Estrone_3-sulfate

Estrone sulfate

Estrone sulfate

Chemical compound


Estrone sulfate, also known as E1S, E1SO4 and estrone 3-sulfate, is a natural, endogenous steroid and an estrogen ester and conjugate.[1][2][3]

Quick Facts Names, Identifiers ...

In addition to its role as a natural hormone, estrone sulfate is used as a medication, for instance in menopausal hormone therapy; for information on estrone sulfate as a medication, see the estrone sulfate (medication) article.

Biological function

E1S itself is biologically inactive, with less than 1% of the relative binding affinity of estradiol for the ERα and ERβ.[3][4] However, it can be transformed by steroid sulfatase, also known as estrogen sulfatase, into estrone, an estrogen.[5] Simultaneously, estrogen sulfotransferases, including SULT1A1 and SULT1E1, convert estrone to E1S, resulting in an equilibrium between the two steroids in various tissues.[1][5] Estrone can also be converted by 17β-hydroxysteroid dehydrogenases into the more potent estrogen estradiol.[1] E1S levels are much higher than those of estrone and estradiol, and it is thought to serve as a long-lasting reservoir for estrone and estradiol in the body.[1][6][7] In accordance, E1S has been found to transactivate the estrogen receptor at physiologically relevant concentrations.[8][9] This was diminished with co-application of irosustat (STX-64), a steroid sulfatase inhibitor, indicating the importance of transformation of estrone sulfate into estrone in the estrogenicity of E1S.[8][9]

Unlike unconjugated estradiol and estrone, which are lipophilic compounds, E1S is an anion and is hydrophilic.[10][11][12] As a result of this, whereas estradiol and estrone are able to readily diffuse through the lipid bilayers of cells, E1S is unable to permeate through cell membranes.[10][11][12] Instead, estrone sulfate is transported into cells in a tissue-specific manner by active transport via organic-anion-transporting polypeptides (OATPs), including OATP1A2, OATP1B1, OATP1B3, OATP1C1, OATP2B1, OATP3A1, OATP4A1, and OATP4C1, as well as by the sodium-dependent organic anion transporter (SOAT; SLC10A6).[11][12][13][14]

E1S, serving as a precursor and intermediate for estrone and estradiol, may be involved in the pathophysiology of estrogen-associated diseases including breast cancer, benign breast disease, endometrial cancer, ovarian cancer, prostate cancer, and colorectal cancer.[1][15][16] For this reason, enzyme inhibitors of steroid sulfatase and 17β-hydroxysteroid dehydrogenase and inhibitors of OATPs, which prevent activation of E1S into estrone and estradiol, are of interest in the potential treatment of such conditions.[1][16][15]

More information Estrogen, Other names ...

Chemistry

E1S, also known as estrone 3-sulfate or as estra-1,3,5(10)-trien-17-one 3-sulfate, is a naturally occurring estrane steroid and a derivative of estrone.[17] It is an estrogen conjugate or ester, and is specifically the C3 sulfate ester of estrone.[17] Related estrogen conjugates include estradiol sulfate, estriol sulfate, estrone glucuronide, estradiol glucuronide, and estriol glucuronide, while related steroid conjugates include dehydroepiandrosterone sulfate and pregnenolone sulfate.

The logP of E1S is 1.4.[15]

Biochemistry

Biosynthesis

E1S is produced via estrogen sulfotransferases from the peripheral metabolism of the estrogens estradiol and estrone.[18][19][20] Estrogen sulfotransferases are expressed minimally or not at all in the gonads.[21] In accordance, E1S is not secreted in meaningful amounts from the gonads in humans.[22][18] However, measurable amounts of estrogen sulfates are said to be secreted by the ovaries in any case.[23]

More information Sex, Sex hormone ...

Distribution

Whereas free steroids like estradiol are lipophilic and can enter cells via passive diffusion, steroid conjugates like E1S are hydrophilic and are unable to do so.[24][25] Instead, steroid conjugates require active transport via membrane transport proteins to enter cells.[24][25]

Studies in animals and humans have had mixed findings on uptake of exogenously administered E1S in normal and tumorous mammary gland tissue.[26][27][28][24][25] This is in contrast to substantial uptake of exogenously administered estradiol and estrone by the mammary glands.[26] Another animal study found that E1S wasn't taken up by the uterus but was taken up by the liver, where it was hydrolyzed into estrone.[29][26]

Metabolism

The elimination half-life of E1S is 10 to 12 hours.[3] Its metabolic clearance rate is 80 L/day/m2.[3]

Ovarian tumors have been found to express steroid sulfatase and have been found to convert E1S into estradiol.[30][31] This may contribute to the often elevated levels of estradiol observed in women with ovarian cancer.[30][31]

The image above contains clickable links
Description: The metabolic pathways involved in the metabolism of estradiol and other natural estrogens (e.g., estrone, estriol) in humans. In addition to the metabolic transformations shown in the diagram, conjugation (e.g., sulfation and glucuronidation) occurs in the case of estradiol and metabolites of estradiol that have one or more available hydroxyl (–OH) groups. Sources: See template page.

Levels

Estrogen levels with radioimmunoassay (RIA) around mid-cycle during the normal menstrual cycle in women.[32][33] The vertical dashed line in the center is mid-cycle.

E1S levels have been characterized in humans.[33][34][35] E1S using radioimmunoassay (RIA) have been reported to be 0.96 ± 0.11 ng/mL in men, 0.96 ± 0.17 ng/mL during the follicular phase in women, 1.74 ± 0.32 ng/mL during the luteal phase in women, 0.74 ± 0.11 ng/mL in women taking oral contraceptives, 0.13 ± 0.03 ng/mL in postmenopausal women, and 2.56 ± 0.47 ng/mL in postmenopausal women on menopausal hormone therapy.[35] In addition, E1S levels in pregnant women were 19 ± 5 ng/mL in the first trimester, 66 ± 31 ng/mL in the second trimester, and 105 ± 22 ng/mL in the third trimester.[35] E1S levels are about 10 to 15 times higher than those of estrone in women.[36]


References

  1. Rezvanpour A, Don-Wauchope AC (March 2017). "Clinical implications of estrone sulfate measurement in laboratory medicine". Crit Rev Clin Lab Sci. 54 (2): 73–86. doi:10.1080/10408363.2016.1252310. PMID 27960570. S2CID 1825531.
  2. Kuiper GG, Carlsson B, Grandien K, Enmark E, Häggblad J, Nilsson S, Gustafsson JA (March 1997). "Comparison of the ligand binding specificity and transcript tissue distribution of estrogen receptors alpha and beta". Endocrinology. 138 (3): 863–70. doi:10.1210/endo.138.3.4979. PMID 9048584.
  3. Falcone T, Hurd WW (22 May 2013). Clinical Reproductive Medicine and Surgery: A Practical Guide. Springer Science & Business Media. pp. 5–6. ISBN 978-1-4614-6837-0.
  4. Melmed S, Polonsky KS, Larsen PR, Kronenberg HM (11 November 2015). Williams Textbook of Endocrinology (13th ed.). Elsevier Health Sciences. pp. 607–. ISBN 978-0-323-34157-8.
  5. Bjerregaard-Olesen C, Ghisari M, Kjeldsen LS, Wielsøe M, Bonefeld-Jørgensen EC (January 2016). "Estrone sulfate and dehydroepiandrosterone sulfate: Transactivation of the estrogen and androgen receptor". Steroids. 105: 50–8. doi:10.1016/j.steroids.2015.11.009. PMID 26666359. S2CID 46663814.
  6. Clark BJ, Prough RA, Klinge CM (2018). "Mechanisms of Action of Dehydroepiandrosterone". Dehydroepiandrosterone. Vitamins and Hormones. Vol. 108. pp. 29–73. doi:10.1016/bs.vh.2018.02.003. ISBN 978-0-12-814361-2. ISSN 0083-6729. PMID 30029731.
  7. Purohit A, Woo LW, Potter BV (July 2011). "Steroid sulfatase: a pivotal player in estrogen synthesis and metabolism" (PDF). Mol. Cell. Endocrinol. 340 (2): 154–60. doi:10.1016/j.mce.2011.06.012. PMID 21693170. S2CID 14296237.
  8. Africander D, Storbeck KH (May 2018). "Steroid metabolism in breast cancer: Where are we and what are we missing?". Mol. Cell. Endocrinol. 466: 86–97. doi:10.1016/j.mce.2017.05.016. PMID 28527781. S2CID 4547808.
  9. Mueller JW, Gilligan LC, Idkowiak J, Arlt W, Foster PA (October 2015). "The Regulation of Steroid Action by Sulfation and Desulfation". Endocr. Rev. 36 (5): 526–63. doi:10.1210/er.2015-1036. PMC 4591525. PMID 26213785.
  10. Banerjee N, Fonge H, Mikhail A, Reilly RM, Bendayan R, Allen C (2013). "Estrone-3-sulphate, a potential novel ligand for targeting breast cancers". PLOS ONE. 8 (5): e64069. Bibcode:2013PLoSO...864069B. doi:10.1371/journal.pone.0064069. PMC 3661587. PMID 23717534.
  11. Gilligan LC, Gondal A, Tang V, Hussain MT, Arvaniti A, Hewitt AM, Foster PA (2017). "Estrone Sulfate Transport and Steroid Sulfatase Activity in Colorectal Cancer: Implications for Hormone Replacement Therapy". Front Pharmacol. 8: 103. doi:10.3389/fphar.2017.00103. PMC 5339229. PMID 28326039.
  12. Longcope C, Flood C, Tast J (1994). "The metabolism of estrone sulfate in the female rhesus monkey". Steroids. 59 (4): 270–273. doi:10.1016/0039-128X(94)90112-0. ISSN 0039-128X. PMID 8079382. S2CID 42846339. The source of E1SO4 in humans is from the peripheral conversion of E1 and E2, 6,7 [...] In human females there is little evidence for the ovarian secretion of E1SO4. 7 Since most of our monkeys were ovariectomized, we cannot say that the rhesus ovaries do not secrete E1SO4, but it is probably unlikely.
  13. Ruder HJ, Loriaux L, Lipsett MB (1972). "Estrone Sulfate: Production Rate and Metabolism in Man". Journal of Clinical Investigation. 51 (4): 1020–1033. doi:10.1172/JCI106862. ISSN 0021-9738. PMC 302214. PMID 5014608.
  14. Longcope C (1972). "The Metabolism of Estrone Sulfate in Normal Males". The Journal of Clinical Endocrinology & Metabolism. 34 (1): 113–122. doi:10.1210/jcem-34-1-113. ISSN 0021-972X. PMID 5008222.
  15. Hobkirk R (1985). "Steroid sulfotransferases and steroid sulfate sulfatases: characteristics and biological roles". Canadian Journal of Biochemistry and Cell Biology. 63 (11): 1127–1144. doi:10.1139/o85-141. ISSN 0714-7511. PMID 3910206.
  16. Strauss JF (2019). "Steroid Hormones and Other Lipid Molecules Involved in Human Reproduction". In Jerome F. Strauss, Robert L. Barbieri (eds.). Yen & Jaffe's Reproductive Endocrinology: Physiology, Pathophysiology, and Clinical Management (8 ed.). Elsevier Health Sciences. pp. 75–114. doi:10.1016/B978-0-323-47912-7.00004-4. ISBN 978-0-323-58232-2. S2CID 90621016.
  17. Brooks, S. C., Horn, L., Pack, B. A., Rozhin, J., Hansen, E., & Goldberg, R. (1980). Estrogen metabolism and function in vivo and in vitro. In Estrogens in the Environment (Vol. 5, pp. 147-167). Elsevier/North Holland New York.
  18. Reed MJ, Purohit A, Woo LW, Newman SP, Potter BV (April 2005). "Steroid sulfatase: molecular biology, regulation, and inhibition". Endocr. Rev. 26 (2): 171–202. doi:10.1210/er.2004-0003. PMID 15561802.
  19. Geisler J (September 2003). "Breast cancer tissue estrogens and their manipulation with aromatase inhibitors and inactivators". J. Steroid Biochem. Mol. Biol. 86 (3–5): 245–53. doi:10.1016/s0960-0760(03)00364-9. PMID 14623518. S2CID 23065230.
  20. Purohit A, Riaz AA, Ghilchik MW, Reed MJ (November 1992). "The origin of oestrone sulphate in normal and malignant breast tissues in postmenopausal women". Horm. Metab. Res. 24 (11): 532–6. doi:10.1055/s-2007-1003382. PMID 1452119. S2CID 260167615.
  21. Masamura S, Santner SJ, Santen RJ (July 1996). "Evidence of in situ estrogen synthesis in nitrosomethylurea-induced rat mammary tumors via the enzyme estrone sulfatase". J. Steroid Biochem. Mol. Biol. 58 (4): 425–9. doi:10.1016/0960-0760(96)00065-9. PMID 8903427. S2CID 22100628.
  22. Thijssen JH (September 2004). "Local biosynthesis and metabolism of oestrogens in the human breast". Maturitas. 49 (1): 25–33. doi:10.1016/j.maturitas.2004.06.004. PMID 15351093.
  23. Holinka CF, Gurpide E (April 1980). "In vivo uptake of estrone sulfate by rabbit uterus". Endocrinology. 106 (4): 1193–7. doi:10.1210/endo-106-4-1193. PMID 7358033.
  24. Day JM, Purohit A, Tutill HJ, Foster PA, Woo LW, Potter BV, Reed MJ (2009). "The Development of Steroid Sulfatase Inhibitors for Hormone-Dependent Cancer Therapy". Annals of the New York Academy of Sciences. 1155 (1): 80–87. Bibcode:2009NYASA1155...80D. doi:10.1111/j.1749-6632.2008.03677.x. ISSN 0077-8923. PMID 19250195. S2CID 25306673.
  25. Kirilovas D, Schedvins K, Naessén T, Von Schoultz B, Carlström K (2009). "Conversion of circulating estrone sulfate to 17β-estradiol by ovarian tumor tissue: A possible mechanism behind elevated circulating concentrations of 17β-estradiol in postmenopausal women with ovarian tumors". Gynecological Endocrinology. 23 (1): 25–28. doi:10.1080/09513590601058333. ISSN 0951-3590. PMID 17484508. S2CID 25115594.
  26. Pasqualini JR, Gelly C, Nguyen BL (1990). "Metabolism and biologic response of estrogen sulfates in hormone-dependent and hormone-independent mammary cancer cell lines. Effect of antiestrogens". Ann. N. Y. Acad. Sci. 595 (1): 106–16. Bibcode:1990NYASA.595..106P. doi:10.1111/j.1749-6632.1990.tb34286.x. PMID 2375600. S2CID 26940935.
  27. Nuñez M, Aedo AR, Landgren BM, Cekan SZ, Diczfalusy E (November 1977). "Studies on the pattern of circulating steroids in the normal menstrual cycle. 6. Levels of oestrone sulphate and oestradiol sulphate". Acta Endocrinol. 86 (3): 621–33. doi:10.1530/acta.0.0860621. PMID 579025.
  28. Honjo H, Kitawaki J, Itoh M, Yasuda J, Iwasaku K, Urabe M, Naitoh K, Yamamoto T, Okada H, Ohkubo T (1987). "Serum and urinary estrone sulfate during the menstrual cycle, measured by a direct radioimmunoassay, and fate of exogenously injected estrone sulfate". Horm Res. 27 (2): 61–8. doi:10.1159/000180788. PMID 3653846.
  29. Ranadive GN, Mistry JS, Damodaran K, Khosravi MJ, Diamandi A, Gimpel T, Castracane VD, Patel S, Stanczyk FZ (February 1998). "Rapid, convenient radioimmunoassay of estrone sulfate". Clin. Chem. 44 (2): 244–9. doi:10.1093/clinchem/44.2.244. PMID 9474019.
  30. Cowie AT, Forsyth IA, Hart IC (1980). "Growth and Development of the Mammary Gland". Hormonal Control of Lactation. Monographs on Endocrinology. Vol. 15. pp. 58–145. doi:10.1007/978-3-642-81389-4_3. ISBN 978-3-642-81391-7. ISSN 0077-1015.

Further reading


Share this article:

This article uses material from the Wikipedia article Estrone_3-sulfate, and is written by contributors. Text is available under a CC BY-SA 4.0 International License; additional terms may apply. Images, videos and audio are available under their respective licenses.