Fish

Types of fish assays used most often include:

1. Rapid developmental toxicity assays (utilizing fathead minnow, medaka, and zebrafish) are now available. As a complement to the targeted in vitro assays included in the High Throughput and Whole Cell Activity screens, in vivo rapid developmental toxicity screens provide a quick and inexpensive method to detect adverse interactions between test chemicals and a vertebrate whole animal system.

The primary advantage of  assays using lower vertebrates is that the embryos develop rapidly and exercise their complete repertoire of gene expression and molecular signaling during the short transition from fertilization to organogenesis. During this window of development, there is a high probability of detecting an adverse interaction between an EDC and its molecular target that manifests as developmental delays or discrete morphological abnormalities including pericardial and yolk sac edemas, curved body axis, and eye, jaw, craniofacial, fin, and/or pigmentation defects[33-38]. The developing embryo can also be monitored for a series of cardiovascular[39] and behavioral [40] endpoints.

Factors for Consideration in Fish EDC Studies

2. Reproduction assays (using medaka, or fathead minnow). Reproduction tests are well established using this model system, and have been used to assess a number of chemicals suspected of having endocrine activity. Partial life-cycle assays employ short-term exposure during critical windows of sensitivity (i.e. sexual differentiation, gonadal development, active reproduction), whereas full life-cycle assays initiate chronic exposure with newly fertilized eggs.

In adult fish, active reproduction represents a period of sensitivity to chemicals that target the hypothalamo-pituitary-gonadal (HPG) axis. Assays designed to exploit this window of susceptibility assess apical (whole organism) endpoints following short-term (typically 21 days) exposure to a chemical.

Several transgenic zebrafish lines have been engineered to detect direct transcriptional activation of specific endocrine signaling pathways in “reporter gene assays”. These assays specifically and rapidly detect aspects of endocrine disruption. Researchers are limited, however, by the types of reporter line available and should understand that a variety of disruptions to the system may be missed because transcriptional reporter-based models are not capable of detecting non-genomic signaling.[44, 45]

About Mammalian Assays

This tier involves testing in mammalian models, primarily rodents. It is not designed to replace regulatory testing but to be a focused assessment of endpoints/tissues/diseases/pathways that may have been missed by earlier tiers because they lack the complexity of mammalian development. It can also be used to shed additional light on endocrine disrupting actions identified by earlier tiers.

The mammalian models are unique in their capacity to study in utero exposures that involve interactions between endocrine responses in the mother, placenta and embryo/fetus.  Furthermore, certain behavioral repertoires can be studied in mammals that have greater biomedical relevance, such as mating and maternal behaviors, lactation, weaning, and complex adult socio-sexual behaviors.

We assume here that the chemist employing TiPED has run his/her molecule through the other 4 tiers without detecting EDC activity. To be confident the chemical has no endocrine activity or to assess a specific endocrine system in more detail it is essential to consider mammalian whole animal assessment. These are the assays of ‘last resort,’ which would only be used if work in prior tiers revealed no EDC activity.

 

Fish and Amphibian

Earlier tiers test individual cells and tissue samples for hormonal effects; this tier looks at the whole animal for complex systemic effects.

We recommend using whole animal assays with fish and/or amphibia (frogs). In vivo assays – or live animal tests – allow for the examination of multiple endpoints, for multiple hormones, and multiple mechanisms of action.

Assays using fish (eg. zebrafish and medaka) can  cast a wide net to detect multiple potential teratological effects (classical toxicology endpoints, like mutation and death), but can also detect sensitive endocrine-affected endpoints – all of which have relevance for human health.

Because fish assays can detect broad toxicological effects, new compounds can be screened without prior information on suspected activity or expected mechanisms of action. Because these tests can find unexpected systemic effects, the Fish and Amphibian tier is needed for any chemicals that “passed” TiPED’s first three tiers.

References

1. Anastas, P.T., Warner, J.C. , Green Chemistry: Theory and Practice. 1998, New York: Oxford University Press, USA.

2. Vandenberg, L.N., et al., Hormones and Endocrine-Disrupting Chemicals: Low-Dose Effects and Nonmonotonic Dose Responses. Endocr Rev, 2012. 33(3).

3. Zoeller, R.T., et al., Endocrine-Disrupting Chemicals and Public Health Protection: A Statement of Principles from The Endocrine Society. Endocrinology, 2012. 25: p. 25.

4. Welshons, W.V., et al., Large effects from small exposures. I. Mechanisms for endocrine-disrupting chemicals with estrogenic activity. Environ Health Perspect, 2003. 111(8): p. 994-1006.

5. Alonso-Magdalena, P., et al., Bisphenol-A acts as a potent estrogen via non-classical estrogen triggered pathways. Molecular and Cellular Endocrinology, 2012. 355(2): p. 201-207.

6. Myers, J.P., R.T. Zoeller, and F.S. vom Saal, A clash of old and new scientific concepts in toxicity, with important implications for public health. Environ Health Perspect, 2009. 117(11): p. 1652-5.

7. Vandenberg, L.N., et al., A new ‘low dose’ paradigm in research on environmental chemicals. Endocr Rev, submitted.

8. Newbold, R.R., E. Padilla-Banks, and W.N. Jefferson, Adverse effects of the model environmental estrogen diethylstilbestrol are transmitted to subsequent generations. Endocrinology, 2006. 147(6 Suppl): p. S11-7.

9. Newbold, R.R., et al., Developmental exposure to endocrine disruptors and the obesity epidemic. Reprod Toxicol, 2007. 23( 3): p. 290-6.

10. Voutchkova, A.M., T.G. Osimitz, and P.T. Anastas, Toward a comprehensive molecular design framework for reduced hazard. Chem Rev, 2010. 110(10): p. 5845-82.

11. Williams, D.P. and D.J. Naisbitt, Toxicophores: groups and metabolic routes associated with increased safety risk. Curr Opin Drug Discov Devel, 2002. 5(1): p. 104-15.

12. Bottegoni, G., et al., Four-dimensional docking: a fast and accurate account of discrete receptor flexibility in ligand docking. J Med Chem, 2009. 52(2): p. 397-406.

13. Hansch, C., Fujita, T., Rho-sigma-pi analysis. A method for the correlation of biological activity and chemical structure. J Am Chem Soc. , 1964. 86: p. 1616-1626.

14. Papa, E., S. Kovarich, and P. Gramatica, QSAR modeling and prediction of the endocrine-disrupting potencies of brominated flame retardants. Chem Res Toxicol, 2010. 23(5): p. 946-54.

15. Shi, L.M., et al., QSAR models using a large diverse set of estrogens. J Chem Inf Comput Sci, 2001. 41(1): p. 186-95.

16. van Drie, J.H., Computational Medicinal Chemistry for Drug Discovery P. Bultinck, et al., Editors. 2005, CRC Press: New York, USA.

17. Yang, W., et al., Insights into the structural and conformational requirements of polybrominated diphenyl ethers and metabolites as potential estrogens based on molecular docking. Chemosphere, 2011. 84(3): p. 328-35.

18. Yang, W.H., et al., Exploring the binding features of polybrominated diphenyl ethers as estrogen receptor antagonists: docking studies. SAR QSAR Environ Res, 2010. 21(3-4): p. 351-67.

19. Yang, W., et al., Anti-androgen activity of polybrominated diphenyl ethers determined by comparative molecular similarity indices and molecular docking. Chemosphere, 2009. 75(9): p. 1159-64.

20. Park, S.J., I. Kufareva, and R. Abagyan, Improved docking, screening and selectivity prediction for small molecule nuclear receptor modulators using conformational ensembles. J Comput Aided Mol Des, 2010. 24(5): p. 459-71.

21. Dix, D.J., et al., The TOXCAST Program for prioritizing toxicity testing of environmental chemicals. Toxicological Sciences, 2007. 95(1): p. 5-12.

22. Kavlock, R.J., C.P. Austin, and R.R. Tice, Toxicity testing in the 21st century: implications for human health risk assessment. Risk Anal, 2009. 29(4): p. 485-7; discussion 492-7.

23. Malo, N., et al., Experimental design and statistical methods for improved hit detection in high-throughput screening. J Biomol Screen, 2010. 15(8): p. 990-1000.

24. Xia, M., et al., Identification of compounds that potentiate CREB signaling as possible enhancers of long-term memory. Proc Natl Acad Sci U S A, 2009. 106(7): p. 2412-7.

25. Inglese, J., et al., Quantitative high-throughput screening: a titration-based approach that efficiently identifies biological activities in large chemical libraries. Proc Natl Acad Sci U S A, 2006. 103(31): p. 11473-8.

26. Huang, R., et al., Chemical genomics profiling of environmental chemical modulation of human nuclear receptors. Environ Health Perspect, 2011. 119(8): p. 1142-8.

27. Welshons, W.V., et al., Large effects from small exposures: I. Mechanisms for endocrine-disrupting chemicals with estrogenic activity. Environ Health Perspect, 2003. 111: p. 994-1006.

28. Zucco, F., et al., Toxicology investigations with cell culture systems: 20 years after. Toxicol In Vitro., 2004. 18(2): p. 153-63.

29. Okuda, K., et al., Novel pathway of metabolic activation of bisphenol A-related compounds for estrogenic activity. Drug Metab Dispos, 2011. 39(9): p. 1696-703.

30. Yang, C.H., et al., Sulfation of selected mono-hydroxyflavones by sulfotransferases in vitro: a species and gender comparison. J Pharm Pharmacol, 2011. 63(7): p. 967-70.

31. Charles, G.D., et al., Incorporation of S-9 activation into an ER-alpha transactivation assay. Reprod Toxicol, 2000. 14(3): p. 207-16.

32. Miller, K.P., R.K. Gupta, and J.A. Flaws, Methoxychlor metabolites may cause ovarian toxicity through estrogen-regulated pathways. Toxicol Sci, 2006. 93(1): p. 180-8.

33. Shi, M. and E.M. Faustman, Development and characterization of a morphological scoring system for medaka (Oryzias latipes) embryo culture. Aquat Toxicol, 1989. 15(2): p. 127-140.

34. Carney, M.W., et al., Differential developmental toxicity of naphthoic acid isomers in medaka (Oryzias latipes) embryos. Mar Pollut Bull, 2008. 57(6-12): p. 255-66.

35. Marty, G.D., et al., Age-dependent changes in toxicity of N-nitroso compounds to Japanese medaka (Oryzias latipes) embryos. Aquat Toxicol, 1990. 17(1): p. 45-62.

36. Laban, G., et al., The effects of silver nanoparticles on fathead minnow (Pimephales promelas) embryos. Ecotoxicology, 2010. 19(1): p. 185-95.

37. Herrmann, K., Teratogenic effects of retinoic acid and related substances on the early development of the zebrafish (Brachydanio rerio) as assessed by a novel scoring system. Toxicol in Vitro, 1995. 9(3): p. 267-83.

38. Strmac, M. and T. Braunbeck, Effects of triphenyltin acetate on survival, hatching success, and liver ultrastructure of early life stages of zebrafish (Danio rerio). Ecotoxicol Environ Saf, 1999. 44(1): p. 25-39.

39. Asharani, P.V., et al., Comparison of the toxicity of silver, gold and platinum nanoparticles in developing zebrafish embryos. Nanotoxicology, 2011. 5(1): p. 43-54.

40. Chen, J., et al., Trimethyltin chloride (TMT) neurobehavioral toxicity in embryonic zebrafish. Neurotoxicol Teratol, 2011.

41. Chen, H., et al., Generation of a fluorescent transgenic zebrafish for detection of environmental estrogens. Aquat Toxicol, 2010. 96(1): p. 53-61.

42. Gorelick, D.A. and M.E. Halpern, Visualization of estrogen receptor transcriptional activation in zebrafish. Endocrinology, 2011. 152(7): p. 2690-703.

43. Terrien, X., et al., Generation of fluorescent zebrafish to study endocrine disruption and potential crosstalk between thyroid hormone and corticosteroids. Aquat Toxicol, 2011. 105(1-2): p. 13-20.

44. Damdimopoulou, P. and E. Treuter, Reporter zebrafish: endocrine disruption meets estrogen signaling. Endocrinology, 2011. 152(7): p. 2542-5.

45. Silva, E., A. Kabil, and A. Kortenkamp, Cross-talk between non-genomic and genomic signalling pathways–distinct effect profiles of environmental estrogens. Toxicol Appl Pharmacol, 2010. 245(2): p. 160-70.

46. Porazzi, P., et al., Thyroid gland development and function in the zebrafish model. Mol Cell Endocrinol, 2009. 312(1-2): p. 14-23.

47. Hayes, T.B., et al., Atrazine induces complete feminization and chemical castration in male African clawed frogs (Xenopus laevis). Proc Natl Acad Sci U S A, 2010. 107(10): p. 4612-7.

48. van Wyk, J.H., E.J. Pool, and A.J. Leslie, The effects of anti-androgenic and estrogenic disrupting contaminants on breeding gland (nuptial pad) morphology, plasma testosterone levels, and plasma vitellogenin levels in male Xenopus laevis (African clawed frog). Arch Environ Contam Toxicol, 2003. 44(2): p. 247-56.

49. Behrends, T., et al., Mate calling behavior of male South African clawed frogs (Xenopus laevis) is suppressed by the antiandrogenic endocrine disrupting compound flutamide. Gen Comp Endocrinol. 168(2): p. 269-74.

50. Hayes, T.B. and K.P. Menendez, The effect of sex steroids on primary and secondary sex differentiation in the sexually dichromatic reedfrog (Hyperolius argus: Hyperolidae) from the Arabuko Sokoke Forest of Kenya. Gen Comp Endocrinol, 1999. 115(2): p. 188-99.

51. Hayes, T.B., et al., Pesticide mixtures, endocrine disruption, and amphibian declines: are we underestimating the impact? Environ Health Perspect, 2006. 114(Suppl 1): p. 40-50.

52. Weiss, B., Endocrine disruptors and sexually dimorphic behaviors: a question of heads and tails. Neurotoxicology, 1997. 18(2): p. 581-6.

53. Newbold, R.R., et al., Uterine adenocarcinoma in mice treated neonatally with genistein. Cancer Res, 2001. 61(11): p. 4325-8.

54. Anway, M.D., et al., Epigenetic transgenerational actions of endocrine disruptors and male fertility. Science, 2005. 308(5727): p. 1466-9.

55. Armenti, A.E., et al., Developmental methoxychlor exposure affects multiple reproductive parameters and ovarian folliculogenesis and gene expression in adult rats. Toxicol Appl Pharmacol, 2008. 233(2): p. 286-96.

56. Gore, A.C., et al., Early Life Exposure to Endocrine-Disrupting Chemicals Causes Lifelong Molecular Reprogramming of the Hypothalamus and Premature Reproductive Aging. Mol Endocrinol, 2011.

57. Ho, S.M., et al., Developmental exposure to estradiol and bisphenol A increases susceptibility to prostate carcinogenesis and epigenetically regulates phosphodiesterase type 4 variant 4. Cancer Res, 2006. 66(11): p. 5624-32.

58. Anastas, N.D. and J.C. Warner, The incorporation of hazard reduction as a chemical design criterion in green chemistry. Chemical Health and Safety, 2005. 12(2): p. 9-13.

59. Wilson, V.S., et al., A novel cell line, MDA-kb2, that stably expresses an androgen- and glucocorticoid-responsive reporter for the detection of hormone receptor agonists and antagonists. Toxicol Sci, 2002. 66(1): p. 69-81.

60. Szelei, J., et al., Androgen-induced inhibition of proliferation in human breast cancer MCF7 cells transfected with androgen receptor. Endocrinology, 1997. 138(4): p. 1406-12.

61. Sonneveld, E., et al., Development of androgen- and estrogen-responsive bioassays, members of a panel of human cell line-based highly selective steroid-responsive bioassays. Toxicol Sci, 2005. 83(1): p. 136-48.

62. Sanderson, J.T., et al., Effects of chloro-s-triazine herbicides and metabolites on aromatase activity in various human cell lines and on vitellogenin production in male carp hepatocytes. Environ Health Perspect, 2001. 109(10): p. 1027-31.

63. Villalobos, M., et al., The E-screen assay: a comparison of different MCF7 cell stocks. Environ Health Perspect, 1995. 103(9): p. 844-50.

64. Vandenberg, L.N., et al., Perinatal exposure to the xenoestrogen bisphenol-A induces mammary intraductal hyperplasias in adult CD-1 mice. Reprod Toxicol, 2008. 26: p. 210-9.

65. Markey, C.M., et al., In utero exposure to bisphenol A alters the development and tissue organization of the mouse mammary gland. Biology of Reproduction, 2001. 65(4): p. 1215-23.

66. Vandenberg, L.N., et al., The mammary gland response to estradiol: monotonic at the cellular level, non-monotonic at the tissue-level of organization? Journal of Steroid Biochemistry and Molecular Biology, 2006. 101(4-5): p. 263-74.

67. Soto, A.M., K.L. Chung, and C. Sonnenschein, The pesticides endosulfan, toxaphene, and dieldrin have estrogenic effects on human estrogen-sensitive cells. Environ Health Perspect, 1994. 102(4): p. 380-3.

68. Soto, A.M., et al., The E-SCREEN assay as a tool to identify estrogens: an update on estrogenic environmental pollutants. Environ Health Perspect, 1995. 103(Suppl 7): p. 113-22.

69. Kester, M.H., et al., Potent inhibition of estrogen sulfotransferase by hydroxylated PCB metabolites: a novel pathway explaining the estrogenic activity of PCBs. Endocrinology, 2000. 141(5): p. 1897-900.

70. Legler, J., et al., Development of a stably transfected estrogen receptor-mediated luciferase reporter gene assay in the human T47D breast cancer cell line. Toxicol Sci, 1999. 48(1): p. 55-66.

71. Terasaka, S., et al., Using a customized DNA microarray for expression profiling of the estrogen-responsive genes to evaluate estrogen activity among natural estrogens and industrial chemicals. Environ Health Perspect, 2004. 112(7): p. 773-81.

72. Hanna, R.N., et al., Characterization and expression of the nuclear progestin receptor in zebrafish gonads and brain. Biol Reprod, 2010. 82(1): p. 112-22.

73. Schreurs, R.H., et al., Interaction of polycyclic musks and UV filters with the estrogen receptor (ER), androgen receptor (AR), and progesterone receptor (PR) in reporter gene bioassays. Toxicol Sci, 2005. 83(2): p. 264-72.

74. Hecker, M., et al., Human adrenocarcinoma (H295R) cells for rapid in vitro determination of effects on steroidogenesis: hormone production. Toxicol Appl Pharmacol, 2006. 217(1): p. 114-24.

75. Miller, C.A., 3rd, et al., Single plasmids expressing human steroid hormone receptors and a reporter gene for use in yeast signaling assays. Plasmid, 2010. 63(2): p. 73-8.

76. Miller, C.A., 3rd, A human aryl hydrocarbon receptor signaling pathway constructed in yeast displays additive responses to ligand mixtures. Toxicol Appl Pharmacol, 1999. 160(3): p. 297-303.

77. Kimura-Kuroda, J., I. Nagata, and Y. Kuroda, Hydroxylated metabolites of polychlorinated biphenyls inhibit thyroid-hormone-dependent extension of cerebellar Purkinje cell dendrites. Brain Res Dev Brain Res, 2005. 154(2): p. 259-63.

78. Kimura-Kuroda, J., I. Nagata, and Y. Kuroda, Hormonal disruption in cerebellar development by hydroxylated polychlorinated biphenyls. Cerebellum, 2008. 7(3): p. 500-1.

79. Kimura-Kuroda, J., I. Nagata, and Y. Kuroda, Disrupting effects of hydroxy-polychlorinated biphenyl (PCB) congeners on neuronal development of cerebellar Purkinje cells: a possible causal factor for developmental brain disorders? Chemosphere, 2007. 67(9): p. S412-20.

80. Tonacchera, M., et al., Relative potencies and additivity of perchlorate, thiocyanate, nitrate, and iodide on the inhibition of radioactive iodide uptake by the human sodium iodide symporter. Thyroid, 2004. 14(12): p. 1012-9.

81. Tran, N., et al., Thyroid-stimulating hormone increases active transport of perchlorate into thyroid cells. Am J Physiol Endocrinol Metab, 2008. 294(4): p. E802-6.

82. Ibhazehiebo, K., et al., Suppression of thyroid hormone receptor-mediated transcription and disruption of thyroid hormone-induced cerebellar morphogenesis by the polybrominated biphenyl mixture, BP-6. Neurotoxicology, 2011. 32(4): p. 400-9.

83. Gutleb, A.C., et al., T-Screen as a tool to identify thyroid hormone receptor active compounds. Environ Toxicol Pharmacol, 2005. 19(2): p. 231-8.

84. Schriks, M., et al., T-screen to quantify functional potentiating, antagonistic and thyroid hormone-like activities of poly halogenated aromatic hydrocarbons (PHAHs). Toxicol in Vitro, 2006. 20(4): p. 490-8.

85. Freitas, J., et al., Detection of thyroid hormone receptor disruptors by a novel stable in vitro reporter gene assay. Toxicol in Vitro, 2011. 25(1): p. 257-66.

86. Doerge, D.R. and H.C. Chang, Inactivation of thyroid peroxidase by soy isoflavones, in vitro and in vivo. J Chromatogr B Analyt Technol Biomed Life Sci, 2002. 777(1-2): p. 269-79.

87. Doerge, D.R. and D.M. Sheehan, Goitrogenic and estrogenic activity of soy isoflavones. Environ Health Perspect, 2002. 110 Suppl 3: p. 349-53.

88. Feige, J.N., et al., The endocrine disruptor monoethyl-hexyl-phthalate is a selective peroxisome proliferator-activated receptor gamma modulator that promotes adipogenesis. Journal of Biological Chemistry, 2007. 282(26): p. 19152-66.

89. Grun, F., et al., Endocrine-disrupting organotin compounds are potent inducers of adipogenesis in vertebrates. Mol Endocrinol, 2006. 20(9): p. 2141-55.

90. Kanayama, T., et al., Organotin compounds promote adipocyte differentiation as agonists of the peroxisome proliferator-activated receptor gamma/retinoid X receptor pathway. Mol Pharmacol, 2005. 67(3): p. 766-74.

91. Li, X., J. Ycaza, and B. Blumberg, The environmental obesogen tributyltin chloride acts via peroxisome proliferator activated receptor gamma to induce adipogenesis in murine 3T3-L1 preadipocytes. J Steroid Biochem Mol Biol, 2011. 127(1-2): p. 9-15.

92. Garrison, P.M., et al., Species-specific recombinant cell lines as bioassay systems for the detection of 2,3,7,8-tetrachlorodibenzo-p-dioxin-like chemicals. Fundam Appl Toxicol, 1996. 30(2): p. 194-203.

93. Murk, A.J., et al., Chemical-activated luciferase gene expression (CALUX): a novel in vitro bioassay for Ah receptor active compounds in sediments and pore water. Fundam Appl Toxicol, 1996. 33(1): p. 149-60.

94. Colliar, L., A. Sturm, and M.J. Leaver, Tributyltin is a potent inhibitor of piscine peroxisome proliferator-activated receptor alpha and beta. Comp Biochem Physiol C Toxicol Pharmacol, 2011. 153(1): p. 168-173.

95. Kumagai, T., et al., Polycyclic aromatic hydrocarbons activate CYP3A4 gene transcription through human pregnane X receptor. Drug Metab Pharmacokinet, 2011.

96. Idres, N., et al., Activation of retinoic acid receptor-dependent transcription by all-trans-retinoic acid metabolites and isomers. J Biol Chem, 2002. 277(35): p. 31491-8.

97. Li, J., M. Ma, and Z. Wang, A two-hybrid yeast assay to quantify the effects of xenobiotics on retinoid X receptor-mediated gene expression. Toxicol Lett, 2008. 176(3): p. 198-206.

98. Schoff, P.K. and G.T. Ankley, Effects of methoprene, its metabolites, and breakdown products on retinoid-activated pathways in transfected cell lines. Environ Toxicol Chem, 2004. 23(5): p. 1305-10.

99. Jeng, Y.J., M. Kochukov, and C.S. Watson, Combinations of physiologic estrogens with xenoestrogens alter calcium and kinase responses, prolactin release, and membrane estrogen receptor trafficking in rat pituitary cells. Environ Health, 2010. 9: p. 61.

100. Bulayeva, N.N. and C.S. Watson, Xenoestrogen-induced ERK-1 and ERK-2 activation via multiple membrane-initiated signaling pathways. Environ Health Perspect, 2004. 112(15): p. 1481-7.

101. Jeng, Y.J. and C.S. Watson, Combinations of physiologic estrogens with xenoestrogens alter ERK phosphorylation profiles in rat pituitary cells. Environ Health Perspect, 2011. 119(1): p. 104-12.

102. Zivadinovic, D. and C.S. Watson, Membrane estrogen receptor-alpha levels predict estrogen-induced ERK1/2 activation in MCF-7 cells. Breast Cancer Res, 2005. 7(1): p. R130-44.

103. Kelder, J., et al., Comparison between steroid binding to membrane progesterone receptor alpha (mPRalpha) and to nuclear progesterone receptor: correlation with physicochemical properties assessed by comparative molecular field analysis and identification of mPRalpha-specific agonists. Steroids, 2010. 75(4-5): p. 314-22.

104. Thomas, P., et al., Steroid and G protein binding characteristics of the seatrout and human progestin membrane receptor alpha subtypes and their evolutionary origins. Endocrinology, 2007. 148(2): p. 705-18.

105. Pang, Y. and P. Thomas, Progesterone signals through membrane progesterone receptors (mPRs) in MDA-MB-468 and mPR-transfected MDA-MB-231 breast cancer cells which lack full-length and N-terminally truncated isoforms of the nuclear progesterone receptor. Steroids, 2011. 76(9): p. 921-8.

106. Hontela, A., V.S. Leblond, and J.P. Chang, Purification and isolation of corticosteroidogenic cells from head kidney of rainbow trout (Oncorhynchus mykiss) for testing cell-specific effects of a pesticide. Comp Biochem Physiol C Toxicol Pharmacol, 2008. 147(1): p. 52-60.

107. McCormick, J.M., et al., Microbially mediated O-methylation of bisphenol A results in metabolites with increased toxicity to the developing zebrafish (Danio rerio) embryo. Environ Sci Technol, 2011. 45(15): p. 6567-74.

108. Truong, L., et al., Persistent adult zebrafish behavioral deficits results from acute embryonic exposure to gold nanoparticles. Comp Biochem Physiol C Toxicol Pharmacol, 2011.

109. Gorge, G. and R. Nagel, Toxicity of lindane, atrazine, and deltamethrin to early life stages of zebrafish (Brachydanio rerio). Ecotoxicol Environ Saf, 1990. 20(3): p. 246-55.

110. Shi, X., et al., Developmental toxicity and alteration of gene expression in zebrafish embryos exposed to PFOS. Toxicol Appl Pharmacol, 2008. 230(1): p. 23-32.

111. Hagenaars, A., et al., Structure-activity relationship assessment of four perfluorinated chemicals using a prolonged zebrafish early life stage test. Chemosphere, 2011. 82(5): p. 764-72.

112. Panter, G.H., et al., Development of chronic tests for endocrine active chemicals. Part 1. An extended fish early-life stage test for oestrogenic active chemicals in the fathead minnow (Pimephales promelas). Aquat Toxicol, 2006. 77(3): p. 279-90.

113. Urushitani, H., et al., Medaka (Oryzias latipes) for use in evaluating developmental effects of endocrine active chemicals with special reference to gonadal intersex (testis-ova). Environ Sci, 2007. 14(5): p. 211-33.

114. Andersen, L., et al., Effects of exposure to 17alpha-ethinylestradiol during early development on sexual differentiation and induction of vitellogenin in zebrafish (Danio rerio). Comp Biochem Physiol C Toxicol Pharmacol, 2003. 134(3): p. 365-74.

115. Ankley, G.T., et al., Description and evaluation of a short-term reproduction test with the fathead minnow (Pimephales promelas). Environ Toxicol Chem, 2001. 20(6): p. 1276-90.

116. Saili, K.S., et al., Neurodevelopmental low-dose bisphenol A exposure leads to early life-stage hyperactivity and learning deficits in adult zebrafish. Toxicology, 2012.

117. Colman, J.R., et al., Effects of the synthetic estrogen, 17alpha-ethinylestradiol, on aggression and courtship behavior in male zebrafish (Danio rerio). Aquat Toxicol, 2009. 91(4): p. 346-54.

118. Ankley, G.T., et al., Evaluation of the aromatase inhibitor fadrozole in a short-term reproduction assay with the fathead minnow (Pimephales promelas). Toxicol Sci, 2002. 67(1): p. 121-30.

119. Dang, Z., et al., Endpoint sensitivity in fish endocrine disruption assays: regulatory implications. Toxicol Lett, 2011. 202(1): p. 36-46.

120. Bringolf, R.B., J.B. Belden, and R.C. Summerfelt, Effects of atrazine on fathead minnow in a short-term reproduction assay. Environ Toxicol Chem, 2004. 23(4): p. 1019-25.

121. Mouriec, K., et al., Early regulation of brain aromatase (cyp19a1b) by estrogen receptors during zebrafish development. Dev Dyn, 2009. 238(10): p. 2641-51.

122. Sassi-Messai, S., et al., The phytoestrogen genistein affects zebrafish development through two different pathways. PLoS One, 2009. 4(3): p. e4935.

123. Goulet, B.N. and A. Hontela, Toxicity of cadmium, endosulfan, and atrazine in adrenal steroidogenic cells of two amphibian species, Xenopus laevis and Rana catesbeiana. Environ Toxicol Chem, 2003. 22(9): p. 2106-13.

124. Lutz, I., et al., Development, standardization and refinement of procedures for evaluating effects of endocrine active compounds on development and sexual differentiation of Xenopus laevis. Anal Bioanal Chem, 2008. 390(8): p. 2031-48.

125. Kloas, W., et al., Endocrine disruption in aquatic vertebrates. Ann N Y Acad Sci, 2009. 1163: p. 187-200.

126. Bates, J.H., M. Rincon, and C.G. Irvin, Animal models of asthma. Am J Physiol Lung Cell Mol Physiol, 2009. 297(3): p. L401-10.

127. Hagiwara, H., et al., Effects of alkylphenols on bone metabolism in vivo and in vitro. Toxicol Lett, 2008. 181(1): p. 13-8.

128. Dickerson, S.M., et al., Endocrine disruption of brain sexual differentiation by developmental PCB exposure. Endocrinology, 2011. 152(2): p. 581-94.

129. Rubin, B.S., et al., Evidence of altered brain sexual differentiation in mice exposed perinatally to low, environmentally relevant levels of bisphenol A. Endocrinology, 2006. 147(8): p. 3681-91.

130. Cao, J. and H.B. Patisaul, Sexually dimorphic expression of hypothalamic estrogen receptors alpha and beta and Kiss1 in neonatal male and female rats. J Comp Neurol, 2011. 519(15): p. 2954-77.

131. Cabaton, N.J., et al., Perinatal exposure to environmentally relevant levels of bisphenol A decreases fertility and fecundity in CD-1 mice. Environ Health Perspect, 2011. 119(4): p. 547-52.

132. McLachlan, J.A., et al., Reduced fertility in female mice exposed transplacentally to diethylstilbestrol (DES). Fertil Steril, 1982. 38: p. 364-71.

133. David, R.M., et al., Chronic toxicity of di(2-ethylhexyl)phthalate in rats. Toxicol Sci, 2000. 55(2): p. 433-43.

134. Korkmaz, A., et al., Vitamin C coadministration augments bisphenol A, nonylphenol, and octylphenol induced oxidative damage on kidney of rats. Environ Toxicol, 2011. 26(4): p. 325-37.

135. Jenkins, S., et al., Oral exposure to bisphenol A increases dimethylbenzanthracene-induced mammary cancer in rats. Environ Health Perspect, 2009. 117(6): p. 910-5.

136. Lamartiniere, C.A., et al., Exposure to the Endocrine Disruptor Bisphenol A Alters Susceptibility for Mammary Cancer. Horm Mol Biol Clin Investig, 2011. 5(2): p. 45-52.

137. Murray, T.J., et al., Induction of mammary gland ductal hyperplasias and carcinoma in situ following fetal bisphenol A exposure. Reprod Toxicol, 2007. 23(3): p. 383-90.

138. Durando, M., et al., Prenatal bisphenol A exposure induces preneoplastic lesions in the mammary gland in Wistar rats. Environ Health Perspect, 2007. 115: p. 80-86.

139. Palanza, P., et al., Exposure to a low dose of bisphenol A during fetal life or in adulthood alters maternal behavior in mice. Environ Health Perspect, 2002. 110(Suppl 3): p. 415-422.

140. Hariri, N. and L. Thibault, High-fat diet-induced obesity in animal models. Nutr Res Rev, 2010. 23(2): p. 270-99.

141. Varga, O., et al., Contribution of animal models to the understanding of the metabolic syndrome: a systematic overview. Obes Rev, 2010. 11(11): p. 792-807.

142. Ho, S.M., et al., Developmental exposure to estradiol and bisphenol A increases susceptibility to prostate carcinogenesis and epigenetically regulates phosphodiesterase type 4 variant 4. Cancer Research, 2006. 66(11): p. 5624-32.

143. Prins, G.S., et al., Serum bisphenol A pharmacokinetics and prostate neoplastic responses following oral and subcutaneous exposures in neonatal Sprague-Dawley rats. Reprod Toxicol, 2011. 31(1): p. 1-9.

144. Cox, K.H. and E.F. Rissman, Sex differences in juvenile mouse social behavior are influenced by sex chromosomes and social context. Genes Brain Behav, 2011. 10(4): p. 465-72.

145. Kokras, N., et al., Behavioral sexual dimorphism in models of anxiety and depression due to changes in HPA axis activity. Neuropharmacology, 2012. 62(1): p. 436-45.

146. Turgeon, S.M., et al., The effects of phencyclidine (PCP) on anxiety-like behavior in the elevated plus maze and the light-dark exploration test are age dependent, sexually dimorphic, and task dependent. Pharmacol Biochem Behav, 2011. 100(1): p. 191-8.

147. Wolstenholme, J.T., et al., Gestational exposure to low dose bisphenol a alters social behavior in juvenile mice. PLoS One, 2011. 6(9): p. e25448.

148. Moggs, J.G., Molecular responses to xenoestrogens: mechanistic insights from toxicogenomics. Toxicology, 2005. 213(3): p. 177-93.

149. Iguchi, T., H. Watanabe, and Y. Katsu, Toxicogenomics and ecotoxicogenomics for studying endocrine disruption and basic biology. Gen Comp Endocrinol, 2007. 153(1-3): p. 25-9.

150. Zama, A.M. and M. Uzumcu, Epigenetic effects of endocrine-disrupting chemicals on female reproduction: an ovarian perspective. Front Neuroendocrinol, 2010. 31(4): p. 420-39.

151. Sacan, M.T., M. Ozkul, and S.S. Erdem, Physico-chemical properties of PCDD/PCDFs and phthalate esters. SAR QSAR Environ Res, 2005. 16(5): p. 443-59.

152. Nose, T. and Y. Shimohigashi, A docking modelling rationally predicts strong binding of bisphenol A to estrogen-related receptor gamma. Protein Pept Lett, 2008. 15(3): p. 290-6.

153. Wu, B., et al., Computational studies of interactions between endocrine disrupting chemicals and androgen receptor of different vertebrate species. Chemosphere, 2010. 80(5): p. 535-41.

154. Gumy, C., et al., Dibutyltin disrupts glucocorticoid receptor function and impairs glucocorticoid-induced suppression of cytokine production. PLoS ONE, 2008. 3(10): p. e3545.

155. Hu, J.Y. and T. Aizawa, Quantitative structure-activity relationships for estrogen receptor binding affinity of phenolic chemicals. Water Res, 2003. 37(6): p. 1213-22.

156. Basu Baul, T.S., et al., Synthesis, crystal structures, cytotoxicity and qualitative structure-activity relationship (QSAR) of cis-bis{5-[(E)-2-(aryl)-1-diazenyl]quinolinolato}di-n-butyltin(IV) complexes, (n)Bu2Sn(L)2. J Inorg Biochem, 2008. 102(9): p. 1719-30.

157. National Toxicology Program. High throughput screening initiative. 2011  [cited 2011 Oct 17 2011]; Available from: http://ntp.niehs.nih.gov/?objectid=05F80E15-F1F6-975E-77DDEDBDF3B941CD#6.

158. Kanai, H., et al., Cell-transforming activity and estrogenicity of bisphenol-A and 4 of its analogs in mammalian cells. Int J Cancer, 2001. 93(1): p. 20-5.

159. Masas, M., et al., Estrogen-like properties of fluorotelomer alcohols as revealed by mcf-7 breast cancer cell proliferation. Environ Health Perspect, 2006. 114(1): p. 100-5.

160. Okubo, T., et al., Estimation of estrogenic and anti-estrogenic activities of some phthalate diesters and monoesters by MCF-7 cell proliferation assay in vitro. Biol Pharm Bull, 2003. 26(8): p. 1219-24.

161. Wetherill, Y.B., et al., The xenoestrogen Bisphenol A induces inappropriate androgen receptor activation and mitogenesis in prostatic adenocarcinoma cells. Molecular Cancer Therapeutics, 2002. 1: p. 515-524.

162. Yamabe, Y., et al., Enhancement of androgen-dependent transcription and cell proliferation by tributyltin and triphenyltin in human prostate cancer cells. Toxicol Appl Pharmacol, 2000. 169(2): p. 177-84.

163. Sargis, R.M., et al., Environmental endocrine disruptors promote adipogenesis in the 3T3-L1 cell line through glucocorticoid receptor activation. Obesity (Silver Spring), 2010. 18(7): p. 1283-8.

164. Ghisari, M. and E.C. Bonefeld-Jorgensen, Impact of environmental chemicals on the thyroid hormone function in pituitary rat GH3 cells. Mol Cell Endocrinol, 2005. 244(1-2): p. 31-41.

165. Ghisari, M. and E.C. Bonefeld-Jorgensen, Effects of plasticizers and their mixtures on estrogen receptor and thyroid hormone functions. Toxicol Lett, 2009. 189(1): p. 67-77.

166. Zeng, Z., et al., Development of estrogen-responsive transgenic medaka for environmental monitoring of endocrine disrupters. Environ Sci Technol, 2005. 39(22): p. 9001-8.

167. Kurauchi, K., et al., In vivo visual reporter system for detection of estrogen-like substances by transgenic medaka. Environ Sci Technol, 2005. 39(8): p. 2762-8.

168. Tabata, A., et al., The effect of bisphenol A and chlorinated derivatives of bisphenol A on the level of serum vitellogenin in Japanese medaka (Oryzias latipes). Water Sci Technol, 2004. 50(5): p. 125-32.

169. Tillitt, D.E., et al., Atrazine reduces reproduction in fathead minnow (Pimephales promelas). Aquat Toxicol. 99(2): p. 149-59.

170. Ji, K., et al., Toxicity of perfluorooctane sulfonic acid and perfluorooctanoic acid on freshwater macroinvertebrates (Daphnia magna and Moina macrocopa) and fish (Oryzias latipes). Environ Toxicol Chem, 2008. 27(10): p. 2159-68.

171. Kim, E.J., J.W. Kim, and S.K. Lee, Inhibition of oocyte development in Japanese medaka (Oryzias latipes) exposed to di-2-ethylhexyl phthalate. Environ Int, 2002. 28(5): p. 359-65.

172. Zhang, Z., et al., Reproductive inhibition and transgenerational toxicity of triphenyltin on medaka (Oiyzias latipes) at environmentally relevant levels. Environ Sci Technol, 2008. 42(21): p. 8133-9.

173. Iwamuro, S., et al., Teratogenic and anti-metamorphic effects of bisphenol A on embryonic and larval Xenopus laevis. Gen Comp Endocrinol, 2003. 133(2): p. 189-98.

174. Shen, O., et al., Thyroid disruption by Di-n-butyl phthalate (DBP) and mono-n-butyl phthalate (MBP) in Xenopus laevis. PLoS ONE, 2011. 6(4): p. e19159.

175. Tietge, J.E., et al., Metamorphic inhibition of Xenopus laevis by sodium perchlorate: effects on development and thyroid histology. Environ Toxicol Chem, 2005. 24(4): p. 926-33.

176. Levy, G., et al., Bisphenol A induces feminization in Xenopus laevis tadpoles. Environ Res, 2004. 94(1): p. 102-11.

177. Hayes, T.B., et al., Hermaphroditic, demasculinized frogs after exposure to the herbicide atrazine at low ecologically relevant doses. Proc Natl Acad Sci U S A, 2002. 99(8): p. 5476-80.

178. Lee, S.K. and D.N. Veeramachaneni, Subchronic exposure to low concentrations of di-n-butyl phthalate disrupts spermatogenesis in Xenopus laevis frogs. Toxicol Sci, 2005. 84(2): p. 394-407.

179. Goleman, W.L., J.A. Carr, and T.A. Anderson, Environmentally relevant concentrations of ammonium perchlorate inhibit thyroid function and alter sex ratios in developing Xenopus laevis. Environ Toxicol Chem, 2002. 21(3): p. 590-7.

180. Midoro-Horiuti, T., et al., Maternal bisphenol a exposure promotes the development of experimental asthma in mouse pups. Environ Health Perspect, 2011. 118(2): p. 273-7.

181. Takagi, H., et al., Impact of maternal dietary exposure to endocrine-acting chemicals on progesterone receptor expression in microdissected hypothalamic medial preoptic areas of rat offspring. Toxicol Appl Pharmacol, 2005. 208(2): p. 127-36.

182. Hobler, C., et al., Sex-dependent aromatase activity in rat offspring after pre- and postnatal exposure to triphenyltin chloride. Toxicology, 2010. 276(3): p. 198-205.

183. Ueda, M., et al., Possible enhancing effects of atrazine on growth of 7,12-dimethylbenz(a) anthracene-induced mammary tumors in ovariectomized Sprague-Dawley rats. Cancer Sci, 2005. 96(1): p. 19-25.

184. Rayner, J.L., R.R. Enoch, and S.E. Fenton, Adverse effects of prenatal exposure to atrazine during a critical period of mammary gland growth. Toxicol Sci, 2005. 87: p. 255-266.

185. White, S.S., et al., Gestational and chronic low-dose PFOA exposures and mammary gland growth and differentiation in three generations of CD-1 mice. Environ Health Perspect, 2011. 119(8): p. 1070-6.

186. Zhao, Y., et al., Perfluorooctanoic acid effects on steroid hormone and growth factor levels mediate stimulation of peripubertal mammary gland development in C57BL/6 mice. Toxicol Sci, 2010. 115(1): p. 214-24.

187. Moral, R., et al., In utero exposure to butyl benzyl phthalate induces modifications in the morphology and the gene expression profile of the mammary gland: an experimental study in rats. Environ Health, 2011. 10(1): p. 5.

188. Baroncelli, S., D. Karrer, and P.G. Turillazzi, Oral bis(tri-n-butyltin) oxide in pregnant mice. I. Potential influence of maternal behavior on postnatal mortality. J Toxicol Environ Health, 1995. 46(3): p. 355-67.

189. Alonso-Magdalena, P., et al., Bisphenol A exposure during pregnancy disrupts glucose homeostasis in mothers and adult male offspring. Environ Health Perspect, 2010. 118(9): p. 1243-50.

190. Ryan, K.K., et al., Perinatal exposure to bisphenol-a and the development of metabolic syndrome in CD-1 mice. Endocrinology, 2010. 151(6): p. 2603-12.

191. Lim, S., et al., Chronic exposure to the herbicide, atrazine, causes mitochondrial dysfunction and insulin resistance. PLoS ONE, 2009. 4(4): p. e5186.

192. Boberg, J., et al., Impact of diisobutyl phthalate and other PPAR agonists on steroidogenesis and plasma insulin and leptin levels in fetal rats. Toxicology, 2008. 250(2-3): p. 75-81.

193. Prins, G.S., et al., Developmental estrogen exposures predispose to prostate carcinogenesis with aging. Reprod Toxicol, 2007. 23(3): p. 374-82.

194. Somm, E., et al., Perinatal exposure to bisphenol a alters early adipogenesis in the rat. Environ Health Perspect, 2009. 117(10): p. 1549-55.

195. Victor-Costa, A.B., et al., Changes in testicular morphology and steroidogenesis in adult rats exposed to Atrazine. Reprod Toxicol, 2010. 29(3): p. 323-31.

196. Shi, Z., et al., Chronic exposure to perfluorododecanoic acid disrupts testicular steroidogenesis and the expression of related genes in male rats. Toxicol Lett, 2009. 188(3): p. 192-200.

197. Moral, R., et al., The plasticizer butyl benzyl phthalate induces genomic changes in rat mammary gland after neonatal/prepubertal exposure. BMC Genomics, 2007. 8: p. 453.

198. Kim, S.K., et al., Inhibitory effect of tributyltin on expression of steroidogenic enzymes in mouse testis. Int J Toxicol, 2008. 27(2): p. 175-82.

199. McDougal, J.N., et al., The effects of perchlorate on thyroidal gene expression are different from the effects of iodide deficiency. J Toxicol Environ Health A, 2011. 74(14): p. 917-26.

200. Watson, C.S., et al., The involvement of ERαlGαi interactions in estrogen- and xenoestrogen-induced ERK signaling in pituitary tumor cells. Steroids, 2012 77(5): p.424-432.

 

 

What is Green Chemistry

Green chemistry is a powerful innovation tool. There has been a sea change in our understanding of chemicals and their effect on the environment and human health.  This new awareness requires an entirely new way of approaching chemical design. That change in chemical design is Green Chemistry.

Definition & Principles

“Green Chemistry is the utilization of a set of principles that reduces or eliminates the use or generation of hazardous substances in the design, manufacture and application of chemical products.” [1]

Green Chemistry is framed by 12 Principles which guide chemists in the design of materials and processes.

The Twelve Principles of Green Chemistry

  1. Prevention
    It is better to prevent waste than to treat or clean up waste after it has been created.
  2. Atom Economy
    Synthetic methods should be designed to maximize the incorporation of all materials used in the process into the final product.
  3. Less Hazardous Chemical Syntheses
    Wherever practicable, synthetic methods should be designed to use and generate substances that possess little or no toxicity to human health and the environment.
  4. Designing Safer Chemicals
    Chemical products should be designed to effect their desired function while minimizing their toxicity.
  5. Safer Solvents and Auxiliaries
    The use of auxiliary substances (e.g., solvents, separation agents, etc.) should be made unnecessary wherever possible and innocuous when used.
  6. Design for Energy Efficiency
    Energy requirements of chemical processes should be recognized for their environmental and economic impacts and should be minimized. If possible, synthetic methods should be conducted at ambient temperature and pressure.
  7. Use of Renewable Feedstocks
    A raw material or feedstock should be renewable rather than depleting whenever technically and economically practicable.
  8. Reduce Derivatives
    Unnecessary derivatization (use of blocking groups, protection/ deprotection, temporary modification of physical/chemical processes) should be minimized or avoided if possible, because such steps require additional reagents and can generate waste.
  9. Catalysis
    Catalytic reagents (as selective as possible) are superior to stoichiometric reagents.
  10. Design for Degradation
    Chemical products should be designed so that at the end of their function they break down into innocuous degradation products and do not persist in the environment.
  11. Real-time analysis for Pollution Prevention
    Analytical methodologies need to be further developed to allow for real-time, in-process monitoring and control prior to the formation of hazardous substances.
  12. Inherently Safer Chemistry for Accident Prevention
    Substances and the form of a substance used in a chemical process should be chosen to minimize the potential for chemical accidents, including releases, explosions, and fires.

Visit Beyond Benign to learn more about: “What is Green Chemistry?”

What is Endocrine Disruption

The endocrine system uses chemical signals—hormones—to direct development and reproduction, regulate body function and metabolism, and influence behavior and immunity [2]. In its broadest sense, endocrine disruption takes place when an agent alters hormone signaling or the response to hormone signaling, and in so doing alters some aspect of the organism under hormonal control. According to the Endocrine Society, the world’s authoritative scientific association of clinical and research endocrinologists, an endocrine-disrupting chemical (EDC) is an exogenous chemical, or mixture of chemicals, that can interfere with any aspect of hormone action [3].

Endocrine disruption can be caused by diverse mechanisms. Hormones work by binding with protein receptors in the cell membrane, the cytoplasm or the nucleus. Binding initiates gene activity or physiological processes (depending upon the receptor, its location, hormone concentration, and the developmental state of the cell/tissue/organism) that are part of and essential to normal organismal function. EDCs work by interfering with that signaling process. They are not necessarily structurally similar to hormones; many, but not all, are lipophilic.

Mechanisms of action include: the EDC binds to the receptor and adds to the normal signal; the EDC binds to the receptor and blocks the normal signal; the EDC affects hormone synthesis (increasing or decreasing the amount of natural hormone that is available for signaling); the EDC alters hormone metabolism or hormone transport and storage within bodily tissue (again, increasing or decreasing hormone amount); and/or the EDC affects the levels of mature hormone receptor via disruption or modulation of gene expression, folding, or transport.

A central part of the phenomenon of endocrine disruption is receptor binding, which depends upon the molecular conformation of the hormone and its receptors. Molecular structure is a good, but imperfect predictor of whether binding will occur; chemists can use information about structure both to predict potential hazard (described below) as well as to guide manipulation of a chemical’s structure to avoid hazard.

A crucial aspect of hormone action is that it takes place at extremely low concentrations.  For an estrogen, for example, typical physiological levels of the biologically-active form of an estrogen are extremely low, in the range of 10-900 pg/ml (high parts per quadrillion to low parts per trillion). This is possible because of the specificity of hormone binding to its receptor, and is biologically necessary because of the large number of signaling molecules present at any one time. Specificity and extreme sensitivity make it possible for an enormous number of signaling molecules to co-exist in circulation [4] without disrupting each other’s signaling. The specificity also evolved, presumably, to reduce or avoid disruption by exogenous compounds with which organisms have had evolutionary experience.

Within the past century, over 80,000 new chemicals have been synthesized and used in ways that have resulted in widespread human exposures.  A subset of these chemicals are toxic; a subset of these toxic chemicals are toxic due to endocrine disruption.  A small number of these chemicals have been created explicitly to alter hormone signaling, e.g., the estrogenic drug diethylstilbestrol and many pesticides (for target species). Other chemicals have molecular structures that unintentionally bear sufficient resemblance to hormones such that they are capable of binding, with varying degrees of affinity, to hormone receptors, or of interacting at the molecular level with other molecules involved in hormonal activity. Often EDCs are much less potent than the endogenous hormones in binding with receptors. An increasing number of examples appearing in the peer-reviewed literature, however, show that in some signaling pathways exogenous hormone-mimics can be equipotent and capable of provoking biological responses at picomolar (pM) levels or lower.[5]

Most early research on EDCs focused on the effects of disruption of sexual reproduction via interactions with the estrogen and androgen nuclear receptors. Evidence gathered over the past decade now shows that the mechanisms and endpoints vulnerable to endocrine disruption are much broader than originally understood. Indeed, EDCs are now known to affect metabolism, diabetes, obesity, liver function, bone function, immune function, learning and behavior via a panoply of receptor systems and signaling pathways. In addition, the actions of EDCs on reproduction are now known to go far beyond nuclear sex steroid hormone receptors. In principal, there is virtually no endocrine signaling system or hormone pathway immune to disruption.

The majority of research on EDCs has examined the consequences of their interactions with nuclear hormone receptors (NRs), especially estrogen receptors alpha and beta (ERa, ERb), the androgen receptor (AR), among others. NRs are a superfamily of transcription factors, proteins that can bind to DNA and influence the expression of nearby genes. NRs play central roles in development, physiology and disease. In humans, there are some 48 identified NRs. Many others remain “orphans,” meaning that their endogenous ligands have not yet been identified. When activated, NRs undergo conformational changes that allow recruitment of co-regulatory molecules and the chromatin-modifying machinery of the cell. The ultimate action of NRs is to influence the transcriptional machinery of target genes. NRs also interact with other intracellular signaling pathways. Examining how chemicals bind to these receptors can provide important information concerning their endocrine disrupting potential. There are in vitro assays, some of which can be performed as part of high throughput, screening systems that can confirm chemical binding to the majority of NRs. The strengths and weaknesses of in vitro tools in predicting hazard is discussed in the section on Tier 2.

Endocrine disruption also takes place outside the cell nucleus. Many natural steroid hormones bind to cell membrane-bound receptors, which in turn partner with a variety of well-known signaling cascade proteins. Recent evidence demonstrate that EDCs may exert hormonal effects via these non-nuclear hormone receptors as well. Rather than acting as transcription factors, membrane hormone receptors act via intracellular signaling molecules to affect phosphorylation and calcium flux within a cell. Disruption of this pathway is another way by which EDCs may alter endogenous hormone actions.

Thus, EDCs can act via multiple pathways and receptor-based mechanisms. At higher doses they may also exert receptor-independent actions via more traditional mechanisms of toxicity. Their effects are species, tissue- and cell-specific, and are influenced by metabolism.