoculated into the flanks of ovariectomized nude mice (108). The incidence of proliferating tumors in mice bearing MCF-7STS cells, supplemented with E2S (71%), was significantly higher than in animals bearing this cell line but not supplemented with E2S (22%). Supplementation with E2S and subsequent hepatic hydrolysis were not sufficient to stimulate the growth of MCF-7v cells. This finding demonstrates the importance of in situ estrogen synthesis, compared with that occurring in peripheral tissues, in supporting tumor growth. E2S was used for these studies because, unlike E1S, it does not require the liberated steroid to be reduced by estradiol dehydrogenase (type 1) before being biologically active.

    Interestingly, results from both in vitro and in vivo experiments have suggested the possibility that estrogen sulfates may have different biological activity than their unconjugated counterparts in cells expressing high STS activity. E2S was found to be more mitogenic than E2 in vitro producing a greater increase in anchorage-independent colony formation in the MCF-7STS clones (108). In vivo the volumes of tumors of animals supplemented with E2S (138 mm3) were greater than those in animals supplemented with E2 (51 mm3). One possible explanation for this observation is that some STS activity may reside in the nucleus (46). Evidence for a nuclear STS isozyme has been obtained, and it is possible that the formation of active estrogen by STS within the nucleus may not be subjected to the same degree of inactivation by 17?HSD type II or sulfotransferase before exerting their action.

    2. Hydrolysis of DHEAS.

    Evidence for the role that DHEAS, and its unconjugated metabolite DHEA, may have in breast cancer stems from two sources. First, steroid dynamic studies have revealed that these steroids can act as precursors for the formation of steroids with estrogenic properties, such as 5-androstenediol (Adiol). Second, studies in cells and animals have revealed that DHEAS, DHEA, and Adiol can stimulate the proliferation of breast cancer cells in vitro and induced mammary tumors in vivo. DHEAS is the most abundant steroid secreted by the adrenal cortex and, like estrogen sulfates, its half-life in plasma (10–20 h) is considerably longer than that of unconjugated DHEA (1–3 h) (109, 110). Isotopic infusion studies have revealed that, in women, as much as 75% of the daily production rate of DHEAS is converted to DHEA in peripheral tissues (111). After removal of the sulfate group by STS, the resulting DHEA can undergo reduction to Adiol, a steroid of particular importance with regard to breast cancer development. In postmenopausal women, the major proportion of Adiol formed is derived in peripheral tissues from DHEAS and DHEA (112). DHEAS can also be converted to Adiol-sulfate, but the contribution that this pathway makes to Adiol production remains to be resolved. Adiol, although an androgen, can bind to the ER with a somewhat lower affinity than that of E2. However, as the plasma concentrations of Adiol are at least 100-fold higher than those of E2 in postmenopausal women, it is considered to be equipotent with E2 as an estrogen in this group of women (113).

    It has been known for many years that Adiol can stimulate the growth of ER+ breast cancer cells in vitro (114, 115). In addition, in vivo studies employing 7,12-dimethylbenz[a]-anthracene-induced mammary tumors in rats revealed that Adiol could stimulate tumor growth (116). Importantly, the aromatase inhibitor 4-hydroxyandrostenedione did not block the ability of Adiol to stimulate tumor growth. This finding showed that Adiol did not need to be converted to an estrogen in order to stimulate tumor growth. More recent studies have revealed that DHEA and Adiol can directly activate the ER and stimulate the proliferation of breast cancer cells (117). Coincubation of these steroids with an aromatase inhibitor did not block their ability to activate the ER. Using a physiological concentration of DHEAS, mass spectrometry analysis has revealed that it can be converted to estrogens and Adiol in MCF-7 breast cancer cells (118).

    Further evidence for an important role of adrenal androgens and the sulfatase pathway in breast cancer was obtained from a study in which their effects on MCF-7 breast cancer cell proliferation were examined (119). DHEAS, DHEA, and Adiol were all found to stimulate cell proliferation, but their ability to do so was blocked by the ER antagonist nafoxidene, but not by aromatase inhibitors. In contrast, a potent STS inhibitor completely blocked that ability of DHEAS to stimulate cell growth. These results provide strong evidence that the stimulation of cell growth by DHEAS occurs via an aromatase-independent pathway that can be blocked by a STS inhibitor.

    There is, therefore, convincing evidence that adrenal androgens and their metabolites can stimulate breast cancer cell growth in vitro and induced mammary tumors in rodents. Convincing clinical evidence was obtained recently in support of a role for DHEAS in stimulating breast tumor growth in humans (120). In a study carried out to monitor serum DHEAS concentrations in women being treated with third-generation aromatase inhibitors, the important observation was made that, whereas those with stable disease had low (0.6 μM) levels of DHEAS, levels were elevated (3.8 μM) in women in whom tumor progression occurred. Serum levels of E1 and E2 in all subjects remained suppressed to minimal detectable levels. It was concluded from this study that, in patients with progressive disease, DHEAS appeared to stimulate tumor progression and led to the suggestion that this finding had serious implications for the use of aromatase inhibitors on their own. A likely explanation for this observation is that DHEAS is converted to DHEA by STS. The subsequent reduction of DHEA will yield a steroid, Adiol, for which there is now convincing evidence that it can stimulate breast cancer cell growth. Inhibition of STS, in addition to blocking the formation of E1 from E1S, should also reduce the production of Adiol, by blocking the conversion of DHEAS to DHEA (Fig. 2).

    B. STS in skin

    STS is also found in the epidermis, and there is increasing evidence that its action within skin may make an important contribution to androgen production in this tissue. It has been known for some time, since the description of a deficiency of STS in X-LI (121, 122), that STS has an important role in skin function. Clinically, X-LI is characterized by scaling of the skin with large, dark-brown scales and an increase in stratum corneum thickness (123). Lipids are important for normal stratum corneum structure and function and may be important for the process of normal desquamation. Concentration of cholesterol sulfate in stratum corneum, and the scales associated with X-LI, are increased (5-fold) compared with levels in stratum corneum from normal subjects (124). Because STS inhibitors currently in development could severely reduce STS activity in skin, it is reassuring to note that ichthyosis can be readily treated by the topical applications of keratolytic agents, such as ammonium lactate or cholesterol cream (125).

    Plasma concentrations of DHEAS can be increased in subjects with androgenic alopecia or hirsutism (126, 127). It is therefore possible that this steroid sulfate may be an important precursor for the formation of more active steroids within the skin. DHEAS can be converted to 5-dihydrotestosterone, the androgen that activates the androgen receptor, in axillary hair follicles (128). Using an immunohistochemical technique, STS was found to be predominantly expressed in the dermal papilla of hair follicles (129, 130). STS activity was also highest in the dermal papilla fraction of hair follicles. Its activity could be effectively inhibited with 1 nM of the potent STS inhibitor estrone-3-O-sulfamate (EMATE) (130). In patients with acne vulgaris, there is some evidence of increased STS immunoreactivity in affected skin areas (131). Thus, STS inhibitors may be of value in treating skin and/or hair conditions in which the action of the enzyme may be increasing local production of biologically active androgens.

    C. STS in the immune system

    Although DHEAS is secreted in large amounts by the adrenal cortex, it has remained controversial as to whether it has a specific biological role apart from serving as a precursor for the formation of active androgens and estrogens. Studies by Daynes et al. (132, 133) and Rook et al. (134) have suggested that DHEAS/DHEA may have an important role in regulating T-helper (Th) cell maturation. Th cells can progress to either a Th1 or Th2 phenotype, each of which secretes a characteristic profile of cytokines (e.g., Th1 cells secrete IL-2 and interferon-; Th2 cells secrete IL-6 and IL-10). The response of Th cells is mutually exclusive, with interferon- inhibiting the formation of Th2 cells and IL-10 inhibiting the formation of Th1 cells (135, 136).

    Plasma IL-6 concentrations were found to be elevated in elderly human subjects, reflecting the increased production of this cytokine by Th2 cells that occurs with aging. In aged mice, in which IL-6 plasma concentrations were also increased, it was possible to correct the elevated levels by the acute or chronic administration of DHEA or DHEAS (137). These studies also revealed that in vitro DHEA, but not DHEAS, was able to suppress the release of Th2 cytokines. Thus, STS, which is present in macrophages within the lymphoid tissues where Th cell maturation occurs and which converts DHEAS to DHEA, has a crucial role in regulating part of the immune response. From such investigations it has emerged that the balance of DHEA to glucocorticoid determines whether Th cells progress to either a Th1 or Th2 phenotype, i.e., DHEA favors development to Th1 cells whereas cortisol promotes a Th2 response.

    Using a contact sensitization model, convincing evidence has been obtained that in vivo DHEA and DHEAS have an immunostimulatory role (138). However, the ability of DHEAS, but not DHEA, to act as an immunostimulant was completely blocked by the coadministration of the potent STS inhibitor EMATE. Because a number of pathological conditions, such as rheumatoid arthritis, may result from an inappropriate immune response and increased production of Th1 cytokines, inhibition of STS could be of therapeutic benefit in such conditions. Using a collagen-induced model of arthritis, evidence has been obtained showing that the progression of arthritis was markedly altered by the STS inhibitor EMATE (139).

    The finding that DHEA has a role in regulating the Th1/Th2 immune response has provided an important insight into the regulation of estrogen synthesis in women (140, 141). IL-6 has a major role in regulating peripheral aromatase activity (142). It has been known for many years that the peripheral aromatase activity increases upon aging and is also higher in obese subjects (143). It is also known that levels of plasma IL-6 increase with aging and its production is increased in obese subjects (137, 144). It is well established that the production of DHEAS starts to decrease from the mid-20s (145). This reduction in the production of DHEAS will favor a Th2-type cytokine response with increased production of IL-6. Thus, increased production of IL-6 is the most likely explanation to account for the increase in aromatase activity detected in aging and obese subjects.

    D. STS, neurofunction, and memory

    In addition to being synthesized in the adrenal cortex, steroids such as DHEAS and DHEA are also formed in parts of the central nervous system and are therefore classified as neurosteroids (146, 147). These neurosteroids have important roles in regulating brain function. Sulfated steroids, e.g., DHEAS and pregnenolone sulfate, are considered to act as -aminobutyric acidA receptor antagonists, whereas their unconjugated analogs act as -aminobutyric acidA receptor agonists (148). In addition, both the sulfated and unsulfated forms of these steroids act positively to modulate N-methyl-D-aspartate receptor function (149).

    Because blood levels of DHEAS and DHEA decrease with aging, experiments were performed to examine the possibility that administration of these neurosteroids to rodents could improve memory. Intracerebroventricular or sc administration of DHEAS produced significant memory-enhancing effects in mice when tested using a foot-shock active avoidance training method (150, 151). Although there is convincing evidence that increasing blood levels of DHEAS and DHEA in rodents can result in memory-enhancing effects, it is not known whether such effects result from the sulfated or nonsulfated form of the neurosteroid. Because STS activity is present in brain tissues, it is possible that DHEAS could be hydrolyzed to DHEA by the action of this enzyme (152). With the advent of potent STS inhibitors, such as EMATE, it became possible to test whether the sulfated or unsulfated form of DHEA was responsible for the memory-enhancing effects of these neurosteroids (153). DHEAS is known to reverse scopolamine-induced amnesia in rodents. Blocking the hydrolysis of DHEAS with EMATE potentiated the ability of this sulfated neurosteroid to reverse scopolamine-induced amnesia. Similar results were obtained in this model using the nonsteroidal STS inhibitor (p-O-sulfamoyl)-N-tetradecanoyl tyramine (154). These findings strongly suggest that it is the sulfated form of DHEA that is responsible for the memory-enhancing effects of this steroid in rodents. Although results from these studies suggest that STS inhibitors may have a role in modulating the neuroexcitatory effects of steroid sulfates in rodents, there is, as yet, no information as to their possible affects in humans. In subjects with sulfatase deficiency, there is no evidence to suggest any abnormality in cognitive function. This suggests that the long-term therapeutic use of STS inhibitors should not have any adverse neurological effects. Whereas decreases in blood levels of DHEA and DHEAS occur in humans with aging, there is no evidence for a decrease in rodents. Therefore, experiments in rodents employing DHEA or DHEAS must be interpreted with caution.

    Inhibition of STS has recently been shown to increase aggressive behavior in CBA/H mice (155). Experimental evidence had previously indicated a possible link between attack behavior and the pseudoautosomal region of the Y-chromosome, which contains the sulfatase gene (156). The finding of a correlation between the initiation of aggressive behavior and liver STS activity in mice also suggests that the sulfatase gene could be a candidate for attack behavior in mice (157). Using a nonsteroidal inhibitor, a single oral dose was found to significantly inhibit brain STS activity and increase the effect of DHEAS on aggressive behavior in CBA/H mice (155).

    E. STS in reproductive tract tissues

    1. Female.

    STS activity has been detected in most tissues of the female reproductive tract. It is present in ovarian tissues from pre- and postmenopausal women, suggesting that in the ovary sulfated precursors, such as DHEAS, could be used as precursors for the formation of androgens and estrogens (158). Support for this concept was obtained from the finding that relatively high STS activity was detected in ovarian follicles, stroma, and corpus luteum, which were capable of utilizing DHEAS as a substrate for the production of DHEA, androstenedione, and testosterone (159). DHEAS is present in high concentrations in follicular fluid in close proximity to the ovarian cells involved in steroidogenesis (160). Using human granulosa cells obtained from women undergoing treatment for in vitro fertility, significant conversion of DHEAS to DHEA was detected, confirming the presence of STS activity in these cells (32). Such conversion was effectively inhibited by the STS inhibitor EMATE. Addition of DHEAS to cultured granulosa cells stimulated estrogen production in a dose-dependent manner, demonstrating that granulosa cells can utilize DHEAS as a substrate for estrogen production. STS is also expressed in human fallopian tubes, which are involved in gamete transport and fertilization (22). Expression of STS was higher in fallopian tubes obtained from the early luteal phase than from the follicular phase of the menstrual cycle.

    In addition to the role that steroid sulfates may have in breast cancer development, it is also likely that they may support the growth of hormone-dependent tumors in the reproductive tract of women. STS activity has been detected in normal and hyperplastic endometrial tissues (161). It has been suggested that uterine STS activity may have an important role in regulating the uterotropic activity of E2S (162). In a comparison of STS activities in malignant and normal endometrial tissues, activity was found to be 12-fold higher in malignant endometrial tissue (87). Sulfotransferase activity was also measured in this study and found to be significantly lower than STS activity, with no difference being detected between normal and malignant tissues. STS activity has also been detected in cultured cells derived from carcinomas of the ovary and vagina (163, 164). Using an immunohistochemical technique, positive STS expression was detected in 70% of ovarian clear cell adenocarcinoma tissue samples (19). Evidence showing that STS activity is present in hormone-sensitive tissues from the reproductive tract of women suggests that this enzyme may have an important role in regulating estrogen production in these tissues. With the development of potent STS inhibitors it will be possible to explore their therapeutic potential for the treatment of malignancies in the female reproductive tract.

    2. Male, including the prostate gland.

    STS is present in the testes of mammals, and it is likely that hydrolysis of steroid sulfates contributes to overall androgen production in this gland (165). An important role for STS has been postulated in the biochemical process of sperm maturation and capacitation (166). High levels of radiolabeled cholesterol sulfate are taken up by spermatozoa, and this was localized mainly within the plasma membrane of the acrosome region. It is thought that the cholesterol sulfate may act as a stabilizing factor that is associated with sperm membranes during transit or storage, inhibiting the release of acrosomal enzymes while sperm remain in the male reproductive tract. STS is present in the female reproductive tract, and hydrolysis of cholesterol sulfate by the STS may allow the release of acrosomal enzymes that facilitate penetration of the ovum by spermatozoa.

    In males, the prostate gland is likely to be the major peripheral site where STS activity makes an important contribution to the production of biologically active androgens. It has been known for many years that men who have been castrated as part of their treatment for prostate cancer can have a further period of remission after adrenalectomy (167). The reason for this is thought to be due to the production by the adrenal cortex of weak androgens, such as DHEAS, that can be converted to testosterone and dihydrotestosterone in prostatic tissues (168). More recently, the combination of castration or LHRH agonist with an antiandrogen has been shown to result in an improved therapeutic response in subjects with prostate cancer (169). Whereas castration/LHRH agonist treatment removes the testicular source of androgen, the use of an antiandrogen is thought to block the action of androgen derived from the adrenal cortex.

    STS activity has been detected in prostatic tissue (170). In studies in which the epithelial and stromal components of the prostate were separated, the highest STS activity was found to reside in the epithelial compartment (171, 172). LNCaP cells, which are derived from prostatic cancer, also possess STS activity, although at a somewhat lower level than that found in breast cancer cells (173). DHEAS was efficiently converted to DHEA in LNCaP cells, and hydrolysis of this steroid sulfate was almost completely blocked by the STS inhibitor, EMATE. The nonsteroidal inhibitor (p-O-sulfamoyl)-tetradecanoyl tyramine also inhibited the hydrolysis of DHEAS by these cells but was considerably less potent than EMATE. In view of the evidence supporting a role for STS in transforming weak androgens into biologically active androgen in the prostate, STS inhibitors could have considerable therapeutic potential for the treatment of prostate cancer.

    F. STS activity in osteoblast cells

    The reduction in ovarian estrogen production that occurs at the menopause has been implicated as an important factor in the development of osteoporosis. Because studies have generally failed to detect any consistent reduction in plasma estrogen concentrations in women with osteoporosis compared with women of similar age without fractures, the possibility of in situ estrogen synthesis by bone cells was considered (174). In three human osteoblast cell lines, HOS, MG 63, and U2 OS, the principal enzyme activities for estrogen synthesis, i.e., aromatase, 17?HSD type 1, and STS, were all detected (29). STS activity in the MG 63 osteoblasts was 1000-fold higher than aromatase activity, suggesting that the local formation of E1 from E1S could be an important source of estrogen for regulating bone formation. In similar investigations, HOS and MG 63 osteoblasts were shown to express STS mRNA and to be capable of utilizing both E1S and DHEAS as substrates for STS activity (30).

    In view of the potential importance of local formation of estrogens by osteoblasts for the maintenance of bone structure, it is possible that inhibition of STS could result in an increased rate of bone loss in treated subjects. However, studies with tibolone, which is used for hormone replacement therapy, have suggested that tibolone or its metabolites may have tissue-specific inhibitory effects on STS activity (175). After ingestion, tibolone is rapidly converted to metabolites that exert estrogenic effects or have progestogenic/androgenic properties (176). Although tibolone or its metabolites have estrogenic effects on bone and the central nervous system, no estrogenic stimulation of breast tissue occurs (177). In an attempt to find an explanation for these important effects, the ability of tibolone, its metabolites, or EMATE to inhibit STS activity in breast cancer cells, endometrial cells, or osteoblast cells was compared (175). All the compounds tested inhibited STS activity strongly in breast cancer cells and moderately in endometrial cells. In contrast, no significant inhibition of STS activity was detected in osteoblast cells. This study therefore raises the intriguing possibility that different tissues may express different isoforms of STS or may be subjected to different modes of regulation. Because STS inhibitors have now entered clinical trials, it will be important to confirm that compounds such as EMATE can act to inhibit STS activity in a tissue-specific manner.

    G. STS activity in leukocytes and thrombocytes

    In addition to the widespread distribution of STS in body tissues, the enzyme is also found in peripheral blood leukocytes (PBLs) and thrombocytes (26, 178). PBLs are capable of metabolizing steroid sulfates, and STS activity measurements, using these cells, have been used for the detection of the STS deficiency related to X-LI (179). Using [3H]E1S as a substrate, STS activity in PBLs from women in the follicular phase of their menstrual cycles was found to be almost twice as high as in cells collected from luteal phase subjects (26). This finding suggests that the high levels of progesterone present during the luteal phase may be involved in regulating STS activity. STS activity in PBLs obtained from men is lower than that in cells from female subjects (26, 180). Because many breast tumors are infiltrated by macrophages and lymphocytes (181), it is possible that the STS activity of these cells may make an important contribution to estrogen synthesis within tumors.

    The presence of STS activity in a readily available tissue, such as PBLs, suggested that these cells could be used to provide a relatively simple method to monitor the extent and duration of STS inhibition when these drugs became available. This contrasts with the complex double-isotope infusion technique that is currently used to monitor aromatase inhibition in postmenopausal women (182). The discovery of the first potent STS inhibitor, EMATE, led to a preclinical study to evaluate the use of measuring STS activity in PBLs to determine the effectiveness of this inhibitor (183). Two hours after the oral administration of EMATE to rats, the extent of STS inhibition was similar in PBLs and liver. STS activity measurements in PBLs were also used to monitor inhibition of this enzyme in a preliminary male volunteer study in two subjects receiving 0.5 mg/kg EMATE (183). Assays of STS activity in PBLs from these subjects revealed that inhibition was almost complete by 4 h after dosing and was maintained for at least 1 wk. The ability to monitor the extent and duration of STS inhibition should be of considerable value when carrying out clinical trials to test the efficacy of this new form of therapy.

    VI. Tissue Availability of Steroid Sulfates

    A central question with regard to the ability of steroid sulfates to exert physiological or pathological effects is whether they are taken up by cells, as such, or whether hydrolysis is a prerequisite for their entry into cells. It has generally been considered that although lipophilic unconjugated steroids are able to diffuse across cell membranes, polar hydrophilic steroid conjugates are unable to do so. Several studies have been carried out to investigate the uptake of steroid sulfates by cells and tissues using radiolabeled substrates, but these results have been difficult to interpret (103, 184).

    In the last few years, convincing evidence has emerged for the existence of a super family of membrane transporter proteins (185). Some of these proteins appear to be involved in the specific uptake of organic anions, such as steroid sulfates, which have a negative charge and hydrophobic backbone. These transporters include the organic anion transporter and organic anion transporter polypeptide (OATP) proteins. Oatp1 was first identified in rats as a multispecific, sodium ion-independent transporter for a range of xenobiotics, bile acids, and conjugated metabolites (186). Subsequently, other structurally related homologs of Oatp1, Oatp2, and Oatp3 were isolated (187, 188). In humans OATP was originally cloned from a human liver-derived cDNA library as a homolog of rat Oatp1 (189). A series of related homologs was subsequently identified in humans (OATP-B, OATP-C, OATP-D, and OATP-E), which were expressed at varying levels in a wide range of tissues (190). OATP-E was expressed in several different cancer cell lines, whereas OATP-D was not expressed in G1–101 breast cancer cells. Functional studies with human embryonic kidney (HEK)-293 cells transfected with the cDNAs for the different transporters revealed that whereas OATP-B, -C, -D, and -E all transported E1S, the highest activity was observed for OATP-B and OATP-C. Results from these studies show that OATPs are active transporters for E1S.

    Using an immunohistochemical technique, OATP-B was recently found to be highly expressed in the human mammary gland (191). Because the main substrate for OATP-B is E1S, it was suggested that the major physiological function of this carrier in the mammary gland is the uptake of E1S. In the normal mammary gland, OATP-B expression was confined to myoepithelial cells. Because these cells have been found to possess STS activity (192), it was postulated that the myoepithelial cells may be responsible for supplying nonsulfated estrogen to the adjacent epithelial cells. The major finding to emerge from this study was that OATP-B is strongly expressed in the majority of epithelial cells in invasive ductal carcinomas. In related in vitro studies, uptake of E1S and DHEAS by OATP-B was found to be stimulated by prostaglandin A1 and prostaglandin A2, suggesting that the uptake of steroid sulfates could be regulated locally at the plasma membrane.

    Because the plasma concentrations of E1S are much higher than those of unconjugated E1 or E2, the finding of a specific transporter for steroid sulfates in malignant breast tissues is of considerable importance. STS activity and expression, as previously discussed, are elevated in breast tumors. Thus, all the elements are in place in breast tumors to ensure the efficient uptake of E1S from the circulation and its rapid hydrolysis to a biologically active estrogen. This combination of an effective transporter for E1S and high STS activity offers a likely explanation for the high concentrations of E1 and E2 that are found in breast tumors.

    VII. STS Inhibitors

    In view of the important roles of STS in physiological and pathological conditions, considerable research has been carried out to develop potent inhibitors of this enzyme (78, 79, 80, 81, 193, 194, 195).

    A. Alternative substrates

    This type of compound, which contains at least one sulfate group in the structure, is designed to compete with E1S for binding to the STS enzyme active site and, as a consequence, impede the hydrolysis of the natural substrate to E1. These inhibitors, in principle, are alternative substrates for STS, the sulfate group(s) of which are expected to be hydrolyzed by the enzyme. The very first example of such a class of STS inhibitor was a series of 2-(hydroxyphenyl) indole sulfates, one of which (Fig. 4, compound 1) showed an IC50 value of 80 μM (196). Several synthetic and naturally occurring steroids were also investigated for STS-inhibitory activity, of which 5-androstene-3?,17?-diol-3-sulfate (Fig. 4, compound 2) was found to be the most potent [inhibition constant (Ki) = 2.0 μM (197)]. Flavonoids daidzein 4'-O-sulfate (Fig. 4, compound 4) and daidzein 4',7-di-O-sulfate (Fig. 4, compound 5) were synthesized and found to inhibit STS competitively with Ki values of 5.9 and 1 μM, respectively (198). However, inhibitors such as compounds 2, 4, and 5 could potentially be problematic because their corresponding metabolites, androstenediol (Fig. 4, compound 3) and daidzein (Fig. 4, compound 6), are known estrogens, which renders them of little value clinically for the treatment of hormone-dependent breast cancer.

    B. Reversible inhibitors

    The initial strategy employed for generating a lead STS inhibitor involved the replacement of the sulfate group (OSO3–) of E1S with surrogates or mimics such as phos-phate (199), phosphonates [-OP(=X)(OH)Me] (200, 201, 202, 203), sulfonates (-OSO2R) (204, 205), sodium methylenesulfonate (-CH2SO3–Na+) (205), sulfonyl halides (-SO2Cl and -SO2F) (206), sulfonamide (-SO2NH2) (206, 207), and the methylsulfonyl group (-SO2CH3) (206, 207). Most of these E1 derivatives were designed to compete with E1S for the enzyme active site but remain metabolically stable by not acting as substrates. Phosphate esters of p-acylphenols and p-alkylphenols were also prepared, and one derivative, n-lauroyl tyramine phosphate (Fig. 5, compound 7) inhibited STS with Ki values of 3.6 μM and 520 nM at pH 7.5 and 7.0, respectively (208).

    After the discovery of EMATE and as a result of the subsequent synthetic efforts that followed, its N-monomethyl- (Fig. 5, compound 8) and N,N-dimethyl (compound 9) derivatives were found to be weak reversible STS inhibitors (209, 210). Replacing the 3-O-atom of EMATE with other heteroatoms (Fig. 5; S, compound 10; and N, compound 11) gave analogs that were also weak reversible inhibitors of STS (211). Several estrone 3-amino derivatives (e.g., CF3CONH-E1) were prepared, but these were only weak inhibitors (212).

    Derivatives of 17-benzylestradiol (Fig. 5, compound 12) bearing a 4'-t-butyl (Fig. 5, compound 13), 3'-bromo (Fig. 5, compound 14), or 4'-benzyloxy (Fig. 5, compound 15) substituent were among the most potent reversible inhibitors reported to date, showing IC50 values (JEG-3 cells) between 22 and 28 nM (213, 214). It was found that compound 13 was about 7-fold weaker than EMATE as an STS inhibitor when tested in a transfected HEK-293 cell preparation (214). A series of 17-alkan- or alkynamide derivatives of E2 were prepared, and the propanamide 16 (Fig. 5) gave an IC50 of 80 nM in JEG-3 cells (215). The relatively high potency against STS observed for compound 16 is evidence of exploitation by the hydrophobic substituent of the hydrophobic binding area(s) that have been postulated to be in the vicinity of the D ring of EMATE.

    In an attempt to overcome the unwanted estrogenicity of some 17-substituted derivatives of EMATE (Fig. 6, compounds 36 and 37) (see Section VII.C), several sulfamates of C19 (androstene) or C21 (pregnene) derivative were prepared (Fig. 5, compounds 17–19) (216). 17-t-Butylbenzyl-5-androsten-17?-ol (compound 19) was the best reversible inhibitor (IC50 = 46 nM) in a homogenate preparation of HEK-293 cells and showed no estrogenic or androgenic activities in vitro (216).

    The sulfamate derivatives of (E)- and (Z)-4-hydroxytamoxifen (Fig. 5, compounds 20 and 21, respectively) were prepared and found to competitively inhibit STS in a rat liver microsome preparation with an apparent Ki of 35.9 μM for the (E)-isomer (compound 20) and more than 500 μM for the (Z)-isomer (compound 21) (217). It appears that their sulfamate group is not activated to inhibit the enzyme in an EMATE-like manner.

    Sulfamoyloxy-substituted 2-phenylindoles have recently been synthesized as antiestrogen-based STS inhibitors (218). Compounds 22 and 23 (Fig. 5) inhibited the conversion of E1S to E1 in MCF-7 breast cancer cells with IC50 values of 0.3 and 0.2 μM, respectively. Despite bearing a sulfamate moiety, the mechanism of action was not reported for this class of inhibitor.

    Recently, a series of thiosemicarbazone derivatives of madurahydroxylactone were studied, and the best agent, the cyclohexylthiosemicarbazone derivative (Fig. 5, compound 24), inhibited STS noncompetitively with a Ki value of 0.35 μM and an IC50 value of 460 nM in a placental microsome preparation (219). A series of nortropinyl-arylsulfonylurea derivatives were prepared, of which compound 25 (Fig. 5) inhibited STS in a purified enzyme assay with an IC50 value of 0.084 μM (cf. EMATE 0.056 μM) (220). Researchers at Bayer (221) identified from their compound library aryl piperazines 26 and 27 (Fig. 5), which inhibited STS in a STS protein preparation with IC50 values of 48 and 78 nM, respectively.

    In addition to those agents specifically designed to inhibit STS, other experimental or clinically used endocrine agents including danazol (61, 222), nomegestrol acetate (223), demegestone and chlormadinone acetate (224), ethinylestradiol (61), tibolone (Org OD14) and its metabolites (225), the "pure" antiestrogen ICI 164384 (226), tamoxifen and its metabolites (226), and pregnenolone 16-carbonitrile (227) have all been shown to exhibit STS-inhibitory activities.

    C. Irreversible inhibitors

    The bulk of STS inhibitors reported to date belong to this class of inhibitor. EMATE (Fig. 6), the very first inhibitor displaying such a mechanism of action, was originally designed to act as a surrogate of E1S. However, it was found to inhibit STS not only potently but also uniquely, in a time- and concentration-dependent manner, indicating that EMATE differed mechanistically from its contemporary E1S surrogates and acted as an irreversible active site-directed inhibitor (78, 209). Although EMATE is orally active and highly potent, it may not suitable for use as a therapeutic agent in the treatment of hormone-dependent breast cancer because it has been shown subsequently to be five times more estrogenic than ethinylestradiol when administered orally in the rat (228).

    1. Irreversible steroidal STS inhibitors.

    After the discovery of EMATE, all steroidal irreversible inhibitors that followed were analogs of EMATE designed to be less estrogenic than the parent, but to still possess similar or superior inhibitory activity to that of EMATE.

    The initial strategy was to introduce substituents such as 2-propenyl, n-propyl, nitro, methoxy, cyano, and halogens to the A ring of EMATE at the 2- and/or 4-positions (229). Analogs of EMATE with electron-withdrawing substituents on the A ring showed comparable or higher potency than EMATE in vitro (e.g., the 2-nitro derivative, compound 28, Fig. 6; IC50 in a placental microsomes preparation = 30 nM). In comparison, those analogs with bulkier aliphatic substituents were found to be weaker STS inhibitors. Overall, the most successful A ring-modified analogs of EMATE were 2-methoxyestrone 3-O-sulfamate (2-MeOEMATE, Fig. 6) and 2-methoxyestradiol-3,17?-bis-O,O-sulfamate (2-MeOE2bisMATE, Fig. 6), the IC50 values of which using placental microsomes were 30 nM (230) and 39 nM (231), respectively, indicating that these two derivatives were equipotent to EMATE in inhibiting STS in vitro.

    A recent paper (232) described the synthesis of a series of steroidal 2',3'- and 3',4'-oxathiazines as inhibitors of estrone sulfatase. The most active compound in the series and with reduced estrogenic activity was compound 29 (Fig. 6), which showed an IC50 value of 9 nM against the STS activity in an intact MCF-7 human breast cancer cells preparation. In vivo, this agent showed moderate antitumor activity against MCF-7 breast cancer xenografts in BALB/c athymic nude mice. The mechanism of action for compound 29 has not been reported. Although it could simply act as a reversible inhibitor of STS, it is possible that compound 29, a Schiff base, could be hydrolyzed/metabolized in situ to 2-formyl-EMATE (Fig. 6, compound 30), which then acts as the active species inactivating the enzyme in a similar manner to EMATE.

    In an attempt to understand the structure-activity relationships (SARs) for the sulfamate group of EMATE, two N-acylated derivatives were prepared of which N-acetyl-EMATE (Fig. 6, compound 31), but not the benzoyl derivative (Fig. 6, compound 32), inhibited STS irreversibly, albeit less potently and efficiently than EMATE (210).

    Because it is accessible to synthetic modifications, the D ring of EMATE has also been targeted for reducing the estrogenicity of the inhibitor. Early work had seen the reduction of the 17-carbonyl of EMATE to the methylene (CH2) derivative, NOMATE (Fig. 6), which was as potent as EMATE but less estrogenic (230). In contrast, the (E)-17-oximino derivative (Fig. 6, compound 33), inhibited STS equipotently to EMATE in vitro (>99% inhibition at 0.1 μM in MCF-7 breast cancer cells) and in vivo but stimulated uterine growth in ovariectomized rats about 1.5-fold greater than that achieved by EMATE, suggesting that this agent had enhanced estrogenicity (233).

    Introduction of hydrophobic substituents to the D ring of EMATE was shown to increase its potency and significantly reduce its estrogenicity. Therefore, 17?-(N-alkylcarbamoyl)estradiol-3-O-sulfamates and 17?-(N-alkanoyl)estradiol-3-O-sulfamates were highly potent STS inhibitors with optimal inhibitory activities shown by their respective n-heptyl derivatives (Fig. 6, compounds 34 and 35), both showing IC50 values of 0.4 nM in MDA-MB-231 cells (234). Using the estrogen-sensitive MCF-7 cell line, which proliferates upon stimulation by estrogens, no significant estrogenic potential of both inhibitors was observed at a concentration of 1 μM, a dose that was about 2000-fold higher than their IC50 values against STS (234). The hydrophobic side chains of these inhibitors were designed as a membrane insertion region and to help anchor the inhibitors in cellular membranes, where STS resides.

    It has already been illustrated that the introduction of hydrophobic substituents at the 17-position of estradiol led to derivatives that exhibited potent reversible STS inhibition (see Section VII.B). When two of these derivatives, namely 17-benzyl-E2 (Fig. 5, compound 12) and 17-4'-t-butylbenzyl-E2 (Fig. 5, compound 13) were sulfamoylated, their corresponding sulfamates (Fig. 6, compounds 36 and 37, respectively) were found to inhibit STS irreversibly in the same manner as EMATE. They were 5- to 14-fold more potent than EMATE in vitro at inhibiting the conversion of E1S to E1 in homogenates of HEK-293 cells transfected with STS (235). More significantly, sulfamates 36 and 37 were found to be nearly 600 times and 55 times more potent, respectively, than their corresponding phenolic parent compounds (Fig. 5, compounds 12 and 13) as STS inhibitors in the same enzyme preparation. These results clearly demonstrate that a sulfamate group is crucial for potent inactivation of STS. Because of the potency exhibited by compound 13 against STS in vitro, it has been reasoned that this phenolic steroid released after the inactivation of enzyme by the irreversible STS inhibitor (compound 37) would still exert reversible inhibition against any unreacted STS. Unfortunately, 17-4'-t-butylbenzyl-E2MATE (compound 37) was subsequently shown to be estrogenic in vivo, rendering this inhibitor unsuitable for further therapeutic exploitation (216, 236). On recognition that substitution at the 2-position of EMATE with a methoxy group abolished estrogenicity but retained potency against STS (230), 2-methoxy-17-benzylestradiol-3-O-sulfamate (Fig. 6, compound 38) and 2-methoxy-17-4'-t-butylbenzylestradiol-3-O-sulfamate (Fig. 6, compound 39) were prepared (237). As anticipated, both agents were found to be as potent as their corresponding 2-unsubstituted counterparts in the inhibition of STS activity in homogenates of transfected HEK-293 cells. In vivo, the 2-methoxy-17-benzyl-derivative (compound 38) showed no estrogenic activity in ovariectomized mice and efficiently blocked uterine growth induced by E1S.

    Continual investigation into 17-substituted analogs of estradiol as STS inhibitors gave two separate libraries of the N-derivatized 17-piperazinomethyl derivatives of estradiol and estradiol 3-O-sulfamate (238). The best STS inhibitors in both the phenol and sulfamate series (Fig. 6, compounds 40 and 41, respectively), as assayed in homogenates of HEK-293 cells transfected with STS, shared the same N-derivatization sequence, i.e., the secondary piperazino N-atom was acylated first with the amino acid phenylalanine, the primary amine of which was then amidated with 3-cyclopentyl propionic acid. At a 1-nM test concentration, sulfamate (compound 41) inhibited STS by 94%, which was close to that obtained with 17-benzyl-E2MATE (Fig. 6, compound 36) and 17-4'-t-butylbenzyl-E2MATE (Fig. 6, compound 37). The corresponding 3-hydroxy analog (compound 40) inhibited STS by 50% at 1 μM and hence was a weaker STS inhibitor than 17-benzyl-E2 (Fig. 5, compound 12) and 17-4'-t-butylbenzyl-E2 (Fig. 5, compound 13), which inhibited STS by 71 and 98%, respectively, at the same concentration.

    A novel D ring enlargement of EMATE led to a series of N-substituted piperidinedione derivatives. Two compounds, the N-(propyl) (Fig. 6, compound 42) and N-(1-pyridin-3-ylmethyl) (Fig. 6, compound 43) derivatives, showed exceptionally high potency, with both sharing the same IC50 value of 1 nM in a human placental microsome preparation (cf. EMATE, 18 nM) (256). The N-unsubstituted derivative (Fig. 6, compound 44) showed similar potency (IC50 = 20 nM) to EMATE, indicating that the six-membered piperidinedione ring is a good mimic of the D ring of EMATE. After an oral dose of 10 mg/kg/d for 5 d, compounds 42 and 43 were found to inhibit rat liver STS by 99% (239). Both compounds were devoid of estrogenic activity in the rat uterine weight gain assay.

    2. Irreversible nonsteroidal STS inhibitors.

    It has long been recognized that nonsteroidal agents themselves and their metabolites are less likely to exhibit unwanted endocrine effects in vivo than their steroidal counterparts. The initial development of nonsteroidal irreversible STS inhibitors resulted in the preparation of the A/B ring mimic of EMATE, tetrahydronaphthalene 2-O-sulfamate (Fig. 7, compound 45), and the mono- (Fig. 7, compound 46) and bis-sulfamate (Fig. 7, compound 47) derivatives of diethylstilbestrol. Whereas the bicyclic sulfamate (compound 45) was a much weaker inhibitor than EMATE (209), diethylstilbestrol bis-sulfamate (compound 47) was a moderate STS inhibitor (IC50 = 10 nM, MCF-7 cells) (79, 230). This finding indicated that it is not necessary to have a fused ring system for STS inhibition.

    The finding that n-lauroyl tyramine phosphate (Fig. 5, compound 7; see Section VII.B) was a significantly reversible STS inhibitor led to the development of a series of (p-O-sulfamoyl)-N-alkanoyl tyramines, of which the N-tetradecanoyl derivative (Fig. 7, compound 48) was found to have an IC50 value of 55.8 nM against STS derived from human placental microsomes (240). The IC50 value for this agent in MDA-MB-231 human breast cancer cells was found to be 350 nM, and the inhibition of STS was irreversible (241). When this agent was studied for inhibition of the proliferation of MCF-7 human breast cancer cells in the presence of E1S at 1 μM, the IC50 value obtained was 38 nM. It was postulated that the phenyl ring of these tyramine sulfamates mimics the A ring of EMATE, whereas the amido functionality, which was shown to be crucial for inhibition by this class of inhibitors, participates in essential hydrogen bonding(s) (242). The optimal distance between the p-sulfamoyloxyphenyl ring and the amido group was found to be one to two methylene units (243). Analogs of this class of STS inhibitor (Fig. 7, compounds 49–52) were prepared, but they were only weak inhibitors with IC50 values against the STS activity in placental microsomes ranging from 10–40 μM (230). In vivo, whereas E-capsaicin sulfamate (compound 52) showed a modest degree of STS inhibition in rat uterine and liver tissues, other analogs were inactive (230).

    Simple p-sulfamates of benzoic acid esters, substituted benzene, and phenyl ketones have also been pursued as STS inhibitors. One compound, the cyclooctyl derivative (Fig. 7, compound 53) was claimed to be the most active with an IC50 value of 170 nM in a placental microsome preparation (cf. 500 nM for EMATE) (244). However, this IC50 value of EMATE reported is much higher than values (18–80 nM) obtained by other groups using similar placental microsome preparations (209, 219, 245).

    A series of bicyclic coumarin sulfamates was synthesized as alternative A/B ring mimics of EMATE, of which 4-methylcoumarin 7-O-sulfamate (COUMATE, Fig. 7) showed an IC50 of 380 nM in an MCF-7 cell preparation (246), about 3-fold more potent than the bicyclic compound 45 (Fig. 7), but was still much weaker than EMATE (IC50 = 65 pM in the same assay). This finding, for the first time, highlighted a relationship between the pKa value (or leaving group ability) of a parent phenol and the inhibitory activity of its sulfamate. For coumarin sulfamates, it was reasoned that the extended conjugation of the coumarin core structure, as a result of its ,?-unsaturated lactone over the saturated cyclic hydrocarbon of tetrahydronaphthalene, enhances the overall potency of a coumarin sulfamate by virtue of lowering the relative pKa of the leaving phenol released during enzyme inactivation (246). This SAR was supported by the weaker inhibition exhibited by analogs of COUMATE in which 1) the sulfamate group was relocated to the 6-position of the ring or 2) the extended conjugation in the coumarin motif was disrupted by either the reduction of the double bond or replacement of the carbonyl group with a methylene group (247).

    Further extension of the coumarin sulfamate series has established that derivatives with hydrophobic substituents introduced at the 3- and/or 4-positions of COUMATE were more potent STS inhibitors (245, 247). Hence, a series of tricyclic coumarin sulfamates was developed, of which 667COUMATE (Fig. 7) was found to inhibit STS in a placental microsome preparation with an IC50 value of 8 nM, some 3-fold more potent than EMATE (245, 248). The apparent Ki value for 667COUMATE was found to be 40 nM, which was significantly lower than that for EMATE (670 nM) (245). This finding suggests that the lower IC50 value observed for this nonsteroidal inhibitor in comparison with EMATE could be attributed to a higher affinity of 667COUMATE for the enzyme active site in addition to an enhanced "sulfamoylation potential" of 667COUMATE as a result of the better leaving ability of its phenolic coumarin precursor (Fig. 7, compound 54) (pKa 8.5 for compound 54 vs. 10 for E1).

    Upon extension of the tricyclic coumarin sulfamates series, it was shown that the in vitro inhibitory activity was the highest with 6610COUMATE (Fig. 7) (IC50 = 1 nM, placental microsomes) (249). In vivo, this compound was found to be marginally more potent than 667COUMATE in inhibiting the STS activity in rat liver tissues (89 vs. 86%, at 1 mg/kg, per os). Surprisingly, the analog 6615COUMATE (Fig. 7), the IC50 value in vitro of which was 370-fold higher than that of 6610COUMATE, was the most potent compound of the series in vivo (94%, at 1 mg/kg, per os) (249).

    Because of the structural resemblance of the naturally occurring flavonoids to estrogens, sulfamates of flavones, isoflavones, and flavanones were prepared, and compounds such as compounds 55–57 (Fig. 7) were shown to be moderate STS inhibitors (230, 250). Additional exploitation of this class of compounds led to a series of chromenone- and thiochromenone-based sulfamates (118, 251) of which compound 58 showed an IC50 value of 0.34 nM in purified STS, rendering this compound about 170-fold superior to EMATE (IC50 = 56 nM from the same assay). However, compound 58 was estrogenic, stimulating the growth of MCF-7 breast cancer cells by 99% at 100 nM (252).

    Because 17-4'-t-butylbenzyl-E2MATE (Fig. 6, compound 37) was a highly potent STS inhibitor, a series of structurally related 4-substituted monoaryl sulfamates were designed as nonsteroidal mimics of sulfamate (compound 37). The optimal inhibitor of the series, compound 59 (IC50 = 0.4 nM) was more potent than EMATE (IC50 = 0.9 nM) in homogenates of HEK-293 cells transfected with STS (236).

    It has been observed that 4,4'-benzophenone-O,O'-bissulfamate (BENZOMATE, Fig. 7, compound 60) was a potent inhibitor (IC50 = 190 nM, cf. 56 nM for EMATE) against recombinant human STS (253). SAR studies have shown that the bis-sulfamate motif is crucial to the high potency because the monosulfamate derivative, i.e., benzophenone 4-sulfamate (Fig. 7, compound 61) was about 25 times less active. These results have been confirmed independently and expanded in a recent report (254).

    Some sulfamic acid biphenyl esters were also reported to be STS inhibitors. For example, propyl 4'-sulfamoyloxy-biphenyl-4-carboxylate (Fig. 7, compound 62) (255) inhibited STS in a human placental microsome preparation with an IC50 value of 3.5 μM, suggesting that this class of inhibitor inhibited STS weakly.

    VIII. Active Pharmacophore Required for Potent Inhibition

    All irreversible STS inhibitors reported to date share a common pharmacophore, i.e., a phenol sulfamate ester (246). The potency of this class of inhibitor is apparently increased by substituent(s) that exploit favorable hydrophobic interactions with the enzyme active site (234, 235, 240, 245, 246, 256) and/or enhance the "sulfamoylation potential" of the sulfamate group (229, 246, 257). However, it is important to note that the presence of a phenol sulfamate ester moiety does not always result in an irreversible inhibitor. Despite the higher affinity of the sulfamate derivative of (E)-4-hydroxytamoxifen (Fig. 5, compound 20) for STS, this compound only inhibited the enzyme competitively (217), suggesting that its sulfamate group was not activated to inhibit the enzyme in an EMATE-like manner.

    The phenolic component of the pharmacophore is crucial for high potency because sulfamate esters of aliphatic alcohols are weak STS inhibitors. Hence, alkyl O-sulfamates (258) and 3?-sulfamate derivatives of the aliphatic C19 and C21 steroids, such as dehydroepiandrosterone 3?-O-sulfamate (79), and compounds 17–19 (Fig. 5) (216) are significantly weaker STS inhibitors than EMATE and its congeners. The lack of potent inhibitory activity observed for these aliphatic compounds, despite the presence of a sulfamate group, could be attributed to the fact that their sulfamate group is not activated in the same manner as those of EMATE-like phenol sulfamate esters. The poorer leaving ability of their parent aliphatic alcohols (pKa > 16 for most primary and secondary alcohols vs. pKa 10 for phenol) in effect precludes the cleavage of the S-O bond of the sulfamate group, which is thought to be crucial mechanistically for the inactivation of the enzyme by EMATE-like inhibitors via sulfamoylation (i.e., enzyme-SO2NH2).

    Although nonsulfamoylated phenolic compounds such as compounds 12 and 13 (Fig. 5) were good reversible inhibitors of STS in vitro, the sulfamate (Fig. 6, compound 37) was a time-dependent inactivator with superior potency to its phenolic counterpart (compound 13) (see Section VII.C) (235). The N-monomethyl (e.g., Fig. 5, compound 8) and N,N-dimethyl derivatives of EMATE (e.g., Fig. 5, compound 9) were weak reversible inhibitors of STS although N-acetyl-EMATE (Fig. 6, compound 31), but not the benzoyl derivative (Fig. 6, compound 32) (209, 210), inhibited the enzyme irreversibly, albeit less potently than EMATE. Analogs of EMATE in which the 3-O-atom was replaced by other heteroatoms (Fig. 5; S, compound 10; and NH, compound 11) were only weak reversible inhibitors of STS (211). All these findings demonstrate that a phenol sulfamate ester with no substitutions at the N-atom (i.e., H2NSO2O-Ar) is the prerequisite for highly potent irreversible STS inhibition.

    The pharmacophore for reversible STS inhibitors has not yet been established. It is anticipated that the SAR studies of several new classes of reversible inhibitors will provide this information in the future. The activity of this class of inhibitor is presumably dependent upon the inhibitor acting so as to mimic the steroid structure, in a similar way to the prototype danazol.

    IX. In Vivo Activity of STS Inhibitors and Efficacy in Tumor Models

    Danazol, a synthetic derivative of 17-ethinyl testosterone, was one of the first compounds shown to be active in vivo as a STS inhibitor. This compound is used in the treatment of endometriosis and was found to increase the ratio of DHEAS to DHEA in plasma, suggesting that it was acting as a STS inhibitor (222). Studies with the first compound that was specifically designed and synthesized as a STS inhibitor, estrone-methylthiophosphonate (E1-MPT, see Section VII.B), revealed that it could reduce plasma E2 levels in rats, indicating that it was active in vivo (79). Because danazol and E1-MPT are both reversible inhibitors, their effects on in vivo STS activity can only be measured indirectly via changes in circulating hormone levels.

    The identification of EMATE as the first potent STS inhibitor in in vitro screening assays was rapidly followed with studies to test its efficacy in vivo as an inhibitor. Daily sc administration of EMATE (10 mg/kg) to rats was found to almost completely (>99%) inhibit the ability of liver STS to hydrolyze both E1S and DHEAS (259). To examine the duration of STS inhibition, EMATE was administered as a single 10 mg/kg sc dose or daily at this level for 7 d. A single dose was found to inhibit STS activity in liver, brain, adrenal, uterine, and ovarian tissues for up to 3 d with only 10–15% recovery being detected 7 d after dosing. Multiple dosing with EMATE resulted in complete inhibition of STS activity in tissues for up to 10 d after the end of dosing. EMATE was also found to be active as a STS inhibitor after oral dosing. The reason for the long duration of steroid-based STS inhibitors, such as EMATE, remains unclear. The half-life of STS has been reported to be of the order of 4 d, but as STS activity continues to be inhibited by EMATE for a much greater length of time, it is likely that EMATE forms a depot in tissues, with its sustained release contributing to its prolonged inhibition of STS activity. Alternatively, EMATE and related sulfamate-based inhibitors may have a long half-life in blood. The NMU-induced mammary tumor model in rats was used to test the ability of EMATE to inhibit the growth of E1S-stimulated tumors in ovariectomized rats. Over a 12-d period, EMATE (10 mg/kg·d) induced significant tumor regression and also completely inhibited tumor STS activity (259). Recently, 2-methoxy derivatives of EMATE have also been shown to be equipotent to EMATE as STS inhibitors in in vivo studies (231).

    Similar studies have also been carried out to assess the in vivo potency of COUMATE and a series of tricyclic coumarin sulfamates, including 667COUMATE. Daily dosing with the two-ring coumarin sulfamate, COUMATE, at 10 mg/kg, resulted in only 85% inhibition of STS activity with full restoration of STS activity occurring by 7 d (260). Thus, COUMATE was less active in vivo than EMATE. The development of a series of tricyclic coumarin sulfamates, including 667- and 6615-COUMATE, resulted in the identification of nonsteroidal STS inhibitors that were equipotent in vivo with EMATE (248, 249, 261). In the NMU-induced mammary tumor model, 667COUMATE caused significant regression of E1S-stimulated tumor growth at 10 mg/kg (85 ± 5%) and at 2 mg/kg (56 ± 13%). These in vivo studies in ovariectomized rats confirmed that 667COUMATE was not estrogenic and, together with its ability to inhibit E1S-stimulated mammary tumor growth in rats, led to this compound being selected for therapeutic development.

    X. Dual-Function Inhibitors

    A. Dual aromatase-sulfatase inhibitors

    As noted earlier, there are two pathways by which estrogens can be synthesized in postmenopausal women, i.e., the aromatase and sulfatase routes. If STS inhibitors prove to have clinical efficacy in women with breast cancer, it would be logical as a next step to test them in combination with an aromatase inhibitor. Such an approach would maximize the deprivation of estrone and steroids such as Adiol, and could enhance the response rates to this form of endocrine inhibitor therapy. Although it would be possible to administer aromatase and STS inhibitors as separate agents, an alternative approach would be to develop single molecule compounds with dual aromatase-sulfatase inhibitor (DASI) properties.

    Initial studies in this area took advantage of the fact that a number of flavonoids can inhibit aromatase activity (262, 263). It was reasoned that sulfamoylation of this class of compound could give rise to molecules with DASI properties. Sulfamoylation of 4'-hydroxy and 4',7-dihydroxyisoflavone to give the 4'-mono- and 4',7-bis-sulfamates revealed that these compounds could inhibit STS activity in vitro. At 1 μM, when tested using intact MCF-7 breast cancer cells, the mono- and bis-sulfamate derivatives inhibited STS activity by 83 and 90%, respectively (230). Both compounds were active in vivo as STS inhibitors but were considerably less potent than EMATE. After a single 10 mg/kg dose per os to adult female Wistar rats, liver STS activity was inhibited by 62 and 81% by the mono- and bis-sulfamate derivatives, respectively. Although this study indicated that it should be possible to employ a compound with aromatase-inhibitory properties to develop a DASI, it was apparent that if this research was to lead to a therapeutic agent, it would be essential to use a molecule with more potent aromatase-inhibitory properties than the isoflavones.

    A more recent approach has been to sulfamoylate a number of third-generation, nonsteroidal, aromatase inhibitors (257, 264). These inhibitors contain a triazole ring that coordinates reversibly to the heme iron of the aromatase. This class of aromatase inhibitors is reversible, in contrast to the steroid-based inhibitors, such as exemestane, which act as irreversible inactivators. Thus, the incorporation of the active pharmacophore required for STS inhibition, i.e., a phenol sulfamate ester, to a triazole-containing aromatase inhibitor, would give rise to an irreversible STS inhibitor, but reversible aromatase inhibitor.

    To test the validity of this concept, the third-generation aromatase inhibitor YM 511 was chosen for initial sulfamoylation (265). YM 511 is a selective, potent aromatase inhibitor, which was recently reported to give an objective response rate of 20.4% when tested in a phase II trial (266). Using JEG-3 choriocarcinoma cells, which possess both STS and aromatase activities, the IC50 value for the inhibition of aromatase activity by YM 511 was 0.5 nM, whereas this compound was inactive against the STS enzyme (257). Synthesis of a p-sulfamoyloxybenzyl derivative of YM 511 yielded a compound that had both moderate aromatase- and STS-inhibitory properties (IC50 values: arom = 100 nM; STS = 227 nM). The incorporation of a halogen to give the m-bromo derivative of this compound significantly increased both its aromatase- and STS-inhibitory properties (IC50 values: arom = 0.82 nM; STS = 39 nM) with a potency against aromatase in the same order of magnitude as YM 511. In vivo in intact rats, using the pregnant mares serum gonadotropin-stimulated ovarian aromatase model, the m-bromo-p-sulfamoyloxybenzyl derivative of YM 511 gave 68% inhibition of aromatase and almost complete (>98%) inhibition of STS activity.

    Thus, these studies have revealed that it will be possible to engineer single molecules that possess both potent aromatase- and STS-inhibitory properties. The development of this class of DASI could offer considerable therapeutic advantage for the treatment of hormone-dependent breast cancer over the use of either an aromatase or STS inhibitor alone.

    B. STS and antiangiogenic microtubule disruptors

    The finding that EMATE, as discussed previously, proved to be a potent estrogen in rodents made it unsuitable for development as an anticancer agent for postmenopausal women. In an attempt to reduce the estrogenicity of EMATE, while retaining the potent STS inhibitory properties associated with this type of molecule, a number of modifications were made to the A ring of the steroid nucleus (229). It had been shown previously that substitution of the aromatic A ring at C-2 and/or C-4 of the steroid nucleus by nitro-, n-propyl, or allyl groups greatly reduces their estrogenicity compared with the parent compound (267, 268). Whereas the addition of a 4-allyl, 4-n-propyl or 4-nitro group to EMATE resulted in derivatives that were active as STS inhibitors in vivo, the 4-nitro EMATE retained some estrogenicity.

    A series of sulfamoylated derivatives of 2-methoxyestrone and 2-methoxyestradiol (2-MeOE2) were also synthesized and tested (229, 269). 2-MeOE2, a natural endogenous estrogen metabolite, had previously been shown to be cytotoxic to MCF-7 breast cancer cells when tested at relatively high concentrations (270). There is currently considerable interest in the use of 2-MeOE2 for cancer therapy (271). Production of 2-MeOE2 appears to be increased in women at low risk of breast cancer, and it has been suggested that it acts as the body’s natural antimitotic metabolite (272, 273). 2-MeOE2 inhibits the proliferation of a wide range of ER+/ER– breast cancer cells. At relatively high doses it is also active in vivo against transplanted Meth-A sarcomas and B16 melanomas in C3H mice and human MDA-MB-435 (ER–) melanoma cells in mice (274, 275). In addition to its antiproliferative effects, 2-MeOE2 is also a potent inhibitor of angiogenesis in vitro and in vivo (274, 275).

    The 2-methoxyestrogen sulfamate derivatives retained potent STS-inhibitory properties. A single oral dose of 10 mg/kg of 2-methoxyestrone 3-O-sulfamate or 2-MeOE2-bisMATE inhibited rat liver STS by more than 90% (229, 264). Additional investigations with the 2-methoxyestrogen sulfamate derivatives revealed that, like 2-MeOE2, they inhibited the proliferation of ER+/ER– breast cancer cells, being considerably more potent than 2-MeOE2 (276, 277). They induced a G2-M cell cycle arrest and induced cells to undergo apoptosis. In contrast to 2-MeOE2, which induces a reversible G2-M arrest, the cell cycle arrest induced by 2-methoxyestrogen sulfamate derivatives was irreversible. In vivo 2-MeOE2-bisMATE (20 mg/kg·d per os) for 28 d almost completely inhibited the growth of tumor xenografts derived from MDA-MB-435 (ER–) melanoma cells in nude mice (278). The sulfamoylated derivatives of 2-MeOE2 have a superior bioavailability and pharmacokinetic profile to that of 2-MeOE2 itself (278). This most likely results from their ability to bind to carbonic anhydrase (CA)II in red blood cells (rbcs) (see Section X.C) and undergo liver transit without first-pass metabolism (279).

    Like 2-MeOE2, the sulfamoylated derivatives are also thought to act by binding to the colchicine site on tubulin. They also effectively inhibit the paclitaxel-induced polymerization of tubulin, suggesting that they act to inhibit microtubule dynamics (277). 2-MeOE2 has a very low affinity for ER/ER?, and it is most likely that it acts via a receptor-independent mechanism to inhibit cell proliferation (280). Recently, 2-MeOE2 was shown to inhibit tumor growth and angiogenesis by reducing the expression of hypoxia-inducible factor-1 (281).

    Thus, this class of 2-methoxyestrogen sulfamates are potent STS inhibitors. In addition, they act to disrupt microtubules and inhibit glucose uptake (282) and are potent angiogenesis inhibitors (283). Attacking tumor growth at multiple points may offer considerable therapeutic advantage over drugs that are only active against a single target.

    C. STS and CA

    During early preclinical evaluation, EMATE was found to inhibit STS activity in rats after oral administration (259). This was unexpected, because derivatives of natural estrogens are usually rapidly inactivated after oral ingestion during their first transit through the liver (284). To date, this problem has been overcome either by preventing metabolic inactivation, e.g., by the introduction of a 17-ethinyl group, or by administering large doses of estrogen. Both approaches result in an increase in the exposure of liver tissues to estrogens that can have a number of adverse effects, including an increase in the production of clotting factors (285). The preclinical development of EMATE, as a STS inhibitor, was halted when it was discovered that its estradiol analog was five times more potent than ethinylestradiol on oral application to rats (228). This finding appeared to render EMATE unsuitable for development for breast cancer therapy in which complete deprivation of estrogen is required. It was subsequently revealed that earlier studies with N,N-dimethylated sulfamoylated derivatives of estrogens had also shown potent estrogenic properties in rodents but that metabolites of such compounds could accumulate in rbcs. Recent studies have confirmed that estradiol-3-O-sulfamate, the C17-reduced form of EMATE, is also taken up by rbcs and 30 min after administration to rats 98% of the dose is present in rbcs (286). Although some aryl sulfamates, such as 667COUMATE, are relatively unstable when added to plasma ex vivo, they are stabilized in vivo by sequestration into rbcs and binding to CAII (279, 287). In general, aryl sulfamates, especially those of a steroidal nature, are stable.

    Estradiol-3-O-sulfamate per se does not bind to the ER and therefore acts as a prodrug for the natural estrogen, estradiol (286). On oral application it does not have an estrogenic effect on the liver, indicating that after absorption it must rapidly enter rbcs and transit the liver without undergoing metabolic inactivation. The finding that sulfamates are able to transit the liver in rbcs without being degraded therefore offers an explanation as to why EMATE is active as a STS inhibitor when administered orally.

    Many sulfonamide drugs, such as acetazolamide, which are structurally similar to the sulfamate-based STS inhibitors, are also transported in rbcs (288). Their transit in rbcs is facilitated by binding to CAII, which is present in the cytosol of rbcs. It was therefore reasoned that the sulfamate-bearing steroidal and nonsteroidal STS inhibitors may also interact with CAII.

    The ability of a number of STS inhibitors, including EMATE and 667COUMATE, to dock into the active site of CAII was initially examined using the known crystal structure for this enzyme (289, 290). Both compounds were found to dock into the active site of CAII. Subsequent studies using human (h)CAII derived from rbcs revealed that EMATE and 667COUMATE are both good inhibitors of CAII activity (IC50 values 42 and 25 nM, respectively). They are equipotent with the established CAII inhibitor, acetazolamide (IC50 25 nM), which is widely used for the treatment of a number of pathological conditions, including glaucoma. EMATE has been cocrystallized with CAII and the ligand-protein complex has been studied by x-ray crystallography (291), showing a good correlation with the structure predicted by docking studies (Fig. 8).

    CAII is a member of a family of 14 CAs that catalyze the reversible hydration of CO2 to HCO3– (292). Whereas CAII has an intracellular location, other CAs, e.g., CAIX and CAXII, have their active site domains located extracellularly. There is convincing evidence that the expression of these CAs is increased in many tumors in which their action to acidify the extracellular milieu may give tumors a growth advantage over normal tissues (293, 294, 295). Expression of CAIX and CAXII is increased under hypoxic condition through the hypoxia inducible factor 1 pathway (296, 297). In squamous cell head and neck cancer, overexpression of CAIX was found to be associated with resistance to radiation and chemotherapy (298). Therefore, in addition to CA inhibitors having a potential therapeutic role in the treatment of some cancers, inhibition of CAIX may render some tumors sensitive to the use of medication or chemotherapy.

    Although the crystal structure of hCAIX has not yet been reported, it has a 41% homology with hCAXII for which the crystal structure is known (299). Docking studies were therefore carried out, with a series of sulfamate-based STS inhibitors, into the extracellular domain of hCAXII to predict whether these compounds might also inhibit the activities of hCAIX and hCAXII (279). These studies revealed that the sulfamate-based STS inhibitors could dock into the active site of hCAXII and should be able to inhibit hCAIX and hCAXII. It was recently confirmed that EMATE and its bis-sulfamate derivative are potent inhibitors of hCAIX (300).

    There is evidence that some CA inhibitors, such as acetazolamide, can inhibit the in vitro invasion of renal cancer cells (301). They can also produce additive delays in tumor growth in vivo when used with other cytotoxic agents (302). The finding that both steroidal and nonsteroidal sulfamates, originally developed as STS inhibitors, are also potent CAII and CAIX inhibitors raises the intriguing possibility that inhibition of CAs may contribute to the overall anticancer efficacy of this class of drug.

    XI. Mechanism of Steroid Sulfate Hydrolysis and STS Inhibition

    STS is bound to the membrane of the endoplasmic reticulum. The highly hydrophobic nature of the enzyme has hampered its purification to homogeneity in quantities sufficient for crystallization. To obtain STS in a pure form, attempts were made 1) to express the protein, using STS cDNA, in the pGEX2T expression system; 2) to express a mutant form of the protein, in which the putative membrane-spanning domain was deleted, in Chinese hamster ovary cells; and 3) to isolate a soluble STS from the snail Helix pomatia (303). Recently, through exploitation of optimal solubilization and detergent conditions to protect the structural and functional integrity of the molecule, thereby preventing nonspecific aggregation and other instabilities, human STS was successfully purified and crystallized (304), and its crystal structure (used in Fig. 9) was reported subsequently (305).

    The overall shape of the protein is "mushroom-like" with the crown protruding toward the lumen side of, and the stalk traversing through, the lipid bilayer of the endoplasmic reticulum (Fig. 9). Similar to its closely related soluble enzymes arylsulfatase A (ARSA) and arylsulfatase B (ARSB), the crystal structures of which were published a few years ago (2, 306), STS shares a similar catalytic site topology and a unique, but universal for all sulfatases, posttranslational modification of a conserved cysteine residue to a formylglycine (FGly, ·-CHO) residue (307, 308, 309). As observed for ARSB, the resting state of human STS at the catalytic site consists of a sulfated gem-diol form of FGly, i.e. [FGlyS, ·-CH(OH)OSO3–] which is, in all likelihood, coordinated to a bivalent Ca2+ cation (Fig. 10). The catalytic site of the STS active site is highly homologous to those in ARSA and ARSB. All three enzymes share nine identical catalytically important residues, namely three aspartic acid residues, two histidine residues, two lysine residues, one arginine residue, and the FGlyS residue. In STS, these residues are Arg35, Arg36, Arg342, His136, His290, Lys134, Lys368, Arg79, and Fgly75 (Fig. 10). The only difference apparently is the 10th residue in STS, which is assigned to be Gln343 instead of an asparagine as in both ARSA and ARSB. When the sulfate group of E1S is superimposed with FGlyS in the crystal structure, amino acid residues Leu74, Arg98, Thr99, Val101, Leu103, Leu167, Val177, Phe178, Thr180, Gly181, Thr484, His485, Val486, and Phe488 surround and interact favorably with the steroid scaffold, suggesting that some of these amino acid residues could be involved in substrate recognition via hydrophobic contacts.

    It has been demonstrated that the gem-diol form of FGly [·- CH(OH)OH] is crucial to the hydrolysis of sulfate substrates by ARSA and ARSB (310). Hence, the putative mechanism of STS for the hydrolysis of E1S to E1 is depicted in Fig. 11. The first step involves the regeneration of the gem-diol form of FGly from FGlyS via 1) desulfation, catalyzed by the nonesterified hydroxyl group, followed by the attack of a molecule of water on the FGly intermediate; or 2) a direct attack of a molecule of water on the sulfur atom of FGlyS. One of the hydroxyl groups of ·-CH(OH)OH then attacks the sulfur atom of E1S, releasing E1 and regenerating FGlyS as a consequence.

    Because the sulfamate group is acting as a mimic of the sulfate group of E1S, it is reasonable to expect that the mechanism of action for sulfamate-based STS inhibitors such as EMATE would also involve the FGlyS residue. One proposed mechanism of STS inhibition by EMATE, which involves a nucleophilic attack on the sulfamoyl group by ·-CH(OH)OH in the enzyme active site, is shown in Fig. 12. It is not clear whether structure I could be a "dead-end" product or will undergo additional modifications to yield a species that irreversibly inactivates the enzyme. The fact that the sulfamoyloxy group of EMATE can exist in an anionic form (OSO2NH–, see below) would suggest the existence of a monoanionic form of structure I that might deactivate its sulfur atom and hence hamper desulfamoylation of structure I in a manner similar to that proposed for the regeneration of ·-CH(OH)OH from FGlyS in Fig. 11.

    Bond et al. (2) proposed that the resting FGlyS observed in the crystals of ARSB could be in equilibrium with the free form, i.e., FGly, which could be an intermediate in the regeneration of ·-CH(OH)OH from FGlyS as depicted in Fig. 11. For these reasons, a potential nucleophilic attack on the formyl group by the lone-pair electrons of the N atom of the sulfamate group of EMATE-like compounds could be envisaged. However, there is now strong evidence to suggest that the nucleophilic attack on the carbonyl group of FGly could well be initiated not by a neutral sulfamate but by its monoanionic form (OSO2NH–). Because the sulfamate moiety is presumably acting as a sulfate surrogate, its anionic form should presumably interact more favorably with the enzyme active site and hence compete more effectively against E1S for binding than the neutral form.

    Anderson et al. (199) demonstrated that the monoanionic form of the phosphate group of estrone phosphate bound most tightly to STS. It has been shown that EMATE binds to the crystal structure of CAII using a coordination of the sulfamate anion to the active site zinc atom (291). Previous studies on various sulfamates have shown that the N-proton is fairly acidic with a pKa value in the range of 7–11, e.g., approximately 9.5 for EMATE (in 70% aqueous MeOH) (311) and about 9.1 for 667COUMATE (in 50% aqueous MeOH) (245). This implies that at physiological pH a proportion of the weakly acidic EMATE and 667COUMATE could be in their conjugate base form. Given that several lysine and histidine residues are among the essential amino acids lining the catalytic site of STS (Fig. 10), it is conceivable that N-deprotonation of EMATE-like compounds by these basic amino acid residues takes place and the resulting anionic species then acts as a nucleophile attacking the FGly. Two putative pathways for such an attack by EMATE are depicted in Fig. 13. The hemiaminal-type intermediate I (Fig. 13) so formed could then be hydrolyzed to give estrone and the intermediate II, which upon dehydration gives an imino structure IV (Fig. 13, path A). Alternatively, structure IV could be formed via hydrolysis of the ester III after the dehydration of the hemiaminal intermediate I (Fig. 13, path B). It has also been suggested that structure IV could be formed via an attack on FGly by sulfamic acid, which is released upon the hydrolysis of the sulfamate group of EMATE-like compounds (255). Structures II, III, and IV, and possibly even structure I, are proposed to be dead-end products, and no regeneration of FGly, as depicted in Fig. 11, is therefore anticipated. The formation of an azomethine adduct similar to structure III or IV is certainly not unprecedented for the nonenzymatic chemical reaction. When a solution of 2-nitrophenol in N,N-dimethyformamide (HCONMe2) was treated with sodium hydride followed by sulfamoyl chloride, an azomethine adduct of 2-nitrophenol-O-sulfamate and N,N-dimethyformamide (i.e., Me2N-CHN-SO2O-Ph-2-NO2) was isolated as a minor product that was shown to be stable and resistant to hydrolysis (247).

    Although the putative mechanisms of action depicted in Figs. 12 and 13 suggest the involvement of ·-CH(OH)OH and FGly, it is also quite possible that irreversibly inhibiting sulfamate esters, like EMATE, could inhibit STS in a more random manner by a specific or nonspecific sulfamoylation of amino acid residues in the active site. Such proposed mechanisms are shown in Fig. 14. Path A involves an attack by a nucleophilic amino acid residue in the active site. This mechanism is analogous to Fig. 12 except that the attacking species on the sulfamate group of EMATE is a nucleophilic amino acid residue other than ·-CH(OH)OH. Path B involves the generation of a highly electrophilic sulfonylamine species via an E1cB process, either from the bound monoanion or possibly initiated by an enzyme catalyzed N-proton abstraction and stimulated by hydrogen bonding to the bridging O-atom or coordination of the bridging O-atom to Ca2+ ion. Such collapse of a sulfamate ester is well precedented in their nonenzymatic chemistry (312, 313).

    Before the publication of the crystal structure of STS, evidence for the involvement of two amino acid residues in the inactivation process was provided by the biphasic nature of inactivation exhibited by EMATE and the fact that two ionizable groups with pKa values of 7.2 and 9.8 were identified upon analysis of the pH dependence of enzyme activity and of enzyme inactivation by EMATE (314). The finding that Rose Bengal inhibited STS activity in a dose-dependent manner strongly suggested that a histidine was involved in the catalytic mechanism. It was originally suggested that the second residue was a tyrosine, but the involvement of this amino acid in the catalytic mechanism of STS was subsequently ruled out by the absence of a tyrosine among the conserved active site residues throughout the entire family of sulfatases, on publication of the first crystal structure of a sulfatase (2). When it was reported that the two lysine residues and a histidine residue are important, inter alia, for sulfate group binding and catalysis, respectively, in ARSA (315), the identity of the second residue was considered to be a lysine, the pKa value of which (as a conjugate acid) of approximately 10 in proteins (316) matches the experimental finding. With what is now known about the topology of the catalytic site of STS as a result of its crystal structure, the attacking nucleophile in mechanism A and the N-proton abstracting amino acid residue in mechanism B of Fig. 10 is most likely to be Lys134, Lys368, His136, or His290. It is highly likely that the reactive sulfonylamine released, as proposed in mechanism B, would attack an immediate nucleophile within the catalytic site but not one located more remotely from the enzyme active site. With the presence of Lys134, Lys368, His136, and His290 in the catalytic site of STS, it could be envisaged that the products of inactivation would most likely be sulfamoylated enzyme intermediates or lysine- or histidine-derived sulfamides (NHSO2NH2 or NSO2NH2), which are proposed to be dead-end products. From literature precedent and also our own experience in the handling of this type of compound, sulfamides such as estrone 3-N-sulfamide (211) are inactive and stable entities, rendering them excellent candidates for a dead-end product of enzyme inactivation.

    The proposed mechanisms above represent the most likely possibilities for the inactivation of STS by an active site-directed EMATE-like inhibitor. However, all of them remain hypothetical because there are no experimental data available yet to support any of these mechanisms. Nonetheless, these hypotheses represent viable models for understanding how inhibition of STS by sulfamate esters may proceed and should promote the design of experiments to test such models. Because technology is now available for isolating and crystallizing the membrane-bound STS, one challenge ahead is the crystallization of an inactivated STS after the enzyme has been incubated with an irreversible inhibitor such as EMATE. Such a crystal structure should help to define the role of the sulfamate group in the mechanism of action of EMATE-like sulfamate-based STS inhibitors and to provide information on the participating essential amino acid(s) in the inactivation process.

    XII. Future Perspectives

    The discovery that steroidal and nonsteroidal aryl sulfamates are potent STS inhibitors will enable the roles that this enzyme has in physiological processes to be explored. Although the enzyme is known to be widely distributed, the lack of potent inhibitors has previously made it difficult to determine its role in the body. More importantly, evidence to support a major role for the enzyme in regulating the growth of some breast cancers has been accumulating for many years. With the advent of potent inhibitors, it will now be possible to explore their therapeutic potential. Indeed, one of the inhibitors detailed in this review, 667COUMATE, an irreversible sulfatase inhibitor, has now entered a phase I trial in postmenopausal women with breast cancer. Preliminary results obtained from a small number of patients are encouraging and have revealed that more than 90% inhibition of STS in PBLs can be achieved with low doses of the drug (317). Additional phase II/III trials will be required to confirm whether such STS inhibitors are to have a place in the armory against breast cancer. There is also growing interest in the pursuit of reversible sulfatase inhibitors, and this is an area that will undoubtedly develop over the next few years as novel templates are synthesized. If the sulfatase inhibitor concept is successful, future trials of STS inhibitors in combination with aromatase inhibitors, or possibly DASI-type compounds, will be required to determine whether such combinations offer any advantages over the use of single-agent therapy. Also, there are many other potential disease targets beyond oncology that could receive attention after an initial clinical success.

    An important property of this class of aryl sulfamates is their ability to be transported in rbcs and to avoid first-pass inactivation during transit through the liver. This property results from their ability to interact with and inhibit a number of CAs that may contribute to the overall efficacy of this class of drug. With the publication of several crystal structures for sulfamates interacting with CA, the molecular features that influence potency are becoming clearer and will support more rational drug design strategies to exploit this idea. It will thus be possible to test whether other classes of drugs can also be delivered by this mechanism upon sulfamoylation. There is already evidence that a sulfamoylated antiestrogen possesses dual antiestrogen/STS-inhibitory properties, suggesting that such an approach should be feasible (318).

    An unexpected outcome of the research to reduce the estrogenicity of EMATE, the first potent STS inhibitor, was the discovery that the 2-substituted estrogen sulfamate derivatives are also potent antitumor/antiangiogenic agents. These derivatives have increased potency compared with their nonsulfamoylated analogs but, in addition, their metabolic stability in vivo is considerably enhanced. The next few years should hold considerable promise for exploring the potential of this new class of sulfamoylated drug that, in addition to exhibiting potent STS inhibition, targets other key steps in the malignant process.

    Acknowledgments

    We thank Mr. J. J. Robinson for molecular modeling support and Miss C. Wilson and Mrs. C. M. Parker for their skilled secretarial assistance in the preparation of this manuscript.

    Footnotes

    First Published Online November 23, 2004

    Abbreviations: Adiol, 5-Androstenediol; Adione, androstenedione; ARSA, arylsulfatase A; ARSB, arylsulfatase B; CA, carbonic anhydrase; COUMATE, 4-methylcoumarin 7-O-sulfamate; 667 COUMATE, 6-oxo-8,9,10,11-tetrahydro-7H-cyclohepta-[c] [1]benzopyran-3-0-sulfamate; DASI, dual aromatase-sulfatase inhibitor; DHEA, dehydroepiandrosterone; DHEAS, DHEA sulfate; E1, estrone; E2, estradiol; E1S, E1 sulfate; E2S, E2 sulfate; EMATE, estrone-3-O-sulfamate; ER, estrogen receptor; FGly, formylglycine; FGlyS, sulfated gem-diol form of FGly; HEK, human embryonic kidney; 17?HSD, 17?-hydroxysteroid dehydrogenase; 2-MeOE2, 2-methoxyestradiol; 2-MeOE2-bisMATE, 2-MeOE2-bis-sulfamate; NMU, nitrosomethyl-urea; OATP, organic anion transporter polypeptide; PBL, peripheral blood leukocyte; rbc, red blood cell; SAR, structure-activity relationship; STS, steroid sulfatase; URE, upstream regulatory element; X-LI, X-linked ichthyosis.

    References

    Schacter B, Marrian GF 1938 The isolation of estrone sulfate from the urine of pregnant mares. J Biol Chem 126:663–669

    Bond CS, Clements PR, Ashby SJ, Collyer CA, Harrop SJ, Hopwood JJ, Guss JM 1997 Structure of a human lysosomal sulfatase. Structure 5:277–289

    Ferrante P, Messali S, Meroni G, Ballabio A 2002 Molecular and biochemical characterisation of a novel sulphatase gene: aryl sulphatase G (ARSG). Eur J Hum Genet 10:813–818

    Yen PH, Marsh B, Allen E, Tsai SP, Ellison J, Connoly L, Neiswanger K, Shapiro LJ 1988 The human X-linked steroid sulfatase gene and a Y-encoded pseudogene: evidence for an inversion of the Y chromosome during primate evolution. Cell 55:1123–1135

    Yen PH, Allen E, Marsh B, Mohandas T, Wang N, Taggart, Shapiro LJ 1987 Cloning and expression of the steroid sulfatase cDNA and the frequent occurrence of deletions in STS deficiency: implications for X-Y interchange. Cell 49:443–454

    Stein C, Hille A, Seidel J, Rijnbout S, Waheed A, Schmidt B, Geuze H, von Figura K 1989 Cloning and expression of human steroid-sulfatase. Membrane topology, glycosylation, and subcellular distribution in BHK-21 cells. J Biol Chem 264:13865–13872

    Shapiro LJ, Yen P, Pomerantz D, Martin E, Rolewic L, Mohandas T 1989 Molecular studies of deletions at the human steroid sulfatase locus. Proc Natl Acad Sci USA 86:8477–8481

    Basler E, Grompe M, Parenti G, Yates J, Ballabio A 1992 Identification of point mutations in the steroid sulfatase gene of three patients with X-linked ichthyosis. Am J Hum Genet 50:483–491

    Alperin ES, Shapiro LJ 1997 Characterization of point mutations in patients with X-linked ichthyosis. Effects on the structure and function of the steroid sulfatase protein. J Biol Chem 272:20756–20763

    Valdes-Flores M, Vaca ALJ, Kofman-Alfaro SH, Cuevas-Covarrubias SA 2001 Characterization of a novel point mutation (Arg432His) in X-linked ichthyosis. Acta Derm Venereol 81:54–78

    Gonzalez-Huerta LM, Riviera-Vega MR, Kofman-Alfeuro SH, Cuevas-Covarrubias SA 2003 Novel missense mutation (Arg432Cys) in a patient with steroid sulphatase-deficiency. Clin Endocrinol (Oxf) 59:263–265

    Peters C, Schmidt B, Rommerskirch W, Rupp K, Zuhlsdorf M, Vingron M, Meyer HE, Pohlmann R, von Figura K 1990 Phylogenetic conservation of arylsulfatases. cDNA cloning and expression of human arylsulfatase B. J Biol Chem 265:3374–3381

    Newman SP, Purohit A, Ghilchik MW, Potter BVL, Reed MJ 2000 Regulation of steroid sulphatase expression and activity in breast cancer. J Steroid Biochem Mol Biol 75:259–264

    Li X-M, Alperin ES, Salido E, Gong Y, Yen P, Shapiro LJ 1996 Characterization of the promoter of human steroid sulfatase: a gene which escapes X inactivation. Somat Cell Mol Genet 22:105–117

    Dodgson KS, Spencer B, Thomas J 1954 Studies on sulphatases. 6. The localization of arylsulphatase in the rat-liver cell. J Biochem 56:177–181

    Partanen S 1985 Histochemistry of estrogen sulfatases in human breast diseases. Virchows Arch (Cell Pathol) 49:53–60

    Willemsen R, Kroos M, Hoogeveen AT, van Dongen JM, Parenti G, Van Der Loos CM, Reuser AJJ 1988 Ultrastructural localization of steroid sulphatase in cultured human fibroblasts by immunocytochemistry: a comparative study with lysosomal enzymes and the mannose-6-phoshphate receptor. Histochem J 20:41–51

    Kawano J-I, Kotani T, Umeki K, Oinuma T, Ohtaki S, Aikawa E 1989 A monoclonal antibody to rat liver arylsulfatase C and its application in immunohistochemistry. J Histochem Cytochem 37:683–690

    Okuda T, Saito H, Sekizawa A, Shimizu Y, Akamatsu T, Kushima M, Yanaihara T, Okai T, Farina A 2001 Steroid sulfatase expression in ovarian clear cell adenocarcinoma: immunohistochemical study. Gynecol Oncol 82:427–434

    Ezaki K, Motoyama H, Sasaki H 2001 Immunohistologic localization of estrone sulfatase in uterine endometrium and adenomyosis. Obstet Gynecol 98:815–819

    Nakamura Y, Miki Y, Susuki T, Nakata T, Darnel AD, Mariya T, Tazawa C, Saito H, Ishibashi TST, Yamada S, Sasano H 2003 Steroid sulfatase and estrogen sulfotransferase in the atherosclerotic human aorta. Am J Pathol 163:1329–1339

    Yanaihara A, Yanaihara T, Toma Y, Shimizu Y, Saito H, Okai T, Higashiyama T, Osawa Y 2001 Localization and expression of steroid sulfatase in human fallopian tubes. Steroids 66:87–91

    Miki Y, Nakata T, Suzuki T, Darnel AD, Moriya T, Kaneko C, Hidaka K, Shiotsu Y, Kusaka H, Sasano H 2002 Systemic distribution of steroid sulfatase and estrogen sulfotransferase in human adult and fetal tissues. J Clin Endocrinol Metab 87:5760–5768

    Suzuki T, Nakata T, Miki Y, Kaneko C, Moriya T, Ishida T, Akinaga S, Hirakawa H, Kimura M, Sasano H 2003 Estrogen sulfotransferase and steroid sulfatase in human breast carcinoma. Cancer Res 63:2762–2770

    Burstein S, Dorfman RI 1963 Determination of mammalian steroid sulfatase with 7-3H-3?-hydroxy androst-5-en-17-one sulfate. J Biol Chem 238:1656–1660

    Han HD, Fencl MM, Tulchinsky D 1987 Variations in estrone sulfatase activity in human leukocytes. J Clin Endocrinol Metab 65:1026–1030

    Iwamori M, Moser HW, Kishimoto Y 1976 Steroid sulfatase in brain: comparison of sulfohydrolase activities for various steroid sulfates in normal and pathological brains, including the various forms of metachromatic leukodystrophy. J Neurochem 27:1389–1395

    Kishimoto Y, Sostek R 1972 Activity of sterol-sulphate sulphohydrolase in rat brain: characterization, localization and change with age. J Neurochem 19:123–130

    Purohit A, Flanagan AM, Reed MJ 1992 Estrogen synthesis by osteoblast cell lines. Endocrinology 131:2027–2029

    Fujikawa H, Okura F, Kuwano Y, Sekizawa A, Chiba H, Shimodaira K, Saito H, Yanaihara T 1997 Steroid sulfatase activity in osteoblast cells. Biochem Biophys Res Commun 231:42–47

    Clemens JW, Kabler HL, Sarrap JL, Beyer AR, Li P-K, Selcer KW 2000 Steroid sulfatase activity in the rat ovary, cultured granulosa cells and a granulosa cell line. J Steroid Biochem Mol Biol 75:245–252

    Bonser J, Walker J, Purohit A, Reed MJ, Potter BVL, Willis DS, Franks S, Mason HD 2000 Human granulosa cells are a site of sulphatase activity and are able to utilize dehydroepiandrosterone sulphate as a precursor for oestradiol production. J Endocrinol 167:465–471

    Payne AH, Mason M, Jaffe RB 1969 Testicular steroid sulfatase: substrate specificity and inhibition. Steroids 14:685–704

    Meyer CH, Grundmann H, Weiss H 1984 Steroid sulfatase = Arylsulfatase C? Chromatographic and electrophoretic properties in extracts from placental microsomes and skin fibroblasts. Dermatologica 169:305–310

    Burns GRJ 1983 Purification and partial characterisation of arylsulphatase C from human placental microsomes. Biochim Biophys Acta 759:199–204

    Noel H, Plante L, Bleau G, Chapdelaine A, Roberts KI 1983 Human placental steroid sulfatase: purification and properties. J Steroid Biochem 19:1591–1598

    Egyed J, Oakey RE 1985 Hydrolysis of deoxycorticosterone-21-yl sulphate and dehydroepiandrosterone sulphate by microsomal preparations of human placenta: evidence for a common enzyme. J Endocrinol 106:295–301

    Dibbelt L, Kuss E 1983 Human placental steroid sulfatase. Hoppe-Seylers Z Physiol Chem 364:187–191

    Kester MHA, Kaptein E, Van Dijk CH, Roest TJ, Tibbod D, Coughtrie MWH, Visser TJ 2002 Characterisation of iodothyronine sulfatase activities in human and rat liver and placenta. Endocrinology 143:814–819

    Jobsis AC, de Groot WP, Tigges AJ, de Bruijn HW, Rijken Y, Meijer AE, Marinkovic-Ilsen A 1980 X-linked ichthyosis and X-linked placental sulfatase deficiency: a disease entity. Histochemical observations. Am J Pathol 99:279–289

    Epstein EH, Benifas JM 1985 Recessive X-linked ichthyosis: lack of immunologically detectable steroid sulfatase enzyme protein. Hum Genet 71:201–205

    Shapiro LJ, Cousins L, Fluharty AL, Stevens RL, Kihara H 1977 Steroid sulfatase deficiency. Pediatr Res 11:894–897

    Lykkesfeldt G 1987 Human steroid sulphatase deficiency: clinical, endocrinology and genetic aspects. Dan Med Bull 34:69–83

    Purohit A, Dauvois S, Parker MG, Potter BVL, Williams GJ, Reed MJ 1994 The hydrolysis of oestrone sulphate and dehydroepiandrosterone sulphate by human steroid sulphatase expressed in transfected COS-1 cells. J Steroid Biochem Mol Biol 50:101–104

    Nelson K, Keinanen BM, Daniel WL 1983 Murine arylsulfatase C: evidence for two isozymes. Experientia 39:740–742

    Zhu BT, Fu J-H, Xu S, Kauffman FC, Conney AH 1998 Different biochemical properties of nuclear and microsomal estrone-3-sulfatases: evidence for the presence of a nuclear isozyme. Biochem Biophys Res Commun 246:45–49

    Gniot-Szulzycka J, Januszewska B 1986 Purifcation of steroid sulphohydrolase from human placenta microsomes. Acta Biochim Pol 33:203–215

    Simard JP, Ameen M, Chang PL 1985 Biochemical characterisation of arylsulfatase-C isozymes in human fibroblasts. Biochem Biophys Res Commun 128:1388–1394

    Chang PL, Mueller OT, Lafrenie RM, Varey PA, Rosa NE, Davidson RG, Henry WM, Shows TB 1990 The human arylsulfatase C isoenzymes: two distinct genes that escape from inactivation. Am J Hum Genet 46:729–737

    Munroe DG, Chang PL 1987 Tissue-specific expression of human arylsulfatase C isozymes and steroid sulfatase. Am J Hum Genet 40:102–114

    Purohit A, Reed MJ, Morris NC, Williams GJ, Potter BVL 1996 Regulation and inhibition of steroid sulfatase activity in breast cancer. Ann NY Acad Sci 784:40–49

    Purohit A, Duncan LJ, Wang DY, Coldham NG, Ghilchik MW, Reed MJ 1997 Paracrine control of oestrogen production in breast cancer. Endocr Rel Cancer 4:323–330

    Matsuoka R, Yanaihara A, Saito H, Furusawa Y, Toma Y, Shimizu Y, Yanaihara T, Okai T 2002 Regulation of estrogen activity in human endometrium: effect of IL-1? on steroid sulfatase activity in human endometrial stromal cells. Steroids 67:655–659

    Purohit A, Budai B, Wang DY, Willemsen EL, de Winkel A, Parish D, Ghilchik MW, Szamel I, Reed MJ 2000 Modulation of oestrone sulphate formation and hydrolysis in breast cancer cells by breast cyst fluid from British and Hungarian women. Br J Cancer 82:492–496

    Purohit A, Chapman O, Duncan L, Reed MJ 1992 Modulation of oestrone sulphatase activity in breast cancer cell lines by growth factors. J Steroid Biochem Mol Biol 41:563–566

    Schneider G, French A, Bullock LP, Bardin CW 1971 Absence of androgen stimulation of hepatic steroid sulfatase in the female rat. Endocrinology 89:308–310

    Lam STS, Polani PE 1985 Hormonal induction of steroid sulphatase in the mouse. Experientia 41:276–278

    Moutaouakkil M, Prost O, Dahan N, Adessi GL 1984 Estrone and dehydroepiandrosterone sulfatase activities in guinea-pig uterus and liver: estrogenic effect of estrone sulfate. J Steroid Biochem 21:321–328

    Barth A, Romer W, Oettel M 2000 Influence of subchronic administration of oestrone-3-O-sulphamate on oestrone sulphatase activity in liver, spleen and white blood cells of ovariectomised rats. Arch Toxicol 74:366–371

    Pasqualini JR, Maloche C, Maroni M, Chetrite G 1994 Effect of the progestagen Promegestone (R-5020) on mRNA of the oestrone sulphatase in the MCF-7 human mammary cancer cells. Anticancer Res 14:1589–1594

    Purohit A, Reed MJ 1992 Oestrogen sulphatase activity in hormone-dependent and hormone-independent breast cancer cells: modulation by steroidal and non-steroidal therapeutic agents. Int J Cancer 50:901–905

    Beck L, Mahfoudi A, Mularoni A, Nicollier M, Adessi GL 1992 Progesterone stimulates sulfate uptake in subcultured endometrial epithelial cells. Mol Cell Endocrinol 90:95–102

    Hughes PJ, Twist LE, Durham J, Choudhry MA, Drayson M, Chandraratna R, Michell RH, Kirk CJ, Brown G 2001 Up-regulation of steroid sulphatase activity in HL60 promyelocytic cells by retinoids and 1, 25-dihydroxyvitamin D3. Biochem J 355:361–371

    James VHT, Reed MJ 1980 Steroid hormones and human cancer. Prog Cancer Res Therapy 14:471–487

    Bernstein L, Ross RK 1993 Endogenous hormones and breast cancer. Epidemiol Rev 15:48–65

    MacDonald PC, Edman CD, Hemsell DL, Porter JC, Siiteri PK 1978 Effect of obesity on conversion of plasma androstenedione to estrone in women with and without endometrial cancer. Am J Obstet Gynecol 130:448–455

    Reed MJ, Hutton JD, Baxendale PM, James VHT, Jacobs HS, Fisher RP 1979 The conversion of androstenedione to oestrone and production of oestrone in women with endometrial cancer. J Steroid Biochem 11:905–911

    Reed MJ, Murray MAF 1979 The oestrogens. In: Gray CH, James VHT, eds, Hormones in blood. 3rd ed. London: Academic Press; 263–353

    Hobkirk R 1993 Steroid sulfation. Trends Endocrinol Metab 4:69–74

    Falany JL, Falany CN 1996 Expression of cytosolic sulfotransferases in normal mammary epithelial cells and breast cancer cell lines. Cancer Res 56:1551–1555

    Falany JL, Macrina N, Falany CN 2002 Regulation of MCF-7 breast cancer cell growth by ?-estradiol sulfonation. Breast Cancer Res Treat 74:167–176

    Strott CA 2002 Sulfonation and molecular action. Endocr Rev 23:703–732

    Noel CT, Reed MJ, Jacobs HS, James VHT 1981 The plasma concentration of oestrone sulphate in postmenopausal women: lack of diurnal variation, effect of ovariectomy, age and weight. J Steroid Biochem 14:1101–1105

    Pasqualini JR, Gelly C, Nguyen BL, Vella C 1989 Importance of oestrogen sulfates in breast cancer. J Steroid Biochem 34:155–163

    Ruder HJ, Loriaux DL, Lipsett MB 1972 Estrone sulfate: production rate and metabolism in man. J Clin Invest 51:1020–1033

    Reed MJ, Purohit A 1993 Sulphatase inhibitors: the rationale for the development of a new endocrine therapy. Rev Endocr Rel Cancer 45:51–62

    Reed MJ, Purohit A 1994 Inhibition of steroid sulphatases. In: Sandler M, Smith HJ, eds. Design of enzyme inhibitors as drugs, vol. 2. Oxford: Oxford University Press; 481–494

    Reed MJ, Purohit A, Howarth NM, Potter BVL 1994 Steroid sulphatase inhibitors: a new endocrine therapy. Drugs Future 19:673–680

    Reed MJ, Purohit A, Woo LWL, Potter BVL 1996 The development of steroid sulphatase inhibitors. Endocr Rel Cancer 3:9–23

    Bonney RC, Reed MJ, Davidson K, Beranek PA, James VHT 1983 The relationship between 17? hydroxysteroid dehydrogenase activity and oestrogen concentrations in human breast tumours and in normal breast tissue. Clin Endocrinol (Oxf) 19:727–739

    Van Landeghem AAJ, Poortman J, Nabuurs M, Thijssen JHH 1985 Endogenous concentrations and sub-cellular distribution of estrogens in normal and malignant breast tissue. Cancer Res 45:2900–2904

    Thijssen JHH, Blankenstein MA, Miller WR, Milewicz A 1987 Estrogens in tissues: uptake from the peripheral circulation or local production. Steroids 50:297–306

    Thijssen JHH, Blankenstein MA 1989 Endogenous oestrogens and androgens in normal and malignant endometrial and mammary tissues. Eur J Cancer Clin Oncol 25:1953–1959

    Fishman J, Nisselbaum JS, Menendez-Botet CJ, Schwartz MK 1977 Estrone and estradiol content in human breast tumours: relationship to estradiol receptors. J Steroid Biochem 8:893–896

    Edery M, Goussard J, Dehennin L, Scholler R, Reiffsteck J, Drosdowsky MA 1981 Endogenous oestradiol-17? concentration in breast tumours determined by mass fragmentography and by radioimmunoassay: relationship to receptor content. Eur J Cancer 17:115–120

    James VHT, McNeill JM, Lai LC, Newton CJ, Ghilchik MW, Reed MJ 1987 Aromatase activity in normal breast and breast tumor tissue: in vivo and in vitro studies. Steroids 50:269–279

    Yamamoto T, Kitawaki J, Urabe M, Honjo H, Tamra T, Noguchi T, Okada H, Sasaki H, Tada A, Terashima Y, Nakamura J, Yoshihama M 1993 Estrogen productivity of endometrium and endometrial cancer tissue; influence of aromatase on proliferation of endometrial cancer cells. J Steroid Biochem Mol Biol 44:463–468

    Tilson-Mallett N, Santner SJ, Feil PD, Santen RJ 1983 Biological significance of aromatase activity in human breast tumors. J Clin Endocrinol Metab 57:1125–1128

    Santner SJ, Feil PD, Santen RJ 1984 In situ estrogen production via estrone sulfatase pathway in breast tumors: relative importance vs. the aromatase pathway. J Clin Endocrinol Metab 59:29–33

    Utsumi T, Yoshimura N, Takeuchi S, Maruta M, Maeda K, Harada N 2000 Elevated steroid sulfatase expression in breast cancer. J Steroid Biochem Mol Biol 73:141–145

    Pasqualini JR, Chetrite G, Blacker C, Feinstein MC, Delalonde M, Talbi M, Maloche C 1996 Concentrations of estrone, estradiol and estrone sulfate and evaluation of sulfatase and aromatase activities in pre- and postmenopausal breast cancer patients. J Clin Endocrinol Metab 81:1460–1464

    Utsumi T, Yoshimura N, Takeuchi S, Ando J, Maruta M, Maeda K, Harada N 1999 Steroid sulfatase expression is an independent predictor of recurrence in human breast cancer. Cancer Res 59:377–381

    Evans TRJ, Rowlands MG, Silva MC, Law M, Coombes RC 1993 Prognostic significance of aromatase and estrone sulfatase enzymes in human breast cancer. J Steroid Biochem Mol Biol 44:583–587

    Evans TRJ, Rowlands MG, Law M, Coombes RC 1994 Intratumoral oestrone sulphatase activity as a prognostic marker in human breast carcinoma. Br J Cancer 69:555–561

    Miyoshi Y, Ando A, Hasegawa S, Ishitobi M, Taguchi T, Tamaki Y, Noguchi S 2003 High expression of steroid sulfatase mRNA predicts poor prognosis in patients with estrogen receptor-positive breast cancer. Clin Cancer Res 9:2288–2293

    Silva MC, Rowlands MG, Dowsett M, Gusterson B, McKinna JA, Fryatt L, Coombes RC 1989 Intratumoral aromatase as a prognostic factor in human breast carcinoma. Cancer Res 49:2588–2591

    Lipton A, Santen RJ, Harvey HA, Sanders SL, Mathews YL 1992 Prognostic value of breast cancer aromatase. Cancer 70:1951–1955

    Yoshimura N, Harada N, Bukholm I, Karesen R, Borresen-Dale A-L, Kristensen VN 2004 Intratumoural mRNA expression of genes from the oestradiol metabolic pathway and clinical and histopathological parameters of breast cancer. Breast Cancer Res 6:R46–R55

    Purohit A, Ghilchik MW, Duncan LJ, Wang DY, Singh A, Walker MM, Reed MJ 1995 Aromatase activity and interleukin 6 production by normal and malignant breast tissues. J Clin Endocrinol Metab 80:3052–3058

    Sasano H, Nagura H, Harada N, Goukou Y, Kimura M 1994 Immunolocalization of aromatase and other steroidogenic enzymes in human breast disorders. Hum Pathol 25:530–535

    Santen RJ, Martel J, Hoaland F, Naftolin F, Roa L, Harada N, Hafer L, Zaino R, Santner SJ 1994 Stromal spindle cells contain aromatase in human breast tumors. J Clin Endocrinol Metab 79:627–632

    Esteban JM, Warsi Z, Haniu M, Hull P, Shively JE, Chen S 1992 Detection of intratumoral aromatase in breast carcinomas. Am J Pathol 140:337–343

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

    Pasqualini JR, Gelly C, Lecerf F 1986 Estrogen sulfates: biological and ultrastructural responses and metabolism in MCF-7 human breast cancer cells. Breast Cancer Res Treat 8:233–240

    Santner SJ, Levin MC, Santen RJ 1990 Estrone sulfate stimulates growth of nitrosomethylurea-induced breast carcinoma in vivo in the rat. Int J Cancer 46:73–78

    Masamura S, Santner SJ, Santen RJ 1996 Evidence of in situ estrogen synthesis in nitrosomethylurea-induced rat mammary tumors via the enzyme estrone sulfatase. J Steroid Biochem Mol Biol 58:425–429

    Reed MJ, Owen AM, Lai LC, Coldham NG, Ghilchik MW, Shaikh NA, James VHT 1989 In situ oestrone synthesis in normal breast and breast tumour tissue: effect of treatment with 4-hydroxyandrostenedione. Int J Cancer 44:233–237

    James MR, Skaar TC, Lee RY, MacPherson A, Zwiebel JA, Ahluwalia BS, Ampy F, Clarke R 2001 Constitutive expression of the steroid sulfatase gene supports the growth of MCF-7 human breast cancer cells in vitro and in vivo. Endocrinology 142:1497–1505

    Kroboth PD, Salek FS, Pittenger AL, Fabian TJ, Frye RF 1999 DHEA and DHEAS: a review. J Clin Pharmacol 39:327–348

    Rosenfeld RS, Rosenberg BJ, Fukushima DK, Hellman L 1975 Twenty four hour secretory pattern of dehydroepiandrosterone and dehydroepiandrosterone sulfate. J Clin Endocrinol Metab 40:850–855

    Bird CE, Masters V, Clark AF 1984 Dehydroepiandrosterone sulfate: kinetics of metabolism in normal young men and women. Clin Invest Med 7:119–122

    Poortman J, Andriesse R, Agema A, Douker GH, Schwarz F, Thijssen JHH 1980 Adrenal androgen secretion and metabolism in postmenopausal women. In: Genazzani AR, Thijssen JHH, Siiteri PK, eds. Adrenal androgens. New York: Raven Press; 219–240

    Bonney RC, Scanlon MJ, Reed MJ, Jones DL, Beranek PA, James VHT 1984 Adrenal androgen concentrations in breast tumours and in normal breast tissue. The relationship to oestradiol metabolism. J Steroid Biochem 20:501–504

    Adams JB, Garcia M, Rochefort H 1981 Estrogenic effects of physiological concentrations of 5-androst-3?, 17?-diol and its metabolism in MCF-7 human breast cancer cells. Cancer Res 41:4720–4726

    Poulin R, Labrie F 1986 Stimulation of cell proliferation and estrogenic response by adrenal C19-5-steroids in the ZR-75–1 human breast cancer cell line. Cancer Res 46:4933–4937

    Dauvois S, Labrie F 1989 Androstenedione and androst-5ene-3?,17?-diol stimulate DMBA-induced rat mammary tumors—role of aromatase. Breast Cancer Res Treat 13:61–69

    Maggiolini M, Donze O, Jeannin E, Ando S, Picard D 1999 Adrenal androgens stimulate the proliferation of breast cancer cells as direct activators of estrogen receptor . Cancer Res 59:4864–4869

    Le Bail JC, Lofti H, Charles L, Pepin D, Habrioux G 2002 Conversion of dehydroepiandrosterone sulfate at physiological plasma concentration into estrogen in MCF-7 cells. Steroids 67:1057–1064

    Billich A, Nussbaumer P, Lehr P 2000 Stimulation of MCF-7 breast cancer cell proliferation by estrone sulfate and dehydroepiandrosterone sulfate: inhibition by novel non-steroidal sulfatase inhibitors. J Steroid Biochem Mol Biol 73:225–235

    Morris KT, Toth-Fejel SE, Schmidt J, Fletcher WS, Pommier RF 2001 High dehydroepiandrosterone-sulfate predicts breast cancer progression during new aromatase inhibitor therapy and stimulates breast cancer cell growth in tissue culture: a renewed role for adrenalectomy. Surgery 130:947–953

    Shapiro LJ, Weiss R, Webster D, France JT 1978 X-linked ichthyosis due to steroid sulphatase deficiency. Lancet 14:70–72

    Hernandez-Martin A, Gonzalez-Sarmiento R, Unamuno P 1999 X-linked ichthyosis: an update. Br J Dermatol 141:617–627

    Hoyer H, Lykkesfeldt G, Ibsen HH, Brandrup F 1986 Ichthyosis of steroid sulphatase deficiency. Clinical study of 76 cases. Dermatologica 172:184–190

    Williams ML, Elias PM 1981 Stratum corneum lipids in disorders of cornification. J Clin Invest 68:1404–1410

    Ballabio A, Shapiro LJ 1995 Steroid sulfatase and X-linked ichthyosis. In: Scriver CR, Beaudet AL, Sly WS, Vallee D, eds. The metabolic and molecular basis of inherited diseases. New York: McGraw-Hill; 2999–3022

    Pitts RL 1987 Serum elevation of dehydroepiandrosterone sulfate associated with male pattern baldness in young men. J Am Acad Dermatol 16:571–573

    Haning RV 1981 Using DHEAS to monitor androgen disorders. Contemp Obstet Gynecol 18:117–132

    Hay JB, Hodgins MB 1973 Metabolism of androgens in vitro by human facial and axillary skin. J Endocrinol 59:475–486

    Hoffman R 2001 Enzymology of the hair follicle. Eur J Dermatol 11:296–300

    Hoffman R, Rot A, Niiyama S, Billich A 2001 Steroid sulfatase in the human hair follicle concentrates in the dermal papilla. J Invest Dermatol 117:1342–1348

    Chen WC, Thiboutot D, Zouboulis CC 2002 Cutaneous androgen metabolism: basic research and clinical perspectives. J Invest Dermatol 119:992–1007

    Daynes RA, Araneo BA, Dowell TA, Huang K, Dudley D 1990 Regulation of murine lymphokine production in vivo. J Exp Med 171:979–996

    Daynes RA, Dudley DJ, Araneo BA 1990 Regulation of murine lymphokine production in vivo. II. Dehydroepiandrosterone is a natural enhancer of interleukin 2 synthesis by helper T cells. Eur J Immunol 20:793–803

    Rook GAW, Hernandez-Pando R, Lightman S 1994 Hormones, peripherally activated hormones and regulation of Th1/Th2 balance. Immunol Today 15:301–303

    Reed MJ, Purohit A, Singh A, Chander SK 2003 Androgenic modulation of the immune response. London: Parthenon Publishing Group

    Watson R, Huls A, Araghinikuam M, Sangbun C 1996 Dehydroepiandrosterone and diseases of aging. Drugs Aging 9:274–291

    Daynes RA, Araneo BA, Ershler WB, Maloney C, Li G-Z, Ryu S-Y 1993 Altered regulation of IL-6 production with normal aging. J Immunol 150:5219–5230

    Suitters AJ, Shaw S, Wales MR, Porter JR, Leonard J, Woodger R, Brand H, Bodmer M, Foulkes R 1997 Immune enhancing effects of dehydroepiandrosterone and dehydroepiandrosterone sulphate and the role of steroid sulphatase. Immunology 91:314–321

    Foulkes R, Shaw S, Suitters A 1997 Immunological consequences of inhibiting dehydroepiandrosterone sulfatase in vivo. In: Rook GAW, Lightman S, eds. Steroid hormones and the T-cell cytokine profile. London: Springer-Verlag; 135–152

    Reed MJ 1995 The discriminant function test: a marker of Th1/Th2 cytokine secretion and breast tumour oestrogen synthesis. Mol Med Today 1:98–103

    Reed MJ, Purohit A 1997 Breast cancer and the role of cytokines in regulating estrogen synthesis: an emerging hypothesis. Endocr Rev 18:701–715

    Singh A, Purohit A, Duncan LJ, Mokbel K, Ghilchik MW, Reed MJ 1997 Control of aromatase in breast tumours: the role of the immune system. J Steroid Biochem Mol Biol 61:185–192

    Siiteri PK, MacDonald PC 1973 Role of extraglandular estrogen in human endocrinology. In: Greep RO, Astwood EB, eds. Handbook of physiology, vol 2. Washington DC: American Physiological Society; 615–629

    Wei J XH, Davies JL, Hemmings GP 1992 Increase of plasma IL-6 concentrations with age in healthy subjects. Life Sci 51:1953–1956

    Orentreich N, Brind JL, Rijer RL, Vogelman JN 1984 Age changes and sex differences in serum dehydroepiandrosterone sulfate concentrations throughout adulthood. J Clin Endocrinol Metab 59:551–555

    Corpechot C, Robel P, Axelson J, Sjovall J, Baulieu EE 1981 Characterization and measurement of dehydroepiandrosterone sulfate in rat brain. Proc Natl Acad Sci USA 78:4704–4707

    Mathur C, Prasad VVK, Raju VS, Welch M, Lieberman S 1993 Steroids and their conjugates in the mammalian brain. Proc Natl Acad Sci USA 90:85–88

    Park-Chung M, Malayev A, Purdy RH, Gibbs TT, Farb DH 1999 Sulfated and unsulfated steroids modulate -aminobutyric A receptor function through distinct sites. Brain Res 830:72–87

    Fahey JM, Lindquist DG, Pritchard GA, Miller LG 1995 Pregnenolone sulfate potentiation of NMDA-mediated increases in intracellular calcium in cultured chick cortical neurons. Brain Res 69:183–188

    Flood JF, Smith GE, Roberts E 1988 Dehydroepiandrosterone and its sulfate enhance memory retention in mice. Brain Res 447:269–278

    Flood JF, Morely JE, Roberts E 1992 Memory-enhancing effects in male mice of pregnenolone and steroids metabolically derived from it. Proc Natl Acad Sci USA 89:1567–1571

    Reed MJ, Purohit A, Duncan LJ, Singh A, Roberts CJ, Williams GJ, Potter BVL 1995 The role of cytokines in regulating oestrogen synthesis in breast tumours. J Steroid Biochem Mol Biol 53:413–420

    Li P-K, Rhodes ME, Jagannathan S, Johnson DA 1995 Reversal of scopolamine induced amnesia in rats by the steroid sulfatase inhibitor estrone-3-O-sulfamate. Brain Res Cogn Brain Res 2:251–254

    Li P-K, Rhodes ME, Burke AM, Johnson DA 1996 Memory enhancement mediated by the steroid sulfatase inhibitor (p-O-sulfamoyl)-N-tetradecanoyl tyramine. Life Sci 60:45–51

    Nicolas LB, Pinoteau W, Papot S, Routier S, Guillamet G, Mortand S 2001 Aggressive behaviour induced by the steroid sulfatase inhibitor COUMATE and by DHEAS in CBA/H mice. Brain Res 922:216–222

    Roubertoux PL, Carlier M, Degrelle H, Haas-Dupertuis MC, Phillips J, Moutier R 1994 Co-segregation of inter-male aggression with the pseudosomal region of the Y chromosome in mice. Genetics 139:225–230

    Le Roy I, Mortand S, Tordjman S, Donez-Darcel E, Carlier M, Degrelle H, Roubertoux PL 1999 Genetic correlation between steroid sulfatase concentration and initiation of attack behaviour in mice. Behav Genet 29:131–136

    Warren JC, French AP 1965 Distribution of steroid sulfatase in human tissues. J Clin Endocrinol Metab 25:278–282

    Hanning RV, Hackett RJ, Boothroid RI, Canick JA 1990 Steroid sulfatase activity in the human ovarian corpus luteum, stroma and follicle: comparison to activity in other tissues and the placenta. J Steroid Biochem 36:175–179

    Dehenin L, Jondet M, Scholler R 1987 Androgen and 19-norsteroid profiles in human preovulatory follicles from stimulated cycles: an isotope dilution mass spectroscopic study. J Steroid Biochem 26:399–405

    Adessi GL, Prost O, Agnani G, Petitjean A, Burnod J 1984 Estrone sulfatase activity in normal and abnormal endometrium. Arch Gynecol 236:13–18

    Zhu BT, Fu JH 1997 Uterine estrogen sulfatase may play a more important role than hepatic sulfatase in mediating the uterotropic action of estrone-3-sulfate. Endocrine 7:191–198

    Milewich L, Porter JC 1987 In situ steroid sulfatase activity in human epithelial carcinoma cells of vaginal, ovarian and endometrial origin. J Clin Endocrinol Metab 65:164–169

    Haning RV, Hackett RJ, Canick JA 1992 Steroid sulfatase in the human ovary and placenta: enzyme kinetics and phosphate inhibition. J Steroid Biochem Mol Biol 41:161–165

    Martel C, Melner MH, Gagne D, Simard J, Labrie F 1994 Widespread distribution of steroid sulfatase, 3?-hydroxysteroid dehydrogenase/5–4 isomerase, 17?-HSD, 5-reductase and aromatase activities in the rhesus monkey. Mol Cell Endocrinol 104:103–111

    Langlais J, Zollinger M, Plante L, Chapelelaine A, Bleau G, Roberts KD 1981 Localization of cholesteryl sulfate in human spermatozoa in support of a hypothesis for the mechanism of capacitation. Proc Natl Acad Sci USA 78:7266–7270

    Huggins C, Scott WW 1945 Bilateral adrenalectomy in prostatic cancer. Ann Surg 122:1031–1041

    Harper ME, Pike A, Peeling WB, Griffiths K 1974 Steroids of adrenal origin metabolized by human prostatic tissue in vivo and in vitro. J Endocrinol 60:117–125

    Labrie F, Dupont A, Belanger A, Lacoursiere Y, Raynaud JP, Husson JM, Gareau J, Fazekas AT, Sandow J, Monfette G, Girard JG, Emond J, Houlc JG 1983 New approach in the treatment of prostate cancer: complete instead of partial withdrawal of androgens. Prostate 4:579–594

    Farnsworth WE 1973 Human prostatic dehydroepiandrosterone sulfate sulfatase. Steroids 21:647–664

    Cowan RA, Cowan SK, Grant JK, Elder HY 1977 Biochemical investigations of separated epithelium and stroma from benign hyperplastic prostatic tissue. J Endocrinol 74:111–120

    Klein H, Molwitz T, Bartsch W 1989 Steroid sulfatase in human benign prostatic hyperplasia: characterization and quantification of the enzyme in epithelium and stroma. J Steroid Biochem 33:195–200

    Selcer KW, Kabler H, Sarap J, Xiao Z, Li P-K 2002 Inhibition of steryl sulfatase activity in LNCaP human prostate cancer cells. Steroids 67:821–826

    Riggs BL, Ryan RJ, Wahner HW, Jiang N, Mattox VR 1973 Serum concentrations of estrogen, testosterone and gonadotrophins in osteoporotic and non-osteoporotic postmenopausal women. J Clin Endocrinol Metab 36:1097–1099

    Gooyer ME, Overlift GT, Kleyn V, Smits KC, Ederveen AGH, Verheul HAM, Kloosterboer HJ 2001 Tibolone: a compound with tissue specific inhibitory effects on sulfatase. Mol Cell Endocrinol 183:55–62

    Kloosterboer HJ 1997 Tibolone and its metabolites: pharmacology, tissue specificity and effects in animal models of tumors. Gynecol Endocrinol 11(Suppl 1):63–68

    Kloosterboer HL 2001 Tibolone: a steroid with a tissue-specific mode of action. J Steroid Biochem Mol Biol 76:231–238

    Soliman HR, Dire D, Bondou P, Julien R, Launay J-M, Brerault J-L, Villette J-M, Fiet J 1993 Characterization of estrone sulfatase activity in human thrombocytes. J Steroid Biochem Mol Biol 46:215–226

    Epstein EH, Leventhal ME 1981 Steroid sulfatase of human leucocytes and epidermis and the diagnosis of recessive X-linked ichthyosis. J Clin Invest 67:1247–1262

    Hirato K, Suzuki T, Hondo T, Saito H, Yanaihara T 1991 Steroid sulfatase activities in human leukocytes: biochemical and clinical aspects. Endocrinol Jpn 38:597–602

    Kelly PM, Davison RS, Bliss E, McGee JO 1988 Macrophages in human breast disease: a quantitative immunohistochemical study. Br J Cancer 57:174–177

    Reed MJ, Lai LC, Owen Am, Singh A, Coldham NG, Ghilchik MW, Shaikh NA, James VHT 1990 Effect of treatment with 4-hydroxyandrostenedione on the peripheral conversion of androstenedione and in vitro tumor aromatase activity in postmenopausal women with breast cancer. Cancer Res 50:193–196

    Purohit A, Froome VA, Wang DY, Potter BVL, Reed MJ 1997 Measurement of oestrone sulfatase activity in white blood cells to monitor in vivo inhibition of steroid sulphatase activity by oestrone-3-O-sulphamate. J Steroid Biochem Mol Biol 62:45–51

    Purohit A, Riaz AA, Ghilchik MW, Reed MJ 1992 The origin of oestrone sulphate in normal and malignant breast tissues in postmenopausal women. Horm Metab Res 24:532–536

    Hargenbuch B, Meier PJ 2003 The super family of organic anion transporting polypeptides. Biochim Biophys Acta 1609:1–18

    Jacquemin E, Hagenbuch R, Stieger B, Wolkoff AW, Meier PJ 1994 Expression cloning of a rat liver Na+-independent organic anion transporter. Proc Natl Acad Sci USA 91:133–137

    Noe IB, Hagenbuch R, Stieger B, Meier PJ 1997 Isolation of a multispecific organic anion and cardiac glycoside transporter from rat brain. Proc Natl Acad Sci USA 94:10346–10350

    Abe T, Kakyo M, Sakagami H, Tokui T, Nishio T, Tanemoto M, Nomura H, Herbert SC, Matsuno S, Kondo H, Yawo H 1998 Molecular characterization and tissue distribution of a new organic anion transporter subtype (Oatp3) that transports thyroid hormones and taurocholate and comparison with Oatp2. J Biol Chem 273:22395–22401

    Kullak-Ublick GA, Hagenbach B, Stieger B, Schteingart CD, Hoffman AF, Wolkoff AW, Meier PJ 1995 Molecular and functional characterization of an organic anion transporting polypeptide cloned from human liver. Gastroenterology 109:1274–1282

    Tamai I, Nezu J-I, Uchino H, Sai Y, Oku A, Shimane M TA 2000 Molecular identification and characterization of novel members of the human organic transporter (OATP) family. Biochem Biophys Res Comm 273:251–260

    Pizzagalli F, Varga Z, Huber RD, Folkers G, Meier PJ, St-Pierre MV 2003 Identification of steroid sulfate transport processes in the human mammary gland. J Clin Endocrinol Metab 88:3902–3912

    Tobacman JK, Hinkhouse M, Khalkhali-Ellis Z 2002 Steroid sulfatase activity and expression in mammary myoepithelial cells. J Steroid Biochem Mol Biol 81:65–68

    Poirier D, Ciobanu LC, Maltais R 1999 Steroid sulfatase inhibitors. Expert Opin Ther Patents 9:1083–1099

    Nussbaumer P, Billich A 2003 Steroid sulfatase inhibitors. Med Res Rev 24:529–576

    Smith HJ, Nicholls PJ, Simons C, LeLain R 2001 Inhibitors of steroidogenesis as agents for the treatment of hormone-dependent cancers. Exp Opin Ther Patents 11:789–824

    Birnb?ck H, von Angerer E 1990 Sulfate derivatives of 2-phenylindoles as novel steroid sulfatase inhibitors. Biochem Pharmacol 39:1709–1713

    Evans TRJ, Rowlands MG, Jarman M, Coombes RC 1991 Inhibition of estrone sulfatase enzyme in human placenta and human breast-carcinoma. J Steroid Biochem Mol Biol 39:493–499

    Wong CK, Keung WM 1997 Daidzein sulfoconjugates are potent inhibitors of sterol sulfatase (EC 3.1.6.2). Biochem Biophys Res Commun 233:579–583

    Anderson CJ, Lucas LJH, Widlanski TS 1995 Molecular recognition in biological systems: phosphate esters vs sulfate esters and the mechanism of action of steroid sulfatases. J Am Chem Soc 117:3889–3890

    Purohit A, Howarth NM, Potter BVL, Reed MJ 1994 Inhibition of steroid sulphatase activity by steroidal methylthiophosphonates: potential therapeutic agents in breast cancer. J Steroid Biochem Mol Biol 48:523–527

    Duncan L, Purohit A, Howarth NM, Potter BVL, Reed MJ 1993 Inhibition of estrone sulfatase activity by estrone-3-methylthiophosphonate—a potential therapeutic agent in breast cancer. Cancer Res 53:298–303

    Howarth NM, Cooper G, Purohit A, Duncan L, Reed MJ, Potter BVL 1993 Phosphonates and thiophosphonates as sulfate surrogates: synthesis of estrone 3-methylthiophosphonate, a potent inhibitor of estrone sulfatase. Bioorg Med Chem Lett 3:313–318

    Howarth NM, Purohit A, Robinson JJ, Vicker N, Reed MJ, Potter BVL 2002 Estrone 3-sulfate mimics, inhibitors of estrone sulfatase activity; homology model construction and docking studies. Biochemistry 41:14801–14814

    Howarth NM, Purohit A, Reed MJ, Potter BVL 1997 Estrone sulfonates as inhibitors of estrone sulfatase. Steroids 62:346–350

    Li P-K, Pillai R, Dibbelt L 1995 Estrone sulfate analogs as estrone sulfatase inhibitors. Steroids 60:299–306

    Li P-K, Pillai R, Young BL, Bender WH, Martino DM, Lin FT 1993 Synthesis and biochemical studies of estrone sulfatase inhibitors. Steroids 58:106–111

    Dibbelt L, Li P-K, Pillai R, Knuppen R 1994 Inhibition of human placental sterylsulfatase by synthetic analogs of estrone sulfate. J Steroid Biochem Mol Biol 50:261–266

    Anderson C, Freeman J, Lucas LH, Farley M, Dalhoumi H, Widlanski TS 1997 Estrone sulfatase: probing structural requirements for substrate and inhibitor recognition. Biochemistry 36:2586–2594

    Howarth NM, Purohit A, Reed MJ, Potter BVL 1994 Estrone sulfamates: potent inhibitors of estrone sulfatase with therapeutic potential. J Med Chem 37:219–221

    Woo LWL, Purohit A, Reed MJ, Potter BVL 1997 Oestrone 3-O-(N-acetyl)sulphamate, a potential molecular probe of the active site of oestrone sulphatase. Bioorg Med Chem Lett 7:3075–3080

    Woo LWL, Lightowler M, Purohit A, Reed MJ, Potter BVL 1996 Heteroatom-substituted analogues of the active site directed inhibitor estra-1,3,5(10)-trien-17-one-3-sulphamate inhibit estrone sulphatase by a different mechanism. J Steroid Biochem Mol Biol 57:79–88

    Selcer KW, Jagannathan S, Rhodes ME, Li PK 1996 Inhibition of placental estrone sulfatase activity and MCF-7 breast cancer cell proliferation by estrone-3-amino derivatives. J Steroid Biochem Mol Biol 59:83–91

    Poirier D, Boivin RP 1998 17-alkyl- or 17--substituted benzyl-17?-estradiols: a new family of estrone sulfatase inhibitors. Bioorg Med Chem Lett 8:1891–1896

    Boivin RP, Luu-The V, Lachance R, Labrie F, Poirier D 2000 Structure-activity relationships of 17-derivatives of estradiol as inhibitors of steroid sulfatase. J Med Chem 43:4465–4478

    Boivin RP, Labrie F, Poirier D 1999 17-Alkan (or alkyn) amide derivatives of estradiol as inhibitors of steroid sulfatase activity. Steroids 64:825–833

    Ciobanu LC, Boivin RP, Luu-The V, Poirier D 2003 3?-Sulfamate derivatives of C19 and C21 steroids bearing a t-butylbenzyl or a benzyl group: synthesis and evaluation as non-estrogenic and non-androgenic steroid sulfatase inhibitors. J Enzyme Inhib Med Chem 18:15–26

    Chu GH, Peters A, Selcer KW, Li PK 1999 Synthesis and sulfatase inhibitory activities of (E)- and (Z)-4-hydroxytamoxifen sulfamates. Bioorg Med Chem Lett 9:141–144

    Golob T, Liebl R, von Angerer E 2002 Sulfamoyloxy-substituted 2-phenylindoles: antiestrogen-based inhibitors of the steroid sulfatase in human breast cancer cells. Bioorg Med Chem 10:3941–3953

    Jütten P, Schumann W, H?rtl A, Heinisch L, Gr?fe U, Werner W, Ulbricht H 2002 A novel type of nonsteroidal estrone sulfatase inhibitors. Bioorg Med Chem Lett 12:1339–1342

    Nussbaumer P, Geyl D, Horvath A, Lehr P, Wolff B, Billich A 2003 Nortropinyl-arylsulfonylureas as novel, reversible inhibitors of human steroid sulfatase. Bioorg Med Chem Lett 13:3673–3677

    Lee W, DeRome M, DeBear J, Noell S, Epstein D, Mahle C, DeCarr L, Woodruff K, Huang Z, Dumas J 2003 Aryl piperazines: a new class of steroid sulfatase inhibitors for the treatment of hormone-dependent breast cancer. J Abstr Papers Amer Chem Soc 226:301-Medi Part 2

    Carlstrom K, Doberl A, Gershagen S, Rannevik G 1984 Peripheral plasma levels of dehydroepiandrosterone sulphate, dehydroepiandrosterone, androstenedione and testosterone following different doses of danazol. Acta Obstet Gynecol Scand 123(Suppl):125–129

    Chetrite G, Paris J, Botella J, Pasqualini JR 1996 Effect of nomegestrol acetate on estrone sulfatase and 17?-hydroxysteroid dehydrogenase activities in human breast cancer cells. J Steroid Biochem Mol Biol 58:525–531

    Prost-Avallet O, Oursin J, Adessi GL 1991 In vitro effect of synthetic progestogens on estrone sulfatase activity in human breast carcinoma. J Steroid Biochem Mol Biol 39:967–973

    Chetrite G, Kloosterboer HJ, Pasqualini JR 1997 Effect of tibolone (ORG OD14) and its metabolites on estrone sulphatase activity in MCF-7 and T-47D mammary cancer cells. Anticancer Res 17:135–140

    Santner SJ, Santen RJ 1993 Inhibition of estrone sulfatase and 17?-hydroxysteroid dehydrogenase by antiestrogens. J Steroid Biochem Mol Biol 45:383–390

    Zhu BT, Kosh JW, Fu J-H, Cai MX, Xu S, Conney AH 2000 Strong inhibition of estrone-3-sulfatase activity by pregnenolone 16-carbonitrile but not by several analogs lacking a 16-nitrile group. Steroids 65:521–527

    Elger W, Schwarz S, Hedden A, Reddersen G, Schneider B 1995 Sulfamates of various estrogens are prodrugs with increased systemic and reduced hepatic estrogenicity at oral application. J Steroid Biochem Mol Biol 55:395–403

    Purohit A, Vernon KA, Hummelinck AEW, Woo LWL, Hejaz HAM, Potter BVL, Reed MJ 1998 The development of A-ring modified analogues of oestrone-3-O-sulphamate as potent steroid sulphatase inhibitors with reduced oestrogenicity. J Steroid Biochem Mol Biol 64:269–275

    Purohit A, Hejaz HAM, Woo LWL, van Strien AE, Potter BVL, Reed MJ 1999 Recent advances in the development of steroid sulphatase inhibitors. J Steroid Biochem Mol Biol 69:227–238

    Raobaikady B, Purohit A, Chander SK, Woo LWL, Leese MP, Potter BVL, Reed MJ 2003 Inhibition of MCF-7 breast cancer cell proliferation and in vivo steroid sulphatase activity by 2-methoxyoestradiol-bis-sulphamate. J Steroid Biochem Mol Biol 84:351–358

    Peters RH, Chao W-R, Sato B, Shigeno K, Zaveri NT, Tanabe M 2003 Steroidal oxathiazine inhibitors of estrone sulfatase. Steroids 68:97–110

    Hejaz HAM, Purohit A, Mahon MF, Reed MJ, Potter BVL 1999 Synthesis and biological activity of the superestrogen (E)-17-oximino-3-O-sulfamoyl-1,3,5(10)-estratriene: x-ray crystal structure of (E)-17-oximino-3-hydroxy-1,3,5(10)-estratriene. J Med Chem 42:3188–3192

    Li P-K, Chu GH, Guo JP, Peters A, Selcer KW 1998 Development of potent non-estrogenic estrone sulfatase inhibitors. Steroids 63:425–432

    Ciobanu LC, Boivin RP, Luu-The V, Labrie F, Poirier D 1999 Potent inhibition of steroid sulfatase activity by 3-O-sulfamate of 17-benzyl (or 4'-tert-butylbenzyl)estra-1,3,5(10)-trienes: combination of two substituents at positions C3 and C17 of estradiol. J Med Chem 42:2280–2286

    Ciobanu LC, Luu-The V, Poirier D 2002 Nonsteroidal compounds designed to mimic potent steroid sulfatase inhibitors. J Steroid Biochem Mol Biol 80:339–353

    Ciobanu LC, Luu-The V, Martel C, Labrie F, Poirier D 2003 Inhibition of estrone sulfate-induced uterine growth by potent nonestrogenic steroidal inhibitors of steroid sulfatase. Cancer Res 63:6442–6446

    Ciobanu LC, Poirier D 2003 Solid-phase parallel synthesis of 17-substituted estradiol sulfamate and phenol libraries using the multidetachable sulfamate linker. J Comb Chem 5:429–440

    Fischer DS, Chander SK, Woo LWL, Fenton JC, Purohit A, Reed MJ, Potter BVL 2003 Novel D-ring modified steroid derivatives as potent, non-estrogenic, steroid sulfatase inhibitors with in vivo activity. J Steroid Biochem Mol Biol 84:343–349

    Li P-K, Milano S, Kluth L, Rhodes ME 1996 Synthesis and sulfatase inhibitory activities of non-steroidal estrone sulfatase inhibitors. J Steroid Biochem Mol Biol 59:41–48

    Selcer KW, Hegde PV, Li P-K 1997 Inhibition of estrone sulfatase and proliferation of human breast cancer cells by nonsteroidal (p-O-sulfamoyl)-N-alkanoyl tyramines. Cancer Res 57:702–707

    Chu G-H, Milano S, Kluth L, Rhodes M, Boni R, Johnson DA, Li P-K 1997 Structure-activity relationship studies of the amide functionality in (p-O-sulfamoyl)-N-alkanoyl tyramines as estrone sulfatase inhibitors. Steroids 62:530–535

    Kolli A, Chu GH, Rhodes ME, Inoue K, Selcer KW, Li P-K 1999 Development of (p-O-sulfamoyl)-N-alkanoyl-phenylalkyl amines as non-steroidal estrone sulfatase inhibitors. J Steroid Biochem Mol Biol 68:31–40

    Patel CK, Owen CP, Ahmed S 2003 The design, synthesis, and in vitro biochemical evaluation of a series of esters of 4-[(aminosulfonyl)oxy]benzoate as novel and highly potent inhibitors of estrone sulfatase. Biochem Biophys Res Commun 307:778–781

    Woo LWL, Purohit A, Malini B, Reed MJ, Potter BVL 2000 Potent active site-directed inhibition of steroid sulfatase by tricyclic coumarin-based sulphamates. Chem Biol 7:773–791

    Woo LWL, Purohit A, Reed MJ, Potter BVL 1996 Active site directed inhibition of estrone sulfatase by nonsteroidal coumarin sulfamates. J Med Chem 39:1349–1351

    Woo LWL, Howarth NM, Purohit A, Hejaz HAM, Reed MJ, Potter BVL 1998 Steroidal and nonsteroidal sulfamates as potent inhibitors of steroid sulfatase. J Med Chem 41:1068–1083

    Purohit A, Woo LWL, Barrow D, Hejaz H, Nicholson RI, Potter BVL, Reed MJ 2001 Non-steroidal and steroidal sulfamates: new drugs for cancer therapy. Mol Cell Endocrinol 171:129–135

    Malini B, Purohit A, Ganeshapillai D, Woo LWL, Potter BVL, Reed MJ 2000 Inhibition of steroid sulphatase activity by tricyclic coumarin sulphamates. J Steroid Biochem Mol Biol 75:253–258

    Hejaz HAM 1998 The synthesis of steroidal and non-steroidal steroid sulphatase inhibitors for endocrine therapy of breast cancer, PhD thesis, University of Bath, Bath, United Kingdom

    Nussbaumer P, Lehr P, Billich A 2002 2-Substituted 4-(thio)chromosome 6-O-sulfamates: potent inhibitors of human steroid sulfatase. J Med Chem 45:4310–4320

    Nussbaumer P, Winiski AP, Billich A 2003 Estrogenic potential of 2-alkyl-4-(thio)chromenone 6-O-sulfamates, potent inhibitors of human steroid sulfatase. J Med Chem 46:5091–5094

    Nussbaumer P, Bilban M, Billich A 2002 4,4'-Benzophenone-O,O'-disulfamate: a potent inhibitor of human steroid sulfatase. Bioorg Med Chem Lett 12:2093–2095

    Hejaz HAM, Woo LWL, Purohit A, Reed MJ, Potter BVL 2004 Synthesis, in vitro and in vivo activity of benzophenone-based inhibitors of steroid sulphatase. Bioorg Med Chem 12:2759–2772

    Ahmed S, James K, Owen CP, Patel CK 2002 Design, synthesis and biochemical evaluation of AC ring mimics as novel inhibitors of the enzyme estrone sulfatase (ES). Bioorg Med Chem Lett 12:1343–1346

    Fischer DS, Woo LWL, Mahon MF, Purohit A, Reed MJ, Potter BVL 2003 D-ring modified estrone derivatives as novel potent inhibitors of steroid sulfatase. Bioorg Med Chem 11:1685–1700

    Woo LWL, Sutcliffe OB, Bubert C, Grasso A, Chander SK, Purohit A, Reed MJ, Potter BVL 2003 First dual aromatase-sulfatase inhibitors. J Med Chem 46:3193–3196

    Ahmed S, James K, Owen CP, Patel CK, Sampson L 2002 The mechanism of the irreversible inhibition of estrone sulfatase (ES) through the consideration of a range of methane- and amino-sulfonate-based compounds. Bioorg Med Chem Lett 12:1279–1282

    Purohit A, Williams GJ, Roberts CJ, Potter BVL, Reed MJ 1995 In vivo inhibition of oestrone sulphatase and dehydroepiandrosterone sulphatase by oestrone-3-O-sulphamate. Int J Cancer 63:106–111

    Purohit A, Woo LWL, Singh A, Winterborn CJ, Potter BVL, Reed MJ 1996 In vivo activity of 4-methylcoumarin-7-O-sulfamate, a nonsteroidal, nonestrogenic steroid sulfatase inhibitor. Cancer Res 56:4950–4955

    Purohit A, Woo LWL, Potter BVL, Reed MJ 2000 In vivo inhibition of estrone sulfatase activity and growth of nitrosomethylurea-induced mammary tumors by 667COUMATE. Cancer Res 60:3394–3396

    Kellis JT, Vickery LE 1984 Inhibition of human estrogen synthase (aromatase) by flavones. Science 225:1032–1034

    Ibrahim AR, Abul-Hajj YJ 1990 Aromatase inhibition by flavonoids. J Steroid Biochem Mol Biol 37:257–260

    Purohit A, Woo LWL, Chander SK, Newman SP, Ireson C, Ho Y, Grasso A, Leese MP, Potter BVL, Reed MJ 2003 Steroid sulphatase inhibitors for breast cancer. J Steroid Biochem Mol Biol 86:423–432

    Kudoh M, Susaki Y, Ideyama Y, Nanya T, Okada M, Shikama H, Fujikura T 1995 The potent and selective inhibition of estrogen production by non-steroidal aromatase inhibitor, YM 511. J Steroid Biochem Mol Biol 54:265–271

    Tominaga T, Suzuki T 2003 Early phase II study with new aromatase inhibitor YM 511 in postmenopausal patients with breast cancer. Difficulty in clinical dose recommendation based on preclinical and phase I findings. Anticancer Res 23:3533–3542

    Patton TL, Dmochowski L 1963 Estrogens. V. Studies on the relationship of estrogenic activity and molecular structure. Arch Biochem Biophys 101:181–185

    Dorfman RI, Kincl FA 1966 Uterotrophic activity of various phenolic steroids. Acta Endocrinol (Copenh) 52:619–626

    Leese ML, Newman SP, Purohit A, Reed MJ, Potter BVL 2004 2-Alkylsulfanyl estrogen derivatives: synthesis of a novel class of multi-targeted anti-tumour agents. Bioorg Med Chem Lett 15:3135–3138

    Seegers J, Aveling M-L, Van Aswegen CH, Cross M, Koch F, Joubert WS 1989 The cytotoxic effects of estradiol-17?, catecholestradiols and methoxyestradiols on dividing MCF-7 and HeLa cells. J Steroid Biochem 32:797–809

    Pribluda VS, Gubish ER, LaVallee TM, Treston A, Swartz GM, Green S 2000 2-Methoxyestradiol: an endogenous antiangiogenic and antiproliferative drug candidate. Cancer Metastasis Rev 19:173–179

    Bradlow HL, Telang NT, Sepkovic DW, Osbourne MP 1996 2-Hydroxyestrone the ’good’ estrogen. J Endocrinol 150:S259–S265

    Zhu BT, Conney AH 1998 Is 2-methoxyestradiol an endogenous estrogen metabolite that inhibits mammary carcinogenesis. Cancer Res 58:2269–2277

    Fotsis T, Zhang Y, Pepper MS, Adlercreutz H, Montesano R, Nawroth PP, Schweigerer L 1994 The endogenous oestrogen metabolite 2-methoxyoestradiol inhibits angiogenesis and suppresses tumour growth. Nature 368:237–239

    Klauber N, Parangi S, Flynn E, Hamel E, D’Amato RJ 1997 Inhibition of angiogenesis and breast cancer in mice by the microtubule inhibitors 2-methoxyestradiol and taxol. Cancer Res 57:81–86

    Purohit A, Hejaz HAM, Walden L, MacCarthy-Morrogh L, Packham G, Potter BVL, Reed MJ 2000 The effect of 2-methoxyestrone-3-O-sulphamate on the growth of breast cancer cells and induced mammary tumours. Int J Cancer 85:584–589

    MacCarthy-Morrogh L, Townsend PA, Purohit A, Hejaz HAM, Potter BVL, Reed MJ, Packham G 2000 Differential effects of estrone and estrone-3-O-sulfamate derivatives on mitotic arrest, apoptosis and microtubule assembly in human breast cancer cells. Cancer Res 60:5441–5450

    Ireson C, Chander SK, Purohit A, Perera S, Newman SP, Parish D, Leese MP, Smith AC, Potter BVL, Reed MJ 2004 Pharmacokinetics and efficacy of 2-methoxyestradiol and 2-methoxyestradiol-bis-sulphamate in vivo in rodents. Br J Cancer 90:932–937

    Ho YT, Purohit A, Vicker N, Newman SP, Robinson JJ, Leese ML, Ganeshapillai D, Woo LWL, Potter BVL, Reed MJ 2003 Inhibition of carbonic anhydrase II by steroidal and non-steroidal sulphamates. Biochem Biophys Res Commun 305:909–914

    LaVallee TM, Zhan XH, Herbstritt CJ, Kough EC, Green SJ, Pribluda VS 2002 2-Methoxyestradiol inhibits proliferation and induces apoptosis independently of estrogen receptors and ?. Cancer Res 62:3691–3697

    Mabjeesh NJ, Escuin D, LaVallee TM, Pribluda VS, Swartz GM, Johnson MS, Willard MT, Zhong H, Simons JW, Giannakakou P 2003 2-ME2 inhibits tumor growth and angiogenesis by disrupting microtubules and dysregulating HIF. Cancer Cell 3:363–375

    Singh A, Purohit A, Hejaz HAM, Potter BVL, Reed MJ 2000 Inhibition of deoxyglucose uptake in MCF-7 breast cancer cells by 2-methoxyestrone and 2-methoxyestrone-3-O-sulfamate. Mol Cell Endocrinol 160:61–66

    Newman SP, Leese MP, Purohit A, James DRC, Rennie CE, Potter BVL, Reed MJ 2004 Inhibition of in vitro angiogenesis by 2-methoxy- and 2-ethyl-estrogen sulfamates. Int J Cancer 109:533–540

    Lobo RA, Cassidenti DL 1992 Pharmacokinetics of oral 17?-estradiol. J Reprod Med 37:77–84

    Mashchak CA, Lobo RA, Dozono-Takano R, Eggena P, Nakamura RM, Brenner PF, Mishell DR 1982 Comparison of the pharmacodynamic properties of various estrogen formulations. Am J Obstet Gynecol 144:511–518

    Elger W, Barth A, Hedden A, Reddersen G, Ritter P, Schneider B, Zuchner J, Krahl E, Muller K, Oettel M, Schwarz S 2001 Estrogen sulfamates: a new approach to oral estrogen therapy. Reprod Fertil Dev 13:297–305

    Ireson CR, Parish D, Purohit A, Woo LWL, Potter BVL, Chander SK, Reed MJ 2003 Development of a sensitive high-performance liquid chromatography method for the detection of 667 COUMATE in vivo. J Steroid Biochem Mol Biol 84:337–342

    Lin JH, Lin TH, Cheng H 1992 Uptake and stereoselective binding of the enantiomers of MK-927, a potent carbonic anhydrase inhibitor, by human erythrocytes in vitro. Pharm Res 9:339–344

    Racacha R, Costanzo MJ, Maryanoff BE, Chattopadhyay D 2002 Crystal structure of human carbonic anhydrase II with an anti-convulsant sugar sulphamate. Biochem J 361:437–441

    Vicker N, Ho YT, Robinson J, Woo LWL, Purohit A, Reed MJ, Potter BVL 2003 Docking studies of sulphamate inhibitors of estrone sulphatase in human carbonic anhydrase II. Bioorg Med Chem Lett 13:863–865

    Abbate F, Winum J-Y, Potter BVL, Casini A, Montero J-L, Scozzafava A, Supuran CT 2004 Carbonic anhydrase inhibitors: x ray crystallographic structure of the adduct of human isozyme II with EMATE, a dual inhibitor of carbonic anhydrases and steroid sulfatase. Bioorg Med Chem Lett 14:231–234

    Sly WS, Hu PY 1995 Human carbonic anhydrases and carbonic anhydrase deficiencies. Annu Rev Biochem 64:375–401

    McKierman JM, Buttyan R, Stifelman MD, Katz AE, Chen MW, Olsson CA, Sawczuk IS 1997 Expression of tumor-associated gene MN: a potential biomarker of human renal cell carcinoma. Cancer Res 57:2362–2365

    Tureci O, Sahin U, Vollmar E, Siemer S, Gottert E, Seitz G, Parkkila A-K, Shah GN, Grubb JH, Pfreundschuh M, Sly WS 1998 Human carbonic anhydrase XII: cDNA cloning, expression and chromosomal localization of a carbonic anhydrase gene that is over expressed in some renal cancers. Proc Natl Acad Sci USA 95:7608–7613

    Potter CPS, Harris AL 2003 Diagnostic, prognostic and therapeutic implications of carbonic anhydrases in cancer. Br J Cancer 89:2–7

    Loncaster JA, Harris AL, Davison SE, Logue JP, Hunter RD, Wycoff CC, Pastorek J, Ratcliffe PJ, Stratford IJ, West CML 2001 Carbonic anhydrase IX expression, a potential new intrinsic marker of hypoxia: correlations with tumor oxygen measurements and prognosis in locally advanced carcinoma of the cervix. Cancer Res 61:6394–6399

    Harris AL 2001 Hypoxia—a key regulatory factor in tumour growth. Nat Cancer Rev 2:38–47

    Koukourakis MI, Giatromanolaki A, Siuridis E, Simopoulos K, Pastorek J, Harris AL 2001 Hypoxia-regulated carbonic anhydrase-9 relates to poor vascularization and resistance of squamous cell head and neck cancer to chemoradiotherapy. Clin Cancer Res 7:3399–3403

    Whittington DA, Waheed A, Ulmasov B, Shah GN, Grubb JH, Sly WS, Christianson DW 2001 Crystal structure of the dimeric extracellular domain of human carbonic anhydrase XII, a bitopic membrane protein over expressed in certain cancer tumor cells. Proc Natl Acad Sci USA 98:9545–9550

    Winum J-Y, Vullo D, Casini A, Montero J-L, Scozzafava A, Supuran CT 2003 Carbonic anhydrase inhibitors. Inhibition of cytosolic isozymes I and II and transmembrane, tumor-associated isozyme IX with sulfamates including EMATE also acting as steroid sulfatase inhibitors. J Med Chem 46:2197–2204

    Parkkila S, Rajaniemi H, Parkkila A-K, Kivela J, Waheed A, Pastorekova S, Pastorek J, Sly WS 2000 Carbonic anhydrase inhibitor suppresses invasion of renal cancer cells in vitro. Proc Natl Acad Sci USA 97:2220–2224

    Teicher BA, Liu S-D, Liu J-T, Holden SA, Herman TS 1993 A carbonic anhydrase inhibitor as a potential modulator of cancer therapies. Anticancer Res 13:1549–1556

    Purohit A, Potter BVL, Parker MG, Reed MJ 1998 Steroid sulphatase: expression, isolation and inhibition for active-site identification studies. Chem Biol Interact 109:183–193

    Hernandez-Guzman FG, Higashiyama T, Osawa Y, Ghosh D 2001 Purification, characterization and crystallisation of human placental estrone/dehydroepiandrosterone sulfatase, a membrane-bound enzyme of the endoplasmic reticulum. J Steroid Biochem Mol Biol 78:441–450

    Hernandez-Guzman FG, Higashiyama T, Pangborn W, Osawa Y, Ghosh D 2003 Structure of human estrone sulfatase suggests functional roles of membrane association. J Biol Chem 278:22989–22997

    Lukatela G, Krauss N, Theis K, Selmer T, Gieselmann V, von Figura K, Saenger W 1998 Crystal structure of human aryl-sulfatase A: the aldehyde function and the metal ion at the active site suggest a novel mechanism for sulfatase ester hydrolysis. Biochemistry 37:3654–3664

    Schmidt B, Selmer T, Ingendoh A, von Figura K 1995 A novel amino acid modification in sulfatases that is defective in multiple sulfatase deficiency. Cell 82:271–278

    Dierks T, Schmidt B, von Figura K 1997 Conversion of cysteine to formylglycine: a protein modification in the endoplasmic reticulum. Proc Natl Acad Sci USA 94:11963–11986

    Selmer T, Hallmann A, Schmidt B, Sumper M, von Figura K 1996 The evolutionary conservation of a novel protein modification, the conversion of cysteine to serine-semialdehyde in arylsulfatase from Volvox carteri. Eur J Biochem 238:341–345

    Recksiek M, Selmer T, Dierks T, Schmidt B, von Figura K 1998 Sulfatases, trapping of the sulphated enzyme intermediate by substituting the active site formylglycine. J Biol Chem 273:6096–6103

    Sahm UG, Williams GJ, Purohit A, Hidalgo Aragones MI, Parish D, Reed MJ, Potter BVL, Poulton CW 1996 Development of an oral formulation for oestrone 3-O-sulphamate, a potent sulphatase inhibitor. Pharm Sci 2:17–20

    Williams A, Douglas KT 1974 Hydrolysis of aryl N-methylaminosulphonates: evidence consistent with an E1cB mechanism. J Chem Soc Perkins Trans 2:1727–1732

    Spillane WJ, Hogan G, McGrath P, King J, Brack C 1996 Aminolysis of sulfamate esters in non-aqueous solvents. Evidence consistent with a concerted E2-type mechanism. J Chem Soc Perkins Trans 2:2099–2104

    Purohit A, Williams GA, Howarth NM, Potter BVL, Reed MJ 1995 Inactivation of steroid sulfatase by an active site-directed inhibitor, estrone-3-O-sulfamate. Biochemistry 34:11508–11514

    Waldow A, Schmidt B, Dierks T, von Bülow R, von Figura K 1999 Amino acid residues forming the active site of arylsulfatase A: role in catalytic activity and substrate binding. J Biol Chem 274:12284–12288

    Fersht A 1984 The pH dependence of enzyme catalysis. In: Enzyme structure and mechanism. Chap 5. New York: W. H. Freeman and Co.; 156

    Stanway S, Purohit A, Woo LWL, Sufi S, Dobbs N, Elliot M, Potter BVL, Reed MJ, Coombes RC2004 First phase I trial of a steroid sulfatase inhibitor in breast cancer. J Clin Oncol 22 (Suppl 14S) Abstract 64S

    Zaveri NT, Chao W-R, Peters R, Tanabe M, Yamada Y, Toko T, Asao T, Eshima K A novel dual-targeted approach for the endocrine therapy and prevention of breast cancer: SR 16158, a novel steroid sulfatase inhibitor/tissue-selective antiestrogen. Proc American Association for Cancer Research, Advances in Breast Cancer Res, Huntingdon Beach, CA, 2003 (Abstract B25)(M. J. Reed, A. Purohit, L)